You are on page 1of 195

Solutions Manual to Accompany

Molecular Thermodynamics of Fluid-


Phase Equilibria

Third Edition

John M. Prausnitz
Rüdiger N. Lichtenthaler
Edmundo Gomes de Azevedo

Prentice Hall PTR


Upper Saddle River, New Jersey 07458
http://www.phptr.com
” 2000 Prentice Hall PTR
Prentice Hall, Inc.
Upper Saddle River, NJ 07458

Electronic Typesetting: Edmundo Gomes de Azevedo

All rights reserved. No part of this book may be


reproduced, in any form or by any means, without
permission in writing from the publisher.

All product names mentioned herein are the trademarks of their respective owners.

Printed in the United States of America


10 9 8 7 6 5 4 3 2

ISBN 0-13-018388-1

Prentice-Hall International (UK) Limited, London


Prentice-Hall of Australia Pty. Limited, Sydney
Prentice-Hall Canada Inc., Toronto
Prentice-Hall Hispanoamericana, S.A., Mexico
Prentice-Hall of India Private Limited, New Delhi
Prentice-Hall of Japan, Inc., Tokyo
Simon & Schuster Asia Pte. Ltd., Singapore
Editora Prentice-Hall do Brasil, Ltda., Rio de Janeiro
C O N T E N T S

Preface i

Solutions to Problems Chapter 2 1

Solutions to Problems Chapter 3 17

Solutions to Problems Chapter 4 29

Solutions to Problems Chapter 5 49

Solutions to Problems Chapter 6 81

Solutions to Problems Chapter 7 107

Solutions to Problems Chapter 8 121

Solutions to Problems Chapter 9 133

Solutions to Problems Chapter 10 153

Solutions to Problems Chapter 11 165

Solutions to Problems Chapter 12 177


S O L U T I O N S T O P R O B L E M S

C H A P T E R 2

1. From problem statement, we want to find wP / wT b g v


.
Using the product-rule,

FG wP IJ FG wv IJ FG wP IJ
H wT K v

H wT K H w v K
P T

By definition,

FG IJ
1 wv
DP
H K
v wT P

and

FG IJ
1 wv
NT 
H K
v wP T

Then,

FG wP IJ DP . u 10 5
18
H wT K v NT 5.32 u 10 6
33.8 bar D C -1

Integrating the above equation and assuming DPand NT constant over the temperature range,
we obtain

DP
'P 'T
NT

For 'T = 1qC, we get


'P 33.8 bar

1
2 Solutions Manual

2. Given the equation of state,


FG V  bIJ
P
Hn K RT

we find:

FG wS IJ FG wP IJ nR
H wV K T H wT K V  nb V

FG wS IJ F wV I  nR
G J
H wP K T H wT K P P

FG wU IJ F wP I
TG J  P 0
H wV K T H wT K V

FG wU IJ FG wU IJ FG wV IJ 0
H wP K T H wV K H wP K
T T

FG wH IJ T G
F wV IJ  V nb
H wP K T H wT K
For an isothermal change,

'S
z FGH
V1
V2 wP
wT
IJ
K V
dV
V  nb
nR ln 2
V1  nb

P
 nR ln 1
P2

'U
z
LMFG  wU IJ FG wV IJ OP dP 0
P1
P2

MNH wV K H wP K PQ T T

'H
z LMT FG wV IJ  V OP dP nbbP  P g
P2

MN H wT K PQ
P1 P
2 1

'H  T'S nbb P  P g  nRT lnG J


FP I
HP K
1
'G 2 1
2

'A 'U  T'S  nRT lnG J


FP I
HP K
1
2
3 Solutions Manual

3. This entropy calculation corresponds to a series of steps as follows:


s3 (vapor,
T = 298.15 K
P = 1 bar)

s1 (saturated liq. s2 (saturated vapor,


T = 298.15 K T = 298.15 K
P = 0.03168 bar) P = 0.03168 bar)

s3 's1o 2  's2o 3  s1

' vap h (2436) u (18.015)


's1o 2 ' vap s 147.19 J K 1 mol 1
T 298.15

's2o 3
z LMMN FGH
P3

P2

wvIJ
K
wT P
dP
OP
PQ
RT
Because v (ideal gas),
P

§P ·
's2 o 3  R ln ¨ 3 ¸
© P2 ¹
§ 1.0 ·
(8.31451) u ln ¨ ¸
© 0.03168 ¹
28.70 J K 1 mol 1

s3 s0 (H2 O, vapor)

147.19  28.70  69.96

188.45 J K1 mol 1


4 Solutions Manual

RT
4. Because D  v,
P

RT RT
P
Dv 2  3 / v2  v

or

RTv 2
P
2v 2  3  v 3

§ wP · RT v (v 3  6)
¨ wv ¸ 
© ¹T (2v 2  3  v 3 )2

As v 2.3 L mol 1 , T = 373.15 K, R = 0.0831451 bar L K-1 mol-1, and molar mass is 100 g
mol-1,

FG wP IJ
H wv K T
3.3245 bar L-1 mol 3.3245 u 10 8 Pa m3 mol

FG wP IJ
w2  g c kv 2
H wv K T

FG IJ u (1.4) u FG 1 IJ F m3 I u FG 3.3245 u 10
2
IJ
K GH JK H
kg m mol N mol
H K H 100 u 10 u 2.3 u 10 3
K
8
 1
N s2 3 kg mol m2 m3

24,621 m2 s 2

w 157 m s 1

5. Assume a three-step process:


5 Solutions Manual

(1) Isothermal expansion to v = f v=


(ideal gas state) V (2)

(2) Isochoric (v is constant) cool-


ing to T2 (3) (1)

(T2 ,v2 )

(3) Isothermal compression to v2


(T1 ,v1 )

For an isentropic process,

's 's1  's2  's3 0

Because s = s(v, T),

FG ws IJ FG ws IJ
ds
H wv K T
dv 
H wT K v
dT

or

FG wP IJ cv
ds
H wT K v
dv 
T
dT

by using the relations

FG wS IJ FG wP IJ
H wv K H wT K
T v
(Maxwell relation)

FG ws IJ 1 wu FG IJ cv
H wT K v T wT H K v T

then,

v f § wP · T2 c 0 v 2 § wP ·
v v dT 
's
³ ¨ wT ¸ d 
³ ¨ ¸ dv ³
v1 © ¹v T1 T v f © wT ¹v

Using van der Waals’ equation of state,

RT a
P 
v  b v2

FG wP IJ R
H wT K v vb

Thus,
6 Solutions Manual

§v b· T2 c 0
R ln ¨ 2 v dT
's
v
© 1  b
¸
¹

T1 T ³
To simplify, assume

cv0 c 0p  R

RT2
v2
P2

Then,
RT2
b § c0p  R · § T ·
P2 ¨ ¸ ln ¨ 1 ¸
ln
v1  b ¨ R ¸ © T2 ¹
© ¹

ª (82.0578) u (T2 )  45 º 623.15


ln « » (3.029) u ln
¬ 600  45 ¼ T2

T2 203 K

6.
LM O
a P
P
RT

a
v  b v2
RT
v
MM b  RTv PP
1

MN1  v PQ
b2
Because  1,
v2

FG1  b IJ 1
b b2
H vK 1 
v v2
"

Thus,

RT LM FG a 1 b2 IJ OP
MN H K PQ
P 1 b   "
v RT v v 2

or

Pv FG IJ
a 1 b2
RT
1 b 
H 
K
RT v v 2
"
7 Solutions Manual

Because
Pv B C
z 1  "
RT v v2

the second virial coefficient for van der Waals equation is given by

a
B b
RT

7. Starting with
du Tds  Pdv

FG wu IJ FG ws IJ FG wv IJ
H wP K T
T
H wP K T
P
H wP K T

FG wv IJ FG wv IJ
T
H wT K P
P
H wP K T

As

RT RT a
v B b
P P T2

FG wv IJ R 2a
H wT K P

P T3

FG wv IJ RT
H wT K T

P2

Then,

FG wu IJ 2a
H wP K T

T2

S § 2a ·
'u
³0  ¨© W2 ¸¹ dP
2a S
'u 
W2
8 Solutions Manual

8. The equation

FG P  n IJ bv  mg
H v2T 1/2 K RT

can be rewritten as

(PT 1/ 2 )v 3  (PmT 1/ 2 )v 2  nv  nm RT 3/2v 2

FG RT 2IJ n FG nm IJ
H
v3  m 
P K
v 
PT 1/ 2
v
HPT 1/ 2 K 0 (1)

At the critical point, there are three equal roots for v = vc, or, equivalently,

FG wP IJ F w PI2

H wv K T Tc
GH wv JK 2
0
T Tc

bv  v g c
3
v 3  3v c v 2  3v c 2v  v c 3 0 (2)

Comparing Eqs. (1) and (2) at the critical point,

RTc
m 3v c (3)
Pc

n
3v 2c (4)
PcTc1/ 2

nm
v c3 (5)
PcTc1/ 2

From Eqs. (3), (4), and (5) we obtain

vc
m (6)
3

3RTc RTc
vc or m
8Pc 8Pc

27 R 2Tc5/ 2
n 3v 2c PcTc1/ 2
64 Pc

The equation of state may be rewritten:


9 Solutions Manual

F I
P
RT GG 1  n JJ
v
GH 1  mv RT 3/2v JK
or
Pv 1 n
z 
RT
1
m RT 3/2 v
v

From critical data,

m 0.0428 L mol 1

n 63.78 bar (L mol 1 )2 K1/2

At 100qC and at v = (6.948)u(44)/1000 = 0.3057 L mol-1,


z = 0.815
This value of z gives P = 82.7 bar. Tables of Din for carbon dioxide at 100qC and v = 6.948
cm3 g-1, give P = 81.1 bar or z = 0.799.

9. We want to find the molar internal energy u(T , v) based on a reference state chosen so that

u(T0 , v o f) 0

Then,

u(T , v) u(T , v)  u(T0 , v o f)

u(T , v)  u(T , v o f)  u(T , v o f)  u(T0 , v o f ) (1)

lim
vof v fz v
FG wu IJ
H wv K T
dv  lim
z FGH
vof T0
T
wu
wT v v
f
IJ
K
dT

Schematically we have:

v =v
1 2 3
Ideal gas

Ref. state Intermediate state State of interest


Ideal gas Ideal gas Real gas
(T0 , v ) (T, v ) (T, v)
10 Solutions Manual

In Eq. (1) we are taking 1 mol of gas from the reference state 1 to the state of interest 3
through an intermediate state 2, characterized by temperature T and volume v o f, in a two-step
process consisting of an isochoric step and an isothermal step.
In the step 1 o 2 the gas is infinitely rarified, and hence exhibits ideal gas behavior. Then,
the second integral in Eq. (1) gives:
T T T
§ wu ·
lim
³¨ ¸
v of T0 © wv ¹
v vf
dT
³ T0
cv0 dT
³ T0
(c0p  R) dT (c0p  R)(T  T0 ) (2)

because for an ideal gas c 0p  cv0 R and because, by the problem statement, the heat capacity at
constant pressure of the gas is temperature independent.
We have now to calculate the first integral in Eq. (1). To make this calculation, we first
transform the derivative involved in the integral to one expressed in terms of volumetric proper-
ties.
By the fundamental equation for internal energy (see Table 2-1 of the text),

FG wu IJ T FG wP IJ  P
H wv K T H wv K T (3)

Making the derivative using the equation of state give we obtain

FG wu IJ RT RT a a
H wv K T  
v  b v  b v(v  b) v(v  b)
(4)

Then,

v § wu · a v ª 1 1º a ª § vb · v º
³ ¨ ¸ dv ³ 
b v f ¬« v  b v ¼»
dv « ln ¨ ¸  ln »
v f © wv ¹ b ¬ © vf  b ¹ vf ¼
(5)
a § v  b vf ·
ln ¨ ¸
b © v vf  b ¹

and

lim
vof v f z FGH
v
wu
wv
IJ dv
KT
a
b
FG
ln
H
vb
v
IJ
K (6)

Combining Eqs. (1), (2) and (6) we obtain the desired expression for the molar internal en-
ergy,

a vb FG IJ
u(T , v) (c 0p  R)(T  T0 )  ln
b v H K
11 Solutions Manual

10.
ln J w A(1  xw )2 such that J w o 1 as xw o 1

Using Gibbs-Duhem equation,

xw d ln J w  xs d ln J s 0

or, because dx w  dx s ( xw  xs 1) ,

d ln J w d ln J s
xw xs
dx w dx s

d ln J w
2 A(1  x w )(1) 2 A(1  xw )
dx w

Then,

2 Axs (1  xs )
d ln J s dxs 2 A(1  xs ) dxs
xs

ln J s xs
³0 d ln J s 2 A ³0 (1  xs ) dxs

§ x2 ·
ln J s 2 A ¨ x s  s ¸
¨ 2 ¸¹
©

ln J s A( x w2  1)

11. Henry’s law for component 1, at constant temperature, is

f1 k1 x1 (for 0  x1  a)

where k1 is Henry’s constant.


