Sie sind auf Seite 1von 74

Chapter 6 Solutions

Engineering and Chemical Thermodynamics

Wyatt Tenhaeff
Milo Koretsky

Department of Chemical Engineering


Oregon State University

koretsm@engr.orst.edu
6.1
(a)
The Clausius-Clapeyron equation:

vap
dPisat !hi dT
" P sat %
i )h vap * 1 1 -
or ln$ '=( i , (
sat = 2 # 101 [kPa] & R + T 373 [K]/.
Pi RT

so
sat * #hivap $ 1 1 '-
Pi = (101 [kPa]) exp+" " .
, R &% T 373 [K])(/
(b) and (c)
Using

⎡ kJ ⎤
Δh vap = 40.626 ⎢ ⎥
⎣ mol ⎦

we obtain the following table

T [K] Eqn 6.24 Steam %


[kPa] Tables [kPa] Difference
273.156 0.84 0.6113 37.30%
278.15 1.16 0.87 33.02%
283.15 1.58 1.23 28.30%
288.15 2.13 1.71 24.51%
293.15 2.84 2.34 21.51%
298.15 3.76 3.17 18.62%
303.15 4.93 4.25 15.94%
308.15 6.40 5.63 13.68%
313.15 8.24 7.38 11.71%
318.15 10.54 9.59 9.86%
323.15 13.36 12.35 8.19%
328.15 16.82 15.76 6.75%
333.15 21.04 19.94 5.50%
338.15 26.13 25.03 4.40%
343.15 32.26 31.19 3.42%
348.15 39.58 38.58 2.58%
353.15 48.28 47.39 1.87%
358.15 58.56 57.84 1.25%
363.15 70.67 70.14 0.75%
368.15 84.84 84.55 0.34%
373.15 101.35 101.35 0.00%

2
The logarithmic trend is well-represented. However, at lower temperatures the Clausius-
⎡ kJ ⎤
Clapeyron equation is up to 37% off. The actual heat of vaporization changes from 2501.3 ⎢ ⎥
⎣ kg ⎦
⎡ kJ ⎤
at 0.01 oC to 2257.0 ⎢ ⎥ at 100 oC, a difference of around 10%.
⎣ kg ⎦
100.00

10.00
Pressure [kPa]

Eqn 6.23
Tables

1.00

0.10
273 283 293 303 313 323 333 343 353 363 373
Temperature [K]

(d)
For 100 ºC to 200 ºC, we obtain the following table:

T [K] Eqn 6.24 Steam Tables %


[kPa] [kPa] Difference
373.15 101.35 101.35 0.00%
378.15 120.51 120.82 0.26%
383.15 142.64 143.28 0.44%
388.15 168.11 169.06 0.56%
393.15 197.30 198.53 0.62%
398.15 230.63 232.1 0.63%
403.15 268.55 270.1 0.57%
408.15 311.54 313 0.47%
413.15 360.11 361.3 0.33%
418.15 414.82 415.5 0.16%
423.15 476.24 475.9 0.07%
428.15 545.00 543.1 0.35%
433.15 621.74 617.8 0.64%
438.15 707.16 700.5 0.95%
443.15 801.99 791.7 1.30%
448.15 906.98 892 1.68%
453.15 1022.93 1002.2 2.07%
458.15 1150.68 1122.7 2.49%
463.15 1291.10 1254.4 2.93%

3
468.15 1445.10 1397.8 3.38%
473.15 1613.62 1553.8 3.85%

10000.00
Pressure [kPa]

Eqn 6.23
1000.00
Tables

100.00
373 383 393 403 413 423 433 443 453 463 473
Temperature [K]

Over this range the Clausius-Clapeyron equation represents the data well and is no more than 4
% off. The actual heat of vaporization changes from 2257.0 [kJ/kg] at 100 oC to 1940.7 [kJ/kg]
at 200 oC, a difference of around 15%.

(e)
The heat of vaporization can be corrected for temperature as follows

Tb T
l
Δhvap (T ) = ∫ cP dT + Δhvap (Tb ) + ∫ c Pv dT
T Tb

We can acquire heat capacity data from Appendix A.2, but to simplify the analysis, we will use
an average heat capacity for the vapor.

Δhvap (T ) = 75.4(373.15 − T ) + 40626 + 34.13(T − 373.15)


Δhvap (T ) = 56026 − 41.27T

Substitute this expression into the Clausius-Clapeyron equation

dPisat (56026 − 41.27T ) dT


=
Pisat RT 2

Integrate:

4
⎡ 1 ⎡ ⎛ 1 1 ⎞ ⎛ T ⎞⎤ ⎤
Pisat = (101.35 kPa )exp⎢ ⎢− 56026⎜ − ⎟ − 41.27 ln⎜ ⎟⎥ ⎥
⎣ R ⎣ ⎝ T 373.15 ⎠ ⎝ 373.15 ⎠⎦ ⎦

Now plot the data as before from 0.01 ºC to 200 ºC.

T [K] Eqn 6.24 Steam Tables %


[kPa] [kPa] Difference
273.16 0.64 0.6113 4.95
278.15 0.91 0.87 4.96
283.15 1.28 1.23 4.24
288.15 1.78 1.71 3.88
293.15 2.43 2.34 3.86
298.15 3.29 3.17 3.65
303.15 4.39 4.25 3.35
308.15 5.81 5.63 3.17
313.15 7.60 7.38 3.03
318.15 9.86 9.59 2.78
323.15 12.66 12.35 2.51
328.15 16.12 15.76 2.28
333.15 20.35 19.94 2.06
338.15 25.49 25.03 1.84
343.15 31.68 31.19 1.58
348.15 39.10 38.58 1.34
353.15 47.91 47.39 1.09
358.15 58.32 57.84 0.82
363.15 70.54 70.14 0.57
368.15 84.80 84.55 0.29
373.15 101.35 101.35 0.00
378.15 120.46 120.82 0.30
383.15 142.40 143.28 0.61
388.15 167.48 169.06 0.94
393.15 196.00 198.53 1.28
398.15 228.29 232.1 1.64
403.15 264.70 270.1 2.00
408.15 305.56 313 2.38
413.15 351.26 361.3 2.78
418.15 402.16 415.5 3.21
423.15 458.64 475.9 3.63
428.15 521.10 543.1 4.05
433.15 589.92 617.8 4.51
438.15 665.51 700.5 4.99
443.15 748.27 791.7 5.49
448.15 838.59 892 5.99
453.15 936.89 1002.2 6.52
458.15 1043.55 1122.7 7.05
463.15 1158.98 1254.4 7.61

5
468.15 1283.56 1397.8 8.17
473.15 1417.67 1553.8 8.76

Water Saturation Pressure as a Function of Temperature

Equation Steam Tables


10000.00

1000.00

100.00
Psat (kPa)

10.00

1.00
250 300 350 400 450 500

0.10
T (K)

The agreement between the two values at lower temperatures improves significantly at lower
temperatures, but actually worsens at higher temperatures. The agreement could potentially be
improved by not averaging the heat capacity.

6
6.2
We can find the required pressure by applying the Clapeyron equation:

dP hl − h s
=
dT (
vl − v s T )
We can find the molar volume of water ice from any number of reference books. At 0 ºC and 1
bar:

l ⎡ kg ⎤ l −5
⎡ m 3 ⎤
ρ = 1000 ⎢ ⎥ ∴ v = 1.80 × 10 ⎢ ⎥
⎣ m 3 ⎦ ⎣⎢ mol ⎦⎥
⎡ kg ⎤ ⎡ m 3 ⎤
ρ s = 917 ⎢ ⎥ ∴ v s = 1.97 × 10 − 5 ⎢ ⎥
⎣ m 3 ⎦ ⎢⎣ mol ⎥⎦

Also at 0 ºC and 1 bar,

⎡ J ⎤
h l − h s = 6010 ⎢ ⎥
⎣ mol ⎦

( ) ( )
If we assume that h l − h s and v l − v s are independent of temperature and pressure, we can
separate variables in the Clapeyron equation and integrate.

h l − h s ⎛ T2 ⎞
P2 = P1 + ln⎜⎜ ⎟⎟
( )
v l − v s ⎝ T1 ⎠
6010 [J/mol ] ⎛ 268.15 K ⎞
P2 = 1 × 10 5 [Pa ] + ln⎜⎜ ⎟⎟
(1.80 ×10 −5
− 1.97 × 10 −5
[ 3
])
m / mol ⎝ 273.15 K ⎠
so
P2 = 66.1 [bar]

7
6.3

(a)
At 1 bar, the gas will act as an ideal gas.

⎛ ⎡ J ⎤ ⎞
⎜⎜ 8.314 ⎢ ⎟⎟(300 K )
RT ⎝ ⎣ mol ⋅ K ⎥⎦ ⎠ ⎡ m 3 ⎤
v= = = 0.0249 ⎢ ⎥
P 1 × 10 5 [Pa ] ⎣ mol ⎦

The number of moles of vapor are found as follows (neglect molar volume of liquid)

vV
n = =
0.001 m 3[ ]
v ⎡ m 3 ⎤
0.0249 ⎢ ⎥
⎣⎢ mol ⎦⎥
n v = 0.0402 mol

(b)
At 21 bar, the gas will not behave ideally. Since we are assuming that the molar volume of
liquid is negligible and the heat of vaporization is independent of temperature, the Clapeyron
equation becomes

dP Δh vap
= v
dT vT
The molar volume using pressure expansion of the virial equation is

vv =
(
RT 1 + B ' P ⎛ 1 ) ⎞
= RT ⎜ + B ' ⎟
P ⎝ P ⎠

Substituting this expression into the Clapeyron equation yields

dP Δh vap
=
dT ⎛ 1 ⎞
R⎜ + B ' ⎟T 2
⎝ P ⎠

Separation of variables yields

P2 = 21 ×105 Par T2
⎛ 1 ' ⎞ Δh vap dT
∫ ⎜ + B ⎟dP = ∫ 2
P1 =1 ×105 Pa ⎝ P ⎠ R T1 =300 K T

and integration results in

8
⎛ P ⎞ − Δh vap ⎛ 1 1 ⎞
ln⎜⎜ 2 ⎟⎟ + B ' (P2 − P1 ) = ⎜⎜ − ⎟⎟
⎝ P1 ⎠ R ⎝ T2 T1 ⎠

We can substitute values for given quantities and constants to solve for T2.