For a liquid phase in equilibrium with its vapor, fi L fiV . If the vapor phase obeys ideal-
gas law, fiV yi P.
Henry’s law can then be written:

y1P k1 x1

Taking logarithms this becomes

ln( y1P) ln k1  ln x1

Differentiation at constant temperature gives


12 Solutions Manual

d ln( y1P) d ln x1 1
dx1 dx1 x1

Using the Gibbs-Duhem equation

d ln P1 d ln P2
x1  x2 0
dx1 dx1

gives

d ln( y2 P)
1  x2 0
dx1

or, because dx 2  dx1 ,

d ln( y2 P)
x2 1
dx2

or,

d ln(y2 P) d ln x2

Integration gives

ln( y2 P) ln x2  ln C

where lnC is the constant of integration.


For x2 1 , y2 1 , and P P2s . This gives C P2s and we may write

ln( y2 P) ln x2  ln P2s ln( x2 P2s )

or

y2 P P2 x 2 P2s [for (1  a)  x 2  1 ]

which is Raoult’s law for component 2.

12. Starting from dgi RTd ln fi ,

f (at P)
'gi RT ln i (P* is a low pressure where
P* o P fi (at P* )
gas i is ideal)

From the Steam Tables we obtain 'h and 's at T and P to calculate 'g from
'g 'h  T 's
13 Solutions Manual

Choose P * = 1 bar.

'h h70 bar  h1 bar 196 J g 1

's s70 bar  s1 bar 2.215 J g-1 D C 1

Then, at 320qC,
'g 1117.8 J g 1 20137 J mol 1

Thus,

20137
b
ln fi 70 bar, 320qC g b g b
8.31451 u 59315
. g 4.08

or
f = 59.1 bar

13. The virial equation for a van der Waals gas can be written (as shown in Problem 6)

RT a
v b (1)
P RT
At the Boyle temperature,

a
B b 0
RT
or
a
b
RT
The Boyle temperature then, is given by

a
TB (2)
bR
The Joule-Thomson coefficient is

FG wT IJ
P
H wP K H
or
14 Solutions Manual

FG wT IJ FG wH IJ
FG wT IJ  H wH K H wP K
H wP K FG wP IJ
P T
P 
cp
H
H wH K T

Because

FG wH IJ FG wv IJ
H wP K T v  T
H wT K P
FG wH IJ
and because cp is never zero, when P 0,
H wP K T 0.

Substitution in Eq. (1) gives

§ wH · RT a § RT a ·
¨ wP ¸  b ¨ 
© ¹T P RT © P RT ¸¹

2a
b 
RT

The inversion temperature is

2a
TJT
Rb
Comparison with Eq. (2) gives

TJT 2TB

14. At equilibrium,

f1G f1L

where subscript 1 stands for the solute.


At constant pressure, a change in temperature may be represented by

F d ln f IG F d ln f I L
GH dT JK 1
dT GH dT JK 1
dT (1)
P P

Since the solvent is nonvolatile, f1G (at constant pressure) depends only on T (gas composi-
tion does not change.) However, f 1L (at constant pressure) depends on T and x1 (or ln x1 ):
15 Solutions Manual

F d ln f1 I
L F w ln f1 IL F w ln f1 I
L
GH dT JK dT GH wT JK dT  GH w ln x1 JK d ln x1 (2)
P P, x T ,P

Further,

F d ln fG I h10  h1G
GH dT 1
JK RT 2
(3)
P

F d ln f L I h10  h1L
GH wT 1
JK RT 2
(4)
P, x

where:
h10 = ideal-gas enthalpy of 1;
h1G = real-gas enthalpy of 1;
h1L = partial molar enthalpy of 1 in the liquid phase.

Assuming Henry’s law,

f1
constant
x1
or

F w ln f1 I L
GH w ln x1 JK 1 (5)
T ,P

Substituting Eqs. (2), (3), (4), and (5) into Eq. (1), we obtain

d ln x1 'h1

d (1 / T ) R

From physical reasoning we expect h1G ! h1L . Therefore x1 falls with rising temperature.
This is true for most cases but not always.
S O L U T I O N S T O P R O B L E M S

C H A P T E R 3

1. The Gibbs energy of a mixture can be related to the partial molar Gibbs energies by

¦ yi e gi  gio j
m
g  go (1)
i 1

Since, at constant temperature, dg RTd ln f , we may integrate to obtain


o
g  go RT ln f mixt  RT ln f mixt

or

g  go RT ln f mixt  RT ln P (2)

where subscript mixt stands for mixture.


For a component in a solution, d gi RTd ln fi . Integration gives

gi  gio RT ln fio

gi  gio RT ln( yi P) (3)

Substituting Eqs. (2) and (3) into Eq. (1) gives


m m
ln f mixt  ln P ¦ yi ln fi  ¦ yi ln( yi P)
i 1 i 1

ln f mixt  ln P
m FfI m
¦ yi lnGH yi JK  ¦ yi ln P (4)
i 1 i i 1

Because

17
Solutions Manual 18

m Fm yI
¦ yi ln P
i 1
ln P GH ¦ iJ
i 1 K
ln P

Eq. (4) becomes

ln f mixt
m FfI
¦ yi lnGH yi JK (5)
i 1 i

Assuming the Lewis fule, fi yi f pure i , Eq. (5) becomes

m
ln f mixt ¦ yi ln fpure i
i 1
or
m
yi
f mixt – fpure i
i=1

2. As shown in Problem 1,

ln f mixt
m FfI
¦ yi lnGH yi JK
i 1 i

This result is rigorous. It does not assume the Lewis fugacity rule.
Using fugacity coefficients,

fi M i yi P

and

ln f mixt y A ln M A  yB ln M B  ln P

y y
fmixt M AA M BB P

(0.65)0.25 u (0.90)0.75 u (50)

fmixt 41.5 bar


Solutions Manual 19

3. Pure-component saturation pressures show that water is relatively nonvolatile at 25qC.


Under these conditions the mole fraction of ethane in the vapor phase (yE) is close to unity.
Henry’s law applies:

fE H (T ) x E

The equilibrium condition is

f EV f EL

or

yE M E P H (T ) x E

At 1 bar, M E | 1 and H (T ) P / xE :

1
H (T ) 3.03 u 10 4 bar
0.33 u 10 4

At 35 bar we must calculate ME:

ln M E z0 P
P
z 1
dP

Using

z 1  7.63 u 10 3 P  7.22 u 10 5 P 2

we obtain

ME 0.733

Because Henry’s constant H is not a strong function of pressure,

fE MEP
xE
H H

(0.733) u (35)
xE x Ethane 8.47 u 10 4
3.03 u 10 4

4. The change in chemical potential can be written,

§ f1 ·
'P1 P1  P10 RT ln ¨ ¸ ( f10 1 bar ) (1)
¨ f10 ¸¹
©
Solutions Manual 20

1
The chemical potential may be defined as :

§ wG · § wG0 ·
P1  P10 ¨ ¸  ¨ ¸
© w n1 ¹T ,P,n2 ¨ w n1 ¸
© ¹T ,n2
(2)
§ wA · § w A0 ·
¨ ¸  ¨ ¸  RT
© w n1 ¹T ,V ,n2 ¨ w n1 ¸
© ¹T ,n2

Combining Eqs. (1) and (2):

1 § w'A ·
ln f1 ¨ ¸  1
RT © w n1 ¹T ,V ,n
2

Using total volume, V nT v , nT n1  n2 ,

'A FG V IJ  n1 lnFG V IJ  n2 lnFG V IJ


RT
nT ln
H V  n b K H n1RT K H n2 RT K
T

Taking the partial derivative and substituting gives

f1 b1
ln
y1 RT vb
vb
or
y1RT b1FG IJ
f1
vb
exp
vb H K
The same expression for the fugacity can be obtained with an alternative (but equivalent)
derivation:

1
A0 U 0  TS 0 ; G0 U 0  PV 0  TS 0 ; PV 0 nT RT

§ wG0 · § wU0 · § w S0 ·
P10 ¨ ¸ ¨ ¸  RT  T ¨ ¸
¨ w n1 ¸ ¨ w n1 ¸ ¨ w n1 ¸
© ¹T ,n j z1 © ¹T ,n j z1 © ¹T ,n j z1
and
§ w A0 · § wU0 · § w S0 ·
¨ ¸ ¨ ¸  RT  T ¨ ¸
¨ w n1 ¸ ¨ w n1 ¸ ¨ w n1 ¸
© ¹T ,n j z1 © ¹T ,n j z1 © ¹T ,n j z1
then
§ w A0 ·
P10 ¨ ¸  RT
¨ w n1 ¸
© ¹T ,n j z1
Solutions Manual 21

§ f ·
P1  P10 RT ln ¨ 1 ¸ ( f10 1 bar)
¨ f0 ¸
© 1 ¹

FG w A IJ
By definition, P1
H wn K
1 T ,V ,n
j z1
. The Helmholtz energy change 'a can be written as

nT 'a A ¦ ni ai0
i

A ¦ ni (P10  RT )
i

Then,

§ wnT 'a · § wA ·
¨ ¸ ¨ ¸  P10  RT
© wn1 ¹n j z1,T ,V wn
© 1 ¹n j z1,T ,V

P1  P10  RT

and

P1  P10 ª w(nT 'a / RT ) º


« »  1
RT ¬ wn1 ¼ n j z1,T ,V

Using the equation for 'a ,

ª w(nT 'a / RT ) º V V n b
« » ln  ln  1 T 1
¬ wn1 ¼ n j z1,T ,V V  nT b n1RT V  nT b

or

§ P  P0 · ª w(nT 'a / RT ) º
ln f1 ln ¨ 1 1 ¸ « » 1
¨ ¸ wn1
© RT ¹ ¬ ¼ n j z1,T ,V

n1 RT n b
ln  T 1
V  nT b V  nT b

n1RT § n b ·
f1 exp ¨ T 1 ¸
V  nT b © V  nT b ¹

Hence,

y1RT § b ·
f1 exp ¨ 1 ¸
vb ©vb¹
Solutions Manual 22

5.
a) Starting with Eq. (3-51):

For a pure component (ni nT ) :

G Gi0
Pi ; Pi0
ni ni

Because
G U  PV  TS

Pi0 Pi0  Tsi0  RT (1)

From Eq. (3-52),



P RT · V PV
Pi
³ ¨ 
n
V © i V
¸ dV  RT ln
¹ ni RT
 Pi0  Tsi0 
ni
(2)

From Eqs. (1) and (2),



P RT · V PV  ni RT
Pi  Pi0
³ ¨ 
V © ni
¸ dV  RT ln
V ¹ ni RT

ni

But,

RT ln fi Pi  Pi0
and
V zi
ni RT P
Substitution gives

§ f· P RT ·
RT ln ¨ ¸
© P ¹i ³ ¨ 
V © ni
¸ dV  RT ln zi  RT ( zi  1)
V ¹

b) Starting with Eq. (3-53):


For a pure component, yi 1. To use Eq. (3-53), we must calculate

FG wP IJ
H wni K T ,V ,pure component i
Pressure P is a function of T, V, and ni and
Solutions Manual 23

§ wP · § wni · § wV ·
¨ ¸ ¨ ¸ ¨ wP ¸ 1
© wni ¹V © wV ¹P © ¹ni

FG wP IJ FG wV IJ FG wP IJ
H wni K V H wni K P H wV K ni
P LM
P V wP FG IJ OP
ni
 
MN
ni ni wV n
i
H K PQ
But,

P V wP FG IJ LM
1 w (PV ) OP

ni ni wV n
i
H K ni NwV n
i Q
Then,
f§ f
wP · P 1 ni RT
³ ¨
w
¸
V © ni ¹T ,V ,n
j z1
³
V ni
dV 
ni PV ³
d ( PV )

f
P PV
³
V ni
dV  RT 
ni

f
P
³ n
V i
dV  RT ( z  1)

Now Eq. (3-54) follows directly.

6. The solubility of water in oil is described by

f1 H (T ) x1

Henry’s constant can be evaluated at 1 bar where f1 1 bar .


Then,

f1 1
H (T ) 286 bar (t 140 D C)
x1 35 u 10 4

To obtain f1 at 410 bar and 140qC, use the Steam Tables (e.g., Keenen and Keyes). Alterna-
tively, get f at saturation (3.615 bar) and use the Poynting factor to correct to 410 bar.
At 140qC,
Solutions Manual 24

RT ln f1 'g1o410 bar 'h1o410 bar  T 's1o410 bar

RT ln f1 282 J g 1 (from Steam Tables)

(282) u (18)
ln f1 1.48
(8.31451) u (413)

Then, f1 = 4.4 bar at 410 bar and 140qC and

f1 4.4
x1 0.0154
H (T ) 286

7.