T2 = 523.3 K

(c)
Using the virial equation,

3
⎛ 1 ⎞ ⎛ ⎡ J ⎤ ⎞ ⎛ 1 −7 ⎡ m ⎤
⎞
v v = RT ⎜ + B ' ⎟ = ⎜⎜ 8.314 ⎢ ⎥ ⎟
⎟(523.3 K )⎜
⎜ 21 × 10 Pa
5
+ −1 × 10 ⎢ ⎥ ⎟
⎟
⎝ P ⎠ ⎝ ⎣ mol ⋅ K ⎦ ⎠ ⎝ ⎣ J ⎦ ⎠
3
⎡ m ⎤
v v = 0.00164 ⎢ ⎥
⎣ mol ⎦

We can assume the volume occupied by the liquid is negligible. Therefore,

nv =
V
=
0.001 m 3[ ]
vv ⎡ m 3 ⎤
0.00164 ⎢ ⎥
⎣ mol ⎦
n v = 0.61 [mol]

9
6.4
We can use the following computational path to solve for pressure at which graphite and
diamond are in equilibrium at 25 oC.

graphite diamond
P = 1 [atm]

$ J &
"g = 2866
% mol'

P 1
"g1 = # v graphdP "g 3 = # v diam dP
1 P

"g 2 = 0 diamond
graphite

Summing together the three steps we get:

⎡ J ⎤
Δg (1[atm]) = 2866 ⎢ ⎥ = Δg1 + Δg 2 + Δg 3
⎣ mol ⎦

To find the change in Gibbs energy with pressure, we apply the fundamental property relation,
Equation 5.9. At constant temperature:

If the solid is assumed incompressible, we can integrate to get

Δgi = ∫ vi dP = vi ΔP
!
Thus the sum of Gibbs energy becomes
P 1
⎡ J ⎤
Δg (1[atm]) = 2866 ⎢ (
⎥ = ∫ v graph dP + 0 + ∫ vdiam dP = v graph − vdiam (P − 1) )
⎣ mol ⎦ 1 P

Solving
⎡ J ⎤ ⎞ ⎧⎪⎛ 1 1 ⎞ ⎡ cm 3 ⎤ ⎫⎪⎛ ⎡ g ⎤ ⎞⎛⎜ 1 ⎡ m 3 ⎤ ⎞⎟
⎛
⎜⎜ 2866 ⎢
⎝
⎥ ⎟
⎟ = ⎨⎜ − ⎟ ⎢ ⎥ ⎬⎜⎜12 ⎢ ⎥ ⎟⎟
6
(
⎣ mol ⎦ ⎠ ⎪⎩⎝ 2.26 3.51 ⎠ ⎣⎢ g ⎦⎥ ⎪⎭⎝ ⎣ mol ⎦ ⎠⎜⎝ 10 ⎣⎢ cm ⎦⎥ ⎟⎠
5
)
⎢ 3 ⎥ P − 1.01 × 10 [Pa ]

or

P = 1,514 [MPa] = 15,143 [bar]

10
6.5
From the Clausius-Clapeyron equation:

dP h lAl − h Al
s ΔhTfus
= = (I)
dT (
v lAl − v Al
s
T ) (
v lAl − v Al
s
T )
where "h Tfus is the enthalpy of fusion at temperature T. We can get the molar volumes from the
densities:

⎡ kg ⎤
0.027 ⎢
MW ⎣ mol ⎥⎦ = 1.17 × 10 − 5 ⎡ m 3 ⎤
v lAl = = ⎢ ⎥
ρl ⎡ kg ⎤
2,300 ⎢ ⎢⎣ mol ⎥⎦
⎥
⎣ m 3 ⎦
and

⎡ kg ⎤
0.027 ⎢
s MW ⎣ mol ⎥⎦ = 1.00 × 10 − 5 ⎡ m 3 ⎤
v Al = = ⎢ ⎥
ρs ⎡ kg ⎤
2,700 ⎢ ⎣⎢ mol ⎦⎥
⎥
⎣ m 3 ⎦

so

⎡ m 3 ⎤
v lAl − v Al
s
= 1.7 × 10 −6 ⎢ ⎥
⎢⎣ mol ⎥⎦

We can use the following path to calculate for "h Tfus .

solid liquid
T

#h Tfus

933 T
" cPs dT " cPl dT
T 933

T = 933 [K]

solid $ J & liquid


#h fus = 10,711
% mol'

11
933.45 T
⎡ J ⎤
ΔhTfus = ∫
s
c P dT + 10,711 + l
cP
∫ dT ⎢ ⎥
T 933.45
⎣ mol ⎦

⎡ J ⎤ ⎡ J ⎤
Using c Ps = 20.608 + 0.0138T ⎢ ⎥
l
and c P = 31.748 ⎢ ⎥ , we get:
⎣ mol K ⎦ ⎣ mol K ⎦

ΔhTfus = 5,819.9 + 11.68T − 0.0069T 2

Back into Equation (I) gives:

⎡ 5,819.9 + 11.68T − 0.0069T 2 ⎤


dP = ⎢ ⎥ dT
⎣⎢ 1 .7 ×(10 −6
T ) ⎦⎥

Integrating:

100 [bar] T ⎡ 5,819.9 + 11.68T − 0.0069T 2 ⎤


∫ dP = ∫ ⎢⎢ 1 .(
7 × 10 −6
T ) ⎥ dT
1 933.45 [K] ⎣ ⎦⎥

or

([100 − 1]×105 )(1.7 ×10 − 6 ) = 5819.9 ln⎛⎜⎝ 933T.45 ⎞⎟⎠ + 11.68(T − 993.45) − 0.00345(T 2 − 993.45 2 )
solving for T gives

T = 934.91 [K]

12
6.6
We can assume that silver acts as an ideal gas at 1500 K. We can also assume the molar volume
of the vapor is much greater than the molar volume of liquid. Therefore, we can use Equation
6.22

dP sat Δh vap dT
=
P sat RT 2

This can be rearranged to show

dP sat Δh vap P sat


=
dT RT 2

This relation assumes


1. vv>>vl
2. Silver acts as an ideal gas

We can differentiate the expression for pressure in the problem statement to obtain

dP sat ⎛ 14260 0.458 ⎞ ⎛ − 14260 ⎞


= ⎜ 2 − ⎟ exp⎜ − 0.458 ln(T ) + 12.23 ⎟
dT ⎝ T T ⎠ ⎝ T ⎠
sat
dP ⎛ 14260 0.458 ⎞ sat
= ⎜ 2
− ⎟ P
dT ⎝ T T ⎠

Therefore,

vap
⎛ 14260 0.458 ⎞ Δh
⎜ 2
− ⎟ =
⎝ T T ⎠ RT 2

Therefore,

Δh vap = R(14260 − 0.458T )

At 1500 K,

⎛ ⎡ J ⎤ ⎞
Δh vap = ⎜⎜ 8.314 ⎢ ⎥ ⎟⎟[14260 − 0.458(1500 [K ])]
⎝ ⎣ mol ⋅ K ⎦ ⎠
⎡ kJ ⎤
Δh vap = 112.8 ⎢ ⎥
⎣ mol ⎦

13
6.7
For a single component system:

µ = Gi = g i = g

From the fundamental property relation given by Equation 5.9:

dg = −sdT + vdP

We can identify a phase transition from the vertical line of the g vs. T plot, as indicated below.
Since this transition is vertical, i.e., the temperature is constant, the pressure must also be
constant. Thus, we can differentiate the Gibbs energy with respect to temperature at constant
pressure to get:

⎛ ∂g ⎞
⎜ ⎟ = −s
⎝ ∂T ⎠ P

Hence the slope of a plot of g (or µ) vs. T at any temperature must be the negative of the value of
entropy on the plot for s vs. T. The resulting curve is sketched below.

14
6.8
The ferrite phase has stronger bonds. At room temperature, iron is in the ferrite phase. The
heating to 912 ºC has the effect of increasing the entropy contribution to the Gibbs energy. At a
high enough temperature, the austenite phase becomes stable, so that its entropy must be greater
than the ferrite phase. If the entropy of the austenite phase is greater, the enthalpy of the ferrite
phase must be greater or else the austenite phase would be stable over the entire temperature
range. Hence, the ferrite phase has stronger bonds.

15
6.9
Since the pressures are low, we can assume ideal gas behavior. We can also assume that the
molar volume of the vapor is much greater than the molar volume of liquid and the heat of
vaporization is independent of temperature. Therefore, we can rearrange Equation 6.24 to obtain

⎛ ⎡ J ⎤ ⎞ ⎛ 760 torr ⎞
− ⎜⎜ 8.314 ⎢ ⎥ ⎟⎟ ln⎜ ⎟
⎝ ⎣ mol ⋅ K ⎦ ⎠ ⎝ 400 torr ⎠
Δh vap =
⎛ 1 1 ⎞
⎜ − ⎟
⎝ 353.25 K 333.75 K ⎠
⎡ kJ ⎤
Δh vap = 32.3 ⎢ ⎥
⎣ mol ⎦

This value is 7.7% smaller than the reported value.

16
6.10
(a)
The freezing point occurs where there is a discontinuity in the g vs. T plot, as indicated below.
The liquid is at a temperature higher than the freezing point and the solid at lower temperature.
These are demarked below. The melting temperature is 250 K, which occurs at a value g = 3,000
[J/mol]

(b)
At constant pressure, the entropy can be found from Equation 5.14. For the solid we have:

⎛ ∂g ⎞ Δg 1,000 ⎡ J ⎤
s = −⎜ ⎟ = − = = 10 ⎢ ⎥
⎝ ∂T ⎠ P ΔT 10 ⎣ mol K ⎦

And for the liquid, we get:

⎛ ∂g ⎞ Δg 2,000 ⎡ J ⎤
s = −⎜ ⎟ = − = = 40 ⎢ ⎥
⎝ ∂T ⎠ P ΔT 50 ⎣ mol K ⎦

(c)
As we change pressure, we can see how the Gibbs energy changes at any given temperature by
Equation 5.14:

17
⎛ ∂g ⎞
⎜ ⎟ = v
⎝ ∂P ⎠T

Assuming the molar volumes of the liquid and vapor stay constant over the temperature range
around the melting point, we see that the Gibbs energy of the liquid increases by 1.2 times the
Gibbs energy of the solid, since the molar volume of the liquid is 20% larger. The Gibbs energy
of the new freezing point at higher pressure is schematically drawn on the plot above. For
convenience, we choose the solid to increase by 1 unit on the plot. Thus, the liquid increases by
1.2 units. As the sketch shows, the freezing point, where the two lines intersect, will shift to
higher temperature.

18
6.11
For a single component system the fundamental property relation, Equation 5.9, gives:

dg = −sdT + vdP

We can identify a phase transition from the vertical line of the g vs. P plot, as indicated below.
Since this transition is vertical, i.e., the pressure is constant, the temperature must also be
constant. Thus, we can differentiate the Gibbs energy with respect to pressure at constant
temperature to get:

⎛ ∂g ⎞
⎜ ⎟ = v
⎝ ∂P ⎠T

Hence the slope of a plot of g vs. P must have a slope that matches the plot for v vs. T. Since the
molar volume of phase α is about twice the value of phase β, its slope should be twice as big.
The resulting curve is sketched below.

19
6.12
The saturation pressure can be found using the Clausius-Clapeyron equation with the assumption
that the heat of vaporization is independent of temperature. First, we need to use the given data
for the 63.5 ºC and 78.4 ºC to find the heat of vaporization.

⎛ P sat ⎞
− R ln⎜ 2 ⎟ ⎛ ⎡ J ⎤ ⎞ ⎛ 760 torr ⎞
⎜ P sat ⎟ − ⎜⎜ 8.314 ⎢ ⎟⎟ ln⎜ ⎟
vap ⎝ 1 ⎠ = ⎝ ⎣ mol ⋅ K ⎥⎦ ⎠ ⎝ 400 torr ⎠
Δh =
⎡ 1 1 ⎤ ⎡ 1 1 ⎤
⎢ − ⎥ ⎢⎣ 351.55 K − 336.65 K ⎥⎦
⎣ T2 T1 ⎦
⎡ kJ ⎤
Δh vap = 42.39 ⎢ ⎥
⎣ mol ⎦

Now we can calculate the vapor pressure at 100 ºC.

⎛ ⎡ kJ ⎤ ⎞
vap ⎜ − 42.39 ⎢ ⎥ ⎡ ⎟
sat sat
⎛ − Δh ⎡ 1 1 ⎤ ⎞ ⎜ ⎣ mol ⎦ 1 1 ⎤ ⎟
P3 = P2 exp⎜ ⎢ − ⎥ ⎟⎟ = (760 torr )exp⎜ ⎢ − ⎥
⎜ R ⎣ T3 T2 ⎦ ⎠ ⎡ kJ ⎤ ⎣ 373.15 K 351.55 K ⎦ ⎟
P3sat = 1760 torr⎝ = 2.32 atm ⎜ 0.008314 ⎢ ⎟
⎝ ⎣ mol ⋅ K ⎥⎦ ⎠
In comparison, ThermoSolver gives a value of 2.23 atm, using the Antoine equation.