To = 300 K Tf = 300 K

Po = ? Pf = 1 bar

Applying an energy balance to this process,

'h hf  ho 0

This may be analyzed on a P-T plot. Assume 1 mole of gas passing through the valve.

P
To ,Po

Tf ,Pf I

III
II

A three-step process applies:

I. Isothermal expansion to the ideal-gas state.

II. Isobaric cooling of the ideal gas.

III. Compression to the final pressure.


For this process,
Solutions Manual 25

'h 'hI  'hII  'hIII 0

Since h = h(T,P),

FG wh IJ d P  FG wh IJ dT
dh
H w P K T H wT K P
T f Pf
'h
³To Po
dh

0 ª § wv · º Tf Pf ª § wv · º
³ «v  T ¨
Po ¬«
¸ » dP 
© wT ¹ P ¼»T
o
³To
c0p d T 
³0
«v  T ¨
¬«
¸ » dP
© wT ¹ P ¼»T
f

But,

RT 10 5
v  50 
P T
Then,

FG wv IJ R 10 5
H wT K P 
P T2

and

c 0p yA c 0p,A  yBc 0p,B

F 50  2 u 10 I b P g  y g FGH I
e jb
5 2 u 10 5
'h 0 GH T o
JK o
0
A c p,A  yBc 0p,B Tf  To  50 
Tf JK
Pf

Substitution gives

(616.7) u Po (33.5 J mol 1 K 1 ) u (10 bar cm 3 J 1 ) u (200  300 K)  (950) u (1 bar )

Po = 55.9 bar

8. From the Gibbs-Helmholtz equation:

FG g IJ
w
HTK 
h
wT T2
Solutions Manual 26

or, alternately,
FG 'g IJ
w
HTK
F 1I
wG J
'h h real  h ideal (1)

HTK
Because, at constant temperature,

b g FG f IJ
d 'g RTd ln
H PK (2)

we may substitute Eq. (2) into Eq. (1) to obtain

FG f IJ
w ln
H PK
R
F 1I
wG J
'h

HTK
From the empirical relation given,

f 30.7 0.416 P 2
ln 0.067P  P  0.0012P 2 
P T T

FG f IJ
w ln
H PK 'h
R
F 1I
wG J
30.7P  0.416 P 2
R
HTK
At P = 30 bar,
'h = (8.31451)u[(-30.7)u(30) + (0.416)u(30)2]

 'h = -4545 J mol-1

9. Consider mixing as a three-step process:

(I) Expand isothermally to ideal-gas state.


(II) Mix ideal gas.
(III) Compress mixture isothermally.
Starting with

LM FG wP IJ  POP dv
du cv dT  T
MN H wT K v PQ
Solutions Manual 27

FG wu IJ FG w P IJ
H wv K T T
H wT K v
P

RT a
Because, P  ,
v  b v2

FG wu IJ a
H wv K T v2

Integration of this equation to the ideal-gas state (v f) gives

'u
z v
f

v
a
2
dv
a
v

Therefore,

x1a1 x 2 a2
'uI  5914 J mol 1
v1 v2

[(1 bar)u(1 cm3) | 0.1 Joule]

'uII 0 (because is the mixing of ideal gases)

amixt [ x12 a11  2 x1x2 a11a22 u (1  0.1)  x22 a22 ]


'uIII 
v mixt x1v1  x2 v 2

5550 J mol 1

' mix h ' mix u 'uI  'uII  'uIII 364 J mol 1


S O L U T I O N S T O P R O B L E M S

C H A P T E R 4

1. At 25 Å we can neglect repulsive forces.


The attractive forces are London forces and induced dipolar forces; we neglect (small) quadrupo-
lar forces. (There are no dipole-dipole forces since N2 is nonpolar.)
Let 1 stand for N2 and 2 stand for NH3.

London force:
3 D 1D 2 I1I 2
*12 
2 r 6 I1  I 2

d* 9D 1D 2 I1I 2
F12  
dr r 7 I1  I 2
Since,

D1 17.6 u 10 25 cm3

D2 22.6 u 10 25 cm3

I1 = 15.5 eV = 2.48 u 10 18 N m

. u 10 18 N m
I 2 = 11.5 eV = 184
then
London
F12 62.0 u 10 18 N

Induced dipole force:

29
Solutions Manual 30

D1P 22
*12 
r6

6D1P 22
F12 
r7

1 D 1 u 10 18 (erg cm 3 )1/ 2

P2 1.47 D 1.47 u 10 18 (erg cm 3 )1/ 2

ind
F12 3.8 u 10 18 (erg cm 3 )1/ 2

Neglecting all forces due to quadrupoles (and higher poles),

F tot F London  F ind

F tot 65.8 u 10 18 N

2. From the Lennard-Jones model:

V6
Attractive potential = 4H *
r6

d* V6
Attractive force =  24 H
dr r7

Assume force of form,

d* FG H IJ V 6
Force = 
dr
(constant )
H kK r 7

Using corresponding states:

FG H JI
H kK (constant ) u Tc V6 (constant u v c ) 2
E v c2
D Tc

where D and E are universal constants.


Solutions Manual 31

(V CH 4 ) 6
force CH 4 (constant ) (H / k ) CH 4 (rCH 4 ) 7
force substance B (constant ) (H / k ) B (V B ) 6
(rB ) 7

E (v cCH )2
4

2 u 10 8 D(TcCH ) (1u 10 7
cm)7
4
u
force substance B D (TcB ) E(v cB ) 2
(2 u 10 7 cm)7

Force substance B = 4 u 10 10 dyne

Force = 4 u 10 15 N

3.

*AA 8 u 10 16 erg

By the molecular theory of corresponding states:

*ii § r ·
f¨ ¸ (r 2V)
Hi © Vi ¹

*BB FG H IJ ML f b2g PO HB
HH K MN f b2g PQ
B
( f is a universal function)
*AA A HA

Since H / k 0.77Tc (taking the generalized function f as the Lennard-Jones (12-6) poten-
tial),

TcB
* BB * AA
TcA

(8 u 10 16 erg) u (180 K/ 120 K)

*BB 12.0 u 10 16 erg

12.0 u 10 23 J
Solutions Manual 32

4. For dipole-dipole interaction:

PiP j
*(dd ) f (T i , T j , I)
(4 SH or 3 )

with

f (Ti , T j , I) 2 cos Ti cos T j  sin Ti sin T j cos(Ii  I j )

For the relative orientation:


i j
i
Ti
Ti = 0º Ii = 0º
Tj j
Tj = 180º Ij = 0º

Ij

PiP j 2P i P j
*(dd ) u [2 u 1 u (1)]
(4 SH o r 3) (4 SH or 3 )

For Pi = Pj = 1.08 D = 3.603u10-30 C m and r = 0.5u10-9 m,


30 2
( 2) u (3.603 u 10 )
* ( dd )
12 9 3
(4 S) u (8.8542 u 10 ) u (0.5 u 10 )

-21
-1.87 × 10 J

For the relative orientation:

j Ti = 0º Ii = 0º
i

Tj = 90º Ij = 0º

For the dipole-induced dipole interaction:

P i2D j P 2j D i
*(ddi )  (3 cos 2 T i  1)  (3 cos 2 T j  1)
2(4 SH o ) 2 r 6 2(4 SH o ) 2 r 6

For the relative orientation:


Solutions Manual 33

i j Ti = 0º Tj = 90º

ª (3.603 u 1030 )2 u (2.60 u 1030 ) º


*(ddi ) 2u « u (4) »
10 9 6
¬« 2 u (1.1124 u 10 ) u (0.5 u 10 ) ¼»

7.77 u 1023 J | 8.0 u 1023 J

For the relative orientation:

j Ti = 0º Tj = 90º
i

(3.603 u 1030 )2 u (2.60 u 1030 )


*(ddi )  u (4  1)
2 u (1.1124 u 1010 ) u (0.5 u109 )6

4.85 u 1023 J | 4.8 u 1023 J

5. The energy required to remove the molecule from the solution is

1 FG Hr  1 IJ P 2
E
a3 H 2Hr  1K
Hr 3.5 a 3.0 u 10-8 cm P 2D

(See, for example, C. J. E. Böttcher, 1952, The Theory of Electric Polarization, Elsevier)

E = 4.61u10-21 J/molecule = 2777 J mol-1

6.
a) The critical temperatures and critical volumes of N2 and CO are very similar, more similar
than those for N2 and argon (see Table J-4 of App. J). Therefore, we expect N2/CO mixtures to
follow Amagat’s law more closely than N2/Ar mixtures.
Solutions Manual 34

b) Using a harmonic oscillator model for CO, F = Kx, where F is the force, x is the displace-
ment (vibration) of nuclei from equilibrium position and K is the force constant. This constant
may be measured by relating it to characteristic frequency X through:

X
1 b
K mC  mO g
2S mC m O

where mC and mO are, respectively, the masses of carbon and oxygen atoms.
Infrared spectrum will show strong absorption at X.
Argon has only translational degrees of freedom while CO has, in addition, rotational and
vibrational degrees of freedom. Therefore, the specific heat of CO is larger than that of argon.

7. Electron affinity is the energy released when an electron is added to a neutral atom (or mole-
cule).
Ionization potential is the energy required to remove an electron from a neutral atom (or
molecule).
Lewis acid = electron acceptor (high electron affinity).

Lewis base = electron donor (low ionization potential).


Aromatics are better Lewis bases than paraffin. To extract aromatics from paraffins we want
a good Lewis acid. SO2 is a better a Lewis acid than ammonia.

8. From Debye’s equation:

ª H 1º 4 4 § P2 ·
v« r » S N AD  SNA ¨ ¸
¬ Hr  2 ¼ ¨ ¸


3

3 © 3kT ¹
Total polarization Static polarization
independent of T

Measurement of molar volume, v , and relative permitivity, H r , in a dilute solution as a


function of T, allows P to be determined (plot total polarization versus 1/T; slope gives P).
Solutions Manual 35

9. We compare the attractive part of the LJ potential (r >> V) with the London formula.
The attractive LJ potential is

FG V11 IJ 6
*11 4 H11
HrK
F V I6
4H 22 G 22 J
*22
HrK
F V I6
4H12 G 12 J
*12
HrK
We assume that V12 1/ 2(V11  V22 ) . The London formula is

3D12 I1
*11 
4r 6

3D 22 I 2
*22 
4r 6

FG
3 D 1D 2 IJ FG I1I2 JI
*12 
H
2 r6 K H I1  I 2 K
Substitution gives
6
ª º ª º
1/ 2 « V11V22 » « I1I2 »
H12 (H11H22 )
«1 V V » «1 I I »
¬« 2 11 22 ¼» ¬« 2 1 2 ¼»

Only when V11 V 22 and I1 I 2 do we obtain

H12 (H11H22 )1/ 2

Notice that both correction factors (in brackets) are equal to or less than unity. Thus, in gen-
eral,

H12 d (H11H22 )1/ 2

10. See Pimentel and McClellan, The Hydrogen Bond, Freeman (1960).
Phenol has a higher boiling point and a higher energy of vaporization than other substituted ben-
zenes such as toluene or chlorobenzene. Phenol is more soluble in water than other substituted
benzenes. Distribution experiments show that phenol is strongly associated when dissolved in
Solutions Manual 36

nonpolar solvents like CCl4. Infrared spectra show absorption at a frequency corresponding to
the –OH " H hydrogen bond.

11. J acetone CCl ! J acetone CHCl because acetone can hydrogen-bond with chloroform but not
4 3

with carbon tetrachloride.

12.
a)

CHCl3 Chloroform is the best solvent due to hydrogen bonding which is not
present in pure chloroform or in the polyether (PPD).

Chlorobenzene is the next best solvent due to its high polarizability and
it is a Lewis acid while PPD is a Lewis base.

Cyclohexane is worst due to its low polarizability.

n-butanol is probably a poor solvent for PPD. Although it can hydrogen-


bond with PPD, this requires breaking the H-bonding network between
n-butanol molecules.

t-butanol is probably better. Steric hindrance prevents it from forming H-


bonding networks; therefore, it readily exchanges one H-bond for an-
other when mixed with PPD. The lower boiling point of t-butanol sup-
ports the view that it exhibits weaker hydrogen bonding with itself than
does n-butanol.

b) Cellulose nitrate (nitrocellulose) has two polar groups: ONO2 and OH. For maximum solu-
bility, we want one solvent that can “hook up” with the ONO2 group (e.g., an aromatic hydro-
carbon) and another one for the OH group (e.g., an alcohol or a ketone).

c) Using the result of Problem 5,


Solutions Manual 37

1 FG Hr  1 IJ P 2
E
a3 H 2Hr  1K
At 20qC, the dielectric constants are

Hr (CCl 4 ) 2.238 Hr (C8 H18 ) 1.948

Thus,

H r (CCl 4 )
117
.
H r (C 8 H18 )

It takes more energy to evaporate HCN from CCl4 than from octane.