20
6.13
We can show using the Chain Rule that

⎡ ⎛ g i ⎞ ⎤ ⎡ ⎛ T∂g i − g i ∂T ⎞ ⎤
⎢ ∂⎜ ⎟ ⎥ ⎢ ⎜⎜ ⎟⎟ ⎥
⎢ ⎝ T ⎠ ⎥ = ⎢ ⎝ T2 ⎠ ⎥ = 1 ⎛ ∂g i ⎞ − g i
⎜ ⎟
⎢ ∂T ⎥ ⎢ ∂T ⎥ T ⎝ ∂T ⎠ P T 2
⎢ ⎥ ⎢ ⎥
⎣ ⎦ P ⎣ ⎦ P

Using fundamental property relations, Equation 5.14 states

⎛ ∂g i ⎞
⎜ ⎟ = − si
⎝ ∂T ⎠ P

Therefore,

⎡ ⎛ g i ⎞ ⎤
⎢ ∂⎜ ⎟ ⎥
⎢ ⎝ T ⎠ ⎥ = − Ts i − g i = − Ts i − (hi − Ts i ) = − hi
⎢ ∂T ⎥ T2 T2 T2
⎢ ⎥
⎣ ⎦ P

21
6.14
Let T1 = 922 K, T2 = 1,300 K

g 2 = h2 − T2 s2
dh = c P dT
1,300 K
h2 − h1 = ∫ c P dT
922 K
⎡ J ⎤
h2 = h1 + c P (T2 − T1 ) = 39,116 ⎢ ⎥
⎣ mol ⎦
c
ds = P dT
T
⎛ T ⎞ ⎡ J ⎤
s2 = s1 + cP ln⎜⎜ 2 ⎟⎟ = 85.10 ⎢
⎝ T1 ⎠ ⎣ mol K ⎥⎦
⎡ J ⎤
g 2 = h2 − T2 s2 = −71,500 ⎢
⎣ mol ⎥⎦

Alternative solution using the result from Problem 6.13:

⎛ d g ⎞
⎜ T ⎟ = − hT = − h1 + c P (T − T1 ) We must leave h as a function of T
⎜ dT
⎝
⎟
⎠ T2 T2

⎛ g ⎞
⎜ ⎟
⎝ T ⎠ 2 1300 K
⎛ g ⎞ ⎡ h + c (T − T )⎤
∫ d ⎜⎝ T ⎟⎠ = − ∫ ⎢⎣ 1 PT 2 1 ⎥⎦ dT
⎛ g ⎞ 922 K
⎜ ⎟
⎝ T ⎠1

⎛ g ⎞ ⎛ g ⎞ ⎡ 1 1 ⎤ ⎛ T ⎞ ⎡ J ⎤
⎜ ⎟ = ⎜ ⎟ + (h1 − c PT1 )⎢ − ⎥ − c P ln⎜⎜ 2 ⎟⎟ = −55.01 ⎢
⎝ T ⎠ 2 ⎝ T ⎠1 ⎣ T2 T1 ⎦ ⎝ T1 ⎠ ⎣ mol K ⎥⎦

⎡ J ⎤
g 2 = −71,500 ⎢
⎣ mol ⎥⎦

22
6.15
A possible hypothetical solution path is presented below:

monoclinic orthorhombic

T = 298 [K]

!g1 !g3

T = 368.3 [K]

monoclinic !g2 = 0 orthorhombic


From the diagram, we see that the Gibbs energy for steps one and three can be calculated as
follows:

368 K
⎛ ∂g m ⎞
Δg1 = ∫ ⎜⎜ ⎟ dT
298 K ⎝
∂T ⎟⎠ P

and
298 K
⎛ ∂g o ⎞
Δg 3 = ∫ ⎜⎜ ∂T ⎟⎟⎠ dT
368 K ⎝ P
respectively. We can apply Equation 5.14 from the thermodynamic web

⎛ ∂g ⎞
⎜ ⎟ = − s
⎝ ∂T ⎠ P

At 368 K, sulfur undergoes a phase transition, so

Δg 3m68→o
K =0

m → o becomes
Using these above relationships, the expression for Δg 298 K

368 K 298 K 368 K 298 K


m →o m o
Δg 298 K =− ∫s dT + 0 − ∫s dT = − ∫ (13.8 + 0.066T )dT − ∫ (11 + 0.071T )dT
298 K 368 K 298 K 368 K
m →o ⎡ J ⎤
Δg 298 K = −79.5 ⎢
⎣ mol ⎥⎦

Therefore, the transition from the monoclinic to orthorhombic state occurs spontaneously. The
orthorhombic state is more stable.

23
6.16
At the phase transition, the following is true

⎛ g ⎞ ⎛ g ⎞
⎜ ⎟ = ⎜ ⎟
⎝ T ⎠ Sr ( s ) ⎝ T ⎠ Sr (l )

Using the thermodynamic web, the following can be shown (see Problem 6.13)

⎛ ∂(g / T ) ⎞ −h
⎜ ⎟ = 2
⎝ ∂T ⎠ P T

The enthalpies can be written as follows

T
h l (T ) = 49179 + ∫ 35.146 dT = 35.146T − 3540
1500 K
T
h s (T ) = 20285 + ∫ 37.656 dT = 37.656T − 16305.4
900 K

g
can also be calculated at 900 K for solid Sr and 1500 K for liquid Sr.
T

⎛ g l ⎞ ⎡ J ⎤
⎜ ⎟ = −83.85 ⎢
⎜ T ⎟ ⎥
⎝ ⎠ ref ⎣ mol ⋅ K ⎦
⎛ g s ⎞ ⎡ J ⎤
⎜ ⎟ = −68.68 ⎢
⎜ T ⎟ ⎥
⎝ ⎠ ref ⎣ mol ⋅ K ⎦

⎛ g ⎞
We can find ⎜ ⎟ at any temperature using the differential equation as follows
⎝ T ⎠
⎛ g ⎞ ⎛ ∂(g / T ) ⎞ h
∫ d ⎜⎝ T ⎟⎠ = ∫ ⎜⎝ ∂T ⎠ P
⎟ dT = − ∫ 2 dT
T

Substituting our expressions, we get

g l /T T
⎛ g ⎞ ⎡ 35.146T − 3540 ⎤
∫ d ⎜⎝ T ⎟⎠ = ∫ − ⎢⎣ T 2 ⎥dT
⎦
− 83.85 1500 K

24
⎛ g l ⎞
⎜ ⎟ = −35.146 ln(T ) − 3540 + 176.04
⎜ T ⎟ T
⎝ ⎠

g s /T T
⎛ g ⎞ ⎡ 37.656T − 16305.4 ⎤
∫ d ⎜⎝ T ⎟⎠ = ∫ − ⎢⎣ T 2 ⎥dT
⎦
− 68.68 900 K
⎛ g s ⎞
∴ ⎜ ⎟ = −37.646 ln(T ) − 16305.4 + 205.5
⎜ T ⎟ T
⎝ ⎠

⎛ g l ⎞ ⎛ g l ⎞
Set ⎜ ⎟ = ⎜ ⎟ and solve for T:
⎜ T ⎟ ⎜ T ⎟
⎝ ⎠ ⎝ ⎠

T melt = 1059.8 K

The enthalpy of melting is defined as

( ) (
Δh fus = h s T melt − h l T melt )
Using the expressions developed above

⎡ kJ ⎤ ⎡ kJ ⎤ ⎡ kJ ⎤
Δh fus = 26.30 ⎢ ⎥ − 33.71 ⎢ ⎥ = −7.41 ⎢ ⎥
⎣ mol ⎦ ⎣ mol ⎦ ⎣ mol ⎦

25
6.17
At the phase transition, the temperature and Gibb’s energy of both phases must be equal.
Mathematically, this is equivalent to

⎛ g ⎞ ⎛ g ⎞
⎜ ⎟ = ⎜ ⎟
⎝ T ⎠ SiO2 ( s ) ⎝ T ⎠ SiO2 (l )

Using the thermodynamic web, the following can be shown (see Problem 6.13)

⎛ ∂(g / T ) ⎞ −h
⎜ ⎟ = 2
⎝ ∂T ⎠ P T

Using the definition of enthalpy, we can write the following

h(T ) T
∫ dh = ∫ c P dT
href Tref
T
∴ h(T ) − href = ∫ c P dT
Tref

The enthalpies can be written as follows

T
l
h (T ) = −738440 + ∫ 85.772 dT
2500 K
T
h s (T ) = −856840 + ∫ [53.466 + 0.02706T − 1.27 × 10 ]
−5 2
T + 2.19 × 10 − 9 T 3 dT
1100 K

g
can also be calculated at 1100 K for solid SiO2 and 2500 K for liquid SiO2.
T

⎛ g l ⎞ ⎡ J ⎤
⎜ ⎟ = −487.3 ⎢
⎜ T ⎟ ⎥
⎝ ⎠ ref ⎣ mol ⋅ K ⎦
⎛ g s ⎞ ⎡ J ⎤
⎜ ⎟ = −903.5 ⎢
⎜ T ⎟ ⎥
⎝ ⎠ ref ⎣ mol ⋅ K ⎦

26
g
We can substitute our expressions for and h(T ) into the above differential equation and
T
separate variables to obtain

⎡ T ⎤
l
g /T
⎢ − 738440 + ∫ 85 . 772 dT ⎥
T ⎢ ⎥
⎛ g ⎞ 2500 K
d
∫ ⎝ T ⎠ ∫ ⎢
⎜ ⎟ = − ⎢ ⎥dT
2
− 487.3 2500 K T ⎥
⎢ ⎥
⎣⎢ ⎦⎥
⎡ T ⎤
s
g /T T
⎢
⎢
− 856840 + ∫[53.466 + 0.02706T − 1.27 × 10 −5 2
T + 2.19 × 10 T]
−9 3
dT ⎥
⎥
⎛ g ⎞ 1100 K
d
∫ ⎝ T ⎠ ∫ ⎢
⎜ ⎟ = − ⎢ ⎥dT
2
− 903.5 1100 K T ⎥
⎢ ⎥
⎣⎢ ⎥⎦

Integration provides

g l − 9.5268 × 10 9 − 85.772 ln(T ) + 1052.4T


= − 487.5
T T
g s − 1.82 × 10 −10 T 4 + 2.12 × 10 −6 T 3 − 0.01353T 2 − 927190.9 − 53.466T ln(T ) + 1230T
= − 903.5
T T

⎛ g l ⎞ ⎛ g s ⎞
If we plot ⎜ ⎟ − ⎜ ⎟ vs. T, we obtain the following:
⎜ T ⎟ ⎜ T ⎟
⎝ ⎠ ⎝ ⎠

27
There are three solutions, but only the solution between 1100 K and 2500 K is physically
meaningful. If we magnify the plot near the middle solution, we find

T = 1983 K
The enthalpy of fusion is defined as

( ) (
Δh fus = h s T melt − h l T melt )
Using the expressions developed above

⎡ kJ ⎤ ⎡ kJ ⎤ ⎡ kJ ⎤
Δh fus = −792.5 ⎢ ⎥ − (− )782.78 ⎢ ⎥ = −9.72 ⎢ ⎥
⎣ mol ⎦ ⎣ mol ⎦ ⎣ mol ⎦

28
6.18
From the Clausius-Clapeyron equation

sat
dPCS Δh vap
2
=
dT (
T vv − vl )
Assuming:

v v >> v l

we get

sat
dPCS Δh vap
2
= (I)
dT v
Tv

The saturation pressure is given by:

sat 4.7063 × 10 3
ln PCS = 62.7839 − − 6.7794 ln T + 8.0194 × 10 − 3 T (II)
2 T
sat
At T = 373 K, PCS = 4.48 "10 5 [Pa] . Taking the derivative of Equation II
2

sat sat
d ln PCS 1 dPCS 4.7063 × 10 3 6.7794
2
= 2
= − + 8.0194 × 10 − 3 (III)
sat
dT PCS dT T2 T
2

Plugging Equation III into Equation I,

⎡ 4.7063 × 10 3 6.7794 ⎤ sat Δh vap


⎢ − + 8.0194 × 10 − 3 ⎥ PCS =
⎢⎣ T2 T ⎥⎦ 2
Tv v

vap ⎡ kJ ⎤
Solving for vv using ΔhCS = 24.050 ⎢ ⎥ gives:
2
⎣ mol ⎦

−1
v Δh vap ⎡ 4.7063 × 10 3 6.7794 ⎤ ⎡ m 3 ⎤
v = ⎢ − + 8.0194 × 10 −3 ⎥ = 6.08 × 10 −3 ⎢ ⎥
sat
TPCS ⎣⎢ T2 T ⎦⎥ ⎣⎢ mol ⎦⎥
2

Pv B
z= = 0.878 = 1 +
RT v

29
or
⎡ m 3 ⎤ ⎡ cm 3 ⎤
B = −7.4 × 10 − 4 ⎢ ⎥ = −740 ⎢ ⎥
⎢⎣ mol ⎥⎦ ⎢⎣ mol ⎥⎦

This value is about 50% higher than the reported value.

Alternative solution:

Following similar development as Problem 6.3:

sat
dPCS "h vap
2
=
dT # &
1
RT % sat + B '(
2

$ PCS 2 '

" %
1 ' sat (h vap
$ sat + B ' dPCS = 2 dT
P
# CS2 &
2 RT

We must be careful about the limits of integration. We need to pick a value of T close so
enthalpy of vaporization is not too different, but far enough away to avoid round off error. If we
sat
choose T = 378 K, Equation I gives PCS = 5.04 "10 5 [Pa] . Integrating:
2

4.48(10 5 [ Pa] " % *h vap


373
1 sat
'
) $ sat + B ' dPCS = ) 2 dT
5 # P
5.04 (10 [ Pa] CS 2 &
2
378 RT

# 4.48 " 105 & ' 5 5 *h vap + 1 1 -


ln% 5(
$ 5.04 " 10 '
+ B (
4.48 "10 ) 5.04 " 10 = ) ) )
R , 373 378 .

Solving for B’ gives:

B' = "2.55 # 10"7 [Pa]

or

$ m3 '
B = B ' RT = "7.9 #10 "4 & )
% mol(

30
6.19
Calculate vA, vB, v, VA, VB, and V from the ideal gas law:

RT
vA = = 0.05 m 3 / mol V A = n A v A = 0.1 m 3
P
RT
vB = = 0.05 m 3 / mol VB = n B v B = 0.15 m 3
P
RT
v = = 0.05 m 3 / mol V = ntot v = 0.25 m 3
P

We calculate the partial molar volumes as follows

⎛ ∂V ⎞
V A = ⎜⎜ ⎟⎟ =
∂ ⎡
⎢ (n A + n B ) RT ⎤⎥ = RT = 0.05 m 3 / mol
⎝ ∂n A ⎠T , P, n B ∂n A ⎣ P ⎦ P
⎛ ∂V
(n A + n B ) RT ⎤⎥ = RT = 0.05 m 3 / mol
⎞ ∂ ⎡
V B = ⎜⎜ ⎟⎟ = ⎢
⎝ ∂n B ⎠T , P, n A ∂n B ⎣ P ⎦ P

To find the remaining quantities, we can apply Equations 6.44 and 6.46

ΔVmix = n A (V A − v A ) + n B (V B − v B )
ΔVmix = 2(0.05 − 0.05) + 3(0.05 − 0.05) = 0

Δv mix = x A (V A − v A ) + x B (V B − v B )
Δvmix = 0

31
6.20

(a)
For a pure species property

va = v( y a = 1)

Substitution yields

⎡ cm 3 ⎤
va = 100(1) + 80(0 ) + 2.5(1)(0) = 100 ⎢ ⎥
⎣⎢ mol ⎦⎥

(b)
From Equation 6.29

⎛ ∂V ⎞
Va = ⎜⎜ ⎟⎟
⎝ ∂na ⎠ n b ,T , P

We can find V by multiplying the given expression for molar volume by the total number of
moles.

⎛ y y ⎞ n n
V = (na + nb )⎜⎜100 ya + 80 yb + 2.5 a b ⎟⎟ = 100na + 80nb + 2.5 a b
⎝ ya yb ⎠ na + nb

Differentiating with respect to na we get,

∂ ⎛ n n ⎞ nb na nb
Va = ⎜⎜100na + 80 nb + 2.5 a b ⎟⎟ = 100 + 2.5 − 2.5
∂na ⎝ na + nb ⎠ n
b
na + nb (na + nb )2
so

Va = 100 + 2.5 yb (1 − ya ) = 100 + 2.5 yb2

To find the molar volume at infinite dilution, we can use the following relation

Va∞ = lim Va
y a →0
⎡ cm 3 ⎤
∴ Va∞ = 102.5 ⎢ ⎥
⎣⎢ mol ⎦⎥

32
(c)
Since species A contributes more to a mixture than to a pure species,

Δvmix > 0

Note: The Gibbs-Duhem equation says that species B also contributes more.

33
6.21
Calculate mole fractions:

n1 1 [mol]
y1 = = = 0.2 y2 = 0.4 y3 = 0.4
ntot 5 [mol]

Calculate v.
Obtain an expression for v:

RT ⎡ 2 ⎡ A B ⎤ ⎤
v=
P ⎢1 + P ⎢ RT ( y1 − y 2 ) + RT ⎥ ⎥
⎣ ⎣ ⎦ ⎦

Substitute values:

⎛ ⎡ 3 ⎤ ⎞
⎜ 82.06 ⎢ cm ⋅ atm ⎥ ⎟(500 [K ])
⎜ ⎢⎣ mol ⋅ K ⎥⎦ ⎟⎠
v= ⎝
50 [atm]
[1 + (50 ) [− 9.0 ×10
2 −5
(0.2 − 0.4) + 3.0 × 10 − 5 ] ]
⎡ cm 3 ⎤
v = 919.0 ⎢ ⎥
⎢⎣ mol ⎥⎦

Calculate V.

V = ntot v
⎛ ⎡ cm 3 ⎤ ⎞
V = (5 [mol])⎜ 919.0 ⎢ ⎥ ⎟
⎜ ⎢⎣ mol ⎥⎦ ⎟⎠
⎝
V = 4595 cm 3 [ ]
Calculate v1.
The value of v1 can be found by substituting y1=1 into the expression for v1.

⎛ ⎡ 3 ⎤ ⎞
⎜ 82.06 ⎢ cm ⋅ atm ⎥ ⎟(500 [K ])
⎜ ⎢⎣ mol ⋅ K ⎥⎦ ⎟⎠
v1 = ⎝
50 [atm]
[1 + (50 ) [− 9.0 ×10
2 −5
]
(1 − 0) + 3.0 × 10 − 5 ]
⎡ cm 3 ⎤
v1 = 698 ⎢ ⎥
⎢⎣ mol ⎥⎦

34
Calculate v2.

⎛ ⎡ 3 ⎤ ⎞
⎜ 82.06 ⎢ cm ⋅ atm ⎥ ⎟(500 [K ])
⎜ ⎢⎣ mol ⋅ K ⎥⎦ ⎟⎠
v2 = ⎝
50 [atm]
[1 + (50 ) [− 9.0 ×10
2 −5
(0 − 1) + 3.0 × 10 − 5 ] ]
⎡ cm 3 ⎤
v2 = 1067 ⎢ ⎥
⎣ mol ⎦

Calculate v3.

⎛ ⎡ 3 ⎤ ⎞
⎜ 82.06 ⎢ cm ⋅ atm ⎥ ⎟(500 [K ])
⎜ ⎢⎣ mol ⋅ K ⎥⎦ ⎟⎠
v3 = ⎝
50 [atm]
[1 + (50 ) [3.0 ×10 ]
2 −5

⎡ cm 3 ⎤
v3 = 882 ⎢ ⎥
⎢⎣ mol ⎥⎦

Calculate V1 .
From Equation 6.29:

⎛ ∂V ⎞ ⎛ ∂ (nv ) ⎞
V1 = ⎜⎜ ⎟⎟ = ⎜⎜ ⎟⎟
⎝ ∂n1 ⎠ n 2 , n 3 ,T , P ⎝ ∂n1 ⎠ n 2 , n 3 ,T , P

We can substitute the expression for V into this derivative and use the fact that
ntot = n1 + n2 + n3 to obtain

∂ ⎛ RT ⎡ 2 ⎡ A B ⎤ ⎤ ⎞⎟
V1 = ⎜ ⎢ (n1 + n 2 + n3 ) + P ⎢⎣ RT (n1 − n 2 ) + (n1 + n 2 + n3 )⎥⎦ ⎥ ⎟
∂n1 ⎜⎝ P ⎣ RT ⎦ ⎠ n 2 , n 3 ,T , P

Differentiating we get

RT ⎡ 2 ⎡ A B ⎤ ⎤
V1 = ⎢1 + P ⎢ RT + RT ⎥ ⎥
P ⎣ ⎣ ⎦ ⎦

Substitute values:

35
⎛ ⎡ 3 ⎤ ⎞
⎜ 82.06 ⎢ cm ⋅ atm ⎥ ⎟(500 [K ])
⎜ ⎢⎣ mol ⋅ K ⎥⎦ ⎟⎠
V1 = ⎝
50 [atm]
[1 + (50 ) [− 9 ×10
2 −5
+ 3 × 10 − 5 ]
⎡ cm 3 ⎤
V1 = 697.5 ⎢ ⎥
⎢⎣ mol ⎥⎦

36
6.22

(a)
By definition:

⎛ ∂H ⎞
H a ≡ ⎜⎜ ⎟⎟
⎝ ∂na ⎠T , P, nb , nc

h = −5,000xa − 3,000xb − 2,200xc − 500xa xb xc [J/mol]

n = na + nb + nc

na nbnc
H = nh = "5,000na " 3,000nb " 2,200nc " 500
( na + nb + nc )2

# "H & * nbn c 2na nbnc -


% ( = )5,000 ) 500, )
2 /
$ "na 'T ,P,n ,n
b c + ( na + n b + nc ) (na + nb + n c )3 .

# "H &
Ha = % ( = )5,000 ) 500xb x c (1) 2xa ) [ J/mol]
$ "na 'T ,P,n ,n
b c

(b)
1
xa = xb = xc =
3

Ha = "5,018.5 [ J/mol]

(c)
xa = 1 , xb = xc = 0

Ha = "5,000 [ J/mol]

(d)
xb = 1 , xa = xc = 0

Hb = hb = "3,000 [ J/mol]

37
6.23
Let the subscript “1” designate CO2, and “2” designate propane. To calculate the partial molar
volumes, the following formulas will be used:

dv
V1 = v − y 2
dy 2
v = y1V1 + y 2V2

Expressions can’t be obtained for the molar volume with the van der Waals EOS; therefore, the
problem will be solved graphically. First, obtain an expression for the pressure that contains the
mole fractions of CO2 and propane:

amix = y12 a1 + 2 y1 y2 a1a2 + y22 a2


bmix = y1b1 + y 2b2

RT y12 a1 + 2 y1 y 2 a1a2 + y 22 a2
P= −
v − ( y1b1 + y 2b2 ) v2

Solve for a and b using data from the appendices.