13. At 170qC and 25 bar:

z H2 is above 1
H2
zamine is well below 1 1

z HCl is slightly below 1 z

HCl

0 1
yamine

a) A mixture of amine and H2 is expected to exhibit positive deviations from Amagat’s law.

b) Since amine and HCl can complex, mixtures will exhibit negative deviations from Amagat’s
law.

c)

1 z argon

z HCl

0 1
yargon
Solutions Manual 38

The strong dipole-dipole attractive forces between HCl molecules cause z HCl  1 , while ar-
gon is nearly ideal. Addition of argon to HCl greatly reduces the attractive forces experienced by
the HCl molecules, and the mixture rapidly approaches ideality with addition of argon. Addition
of HCl to Ar causes induced dipole attractive forces to arise in argon, but these forces are much
smaller than the dipole-dipole forces lost upon addition of Ar to HCl. Thus the curve is convex
upwards.

14.
a) Acetylene has acidic hydrogen atoms while ethane does not. Acetylene can therefore com-
plex with DMF, explaining its higher solubility. No complexing occurs with octane.

b) At the lower pressure (3 bar), the gas-phase is nearly ideal. There are few interactions be-
tween benzene and methane (or hydrogen). Therefore, benzene feels equally “comfortable” in
both gases.
However, at 40 bar there are many more interactions between benzene and methane (or hy-
drogen) in the gas phase. Now benzene does care about the nature of its surroundings. Because
methane has a larger polarizability than hydrogen, benzene feels more “comfortable” with meth-
ane than with hydrogen. Therefore, KB (in methane) > KB (in hydrogen).

c) Under the same conditions, CO2 experiences stronger attractive forces with methane than
with hydrogen due to differences in polarizability. This means that CO2 is more “comfortable” in
methane than in H2 and therefore has a lower fugacity that explains the condensation in H2 but
not in CH4.

d) It is appropriate to look at this from a corresponding-states viewpoint. At 100°C, for ethane


TR | 12. , for helium, TR | 80 .
At lower values of TR (near unity) the molecules have an average thermal (kinetic) energy
on the order of H (because TR is on the order of kT / H ). The colliding molecules (and molecules
near one another, of which there will be many at 50 bar) can therefore be significantly affected
by the attractive portion of the potential, leading to z < 1. At higher TR , the molecules have such
high thermal energies that they are not significantly affected by the attractive part. The mole-
cules look like hard spheres to one another, and only the repulsive part of the potential is impor-
tant. This leads to z > 1.

e) Chlorobenzene would probably be best although cyclohexane might be good too because
both are polar and thus can interact favorably with the polar segment of poly(vinyl chloride).
Ethanol is not good because it hydrogen bonds with itself and n-heptane will be poor because it
is nonpolar.

f) i) Dipole.
ii) Octopole.
iii) Quadrupole.
iv) Octopole.
Solutions Manual 39

g) Lowering the temperature lowers the vapor pressure of heptane and that tends to lower sol-
vent losses due to evaporation. However, at 0qC and at 600 psia, the gas phase is strongly non-
ideal, becoming more nonideal as temperature falls. As the temperature falls, the solubility of
heptane in high-pressure ethane and propane rises due to increased attraction between heptane
and ethane on propane. In this case, the effect of increased gas-phase nonideality is more impor-
tant than the effect of decreased vapor pressure.

15.
a) They are listed in Page 106 of the textbook:

1. Q can be factored so that Qint is independent of density.


2. Classical (rather than quantum) statistical mechanics is applicable.
3. *total ¦ (*Pairs ) (pairwise additivity).
4. * / H F (r / V ) (universal functionality).

b) In general, assumption 4 is violated. But if we fix the core size to be a fixed fraction of the
collision diameter, then Kihara potential is a 2-parameter (V, H) potential that satisfies corre-
sponding states.
c) Hydrogen (at least at low temperatures) has a de Broglie wavelength large enough so that
quantum effects must be considered and therefore assumption 2 is violated. Assumption 1 is
probably pretty good for H2; assumption 4 is violated slightly. All substances violate assumption
3, but H2 isn’t very polarizable so it might be closer than the average substance to pairwise addi-
tivity.
d) Corresponding states (and thermodynamics in general) can only give us functions such as
c p  c 0p . Values of c 0p (for isolated molecules) cannot be computed by these methods, because
the contributions to c 0p (rotation, vibration, translational kinetic energy) appear in Qint and the
kinetic energy factor, not in the configuration integral.

16. Let D represent the phase inside the droplet and E the surrounding phase.
Schematically we have for the initial state and for the final (equilibrium) state:
Solutions Manual 40

NO3
D NO3 NO3 D NO3
+
+ K+ Na
K + Na
+
Na K+
2
Lys Lys2
E E

Initial state Equilibrium state

Because the molar mass of lysozyme is above the membrane’s cut-off point, lysozyme can-
not diffuse across the membrane.
Let:
G represent the change in K+ concentration in D;

M represent the change in Na+ concentration in E.

The final concentrations (f) of all the species in D and E are:

In D: c fD 0 c fD c 0D  G c fD  M c fD  c 0D   G  M
Lys 2 K K Na NO 3 NO 3

In E: c fE c 0E c fE G c fE  c 0E  M c fE c 0E   G  M
Lys2 Lys2 K Na Na NO
3 NO3

The equilibrium equations for the two nitrates are

P D  P D  PE   PE 
K NO 3 K NO 3
(1)
PD   PD  PE   PE 
Na NO3 Na NO3

Similar to the derivation in the text (pages 102-103), Eq. (1) yields

cE cE  c D c D 
K NO 3 K NO 3
(2)
E E
c c  cD cD
Na NO 3 Na NO
3

where, for clarity, superscript f has been removed from all the concentrations.
Substituting the definitions of G and M gives:

FH c0D  GKI FG c0D   G  MIJ FG IJ


K H NO3 K H
(G ) c 0E   G  M
NO 3 K
(3)
FH c0E  MIK FG c0E   G  MIJ FG IJ
Na H NO3 K H
(M) c 0D   G  M
NO 3 K
where
Solutions Manual 41

0.01 mol
c 0D c0D - 9.970 u 10 3 mol L-1
K NO3 1 kg water
(using the mass density of water at 25D C: 0.997 g cm-3 )

and

c 0E  c 0D - 0.01 mol L1


Na NO3

Solving for G and M gives

G 4.985 u 10 3 mol L1


(4)
3 1
M 5.000 u 10 mol L

Because both solutions are dilute, we can replace the activities of the solvent by the corre-
sponding mole fractions. The osmotic pressure is thus given by [cf. Eq. (4-50) of the text]
E
RT x s
S ln (5)
vs x sD

where xs is the mole fraction of the solvent (water) given by

xsD xwD D
1  ( xNO D
 x  x
D
)
3 K Na
(6)
E
xsE E
xw 1  ( xNO  xE +  xE  E
 xLys )
3 K Na

Because solutions are dilute, we expand the logarithmic terms in Eq. (5), making the ap-
proximation ln(1  A) |  A :

RT ª ( x E E E E D D D º
S «¬ NO3  xK   xNa   xLys )  ( xNO3  xK   xNa  ) »¼
vw

Again, because solutions are very dilute,

ci c
xi | i
¦i w
c c
i

with
v w cw | 1

Therefore, with these simplifying assumptions, the osmotic pressure is given by

D D D E E E E
S RT [cNO  c  c  cNO  c  c  cLys ]
3 K Na  3 K Na 

Using the relationships with the original concentrations, we have


Solutions Manual 42

S RT [(c 0D   G  M )  (c 0D  G)  (M)  (c 0E  G M )  (c 0E   M)  (cLys


0E
)]
NO3 K NO3 Na

Because

c 0D  c 0D
NO3 K

c 0E  c 0E 
NO3 Na

we obtain

S RT (2c 0D  2c 0E   cLys


0E
 4G  4 M ) (7)
K Na

The lysozyme concentration is

0E 2g (2 / 14,000) mol
cLys 1.429 u 10 4 mol L-1
1L 1L
Substitution of values in Eq. (7) gives the osmotic pressure

S (8314.51 Pa L mol -1 K -1 ) u (298 K) u [(2 u 9.97 u 10 3 )  (2 u 0.01)  (1.429 u 10 4 )

 (4 u 4.985 u 10 3 )  (4 u 5.000 u 10 3 ) (mol L1 )]

354 Pa

17. Because only water can diffuse across the membrane, we apply directly Eq. (4-41) derived
in the text:

Sv pure w
 ln aw  ln( xw J w ) (1)
RT
where subscript w indicates water.
Since the aqueous solution in part D is dilute in the sense of Raoult’s law, J w | 1 . This re-
duces Eq. (1) to:

Sv pure w
 ln xw  ln(1  xA  xA2 ) | ( xA  xA2 ) (2)
RT
or equivalently,
Solutions Manual 43

RT ( x A  xA 2 )
S
v pure w

At T = 300 K, v pure w 0.018069 mol L1.


Mole fractions xA and xA2 can be calculated from the dimerization constant and the mass
balance on protein A:

aA2 xA2
K 10 5 | (4)
2
aA 2
xA

Mw 18.015
vw 18.069 cm3 mol 1
dw 0.997

(5 / 5,000) mol A
. u 10 5
181 x A  2 x A2 (5)
(1000 / 18.069) mol water

Solving Eqs. (4) and (5) simultaneously gives

xA 7.34 u 10 6

xA2 5.38 u 10 6

Substituting these mole fractions in Eq. (3), we obtain

(0.0831451) u (300) u (5.38 u 10 6  7.34 u 10 6 )


S
(0.018069)

0.01756 bar 1756 Pa

18.
a) From Eq. (4-45):

S RT
 RTB*c2  "
c2 M2

Ploting S / c2 (with S in pascal and c2 in g L-1) as a function of c2, we obtain the protein’s
molecular weight from the intercept and the second virial osmotic coefficient from the slope.
Solutions Manual 44

50

47.5

45
S /c2, Pa L g -1

42.5

40

37.5

35
0 20 40 60
-1
BSA Concentration, g L

From a least-square fitting we obtain:


Intercept = RT/M2 = 35.25 Pa L g-1 = 35.25 Pa m3 kg-1 from which we obtain

RT . ) u (8.314)
(29815
M2
35.25 35.25

70.321 kg mol 1 70,321 g mol -1

Slope = 0.196 = RTB*. Therefore, B* = 7.92u10-8 L mol g-2.


The protein’s specific volume is given by the ratio molecular volume/molecular mass.
The mass per particle is

M2 70,321
m 1.17 u 10 19 g molecule 1
NA 6.022 u 1023

Because protein molecule is considered spherical, the actual volume of the particle is 1/4 of
the excluded volume. Therefore the actual volume of the spherical particle is

. u 10 24
118
2.95 u 10 25 m3 molecule 1
4
which corresponds to a molecular radius of 4.13u10-9 m or 4.13 nm.
For the specific volume:

2.95 u 10 25
2.52 u 10 6 m3 g 1 2.52 cm3 g 1
. u 10 19
117

b) Comparison of this value with the nonsolvated value of 0.75 cm3 g-1, indicates that the par-
ticle is hydrated.
Solutions Manual 45

c) Plotting S / c2 (with S in pascal and c2 in g L-1) as a function of c2 for the data at pH = 7.00,
we obtain:

60

57.5

55
S /c2, Pa L g -1

52.5

50

47.5

45

42.5

40
15 25 35 45 55
BSA Concentration, g L-1

Slope = 0.3317, which is steeper that at pH = 5.34, originating a larger second virial os-
motic coefficient: B* = 1.338u10-7 L mol g-2 = 13.38u10-8 L mol g-2.
At pH = 7.00, the protein is charged. The charged protein particles require counterions so
electroneutrality is obtained. The counterions form an ion atmosphere around a central protein
particle and therefore this particles and its surrounding ion atmosphere have a larger excluded
volume than the uncharged particle.
It is the difference between the value of B* at pH = 7.00 and that at pH = 5.37 for the un-
charged molecule gives the contribution of the charge to B*:

1000 z 2
(13.38  7.92) u10-8 L mol g-2 =
4 M22U1mMX

In this equation, we take the solution mass density U1 | Uwater | 1 g cm 3 . Moreover, M2 =


70,321 g mol-1, and mMX | 0.15 mol kg-1:

4 u (5.46 u 10 8 ) u (70,321) 2 u (10


. ) u (0.15 u 10 3 )
z2 | 162
1000
or z r13. Because pH is higher than the protein’s isoelectric point, the BSA must be nega-
tively charged. Hence, z = -13.
19.
a) From Eq. (4-45) and according to the data:

S RT
 RTB* (c  c0 )  "
c  c0 M2
Solutions Manual 46

Ploting S / (c  c0 ) (with S in pascal and c2 in g L-1) as a function of c  c0, we obtain the


solute’s molecular weight from the intercept and the second virial osmotic coefficient from the
slope.

18.0

17.5
-1
S/(c-c0), Pa L g

17.0

16.5

16.0

30 40 50 60
-1
(c-c 0), g L

From a least-square fitting we obtain:

Intercept = RT/M2 = 14.658 Pa L g-1 = 14.658 Pa m3 kg-1 from which we obtain

RT (298.15) u (8.314)
M2
14.659 14.659

169.109 kg mol1 169,109 g mol -1

Slope = 0.053495 = RTB*. Therefore, B* = 2.16u10-8 L mol g-2.

b) The number of molecules in the aggregate is obtained by comparison of the molecular


weight of the original ether (M = 390 g mol-1) with that obtained in a):

169,109
Number of molecules in the aggregate = | 434
390

Assuming that the colloidal particles are spherical, we obtain the molar volume of the ag-
gregates from the value of the second virial osmotic coefficient B*. It can be shown (see, e.g.,
Principles of Colloid and Surface Chemistry, 1997, P.C. Hiemenz, R. Rajagopalan, 3rd. Ed., Mar-
cel Dekker) that B* is related to the excluded volume Vex through

NA V ex
B*
2 M22
Solutions Manual 47

From a), B* = 2.16u10-8 L mol g-2 = 2.16u10-5 m3 mol kg-2 and M2 = 169,109 g mol-1 =
169.109 kg mol-1. The above equation gives Vex that is 4 times the actual volume of the particles
(we assume that the aggregates are spherical). Calculation gives for the aggregate’s volume,
5.13u10-25 m3 (or 30.90 dm3 mol-1) with a radius of 4.97u10-9 m or 5 nm.
S O L U T I O N S T O P R O B L E M S

C H A P T E R 5

1. Initial pressure Pi:


- for n1 moles of gas 1 at constant T and V:

PiV B
1  n1 11
n1RT V

or
n1RT n12 RTB11
Pi 
V V2
Final pressure Pf:
- after addition of n2 moles of gas 2 at same T and V:

(n1  n2 ) RT (n1  n2 )2 RTB


Pf 
V V2

where (n1  n2 ) 2 B n12 B11  2n1n2 B12  n22 B22


Pressure change 'P :

n2 RT RT
'P Pf  Pi  (2n1n2 B12  n22 B22 )
V V2

Solving for B12:

1 § V 2 'P ·
B12 ¨  n2V  n22 B22 ¸
2n1n2 ¨© RT ¸
¹

49
Solutions Manual 50

2. For precipitation to occur,


V
fCO ! fCO s
2 2

s
To obtain fCO ,
2

fCO s 1 60
ln
(0.1392)
2
(83.1451) u (173) 0.1392 ³
(27.6) dP

fCO s 0.156 bar (at 60 bar, 173 K)


2

Next, find the vapor mole fraction of CO2 that is in equilibrium with the solid at the speci-
fied P and T:
V
fCO
2
yCO2
M CO2 P

Using the virial equation for the vapor,

2
ln M VCO ( yCO2 BCO2  yH2 BH2  CO2 )  ln z
2 v

Because yCO2  1 , we may make the approximations

z H2 RT
z z H2 and v v H2
P

From data for H2 (see App. C) at –100qC,

BH2 8.8 cm3 mol 1

which indicates that z H2 1.


From correlations:

BCO2 460 cm3 mol 1

BCO2  H2 . cm3 mol 1


321

At equilibrium
V
fCO s
fCO
2 2

and then
Solutions Manual 51

s
fCO
2
yCO2
­ 2P ª ½
exp ® ¬ yCO2 BCO2  (1  yCO2 ) BH2 CO2 º¼ ¾ P
¯ RT ¿

yCO2 0.00344

Because yCO2  0.01 at equilibrium, CO2 precipitates.


To find out how much, assume solid is pure CO2. Let n be the number of moles of CO2 left
in the gas phase. From the mass balance and, as basis, 1 mole of mixture,

n
0.00344
n  0.99

n 0.003417

The number of moles precipitating is


0.01  0.003417 0.0066 moles CO2

3.

To = 313 K Tf = ?

Po = 70 bar Pf = 1 bar

V L .
Condensation will occur in the outlet if fCO ! fCO
2 2
First it is necessary to find the outlet temperature, assuming no condensation. Joule-
Thomson throttling is an isenthalpic process that may be analyzed for 1 mole of gas through a 3-
step process:

I, III: Isothermal pres- P


sure changes. To ,Po

II: Isobaric temperature


change.
I
Tf ,Pf

III
II
P=0
T
Solutions Manual 52

Then

'htotal 0 'hI  'hII  'hIII (1)

'hI
z LMMN
0

Po
v  To
FG wv IJ OP dP
H wT K P PQ (2)

Tf
'hII ³T o
c p,mixt dT (3)

'hIII
z LMMN
0
Pf
v  Tf
FG wv IJ OP dP
H wT K P PQ (4)

Assuming that the volumetric properties of the gaseous mixture are given by the virial equa-
tion of state truncated after the second term,

RT
v  Bmixt
P
then,

FG wv IJ R FG
dBmixt IJ
H wT K P P

dT HP K
where Bmixt is the second virial coefficient of the mixture.
Because

Bmixt y12 B11  2 y1 y2 B12  y22 B22 (5)

(1 = CH4; 2 = CO2)

dBmixt dB11 dB dB
y12  2 y1 y2 12  y22 22 (6)
dT dT dT dT
If

c (1) c (2)
B c (0 )  
T T2

then

dB c (1) 2c (2)
  (7)
2
dT T T3

Assume

c 0p, mixt yCH 4 c 0p,CH  yCO2 c 0p,CO (8)


4 2

then,
Solutions Manual 53

0 ª § dBmixt · º Tf Pf ª § dBmixt · º
0
³P «¬ Bmixt  To ¨©
o dT ¸ » dP 
¹P ¼ To ³
c p,mixt dT 
0 ¬ ³
« Bmixt  Tf ¨
© dT
¸ » dP
¹P ¼

§
(0) 2c(1) 3c(2) · §
(0) 2c(1) 3c(2) ·
0 ¨ cmixt  mixt  mixt ¸ ( Po )  c p,mixt (Tf  To )  ¨ cmixt  mixt  mixt ¸ ( Pf ) (9)
¨
©
To To2 ¸
¹
¨
©
Tf Tf2 ¸
¹

From data:

As y1 0.7 , y2 0.3 , c p,mixt 36.22 J K 1 mol1 according to Eq. (8).


From Eqs. (5), (6) and (7),
(0 )
cmixt 41849
.

(1)
cmixt 18683

( 2)
cmixt 34.12 u 10 5

Substitution in Eq. (9) gives Tf 278.4 K .

Second, the fugacities of liquid and vapor phases may be calculated.

ª P vL º
L s Ms exp « CO2 »
fCO
2
xCO2 J CO2 PCO
2 CO2 « PCO
s RT
dP
³»
¬« 2 »¼

This equation may be simplified assuming that xCO2 , J CO2 , M sCO equal to unity.
2

At 278 K, P s 39.8 bar and v L 3 


49.0 cm mol . 1
CO2 CO 2

L LM (49.0) u (1  39.8) OP
fCO
2
(39.8) u exp
N (831451
. ) u (278) Q
36.6 bar

Fugacity of vapor is calculated from


V
fCO M CO2 yCO2 P
2

with

P
ln MCO2 ª 2( yCO BCO  yCH BCO CH )  Bmixt º
¬ 2 2 4 2 4 ¼ RT

P
ln MCO2 >2 u (0.3) u (139)  2 u (0.7) u (77)  (69)@
RT
Solutions Manual 54

M CO2 0.995 | 1

Then
V
fCO 0.3 bar
2

Because
V L
fCO  fCO
2 2

no condensation occurs.

4. The Stockmayer potential is

LMF V I 12 F V I 6 OP P 2
* 4H
MNGH r JK  GH ( r JK PQ  r 3 g(T1, T2 , I2  I1 )
where P is the dipole moment.

T1 T2 I

We can write the potential in dimensionless form:

F r , P2 I
*
H
f GH V HV 3 JK where f is a universal function.

Therefore, we can write the compressibility factor z in terms of the reduced quantities:

z f (T , P , P )

with
~ kT
T
H
Solutions Manual 55

PV3
P
H

~ P
P
HV 3

5.
a) For acetylene: Tc 308.3 K , Z 0.184 . At 0qC, TR 0.886 | 0.90 . Using Lee-Kesler charts
(see, e.g., AIChE J., 21: 510 [1975]):

's (0 ) / R 3.993 zV(0 ) 0.78 z L(0 ) 0.10

's (1) / R 3.856 zV(1) 0.11 z L(1) 0.04

' vap s . J K 1 mol 1


(8.31451) u [3.993  (0.184) u (3.856)] 391

' vap h T ' vap s (273) u (39.1) 10.67 kJ mol 1

' vapu ' vap h  RT (zV  z L ) 10.67  (8.31451) u (273) u (0.76  0.092) u 10 3

' vapu . kJ mol 1


915

b)
C 4 H10 : Tc 425.2 K N 2 : Tc 126.2 K

Pc 38.0 bar Pc 33.7 bar


Z 0.193 Z 0.04
vc 255 cm3 mol-1 vc 89.5 cm3 mol-1

At 461 K: TR 1084
. TR 3.65

Using the Pitzer-Tsonopoulos equation (see Sec. 5.7):

BC 4 H10 267 cm3 mol-1 BN2 15.5 cm3 mol-1

For B12 :
1
Z 12 (Z 1  Z 2 ) 0.1165
2
Solutions Manual 56

Tc12 (Tc1Tc2 )1 2 2316


. K ( TR12 1990
. )

zc12 0.291  0.08Z12 0.2817

zc12 RTc12
Pc12 34.3 bar
1 1/ 3
(v  v1c/ 3 ) 3
8 c1 2

Then, B12 . cm3 mol-1


216

Bmixt y12 B11  2 y1y2 B12  y22 B22 177.2 cm3 mol 1

c)
CH 4 (1): Tc 190.6 K N 2 (2): Tc 126.2 K H 2 (3): Tc 33.2 K

Pc 46.0 bar Pc 33.7 bar Pc 13.0 bar


Z 0.008 Z 0.040 Z 0.22

vc 99.0 cm3 mol 1 vc 89.5 cm3 mol 1 vc 65.0 cm3 mol 1

At 200 K: TR .
158 TR 6.02 TR .
105

At 100 bar: PR 2.97 PR 2.17 PR 7.69

Using the mixing rules suggested by Lee and Kesler:

1
v c,mixt ¦ ¦ x j x k (v1c/j3  v1c/k3 )3
8 j k

1
Tc,mixt ¦ ¦ x j x k (v1c/j3  v1c/k3 )3 (Tc j Tck )1/ 2
8v c j k

Zmixt ¦ x jZ j # 0
j

Pc,mixt (0.291  0.08Z mixt ) RTc,mixt / v c,mixt

v c,mixt 84.1 cm 3 mol 1

Tc,mixt 111.38 K ( TR . )
180

Pc,mixt 31.47 bar ( PR . )


318

1
Enthalpy of mixing = h E hmixture  (h1  h2  h3 ).
3
Using the Lee-Kesler charts,
Solutions Manual 57

1
hE 922  (5385  1383  0) 1334 J mol -1
3

6.

L/G 5 L G

P 40 bar

t 25 qC

40% N2
50% H2
10% C3 H8

From the mass balance for C 3 H 8 :

yinG yout G  x out (L  [ yin  yout ]G )

yout G 0.05 yinG

0.10 0.005  5.005 xC3

x C3 0.01898 mole fraction of C3H8 in effluent oil.

To find the driving force, note that

Pi PC3 yC3 P total (0.10) u (40) 4 bar

Pi* PC*3 fC*3 / M *C3

where

fC*3 x C3 H (0.01898) u (53.3) 1012


. bar

and

2P
ln M *C3 [ yC3 BC3  yN2 BC3  N2  yH2 BC3  H2 ]  ln z *
z * RT

Obtain virial coefficients from one of the generalized correlations:


Solutions Manual 58

BC3 400.8 cm3 mol-1

BC3  N2 73.5 cm3 mol-1

BC3  H2 3.5 cm3 mol-1

We may estimate z * 0.95 (feed value).


* * * *
To find yC 3
, yN 2
, and yH 2
, we know that ( yN / y* ) = 4/5. As first guess, assume
2 H2
*
yC 3
0.10. Then we calculate,

1012
.
M *C3 *
0.855 and yC 3
0.0292
(40) u (0.855)

This gives
*  y*
yN 0.9708 *
yN 0.4308 , y*H2 0.540
2 H2 or 2

With these y * ’s, calculate M *C3 again:

M *C3 0.924 and *


yC 3
0.027

That is close enough. Thus,

1012
.
PC*3 1095
. bar
0.924

Now we must check the assumption z * 0.95 was correct. Using the virial equation, the as-
sumption is close enough.