⎡ J ⋅ m 3 ⎤ ⎡ mol ⎤
a1 = 0.366 ⎢ ⎥ b1 = 4.29 × 10 − 5 ⎢ ⎥
⎢⎣ mol 2 ⎥⎦ ⎣ m 3 ⎦
⎡ J ⋅ m 3 ⎤ ⎡ mol ⎤
a 2 = 0.941 ⎢ ⎥ b2 = 9.06 × 10 − 5 ⎢ ⎥
2
⎣⎢ mol ⎦⎥ ⎣ m 3 ⎦

Now we can create a spreadsheet with the following headings:

y1 y2 amix bmix v

The last column contains the molar volumes obtained by solving the van der Waals equation
with the spreadsheet’s solver function. After the table is completed, we create the following
graph.

38
v vs. y2

1.48E-03

1.46E-03

1.44E-03

1.42E-03
v (m 3 mol -1)

1.40E-03

1.38E-03

1.36E-03
y = -0.00008x 2 - 0.00008x + 0.00147
1.34E-03
R2 = 0.99946
1.32E-03

1.30E-03
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
y2 (Mole Fraction)

From the line of best fit, we find

v = −8 × 10 − 5 y 22 − 8 × 10 − 5 y 2 + 0.00147

Therefore,

dv
= −1.6 × 10 − 4 y 2 − 8 × 10 − 5
dy 2

and

(
V1 = v − y 2 − 1.6 × 10 −4 y 2 − 8 × 10 −5 )
We can find the partial molar volume of propane from the following relationship

v = y1V1 + y 2V2
v − y1V1
V2 =
y2

Tabulate the values of the partial molar volumes in the spreadsheet and create the following
graph

39
Partial Molar Volumes as a Function of Carbon Dioxide Mole
Fraction
0.0016
V1
V2
0.00155
mol -1)

0.0015
3
Partial Molar Volume (m

0.00145

0.0014

0.00135

0.0013

0.00125
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
y1 (CO2 mole fraction)

40
6.24

(a)
kJ
ga = 40
mol

dg
Ga = g ! x b
dx b
dg
= 40 ! 60 + RT (! ln xa ! 1 + ln x b + 1) + 5x a ! 5xb
dx b
Ga = !40xa ! 60x b + RT ( xa ln x a + x b ln xb ) + 5xa x b
! x b[ 40 ! 60 + RT (! ln xa ! 1 + ln x b + 1) + 5x a ! 5xb ]

Ga = !40( xa + xb ) + RT ( xa + xb ) ln x a + 5xb2

Ga = !40 + RT ln x a + 5xb2
8.314(300) kJ kJ
Ga = !40 + ln 0.2 + 5(0.64) = !40.8
1000 mol mol

Ga! = "!

!Gmix = n(g " x a ga " xb gb )


g ! ( xa ga + x bgb ) = RT ( x a ln xa + xb ln x b ) + 5x a xb

!Gmix = nT [ RT ( x a ln xa + xb ln x b ) + 5xa x b ] = "2.2 kJ

(b)

!gmix = !hmix + T!smix

Assume the entropy of mixing is ideal:

!hmix = 5x a xb > 0

so
Δh = Δhmix + Δhsensibleheat = 0

so
!hsensibleheat < 0 and T goes down

41
6.25
To find V1 and V2 , we can read values directly from the graphs. Calculate mole fractions

1
x1 = = 0.2
5

At x1 = 0.2 ,

⎡ cm 3 ⎤
V1 = 46.5 ⎢ ⎥
⎣ mol ⎦
⎡ cm 3 ⎤
V2 = 69.8 ⎢ ⎥
⎣ mol ⎦

The following relationships are employed to calculate the molar volumes of pure species

v1 = lim V1
x1 →1
v2 = lim V2 = lim V2
x 2 →1 x1 → 0

From the graph

⎡ cm 3 ⎤
v1 = 50 ⎢ ⎥
⎣ mol ⎦
⎡ cm 3 ⎤
v2 = 70 ⎢ ⎥
⎣ mol ⎦

Therefore,

[ ]
V1 = 50 cm 3
V2 = 280 [cm ] 3

To calculate the total volume, we can use

V = n1V1 + n2V2

Substituting the values, we find

[ ]
V = 1(46.5) + 4(69.8) = 325.7 cm 3

42
Therefore

V 325.7
v= =
ntot 1+ 4
⎡ cm 3 ⎤
v = 65.14 ⎢ ⎥
⎣ mol ⎦

Using Equation 6.43 we can calculate the change in volume.

ΔVmix = V − (V1 + V2 ) = 325.7 − (50 + 280)

ΔVmix = −4.3 cm3[ ]

43
6.26

(a)
Expression for Δhmix :

Δhmix = ∑ X i Hi −∑ X i hi =(X Cd H Cd + X Sn H Sn ) − (X Cd hCd + X Sn hSn )

Multiply both sides by the total number of moles

ΔH mix = (nCd H Cd + n Sn H Sn ) − (nCd hCd + n Sn hSn )

Therefore,

⎛ ∂ΔH mix ⎞
(ΔH mix )Cd = ⎜⎜ ⎟⎟ = H Cd − hCd
⎝ ∂nCd ⎠ n Sn ,T , P

(b)
We can show by repeating Part (a) for Sn that

(ΔH mix )Sn = H Sn − hSn


Equation 6.65:

dΔhmix
(ΔH mix )Cd = Δhmix − X Sn
dX Sn

(ΔH mix )Sn = Δhmix − X Cd dΔhmix


dX Cd

Since,
Δhmix = 13000 X Cd X Sn

We get,

dΔhmix d
= [13000 X Cd X Sn ] = 13000( X Cd − X Sn )
dX Sn dX Sn

dΔhmix d
= [13000 X Cd X Sn ] = 13000( X Sn − X Cd )
dX Cd dX Cd

Therefore, for 3 moles of cadmium and 2 moles of tin at 500 ºC:

44
2
HCd − hCd = (ΔH mix )Cd = 13000 X Cd X Sn − X Sn13000( X Cd − X Sn ) = 13000 X Sn

⎡ J ⎤
H Cd − hCd = (ΔH mix )Cd = 2080 ⎢
⎣ mol ⎥⎦

and
2 ⎡ J ⎤
H Sn − hSn = (ΔH mix )Sn = 13000 X Cd = 4680 ⎢
⎣ mol ⎥⎦
(c)
Equation 6.37:

nCd d (ΔH mix )Cd + nSn d (ΔH mix )Sn = 0

Differentiate with respect to XCd:

d (ΔH mix )Cd d (ΔH mix )Sn


nCd + n Sn =0
dx Cd dxCd

where

d (ΔH mix )Cd


= 26000 X Sn = −26000 + 26000 X Cd
dX Cd
d (ΔH mix )Sn
= 26000 X Cd
dX Sn

Therefore,

d (ΔH mix )Cd d (ΔH mix )Sn


nCd + nSn = xCd (ntot )(− 26000 + 26000 xCd ) + (1 − xCd )(ntot )26000 xCd
dxCd dxCd

Inspection of the above expression reveals that

d (ΔH mix )Cd d (ΔH mix )Sn


nCd + n Sn =0
dx Cd dxCd

(d)
A graphical solution can be found using the tangent-slope method discussed on pages 285-287:

A plot of a line tangent to the enthalpy of mixing curve at XCd = 0.6, is given below:

45
Heat of mixing in cadmium (Cd)-Tin (Sn) system
6000

5400

4800

4200
" J $
mol %

3600
data
! hmix #

3000

2400

1800

1200

600

0
0 0.2 0.4 0.6 0.8 1
fit to: xCd
" J $
! hmix = 13, 000 xCd x Sn #
mol %

The intercepts give the respective partial molar quantities as follows:

⎡ J ⎤
H Cd − hCd ≅ 2050 ⎢
⎣ mol ⎥⎦
⎡ J ⎤
H Sn − hSn ≅ 4800 ⎢
⎣ mol ⎥⎦

The values using the graphical method are reasonably close to the analytical method.

46
6.27
The following can be shown with the Gibbs-Duhem equation

0 = x1V1 + x2V2

Differentiation with respect to x1:

dV1 dV
0 = x1 + x2 2
dx1 dx1

If the partial molar volume of species 1 is constant, the Gibbs-Duhem equation simplifies to

dV2
0=
dx1

Therefore, the partial molar volume of species 2 is also constant.

Note that in this case, since the partial molar volume of species 1 is constant:

V1 = v1

and similarly for species 2:

` V2 = v2

Hence, the molar volume can be written:

v = x1V1 + x2V2 = x1v1 + x2v2

This is known as Amagat’s law.

47
6.28

(a)
Let species 1 represent HCl and species 2 represent H2O. An expression for the enthalpy of the
solution is

h = x1h1 + x2 h2 + Δhmix

which can be written


~
h = x1h1 + x2h2 + x1Δhs

Therefore,
~
H = n1h1 + n2 h2 + n1Δhs

To use the heat of solution data in Table 6.1, we need to determine the values of n1 and n2
consistent with the convention used in the table. As seen in Example 6.6,

n1 1
x1 = =
(n1 + n2 ) 1 + n
For this problem

x1 = 0.2

Therefore,

n1 = 1
n2 = n = 4

Now we can find expressions for the partial molar enthalpies.

⎛ ∂H ⎞
H H 2 O = H 2 = ⎜⎜ ⎟⎟
∂n
⎝ 2 ⎠ n1 ,T , P
~
dΔhs
H H 2 O = H 2 = h2 + n1
dn 2
~
dΔhs
∴ H H 2 O − hH 2 O = H 2 − h2 = n1
dn 2

Using the data in Table 6.1 for n = 4 ,

48
~ ⎡ J ⎤
Δhs = −61,204 ⎢
⎣ mol solute ⎥⎦

⎡ J ⎤ ⎡ J ⎤
~ ~ ~ − 64049 ⎢ ⎥ − (− )56852 ⎢
dΔhs Δhs (n = 5) − Δhs (n = 3)
≅ = ⎣ mol solute ⎦ ⎣ mol solute ⎥⎦
~ (5 [mol⎡ ] − 3 [mol
ddn
Δh2s J ]) ⎤ (5 [mol] − 3 [mol])
= −3598.5 ⎢
dn2 ⎣ mol ⋅ mol solute ⎥⎦

Therefore,

⎛ ⎡ J ⎤ ⎞ ⎡ J ⎤
H H 2O − hH 2 O = (1 [mol solute ])⎜⎜ − 3598.5 ⎢ ⎥ ⎟⎟ = −3599 ⎢
⎝ ⎣ mol ⋅ mol solute ⎦ ⎠ ⎣ mol ⎥⎦

We can calculate H HCl − hHCl using Equation 6.46

H HCl − hHCl =
(
Δhmix − xH 2 O H H 2 O − hH 2 O ) = xHCl Δh~s − xH O (H H O − hH O )
2 2 2

xHCl xHCl

⎡ ⎛ ⎡ J ⎤ ⎞ ⎛ ⎡ J ⎤ ⎞⎤
⎢0.2⎜⎜ − 61204 ⎢ ⎥ ⎟⎟ − 0.8⎜⎜ − 3599 ⎢ ⎟⎟⎥
⎣ ⎝ ⎣ mol ⎦ ⎠ ⎝ ⎣ mol ⎥⎦ ⎠⎦ ⎡ J ⎤
∴ H HCl − hHCl = = −46808 ⎢
0.2 ⎣ mol ⎥⎦

(b)
For n1 = 2 and n2 = 80 ,

⎛ n + n2 ⎞ 82
n = ⎜⎜ 1 ⎟⎟ − 1 = − 1 = 40
⎝ n1 ⎠ 2

The new values for the number of moles consistent with Table 6.1

n1 = 1 [mol]
n2 = 40 [mol]

Using the data in the table for n = 40 ,


~ ~ ~
dΔhs Δhs (n = 50 ) − Δhs (n = 30 )

dn2 (50 [mol] − 30 [mol])
Interpolating the data in Table 6.1

49
~ ⎡ J ⎤
Δhs (n = 30 ) = −72428 ⎢
⎣ mol solute ⎥⎦

Therefore,

⎡ J ⎤ ⎡ J ⎤
~ − 73729 ⎢ ⎥ − (− )72428 ⎢
dΔhs
≅ ⎣ mol solute ⎦ ⎣ mol solute ⎥⎦ = −65.05 ⎡ J ⎤ and
dn2 (50 [mol H 2O] − 30 [mol H 2O]) ⎢
⎣ mol ⋅ mol solute ⎥⎦

⎛ ⎡ J ⎤ ⎞ ⎡ J ⎤
H H 2O − hH 2O = (1 [mol solute])⎜⎜ − 65.05 ⎢ ⎥ ⎟⎟ = −65.05 ⎢ ⎥
⎝ ⎣ mol ⋅ mol solute ⎦ ⎠ ⎣ mol H 2O ⎦

50
6.29
First perform an energy balance on the mixing process.