Driving force = (PC3  PC*3 ) (4.00  1095


. ) 2.91 bar

7. Since fiV fiL ,

v f ( P  Psolv
s
)
yi Mi P xi Hi,solv exp i
RT
s
As Psolv 0,

yi Hi exp(vif P / RT )
xi Mi P

Using the virial equation,


Solutions Manual 59

2
ln M1 ( y1B11  y2 B12 )  ln z mixt
v

2
ln M 2 ( y2 B22  y1B12 )  ln z mixt
v
Because

Pv B
1  mixt
RT v

RT B RT
v2  v  mixt 0
P P

Bmixt 3.23 cm3 mol-1

v 524 cm3 mol-1 , z mixt 10067


.

Thus,

M1 0.8316 M2 0.9845

LM (60) u (50) OP
y1
(100) u exp
N (313) u (8314
. )Q
2.70
x1 (50) u (0.8316)

and

y2
2131
.
x2

FG y2 IJ FG x1 IJ
D 2,1
H x2 K H y1 K 7.89

8. For methane (1) and methanol (2), we may write

f 1V f 1L

f 2V f 2L

Neglecting the Poynting corrections [Note: the Poynting Correction is 1.035. Including this,
we get y2 = 0.00268],

y1M1P x1H1,2
Solutions Manual 60

y2 M 2 P x 2 J 2 P2s M s2

Because x2 = 1, assume J 2 1 .
Use virial equation to get fugacity coefficients:

B22 P2s (4068) u (0.0401)


M s2 exp exp 0.993
RT . ) u (273)
(8314

Assuming y1 1 , y2 0 as first estimate,

M1 0.954 M2 0.783

Thus,

M 1P
x1 0.0187
H1,2

(1  x1 )P2s M s2
y2 0.00250
M2P

Using now y2 0.00250 , get

M1 0.954 M2 0.770

and

x1 0.01867 y2 0.00255

This calculation is important to determine solvent losses in natural gas absorbers using
methanol as solvent.

9.

2(monomer) dimer

The equilibrium constant is


2
ad fd / fd0 fd §¨ fm0 ·
¸
Ka
(am )2 ( fm / fm0 )2 fm2 ¨© fd0 ¸
¹

where ad is the activity of the dimer, and am is the activity of the monomer.
2
The quantity f m0 / fd0 is a constant that depends on T, but not on P or y.
Then,
Solutions Manual 61

fd k f m2

where k is a constant.

10.
CH3 CH3

CH O CH

CH3 CH3 CH3CH2 O CH2CH2CH2CH3

Di-isopropyl ether Ethyl butyl ether

HCl can associate with the ether’s non-bonded electron pairs. However, the di-isopropyl
ether will offer some steric hindrance. The cross-coefficient, B12 , is a measure of association.
Both virial coefficients will be negative; B12 for ethyl butyl ether/HCl will be more negative.

11. Let D be the fraction of molecules that are dimerized at equilibrium.

2A A2
1 D D/2

D D
nT 1 D  1
2 2

D/2
yA
2 1 D / 2

1 D
yA
1 D / 2

By assuming the vapor to be an ideal gas, we may write

PA2 yA2 P (D / 2)(1  D / 2)


K
2 2
( PA ) ( yA P) (1  D)2 P

At the saturation pressure, P = 2.026 bar and

yA 0.493 D 0.6726
Solutions Manual 62

Then

fAV fAL yA P (0.493) u (2.026) 0.999 bar

The pressure effect on fugacity is given by

FG w ln f IJ vi
H wP K
i
T RT

Assuming the liquid to be incompressible in the range Ps to 50 bar,

fA (50 bar) v A 'P


ln
fA (2.026 bar) RT

fA (50 bar) 1.1 bar

12.
a) The Redlich-Kwong equation is

Pv v a 1 FG IJ
z
RT

v  b RT 1.5 vb H K
If z is expanded in powers of 1/ v :

B C
z 1  "
v v2
This gives
a
B b
RT 1.5

ab
C b2 
RT 1.5
But,
B' B / RT

C  B2
C'
(RT )2

Substitution gives

1 FG
a IJ
B'
RT
b
H
RT 1.5 K
Solutions Manual 63

a FG 3b  a IJ
C'
R T 3 3.5 H RT K 1.5

b) Using an equation that gives fugacities from volumetric data, we obtain

P ª y2 a  2 y1a1  y22 a2  2 y22 (a1a2 )1/ 2 º


ln M1 « b1  1 1 »
RT «¬ RT 1.5 »¼

Evaluate a and b using critical data:


Ethylene (1): Tc = 282.4 K Pc = 50.4 bar

Nitrogen (2): Tc = 126.2 K Pc = 33.7 bar

a1 = 7.86u107 bar cm6 K1/2 mol-2

a2 = 1.57u107

b1 = 40.4 cm3 mol-1

Substitution gives M1 0.845 ,

f1 y1M1P 8.44 bar

13. Using the virial equation,

RT RTBmixt
P 
v v2
with

Bmixt y12 B11  2 y1 y2 B12  y22 B22 (1)

For maximum pressure,

FG wP IJ FG
RT wBmixt IJ
H wy K T
1 ,v v2 H
wy1 T K 0 (2)

Substituting Eq. (1) into Eq. (2) gives ( y2 1  y1 ) :

§ wBmixt ·
¨ ¸ 2 y1 B11  2 y2 B12  2 y1 B22  2 B22 0
© wy1 ¹T

At maximum,
Solutions Manual 64

B22
y1 (3)
B11  B22  B12

Using the correlations,

B11 126.7 cm3 mol-1 (ethylene)

B22 12.5 cm3 mol-1 (argon)

B12 45.9 cm3 mol-1

Substitution in Eq. (3) gives

y1 = 0.134

14. Consider a 3-step process:

Pure ideal gases Mixed ideal gases


Tf Tf

P = 20.7 bar
I

Pure gases Mixed gases


Ti Tf

The overall enthalpy change is zero:

'H I  'H II  'H III 0

'H I n1'H1  n2 'H 2 (1 = hydrogen; 2 = ethylene)

H0  H
'H1 c 0p1 (Tf  Ti )  (RTc1 )
RTc1

H0  H
'H 2 c 0p2 (Tf  Ti )  (RTc2 )
RTc2

H0  H
where is evaluated using Lee-Kesler Tables.
RTc
Solutions Manual 65

For heat capacities, we can estimate

c 0p1 28.6 J K-1 mol-1

c 0p2 43.7 J K-1 mol-1

'H II 0 (because we are mixing ideal gas)

F H  H I RT
0
'HIII  GH RTc JK c
,mixt
, mixt evaluated at TR = Tf / Tc,mixt, PR = 20.7/Pc,mixt

Find Tf by trial and error:

Tf 247 K

15. At equilibrium,

fAV s
fA

yA M A P (1  xCO2 )PAs M sA exp


z s
PA
P vA
RT
s
dP

Assuming xCO2 | 0 ,

( y A P)MA s
ª v ( P  Ps ) º
exp « A A »
PAs MsA «
¬
RT »
¼

But

§ B Ps ·
MsA exp ¨ AA A ¸ 1.00
¨ RT ¸
© ¹

Because yA  yCO2 , yCO2 # 1

ª P º
MA exp «(2 BA CO2  BCO2 )
¬ RT »¼

yA P s
ª v ( P  P s ) (2 BA CO  BCO )P º
exp « A A
 2 2
»
PAs «
¬
RT RT »
¼

or
Solutions Manual 66

P
ª s s y Pº s
« PA v A  RT ln As » /(v A  2 BA CO2  BCO2 )
¬« PA ¼»

Substitution gives
P = 68.7 bar

For this pressure, yA . u 10 4  1 and assumption yA | 0 is correct.


19

16. Let 1 = ethylene and 2 = naphthalene.


As, at equilibrium,

f2 s f2V

and as

s
f2 P2s M s2 exp
z
P22
P s
v2
RT
dP

f2V y2 M 2 P

we may write

y2 s
(1  x1 )P2s M s2 exp[v 2 (P  P2s ) / RT ] / M V2 P

a) Using ideal-gas law:

M s2 M V2 1

s
As v 2
128.174
111.94 cm 3mol 1 ,
1.145

s
ª v (P  Ps ) º
y2 P2s exp « 2 2 »
/ P 1.1u 10 5
« RT »
¬ ¼

b) Using VDW constants:

a 27 R 2 Tc2 / 64 Pc b RTc / 8 Pc

a1 4.62 u106 bar cm6 mol-2 b1 58.23 cm3 mol-1

a2 4.03 u107 bar cm6 mol-2 b2 192.05 cm3 mol-1


Solutions Manual 67

B11 b1  a1 / RT 122.2 cm3 mol-1

B22 1382 cm3 mol-1

Because y2 << y1, y1 # 1 and Bmixt # B11

B Ps
M2s exp 22 2 # 1
RT

P P
ln MV2 >2( y1B12  y2 B22 )  Bmixt @ # (2 B12  B11 )
RT RT

B12 b12  a12 / RT with b12 1 / 2(b1  b2 ) and a12 (a1a2 )

b12 . cm3 mol-1


12514 and a12 = 1.36u107 bar cm6 mol-2

Then B12 = -406 cm3 mol-1.

ª P º
MV2 exp «(2 B12  B11 ) 0.446
¬ RT »¼

y2 (2.80u10-4) u exp [(111.9u30)/(83.1451u308)]/(0.446u30)

y2 = 2.4u10-5

17. Water will condense if f HV O ! f H O . s


2 2
Thus, the maximum moisture content, yH2O , is given by

f HV s
fH
2O 2O

s
yH2O M VH O P
2
x H2O PHs O
2
exp
z Ps
P

H 2O
vH O
2
RT
dP

Assuming that the condensate is pure (solid) water,

x H2O 1 and M sH 1
2O

Then
Solutions Manual 68

ª § 1.95 · º
(18 / 0.92) u ¨ 30 
(1.95 torr) «
© 750.06 ¸¹ »
( yH2O MVH ) u (30 bar) exp « »
2O (750.06 torr/bar) « (83.1451) u (263.15) »
« »
¬ ¼

8.90 u 10 5
y H 2O
M VH
2O

Let 1 = N2, 2 = O2, and 3 = H2O.


To get M VH use the virial equation of state:
2O

P
ln MVH O ln M3 [2( y1 B13  y2 B23  y3 B33 )  Bmixt ]
2 RT
with

Bmixt ¦ ¦ yi y j Bij
i j

Assume y3 << y1 where y1 = 0.80 and y2 = 0.20.


Then,

Bmixt y12 B11  y22 B22  2 y1 y2 B12 21.8 cm3 mol-1

Substitution in the equation for ln M 3 gives M 3 0.871.


As

8.90 u 10 5
y H 2O
M VH
2O

then

yH 2O 1.0 u 10 4

18. The Joule-Thompson coefficient is defined as

FG wT IJ
PH {
H wP K H
(1)

Applying the triple-product rule with T, P and H, we have


Solutions Manual 69

FG wT IJ FG wP IJ FG wH IJ
H wP K H wH K H wT K
H T P
1 (2)

Because

§ wH ·
¨ wT ¸ { c p
© ¹P

and

FG wH IJ 1
H wP K T FG wP IJ
H wH K T
combining Eqs. (1) and (2) gives

1 § wH ·
PH  (3)
c p ¨© wP ¸¹T

From the fundamental equation dH TdS  VdP , we have

FG wH IJ FG wS IJ
H wP K T
T
H wP K T
V (4)

However, we also have Maxwell’s relation:

FG wS IJ FG wV JI
H wP K T

H wT K P

Therefore,

FG wH IJ FG wV IJ
H wP K T
T
H wT K P
V (5)

Substituting Eq. (5) into Eq. (3) gives

1 ª § wV · º
PH  « T ¨ ¸ V» (6)
c p ¬ © wT ¹ P ¼

or, in terms of molar volume (v) and molar heat capacity at constant pressure (c p ) ,

1 wvLM FG IJ  vOP
PH 
cp
T
wT MN H K P QP (7)

Because in this specific problem we want P H of the hydrogen-ethane mixture to be zero,


Eq. (7) yields

FG wv IJ
v T
H wT K P
(8)
Solutions Manual 70

From the truncated-virial equation of state

Pv B
1 (9)
RT v
we have

FG wT IJ FG P IJ FG v IJ FG 2  v IJ FG P IJ v(v  2B)
H wv K H R K H v  B K H v  B K H R K (v  B)
P
2
(10)

Substituting the equation of state Eq. (9) and Eq. (10) into Eq. (8) yields

RT FG
B IJ T LM R (v  B) OP 2

P
1
H
v K MN P v(v  2B) QP (11)

or equivalently,
B 0 (12)
where B is the second virial coefficient of the hydrogen-ethane mixture at 300 K.
Using the McGlashan and Potter equation [Eq. (5-52)],

B FG T IJ 1  0.694FG T IJ 2
vc
0.430  0.866
H Tc K H Tc K (13)

we obtain B11 11.4 cm 3 mol 1 for hydrogen, and B22 173.3 cm 3 mol 1 for ethane, and
B12 B21 11.4 cm3 mol 1 for the cross term, respectively.
Applying

Bmixt ¦ ¦ yi y j Bij
i j

and the material balance y1  y2 1, , we have

Bmixt ( cm 3 mol 1 ) 11.4 y12  22.8 y1 (1  y1 )  173.3(1  y1 ) 2 (14)

Equations (12) and (14) yield y1 0.73.