Δhmix = q

We can calculate Δhmix using data from Table 6.1. Referring to Equation E6.7A, we find

~
Δhmix = x HCl Δhs

Calculate x HCl :

wHCl 0.30
(MW )HCl 36.46
x HCl = = = 0.175
wHCl wH 2O 0.30 0.70
+ +
(MW )HCl (MW )H 2O 36.46 18.0148
Heats of data are tabulated for a solution containing one mole of the solute for various amounts
of water. Thus, we need to calculate how many moles of water must be added to HCl to obtain
the above mole fraction.

1 [mol HCl]
x HCl = ; where n is the number of moles of H2O
1 [mol HCl] + n
n = 4.71[mol H 2O]

By interpolation of data from Table 6.1, we get

~ ⎡ J ⎤
Δhs = −63224 ⎢ ⎥ (for n = 4.71 )
⎣ mol ⎦

Therefore,

⎛ ⎡ J ⎤ ⎞ ⎡ J ⎤
Δhmix = (0.175)⎜⎜ − 63224 ⎢ ⎥ ⎟⎟ = −11064 ⎢ ⎥
⎝ ⎣ mol ⎦ ⎠ ⎣ mol ⎦

and

⎡ J ⎤
q = Δhmix = −11064 ⎢ ⎥
⎣ mol ⎦

51
6.30
To calculate the enthalpy of mixing from Table 6.1, we must use the following expression
~
Δhmix = x H 2 SO4 Δhs

The mole fraction of sulfuric acid is

1
x H 2 SO4 =
1+ n

where n is the number of moles of water.

Equation 6.47 states

(
Δhmix = −74.4 x H 2 SO4 x H 2 O 1 − 0.561x H 2 SO
4
)
For n = 1 , x H SO = 0.5 and x H O = 0.5
2 4 2

~ ⎡ J ⎤
Table 6.1: Δhs = −31087 ⎢ ⎥
⎣ mol ⎦
⎛ ⎡ J ⎤ ⎞ ⎡ J ⎤
∴ hmix = 0.5⎜⎜ − 31087 ⎢ ⎥ ⎟⎟ = −15543.5 ⎢ ⎥
⎝ ⎣ mol ⎦ ⎠ ⎣ mol ⎦

Equation 6.47: Δhmix = −74.4(0.5)(0.5)(1 − 0.561(0.5))


⎡ J ⎤
Δhmix = −13383 ⎢ ⎥
⎣ mol ⎦

The following table was made

Δhmix [kJ/mol] Δhmix [kJ/mol]


n [mol H2O] x H 2 SO4 % Difference
(Table 6.1) (Eq. 6.47)
1 0.5 -15543.5 -13382.7 14.94
2 0.333333 -14978.7 -13441.6 10.82
3 0.25 -13001.8 -11993.5 8.07
4 0.2 -11414 -10568.4 7.69
5 0.166667 -10174.2 -9367.17 8.26
10 0.090909 -6367.27 -5835.17 8.72
20 0.047619 -3548.43 -3284.01 7.74
50 0.019608 -1497.22 -1414.49 5.68
100 0.009901 -762.238 -725.289 4.97

52
As you can see, the percent difference between the two methods decreases as the mole fraction
of sulfuric acid decreases. Although Equation 6.47 fit data at 21 ºC, while the Table 6.1
tabulates data taken at 25 ºC, we do not expect the temperature dependence to account for all the
observed difference. The table and equation come from different experimental data sets, and
also represent measurement uncertainty. Nevertheless, the agreement is reasonable.

53
6.31
To calculate the enthalpy of mixing from Table 6.1, we must use the following expression
~
Δhmix = x HCl Δhs

The mole fraction of HCl is

1
x HCl =
1+ n

where n is the number of moles of water. The following table was made using these two
equations.

~ ⎡ J ⎤ ⎡ J ⎤
n [mol H2O] x1 Δhs ⎢ ⎥ Δhmix ⎢ ⎥
⎣ mol HCl ⎦ ⎣ mol ⎦
1 0.5 -26225 -13112.5
2 0.333 -48819 -16273
3 0.25 -56852 -14213
4 0.2 -61204 -12240.8
5 0.167 -64049 -10674.8
10 0.091 -69488 -6317.09
20 0.048 -71777 -3417.95
50 0.020 -73729 -1445.67
100 0.0099 -73848 -731.168

54
6.32
A schematic for the process is given below. The inlet streams are labeled “1” and “2” and the
exit stream “3”.
q

50 wt% NaOH 10 wt% NaOH


50 wt% H2O 90 wt% H2O

Stream 1 Stream 3

Stream 2 H2O

The energy balance for this process reduces to

Q = H 3 − H 2 − H1

We first convert from weight percentage to mole fraction. For stream 1,

wNaOH 0.50
(MW )NaOH 40
x NaOH ,1 = = = 0.311
wNaOH wH 2O 0.50 0.50
+ +
(MW )NaOH (MW )H 2O 40 18.0148
and for stream 3,

wNaOH 0.10
x NaOH ,3 =
(MW )NaOH = 40 = 0.048
wNaOH wH 2 O 0.10 0.90
+ +
(MW )NaOH (MW )H 2O 40 18.0148
We now calculate the moles of water per mole of NaOH so that we can use Table 6.1:

1
x NaOH =
1 + nH 2 O

Therefore, for every mole of NaOH

nH 2 O,1 = 2.21

55
n H 2 O,3 = 19.8

Since enthalpy is a state function, we can choose any hypothetical path to calculate the change in
enthalpy. One such path is shown below. The box in our original schematic is depicted with
dashed lines below. We pick a basis of 1 mole NaOH. In step A, the inlet stream is separated
into its pure components. In step B, 17.6 additional moles of water are added to the pure water
stream. Finally the H2O and NaOH streams are remixed

The enthalpy change is found by adding each step

H 3 − H 2 − H1 = ΔH A + ΔH B + ΔH C

Since ΔH B represents the mixing of water with water, ΔH B = 0 .

The enthalpies of mixing for steps A and C can be related to enthalpy of solution data from Table
6.1:

~ ⎡ J ⎤
Δhs ,1 = −23906 ⎢
⎣ mol NaOH ⎥⎦
~ ⎡ J ⎤
Δhs ,3 = −42858 ⎢
⎣ mol NaOH ⎥⎦

Note: The enthalpy of solution for Stream 1 is calculated by extrapolation. Generally,


extrapolation should be avoided, but it is necessary to complete this problem, and we are
not extrapolating very far.

For step A, we need the negative value of the heat of solution of stream 1. Thus for a basis of 1
mole NaOH:

ΔH A = −(− 23906) [J]

56
while for step C:

ΔH C = −42858 [J ]

Now adding the enthalpies of each step per 1 mole of NaOH:

ΔH = 23906 + 0 − 42858 = −18952 [J]

To get the total heat that must be removed per mole of product solution, we divide by the number
of moles of product per mol of NaOH:

ΔH ⎡ J ⎤
q= = −910 ⎢
nNaOH ⎣ mol ⎥⎦

57
6.33
The partial molar property can be written as follows:

⎛ ∂n K ⎞
K1 = ⎜⎜ T ⎟⎟
⎝ ∂n1 ⎠T ,P ,n2 ,n3

Applying the chain rule to the above relationship:

⎛ ∂n ⎞ ⎛ ∂k ⎞
K1 = k ⎜⎜ T ⎟⎟ + nT ⎜⎜ ⎟⎟
⎝ ∂n1 ⎠T , P,n2 ,n3 ⎝ ∂n1 ⎠T ,P,n2 ,n3

⎛ ∂k ⎞
K1 = k + nT ⎜⎜ ⎟⎟ (1)
⎝ ∂n1 ⎠T , P,n2 ,n3

⎛ ∂k ⎞
Now focus on nT ⎜⎜ ⎟⎟ . At constant T and P, we can write,
⎝ ∂n1 ⎠T , P ,n2 ,n3

⎛ ∂k ⎞ ⎛ ∂k ⎞ ⎛ ∂k ⎞
dk = ⎜⎜ ⎟⎟ dx1 + ⎜⎜ ⎟⎟ dx2 + ⎜⎜ ⎟⎟ dx3
⎝ ∂x1 ⎠T , P, x2 , x3 ⎝ ∂x2 ⎠T , P, x1 , x3 ⎝ ∂x3 ⎠T , P, x1 , x2

Therefore,

⎛ ∂k ⎞ ⎛ ∂k ⎞ ⎛ ∂x1 ⎞ ⎛ ∂k ⎞ ⎛ ∂x2 ⎞
⎜⎜ ⎟⎟ = ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟
⎝ ∂n1 ⎠T , P ,n2 ,n3 ⎝ ∂x1 ⎠T , P , x2 , x3 ⎝ ∂n1 ⎠T , P ,n2 ,n3 ⎝ ∂x2 ⎠T , P , x1 , x3 ⎝ ∂n1 ⎠T , P ,n2 ,n3
(2)
⎛ ∂k ⎞ ⎛ ∂x3 ⎞
+ ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟
⎝ ∂x 3 ⎠T , P , x1 , x2 ⎝ ∂n1 ⎠T , P ,n2 ,n3
but,

n1
x1 =
n1 + n2 + n3

so

⎛ ∂x1 ⎞ 1 n1 1
⎜⎜ ⎟⎟ = − 2
= (1 − x1 ) (3)
⎝ ∂n1 ⎠T ,P,n2 ,n3 n1 + n2 + n3 (n1 + n2 + n3 ) nT

Similarly,

58
⎛ ∂x2 ⎞ n2 x
⎜⎜ ⎟⎟ =− 2
=− 2 (4)
⎝ ∂n1 ⎠T , P,n2 ,n3 (n1 + n2 + n3 ) nT
and

⎛ ∂x3 ⎞ n3 x
⎜⎜ ⎟⎟ =− 2
=− 3 (5)
⎝ ∂n1 ⎠T ,P,n2 ,n3 (n1 + n2 + n3 ) nT

Substituting Equations 2, 3, 4, and 5 into the expression 1 for K1 and simplifying:

⎛ ∂k ⎞
(1 − x1 ) + ⎜⎜ ∂k ⎟⎟ (− x2 ) + ⎜⎜ ∂k ⎟⎟
⎛ ⎞ ⎛ ⎞
K1 = k + ⎜⎜ ⎟⎟ (− x3 )
⎝ ∂x1 ⎠T , P, x 2 , x 3 ⎝ ∂x 2 ⎠T , P, x1 , x 3 ⎝ ∂ x 3 ⎠T , P, x1 , x 2

Utilize the fact that x1 + x2 + x3 = 1


⎡⎛ ∂k ⎞ ⎛ ∂k ⎞ ⎤ ⎡⎛ ⎤
⎥ + x3 ⎢⎜ ∂k ⎟
⎞ ⎛ ∂k ⎞
K1 = k + x2 ⎢⎜⎜ ⎟⎟ − ⎜⎜ ⎟⎟ − ⎜ ⎟ ⎥ (6)
⎢⎝ ∂x1 ⎠T , P, x , x ⎝ ∂x2 ⎠T , P, x , x ⎥ ⎢⎜⎝ ∂x1 ⎟⎠T , P, x , x ⎜⎝ ∂x3 ⎟⎠T , P, x , x ⎥
⎣ 2 3 1 3 ⎦ ⎣ 2 3 1 2 ⎦

When we hold species 3 constant:

⎛ ∂k ⎞ ⎛ ∂k ⎞
dk = ⎜⎜ ⎟⎟ dx1 + ⎜⎜ ⎟⎟ dx 2
⎝ ∂x1 ⎠T , P, x2 , x3 ⎝ ∂x 2 ⎠T , P, x1 , x3

⎛ ∂k ⎞ ⎛ ∂k ⎞ ⎛ dx1 ⎞ ⎛ ∂k ⎞ ⎛ dx 2 ⎞
⎜⎜ ⎟⎟ = ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟
⎝ ∂x2 ⎠T , P, x3 ⎝ ∂x1 ⎠T , P, x2 , x3 ⎝ dx 2 ⎠ ⎝ ∂x2 ⎠T ,P, x1 , x3 ⎝ dx 2 ⎠

Thus,

⎛ ∂k ⎞ ⎛ ∂k ⎞ ⎛ ∂k ⎞
⎜⎜ ⎟⎟ = −⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟ (7)
∂x
⎝ 2 ⎠T ,P,x3 ∂x ∂x
⎝ 1 ⎠T ,P,x2 ,x3 ⎝ 2 ⎠T ,P,x1 ,x3

When we hold species 2 constant, a similar analysis shows:

⎛ ∂k ⎞ ⎛ ∂k ⎞ ⎛ ∂k ⎞
⎜⎜ ⎟⎟ = −⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟ (8)
⎝ ∂x3 ⎠T , P, x2 ⎝ ∂x1 ⎠T ,P , x2 , x3 ⎝ ∂x3 ⎠T , P, x1 , x2

Substituting Equations 7 and 8 into Equation 6 gives

59
⎛ ∂k ⎞ ⎛ ∂k ⎞
K1 = k − x2 ⎜⎜ ⎟⎟ − x3 ⎜⎜ ⎟⎟
⎝ ∂x2 ⎠T , P, x3 ⎝ ∂x3 ⎠T , P, x2

Furthermore, the above analysis can be extended to m components. In general,

⎛ ∂k ⎞
K i = k − ∑ xm ⎜⎜ ⎟⎟
m ≠i ∂x
⎝ m ⎠T , P , x j≠i ,m

60
6.34
The expression can be found by employing the Gibbs-Duhem equation:

0 = n1V1 + n2V2

Differentiate with respect to x1 and then divide by the total number of moles:

dV1 dV dV dV
0 = x1 + x2 2 = x1 1 + (1 − x1 ) 2
dx1 dx1 dx1 dx1

Differentiate the expression given in the problem statement.

dV1
= −5.28 + 5.28 x1
dx1

Substitution of this result into the Gibbs-Duhem equation and rearrangement yields

dV2 5.28 x12 − 5.28 x1


= = 5.28 x1
dx1 x1 − 1

Integrate:

⎡ cm 3 ⎤
V2 = 2.64 x12 + C ⎢ ⎥
⎢⎣ mol ⎥⎦

To determine C, we can use the density information given in the problem statement.

V2 (x2 = 1) = V2 (x1 = 0) = C = v2

where

1 1 ⎡ cm 3 ⎤
C = v2 = = = 109.58 ⎢ ⎥
⎛ ρ 2 ⎞
⎜⎜ ⎟⎟ ⎜ [
⎛ 0.768 g/cm3 ⎞]
⎟ ⎢
⎣ mol ⎥⎦
⎝ MW2 ⎠ ⎜ 84.16 [g/mol] ⎟
⎝ ⎠

Therefore,

⎡ cm 3 ⎤
V2 = 2.64 x12 + 109.58 ⎢ ⎥
⎢⎣ mol ⎥⎦

61
6.35
An expression for the enthalpy of the solution is

h = x1h1 + x2 h2 + Δhmix

which is equivalent to
~
h = x1h1 + x2 h2 + x1Δhs

Multiplication by the total number of moles yields


~
H = n1h1 + n2 h2 + n1Δhs

To use the heat of solution data in Table 6.1, we need to determine the values of n1 and n2
consistent with the convention used in the table. As seen in Example 6.6,

n1 1
x1 = =
(n1 + n2 ) 1 + n
For this problem

x1 = 0.33

Therefore,

n1 = 1
n2 = n = 2

Now we can find expressions for the partial molar enthalpies.

⎛ ∂H ⎞
H H 2 O = H 2 = ⎜⎜ ⎟⎟
⎝ ∂n2 ⎠ n1 ,T , P
~
dΔhs
H H 2 O = H 2 = h2 + n1
dn 2
~
dΔhs
∴ H H 2 O − hH 2 O = H 2 − h2 = n1
dn 2

Using the data in Table 6.1 for n = 2 ,

62
⎡ J ⎤ ⎡ J ⎤
~ ~ ~ − 2,787 ⎢ ⎥ − (− )812 ⎢
dΔhs Δhs (n = 3) − Δhs (n = 1)
≅ = ⎣ mol solute ⎦ ⎣ mol solute ⎥⎦
~ (3 [mol
ddn
Δh2s ⎡ ] − 1 [mol
J ]) ⎤ (3 [mol H 2O] − 1 [mol H 2O])
= −987.5 ⎢
dn2 ⎣ mol ⋅ mol solute ⎥⎦

Therefore,

⎛ ⎡ J ⎤ ⎞ ⎡ J ⎤
H H 2 O − hH 2 O = (1 [mol solute ])⎜⎜ − 987.5 ⎢ ⎥ ⎟⎟ = −987.5 ⎢
⎝ ⎣ mol ⋅ mol solute ⎦ ⎠ ⎣ mol ⎥⎦
Calculate the partial molar enthalpy:

⎡ J ⎤
H H 2O = hH 2O − 987.5 ⎢
⎣ mol ⎥⎦

From the saturated steam tables at 25 ºC:

⎡ kJ ⎤
hˆH 2 O = 104.87 ⎢ ⎥
⎣ kg ⎦
⎡ kJ ⎤
∴ hH 2 O = 1.89 ⎢
⎣ mol ⎥⎦

Now we can find the partial molar enthalpy

⎡ kJ ⎤ ⎡ J ⎤ ⎡ J ⎤
H H 2 O = 1.89 ⎢ ⎥ − 0.988 ⎢ ⎥ = 0.90 ⎢ ⎥
⎣ mol H 2 O ⎦ ⎣ mol H 2 O ⎦ ⎣ mol H 2 O ⎦

63
6.36

(a)
Calculate the mole fraction of sulfuric acid

w1 0.20
x1 =
(MW )1 = 98.078 = 0.044
w1 w2 0.20 0.80
+ +
(MW )1 (MW )2 98.078 18.0148

Calculate n to use in Table 6.1:

1
x1 = = 0.044
1+ n
n = 21.7 [mol H 2O]

Interpolating from Table 6.1

~ ⎡ J ⎤
Δhs = −74621 ⎢ ⎥
⎣ mol ⎦

Now calculate the heat transfer

~ ⎛ ⎡ J ⎤ ⎞
q = Δhmix = x1Δhs = (0.044)⎜⎜ − 74621 ⎢ ⎥ ⎟⎟
⎝ ⎣ mol ⎦ ⎠
⎡ J ⎤
q = −3283 ⎢ ⎥
⎣ mol ⎦

(b)
Calculate the mole fraction of pure sulfuric acid. Consider a mixture of 20 kg of 18 M sulfuric
acid and 80 kg of water. Find the mass of sulfuric acid present.

mH 2 SO4 20 kg
VH 2 SO4 = = = 10.9 L
ρ H 2 SO4 1.84 [kg/L ]
nH 2 SO4 = VM = (10.9 L)(18 mol/L) = 196.2 mol
( )
mH 2 SO4 = nH 2 SO4 MWH 2 SO4 = (196.2 mol)(0.098 kg/mol) = 19.2 kg

Since both the initial (i) and final (f) states contain mixtures, to get the enthalpy of mixing, we
need to calculate the relative differences follows:

64
~ ~
q = Δhmix = x1, f Δhs, f − x1,i Δhs,i

calculate the mole fraction in the final state

w1 0.192
(MW )1 98.078
x1, f = = = 0.042
w1 w2 0.192 0.808
+ +
(MW )1 (MW )2 98.078 18.0148
Calculate n to use in Table 6.1:

1
x1, f = = 0.042
1+ n
n = 22.8 [mol H 2O]

Interpolating from Table 6.1

~ ⎡ J ⎤
Δhs, f = −74689 ⎢
⎣ mol ⎥⎦

For the initial 18 M sulfuric acid:

w1 0.192
(MW )1 98.078
x1,i = = = 0.81
w1 w2 0.192 0.008
+ +
(MW )1 (MW )2 98.078 18.0148
Calculate n to use in Table 6.1:

1
x1,i = = 0.81
1+ n
n = 0.23 [mol H2O]

We must extrapolate from Table 6.1. To do this we wish to extend the trend at low water
concentration. A plot of the data in Table 6.1 is useful. A semi-log plot follows:

65
0
3200 [J/mol]
-10000

enthalp of solution [J/mol solute]


-20000

-30000

-40000

-50000

-60000

-70000

-80000

-90000
0.1 1 10 100
n

~ ⎡ J ⎤
Δhs,i = −3200 ⎢
⎣ mol ⎥⎦

Now calculate the heat transfer

~ ~ ⎛ ⎡ J ⎤ ⎞ ⎛ ⎡ J ⎤ ⎞
q = Δhmix = x1, f Δhs, f − x1,i Δhs,i = (0.042)⎜⎜ − 74689 ⎢ ⎥ ⎟⎟ − (0.81)⎜⎜ − 3200 ⎢ ⎟⎟
⎝ ⎣ mol ⎦ ⎠ ⎝ ⎣ mol ⎥⎦ ⎠
⎡ J ⎤
q = −710 ⎢
⎣ mol ⎥⎦

(c)
Calculate the mole fraction of sodium hydroxide

w1 0.20
x1 =
(MW )1 = 40 = 0.101
w1 w2 0.20 0.80
+ +
(MW )1 (MW )2 40 18.0148

Calculate the n value to use in Table 6.1:

1
x1 = = 0.101
1+ n
n = 8.9 [mol H 2O]

Interpolating from Table 6.1

~ ⎡ J ⎤
Δhs = −41458 ⎢ ⎥
⎣ mol ⎦

66
Now calculate the heat transfer

~ ⎛ ⎡ J ⎤ ⎞
q = Δhmix = x1Δhs = (0.101)⎜⎜ − 41458 ⎢ ⎥ ⎟⎟
⎝ ⎣ mol ⎦ ⎠
⎡ J ⎤
q = −4187 ⎢ ⎥
⎣ mol ⎦

(d)
Calculate the mole fraction of ammonia

w1 0.20
x1 =
(MW )1 = 17.03 = 0.209
w1 w2 0.20 0.80
+ +
(MW )1 (MW )2 17.03 18.0148

Calculate the n value to use in Table 6.1:

1
x1 = = 0.209
1+ n
n = 3.78 [mol H 2O]

Interpolating from Table 6.1

~ ⎡ J ⎤
Δhs = −33153 ⎢ ⎥
⎣ mol ⎦

Now calculate the heat transfer

~ ⎛ ⎡ J ⎤ ⎞
q = Δhmix = x1Δhs = (0.209)⎜⎜ − 33153 ⎢ ⎥ ⎟⎟
⎝ ⎣ mol ⎦ ⎠
⎡ J ⎤
q = −6929 ⎢ ⎥
⎣ mol ⎦

67
6.37
Let species 1 designate ethanol and species 2 designate water. We need to obtain an expression
for the molar volume, so first, convert the given the mass fractions and densities to mole
fractions and molar volumes.

w1

Mole fractions: x1 =
(MW )1
w1 w2
+
(MW )1 (MW )2
Specific molar volumes: v = vˆ(MW )mixture
where (MW )mixture = x1 (MW )1 + x2 (MW )2

Using this set of equations, the following table was made.