Consequently, if we start out with 1 mol of H2, the amount of ethane that must be added to
have a zero P H is 0.37 mol.

19. Because methane does not significantly dissolve in liquid water at moderate pressures, the
equation of equilibrium is

f2V L
f pure 2 (1)
Solutions Manual 71

or equivalently,

LM v2L eP  P2s j OP
y2 M 2 P P2s M s2 exp
MM RT PP (2)
N Q
where subscript 2 denotes water.
The fugacity coefficient from the volume-explicit virial equation of state is given by Eq. (5-
33):

P
ln M 2 2( y1B12  y2 B22 )  Bmixt (3)
RT
with

Bmixt y12 B11  2 y1 y2 B12  y22 B22 (4)

Similarly, we also have

B22 P2s
ln M s2 (5)
RT

(i) At the inlet (60ºC, 20 bar):

Substitute Eqs. (3), (4), and (5) into Eq. (2). Solving Eq. (2) using v 2L 18 cm 3 mol -1 and
P2s 149 mmHg , we obtain

y2i 0.011 (superscript i denotes inlet).

(ii) At the outlet (25ºC, 40 bar):

Similarly, with v 2L 18 cm 3 mol -1 and P2s 24 mmHg , we obtain

y2o 0.000951 (superscript o denotes outlet).

Because the gas phase is primarily methane, the amount of water that must be removed per
mol methane is

mol water removed


y2i  y2o | 0.01
mol methane

20. Assuming negligible changes in potential and kinetic energies, the first law of thermody-
namics for a steady-state flow process is
Solutions Manual 72

'H Q  Ws (1)

where 'H is the change in enthalpy, and Q and Ws are, respectively, the heat and the shaft
work done on the system (by the surroundings).
Because we are given the initial and final state and enthalpy is a state function, 'H is fixed
in this problem. Consequently, minimum Ws corresponds to maximum Q . Maximum Q occurs
when the process is reversible, or equivalently Q T'S .
Hence, Eq. (1) can be rewritten as

Ws 'H  T'S (2)

where Ws is now the minimum amount of work required for the process.
To calculate 'H and 'S from the volume-explicit virial equation of state

PV
1  BP
nT RT

we take an isothermal reversible path from the initial to the final state.
Expressions for enthalpy and entropy are given by Eqs. (3-9) and (3-10) in the text:

H
z P

0
MMLV  T FGH wwVT IJK
N P,nT
OP
PQ dP  n h
0
1 1  n2 h20

z MN
(3)
PLM
nT R wV FG IJ OP
S
0 P

wT P,n
T
H K PQ
dP  R(n1 ln y1P  n2 ln y2 P)  n1s10  n2 s20

Substituting

§ wV · nT R
¨ wT ¸ (1  BP)
© ¹ P,nT P

from the virial equation of state gives

H n1h10  n2 h20
(4)
S nT RBP  R(n1 ln y1P  n2 ln y2 P )  n1s10  n2 s20

Applying Eq. (4) to this specific problem, we have


'H 0
(5)
'S RP(nT Bmixt  n1B11  n2 B22 )  R(n1 ln y1  n2 ln y2 )

At 298 K, second virial coefficients are


Solutions Manual 73

B11 . cm3 mol -1


419

B12 122.2 cm3 mol -1 (6)

B22 66.0 cm 3 mol -1

Hence, at 298 K,

Bmixt ¦ ¦ yi y j Bij 74.0 cm3 mol -1


i j

Taking a basis of nT 1 mol of the mixture initially, then n1 n2 0.5 mol, Eq. (5) yields
at 298 K
'H 0
(7)
T'S 248.5 kJ

Substituting Eq. (7) into Eq. (2), the minimum amount of work required for this process is
248.5 kJ mol 1 of initial mixture.

21. The second virial coefficient for a square-well potential is given by Eq. (5-39) in the text

F R3  1 I
B GH
b0 R 3 1 
R3
exp
H
kT JK (1)

with

2
b0 SN AV 3 (2)
3
Substituting H / k 469 K , V 0.429 nm and R 0.337V gives

B(423 K) 302 cm3 mol -1

Because, for a pure component,

f BP
ln M ln
P RT
fugacity is given by

FG BP IJ
f P exp
H RT K
Solutions Manual 74

Hence, at 150ºC and 30 atm,

LM (30 atm) u (301.8 cm mol ) OP 3 1


f (30 atm) u exp
MN (82.06 atm cm mol ) u b423 Kg PQ
3 1

38.94 atm | 39 atm

To obtain the standard enthalpy and entropy of dimerization of methyl chloride we assume a
small degree of dimerization. In this case, the relation between the second virial coefficient and
the dimerization constant is given by Eq. (5-113):

RTK
B b
P0

Applying P 0 RTc 0 (where c 0 1 mol L1 = 10 -3 mol cm 3 is the standard state) gives

103 (cm3 mol 1 )K Bb

Because we only have a weak dimerization,

2
b b0 SN AV 3 (3)
3
and the second virial coefficient is essentially that of pure methyl chloride [Eq. (1)].
Combining Eqs. (1) and (3) gives

FG 2 SN V 3 IJ LMR3 FG1  R3  1 exp H IJ  1OP


10 3 (cm 3 mol 1 )K
H3 A
K MN H R3 kT K PQ (4)

The following table shows K (T ) calculated from Eq. (4).

T (K) K
100 10.74
200 0.939
300 0.376
400 0.222
500 0.155

Because the standard enthalpy and entropy of dimerization obey [Eq. (5-114)]

'h 0
 R ln K  's 0
T
plotting R ln K as a function of 1/T gives 'h 0 as the slope, and 's 0 as the intercept.
Results are

'h0 4.34 kJ mol 1

's 0 23.2 J K 1 mol 1


Solutions Manual 75

22.
a) Substitution of given *(r) in Eq. (5-19) gives

ª V f n º
2SN A « (1  ef / kT ) r 2 dr  (1  e A / kTr ) r 2 dr »
B
¬ 0
³ V ³ ¼

ª V f n º
2SN A « r 2 dr  (1  e A / kTr ) r 2 dr »
¬ 0
³ V ³ ¼
(1)

ª V3 f n º
(1  e A / kTr ) r 2 dr »
2SN A « 
¬« 3 V ³ ¼»

At high temperatures, A/kTrn is small.


Because

x2
ex | 1 x  " when x is small,
2!
n
we expand the exponencial e A/ kTr :

n A
e A/ kTr | 1  "
kTr n

Substitution of this approximation into Eq. (1) gives:


f f§ A · 2  A f n 2
A / kTr n
) r 2 dr
³V (1  e ³V ¨©1  1  kTr n ¸¹ r dr kT V ³
(r )r dr

f
 A f 2 n  A ª r 3 n º
kT V³(r )dr « »
kT «¬ 3  n »¼
V
(2)

A § V3 n ·  AV3 n
¨ ¸
kT ¨© 3  n ¸¹ kT (n  3)

We substitute now this result in Eq. (1):

2
SN AV 3  2SAN A
V 3 n LM OP
MN PQ
B
3 kT (n  3)

Constant n is large (i.e., n > 3):


Solutions Manual 76

2
SN AV 3  2SAN A
V3 LM OP
MN PQ
B
3 kT (n  3)V n
(3)
2 2
SN AV 3  SN AV 3
3A LM OP
3 3 kT (n  3)V n N Q
b) From Eq. (3) we see that it is the attractive part of the potential that causes negative B
and is responsible for the temperature dependence of B [the first term on the right hand side of
Eq. (3) is independent of temperature].

23. Substitution of the square-well potential [Eq. (5-39)] into Eq. (5-17) gives

f
(1  e*(r ) / kT ) r 2 dr
B 2SN A
³ 0

ª V R' f º
(1  ef / kT )r 2 dr  (1  eH / kT )r 2 dr  (1  e0 / kT )r 2 dr »
2SN A «
¬
³ 0 ³ V ³ R' ¼

ª V3 R' º
(1  eH / kT )r 2 dr  0 »
2SN A « 
¬« 3
³
V ¼»

ª V 3 § R '3 V 3 · º
2SN A «  ¨  ¸ (1  eH / kT ) »
«¬ 3 ¨© 3 3 ¸¹ »¼

2 2
SN AV3  SN A (1  eH / kT )( R '3  V3 )
3 3

In the equation above, R' = RV = 1.55V. For argon, V = 0.2989 nm = 0.2989u10-9 m, H/k
= 141.06 K, and R' = 1.55V = 4.633u10-10 m.
The above equation gives for T = 273.15 K,

B 3.368 u 10 5  (6.202 u 10 5 )

2.834 u 10 5 m 3 mol 1 | 28 cm 3 mol -1

The calculated value compares relatively well with the experimental B for argon at the same
temperature: Bexp = -22.08 cm3 mol-1.
Solutions Manual 77

24. Substitution of Sutherland potential [Eq. (5-37)] in Eq. (5-19) with N = NA gives

ª V f 6 º
(1  ef / kT ) r 2 dr  (1  eK / kTr ) r 2 dr »
B(T ) 2SN A «
¬
³0 ³ V ¼
(1)
2 f 6
SN AV3  (1  eK / kTr ) r 2 dr
3 ³V

§ K · K K2 K3 K4
exp ¨ ¸ | 1    " (2)
© kTr 6 ¹ kTr 6 2(kT )2 r12 6(kT )3 r18 24(kT )4 r 24

x2 x3 x 4
[ ex | 1  x    "]
2! 3! 4!

We now have to replace the approximate result [Eq. (2)] in Eq. (1) and perform the neces-
sary integrations.
The result is:

2 2SN A K 2SN A K 2 2SN A K 3 2SN A K 4


B(T ) SN AV 3     "
3 3kTV 3 18(kT ) 2 V 9 90(kT )3 V15 504(kT ) 4 V 21

This equation is best solved using an appropriate computer software such as Mathematica,
TKSolver, MathCad, etc.
Making the necessary programming we obtain at 373 K,

B(methane) = -20 cm3 mol-1

B(n-pentane) = -634 cm3 mol-1


In both cases, the agreement with experiment is very good.

25. The equation of equilibrium for helium is

f1L f1R (1)

where superscripts L and R stand for left and right compartments, respectively.
Equivalently,
Solutions Manual 78

y1L M1L P L y1R M1R P R (2)

Equation (5-33) of the text gives the fugacity coefficients in both compartments from the
volume-explicit virial equation of state:

P
ln M1 2( y1B11  y2 B12 )  Bmixt (3)
RT
with

Bmixt y12 B11  2 y1 y2 B12  y22 B22 (4)

Further, we also have material balances

y1L  y2L 1
(5)
y1R  y3R 1

Applying Eqs. (3), (4), and (5) to both compartments yields

PL
ln M1R {2 ª¬ y L B
1
L º ª L 2 L L
11  (1  y1 ) B12 ¼  ¬( y1 ) B11  2 y1 (1  y1 ) B12 }
 (1  y1L )2 B22 º
¼ RT
(6)
PR
ln M1L {2 ª¬ y R B
1
R º ª R 2 R R
11  (1  y1 ) B13 ¼ ¬( y1 ) B11  2 y1 (1  y1 ) B13 }
 (1  y1R )2 B33 º
¼ RT

Total mole balance on helium gives

n1L  n1R 0.02 mol (7)

Combining with mass balances on ethane and nitrogen gives

n1L
y1L
0.99  n1L
(8)
n1R 0.02  n1L
y1R
0.99  n1R .  n1L
101

Substituting Eqs. (6) and (8) into Eq. (2) yields

n1L 0.013 mol

Combining this result and Eq. (8) gives

y1L 0.013

y1R 0.007
Solutions Manual 79

26. Because the equilibrium constant is independent of pressure, the more probable reaction is
the one that satisfies this condition.