Mole Frac. EtOH Mole Frac. H2O v (ml/mol)


0.000 1.000 18.05
0.042 0.958 19.54
0.089 0.911 21.18
0.144 0.856 23.11
0.207 0.793 25.47
0.281 0.719 28.34
0.370 0.630 31.85
0.477 0.523 36.19
0.610 0.390 41.65
0.779 0.221 48.73
1.000 0.000 58.36

The following graph plots the data. The trendline relates v to x1.

60.00

50.00

40.00
v (ml/mol)

2
v = 4.5491x1 + 35.918x1 + 17.957
30.00
R2 = 1
20.00

10.00

0.00
0.000 0.200 0.400 0.600 0.800 1.000
Mole Fraction EtOH (x1)

68
Now, we can calculate V1 .

⎛ ∂V ⎞ ⎛ ∂ (nv ) ⎞
V1 = ⎜⎜ ⎟⎟ = ⎜⎜ ⎟⎟
⎝ ∂n1 ⎠ n 2 , n 3 ,T , P ⎝ ∂n1 ⎠ n 2 , n 3 ,T , P

We can substitute the trendline for V into this derivative and use the fact that ntot = n1 + n2 to
obtain

∂ ⎛⎜ n12 ⎞
V1 = 4.5491 + 35.918n1 + 17.957(n1 + n2 )⎟
∂n1 ⎜⎝ (n1 + n2 ) ⎟
⎠ n 2 ,T , P

Differentiating we get

⎡ 2n (n + n ) − n 2 ⎤
V1 = 4.5491⎢ 1 1 2 1 + 53.875 ⎡ ml ⎤
⎥ ⎢ mol ⎥
⎢⎣ (n1 + n2 )2 ⎥⎦ ⎣ ⎦
⎡ ml ⎤
( )
V1 = 4.5491 2 x1 − x12 + 53.875 ⎢ ⎥
⎣ mol ⎦

Calculate V2 :

∂ ⎛⎜ n12 ⎞
V2 = 4.5491 + 35.918n1 + 17.957(n1 + n2 )⎟
∂n2 ⎜⎝ (n1 + n2 ) ⎟
⎠ n1 ,T , P

Differentiating we get

⎡ − n 2 ⎤ ⎡ ml ⎤
V2 = 4.5491⎢ 1
⎥ + 17.957 ⎢ ⎥
⎢⎣ (n1 + n2 )2 ⎥⎦ ⎣ mol ⎦
⎡ ml ⎤
V2 = −4.5491x12 + 17.957 ⎢ ⎥
⎣ mol ⎦

Plotting V1 and V2 vs. x1 we obtain

69
Partial Molar Volumes vs. EtOH Mole Fractions

60 20
59 19
EtOH Partial Molar

H2O Partial Molar


Volume (ml/mol)

Volume (ml/mol)
58 18
57 17
EtOH
56 16
H2O
55 15
54 14
53 13
52 12
0 0.2 0.4 0.6 0.8 1
Mole Fraction EtOH (x1)

(b)
Δvmix can be calculated using Equation 6.46

( )
Δvmix = x H 2 O VH 2 O − v H 2 O + x EtOH (VEtOH − v EtOH )

From the data table in Part (a),

⎡ ml ⎤
v EtOH = v x1 =1 = 58.36 ⎢ ⎥
⎣ mol ⎦
⎡ ml ⎤
v H 2 O = v x1 = 0 = 18.05 ⎢ ⎥
⎣ mol ⎦

Using the expressions for partial molar volumes

⎡ ml ⎤
( )
V1 = 4.5491 2(0.5) − (0.5)2 + 53.875 = 57.29 ⎢ ⎥
⎣ mol ⎦
⎡ ml ⎤
V2 = −4.5491(0.5)2 + 17.957 = 16.82 ⎢ ⎥
⎣ mol ⎦

Therefore,

⎛ ⎡ ml ⎤ ⎞ ⎛ ⎡ ml ⎤ ⎞
Δvmix = (0.5)⎜⎜16.82 − 18.05 ⎢ ⎥ ⎟⎟ + 0.5⎜⎜ 57.29 − 58.36 ⎢ ⎥ ⎟⎟
⎝ ⎣ mol ⎦ ⎠ ⎝ ⎣ mol ⎦ ⎠
⎡ ml ⎤
Δvmix = −1.15 ⎢ ⎥
⎣ mol ⎦

70
6.38
We can use the density data given in the problem statement to determine the pure species
properties. For pure ethanol (x1 = 1):

MW1 46 [g/mol] ⎡ cm 3 ⎤
v1 = = = 58.55 ⎢ ⎥
ρ1 0.7857 g/cm 3 [ ]
⎣⎢ mol ⎦⎥
⎛ ⎡ cm 3 ⎤ ⎞
V1 = n1v1 = (3 [mol])⎜ 58.55 ⎢ ⎥ ⎟ = 175.7 cm
3
[ ]
⎣⎢ mol ⎦⎥ ⎠
⎜ ⎟
⎝

For pure formamide (x1 = 0):

MW2 45 [g/mol] ⎡ cm 3 ⎤
v2 = = = 39.77 ⎢ ⎥
ρ2 1.1314 g/cm 3 [ ] ⎢⎣ mol ⎥⎦
⎛ ⎡ cm 3 ⎤ ⎞
V2 = n2 v2 = (1 [mol])⎜ 39.77 ⎢
⎜ ⎢⎣ mol
⎥ ⎟ = 39.77 cm
⎟
⎥⎦ ⎠
[ ] 3

⎝

To calculate V and v, interpolate in the data table to obtain the density of the mixture
when x1 = 0.75:

[
ρ = 0.8550 g/cm 3 ]
Therefore,

MW
v=
ρ
where
MW = 0.75(46 [g/mol]) + 0.25(45 [g/mol]) = 45.75 [g/mol]

Substitute numerical values:

45.75 [g/mol]
v=
[
0.8550 g/cm 3
] [
= 53.51 cm 3 /mol ]
⎛ ⎡ cm 3 ⎤ ⎞
⎜
V = nv = (4 [mol]) 53.51 ⎢ ⎥ ⎟ = 214.0 cm
3
[ ]
⎣⎢ mol ⎦⎥ ⎠
⎜ ⎟
⎝

Now, we can calculate the volume change of mixing by using Equation 6.43:

71
[ ] [ ]
ΔVmix = V − V1 − V2 = 214 cm 3 − 175.7 cm 3 − 39.77 cm 3 [ ]
ΔVmix = −1.47 cm 3 [ ]
The intensive volume change of mixing:

Δvmix =
ΔVmix − 1.47 cm 3
=
[ ]
= −0.368 cm 3 / mol[ ]
n 4

We can use Equation 6.65 to determine the partial molar volume of formamide:

dv Δv
V2 = v − x1 ≈ v − x1
dx1 Δx1

From the provided data table

45.8 45.7

Δv
= 0.8401 0.8701 = 19.50 cm 3 / mol
Δx1 0.8009 − 0.6986
[ ]
Therefore,

[ ] ( [ ])
V2 = 53.51 cm 3 / mol − 0.75 19.5 cm 3 / mol = 38.89 cm 3 / mol [ ]
Now calculate the partial molar volume of ethanol:

V = n1V1 + n2V2

∴V1 =
[ ] ( [
214.0 cm 3 − (1 [mol]) 38.89 cm 3 /mol ])
= 58.37 cm 3 / mol [ ]
3 [mol]

72
6.39
Using the definition of G, the Gibbs energy of mixing of an ideal gas can be rewritten in terms of
the enthalpy of mixing and the entropy of mixing:

ideal gas ideal gas ideal gas


Δg mix = Δhmix − TΔsmix

Since an ideal gas exerts no intermolecular interactions,

ideal gas
Δhmix =0

From Equation 6.48:

ideal gas
Δsmix = − R(xa ln xa + xb ln xb )

so

ideal gas
Δg mix = RT (xa ln xa + xb ln xb )

To find the partial molar Gibbs energy of mixing of species a, we apply Equation 6.29:

(ΔGmix )a = ⎡⎢ ∂ (n∂Δng mix )⎤⎥


⎣ a ⎦ T , P, nb

Applying the expression above

⎛ na nb ⎞
nΔg mix = RT ⎜⎜ na ln + nb ln ⎟⎟ = RT [na ln na + nb ln nb − (na + nb )ln(na + nb )]
⎝ na + nb na + nb ⎠
where the mathematical relation of logarithms was used. Thus,

(ΔGmix )a = ⎡⎢ ∂ (n∂Δng mix )⎤⎥ ⎡ n (n + n )⎤


= RT ⎢ln na + a + 0 − ln(na + nb ) − a b ⎥ = RT ln xa
na (na + nb )⎦
⎣ a ⎦ T , P, nb ⎣
at infinite dilution xa goes to zero, and the ln term blows up,

(ΔGmix )∞a = µa∞ − ga = −∞


As chemical engineers, we are often interested in the limiting case of infinite dilution. We see
that even for ideal gas mixtures the chemical potential in this limit is not mathematically well-
.behaved, In Chapter 7, we will develop a different function, the fugacity, which behaves better.

73
6.40
At equilibrium

l v
µH O
= µH O
2 2

or
G Hl O = G Hv O
2 2

Since the water in the liquid phase is pure

GHl 2O = g Hl 2O = hHl 2O − TsHl 2O

The enthalpy and entropy of liquids are not sensitive to pressure changes. We can use data from
the saturated steam tables at 25 ºC to determine the Gibbs energy.

l
gˆ H = 104.87 [kJ/kg ] − (298.15 K )(0.3673 [kJ/kg ⋅ K ])
2O
l
gˆ H = −4.640 [kJ/kg ]
2O

gHl
2O
( l
= MWH 2 O gˆ H
2O
)( )
= (0.0180148 [kg/mol])(− 4.640 [kJ/kg]) = 0.0836 [kJ/mol]
Therefore,

v l
µH = gH = 83.6 [J/mol]
2O 2O

74

Das könnte Ihnen auch gefallen