(1) Assuming reaction (a) is more probable:

With this scheme, concentration of (HF) 6 is negligible compared to those of ( HF) and
(HF) 4 .
The equilibrium constant K (a ) is

y( HF ) 4 y( HF ) 4
K (a ) (1)
y(4HF ) P 3 1  y( HF ) 4
4
P3

Total mass balance for ( HF) gives

y( HF) 4 nT (4 u M HF )  1  y( HF) 4 nT M HF V u U HF (2)

where V is the total volume; U HF and M HF are the mass density and the molar mass of hydro-
gen fluoride, respectively; nT is the total number of moles that can be calculated by assuming
that the gas phase is ideal:

PV
nT (3)
RT
Substituting Eqs. (2) and (3) into Eq. (1) yields

FG U RT  1IJ
HPuM K
(1 / 3) HF
HF

LM1  (1 / 3)F U RT  1I OP
K (a ) (4)
4

MN GH P u M JK PQ
HF
HF
P 3

Applying Eq. (4) at the two pressures, 1.42 and 2.84 bar, we obtain for K (a ) :

P (bar) UHF (g/L) K(a)

1.42 1.40 0.0595


2.84 5.45 0.453

Because K (a ) depends on pressure, reaction (a) cannot be the more probable one.

Next we need to check for the pressure independence of K ( b ) .

(2) Assuming reaction (b) is more probable:


Solutions Manual 80

In this case, concentration of (HF) 4 is negligible compared to those of ( HF) and (HF) 6 .
The equilibrium constant K ( b ) is

y( HF )6 y( HF )6
K( b) (5)
y(6HF ) P 5 1  y( HF )6
6
P5

Mass balance for HF in this case is:

y( HF )6 nT (6 u M HF )  1  y( HF )6 nT M HF V u U HF (6)

where all terms are defined in Eq. (2).


Substitution of Eqs. (3) and (6) into Eq. (5) gives

FG U RT  1IJ
HPuM K
(1 / 5) HF
HF

MMNL1  (1 / 5)FGH PUu MRT  1IJK OPQP


K ( b) (7)
6
HF P5
HF

Corresponding values of K ( b ) at 1.42 and 2.84 bar are:

P (bar) UHF (g/L) K(b)

1.42 1.40 0.017


2.84 5.45 0.017

Because K ( b ) is independent of pressure, reaction (b) is the more probable.


S O L U T I O N S T O P R O B L E M S

C H A P T E R 6

1. The three equations of equilibrium (in addition to T L T V and P L P V ) are

y1M1V P x1J 1 f1L

y2M V2 P x2 J 2 f2L

y3M V3 P x3 J 3 f3L

with (assuming the liquid incompressible)

v L (P  P1s )
f1L P1s M1s exp 1
RT
L L
We write similar expressions for f 2 and f 3 .
For M V we may write

RT ln M1V P(2 y1B11  2 y2 B12  2 y3 B13  Bmixt )

and similar expressions for M V2 and M V3 . In these equations,

Bmixt y12 B11  y22 B22  y32 B33  2 y1 y2 B12  2 y1 y3 B13  2 y2 y3 B23

81
Solutions Manual 82

2. Given g E Ax1 x 2 with P1s / P2s 1.649 , and assuming ideal vapor,

y1P
J1
x1P1s

y2 P
J2
x 2 P2s

At azeotrope, x1 y1:

P
ln J 1 ln
P1s

P
ln J 2 ln
P2s

or
J1 P2s
ln ln 0.5
J2 P1s

From the g E expression,

A 2
ln J 1 x
RT 2
and
A 2
ln J 2 x
RT 1
Then
J A 2
ln 1 ( x  x12 )
J2 RT 2

A 0.5 1
RT x22  x12 4 x2  2

or
1 RT
x2 
2 4A

Because 0 d x 2 d 1 ,

1 RT 1
 d d
2 4A 2
1
Thus, if | A| t RT , an azeotrope exists.
2
Solutions Manual 83

3. From the plot P-x-y we can see the unusual behavior of this system:

1. There is a double azeotrope


2. Liquid and vapor curves are very close to each other.

75

74
Pressure, kPa

73

Liquid ( )
72
Vapor ( )

0 0.2 0.4 0.6 0.8 1.0


x1, y1

4. Neglecting vapor phase non-idealities,

P x1J 1P1s  x 2 J 2 P2s (1)

At the maximum,

§ wP · § wJ · § wJ ·
¨ ¸ 0 J1P1s  x1P1s ¨ 1 ¸  x2 P2s ¨ 2 ¸  J 2 P2s (2)
© wx1 ¹T wx
© 1 ¹T © wx2 ¹T

From the Gibbs-Duhem equation (at constant T and low pressure),


Solutions Manual 84

§ w ln J1 · § w ln J 2 ·
x1 ¨ ¸  x2 ¨ ¸ 0
© wx1 ¹T © wx1 ¹T
or
x1 § wJ1 · x2 § wJ 2 ·
¨ ¸ ¨ ¸ (3)
J1 © wx1 ¹T J 2 © wx2 ¹T

Substituting Eq. (3) into Eq. (2) and simplifying,

§ x wJ ·
( J1P1s  J 2 P2s ) ¨ 1  1 1 ¸ 0
© J1 wx1 ¹

There are two possibilities:

(1) J 1P1s  J 2 P2s 0

Then
J 2 P2s ( y2 / x2 )
1 D
J 1P1s ( y1 / x1 )

D 1 corresponds to an azeotrope

x wJ
(2) 1 1 1 0
J 1 wx1

The solution to this differential equation is x1J 1 constant .


To find the constant, use the boundary condition J 1 1 when x1 1.
Hence J 1 x1 1.

As y1P x1J 1P1s if J 1 x1 1, then y1 must be 1.
Hence, the curve P-x goes through a maximum at x1 1. This is also an azeotrope (but a
trivial one).

5. Given

gE A12 x1 x 2  A13 x1 x 3  A14 x1 x 4  A23 x2 x3  A24 x 2 x 4  A34 x3 x 4

where


This may not be immediately obvious. But J1 x1 is the activity, and the activity of component 1 cannot reach unity for any x1 less
than one because the solution will split into two phases of lower activity. See Fig. 6-25 in the text.
Solutions Manual 85

x1 n1 / nT

x2 x2 / nT

#
with nT n1  n2  n3  n4 the total number of moles.
Because

F wn g I E
RT ln J 1 GH wn JKT
1 P,T ,n2 ,n3 ,n4

we find

RT ln J1 A12 x22  A13 x32  A14 x42  x2 x3 ( A12  A13  A23 )

 x2 x4 ( A12  A14  A24 )  x3 x4 ( A13  A14  A34 )

6. Calculate T-y giving pressure and for x = 0.1, 0.2, …, 0.9  bubble-point calculation.
We have to solve the equilibrium equations:

M i yi P J i xi Pis (1)

Because total pressure is low (below atmospheric) we assume vapor phase as ideal: M i | 1 .
The activity coefficients are obtained from the equation for GE given in the data. Using Eq.
(6-47) of the text we obtain

ln J 1 . x 22
21
(2)
ln J 2 . x12
21

As the pressure is fixed, temperature varies along with x1 (and y1) and is bounded by the
saturation temperatures of the two components. These can be easily obtained from the vapor-
pressure equations. They are given in the form,

B
ln P s A (3)
T C
from which we obtain the saturation temperature

B
Ts C (4)
A  ln P s
Solutions Manual 86

For Ps = 30 kPa, we obtain T1s 387.26 K for cyclohexanone (1) and T2s 415.59 K for
phenol (2).
To obtain the T-x1-y1 diagram we assign values for the liquid mole fraction x1. Total pres-
sure is

P J 1 x1P1s  J 2 x2 P2s

or

P
P1s (5)
P2s
J 1 x1  J 1 x1 s
P1

To start the calculation we make an initial estimate of the temperature:

T x1T1s  x 2T2s (6)

For example, let us fix x1 = 0.5: T 0.5 u 387.26  0.5 u 415.59 401.42 K
With this temperature we obtain P1s and P2s from Eq. (3), the pure-component vapor pres-
sure equations: P1s 47.243 kPa and P2s 17.918 kPa Because we fixed x1, Eqs. (2) give
J1 . u 0.52 )
exp(21 0.592 and J 2 0.592 .
Next we recalculate P1s 73.482 kPa from Eq. (5), which in turn gives a new temperature,
T = 416.34 K, from the pure cyclohexanone vapor pressure equation.
The sequence of calculations is now repeated for this new temperature (we assume here that
activity coefficients are independent of temperature), yielding:

P2s 30.786 kPa; P1s 71.426 kPa [from Eq.(5)]

T1s 415.34 K [from Eq.(6)]


x

P1s 71.552 kPa; T1s 415.40 K; P2s 29.798 kPa

After these values, the change in temperature is small and therefore additional iterations
leads to no significant further change in the remaining values.
We can now calculate the vapor phase mole fraction from

x1J 1P1s (0.5) u (0.592) u (71552


. )
y1 0.706
P 30
The whole process is repeated for a new liquid mole fraction.
The following figure shows the computed T-x1-y1 diagram for this system at 30 kPa.
Solutions Manual 87

430

V
420
Temperature/K

410

L+V

400

L
390

380
0.0 0.2 0.4 0.6 0.8 1.0

Cyclohexanone Mole Fraction

Similarly, with the data calculated we can easily draw the corresponding y1-x1 diagram,
shown in the figure below.

1
Vapor Mole Fraction Cyclohexanone

0
0.0 0.2 0.4 0.6 0.8 1.0

Liquid Mole Fracton Cyclohexanone

As both figures show, this system has an azeotrope at T az | 421 K and for the composition
x1az | 0.3 .
Solutions Manual 88

7. Assume g E Ax1 x 2 , where A is a function of temperature. Then,

RT ln J 1f RT ln J 2f A

But

10 A
ln J 1f 0.15 
T  273 RT
Because

w( g E / T ) h E
wT T2

w( g E / T ) 10 Rx1 x 2 h E
wT (T  273) 2 T2

At x1 x2 0.5 and T = 333K,

hE ' mix h 641 J mol 1

8.
a) From the equation for H w we can obtain the infinite dilution partial molar enthalpy of water
in sulfuric acid solutions at 293 K and 1 bar as:

Hwf lim Hw lim Hw -41.44 kJ mol-1


xw o 0 x A o1

b) The mixing process is schematically shown below.

HA HW
A W

W
Cooling
coils
Solutions Manual 89

Taking the liquid in the vessel as the system, a first law balance gives for this flow process:

dU dQ  dW  H Adn A  H w dnw

where work W is dW  PdV , done on environment by the rising liquid level, under constant
pressure.
Then,

d (U  PV ) dH dQ  H Adn A  H w dnw

Integrating between initial state (empty vessel) and final state (full vessel), because
HA H A (T , P) of the pure acid and H w H w (T , P) of the pure water are constant, we obtain

H Q  n A H A  nw H w or Q H  n A H A  nw H w

But H n A H A  nw Hw , and the equation above becomes

Q n A (H A  H A )  nw (H w  H w ) n A 'H A  nw 'H w 'H (1)

where nA = 1 mol and nw = 2 mol in the final state.


In Eq. (1), the quantity (H w  H w ) is given by the equation given in the data, because the
reference state in that equation has been chosen to be pure water at system T and P:

134 x 2A
Hw  Hw  kJ mol -1 (2)
(1  0.7983x A ) 2

We need now to calculate the quantity (H A  H A ) , knowing (Hw  Hw ) . This can be done
by using the Gibbs-Duhem equation.
At constant T and P:

xw 1 xA
x Ad H A  x w d H w 0 Ÿ dHA  d Hw  d Hw (3)
xA xA

Differentiating Eq. (2), at constant T and P, we obtain:

2 x A (1  0.7983x A ) 2  2(0.7983) x 2A (1  0.7983x A )


d Hw (134 kJ mol -1 ) u dx A
(1  0.7983x A ) 4
(4)
268 x A
dx A (kJ mol 1 )
3
(1  0.7983x A )

Therefore, from Eqs. (3) and (4),

268(1  x A )
dH A dx A (kJ mol 1 ) (5)
(1  0.7983x A ) 3

Integrating Eq. (5) between composition xA and composition xA = 1 (pure acid) gives
Solutions Manual 90

H A  H A (x A 1) HA  HA
z1
xA
268(1  x A )
(1  0.7983x A ) 3
dx A

(6)
LM 335.71(x  0.1263) OP xA
74.51(1  x A ) 2
(kJ mol) 1
MN (1  0.7983x ) PQ
A
2
1
(1  0.7983x A ) 2

We can now calculate the heat load Q in Eq. (1). Setting n = nA + nw = 3 mol, and using
Eqs. (2) and (6) in (1),

ª 74.51x A (1  x A )2 134 x 2A (1  x A ) º
Q n«  »
«¬ (1  0.7983x A )2 (1  0.7983x A )2 »¼

ª 74.51x A (1  x A )(1  x A  1.7983x A ) º


n« » (7)
¬« (1  0.7983x A )2 ¼»

ª 74.51x A (1  x A ) º
n« » (kJ mol 1 )
¬ 1  0.7983x A ¼

Substitution of n = 3 mol and xA = 1/3 gives the desired heat load:


Q 39.23 kJ

Q is negative because heat is removed from the system.

9.
a) Yes, it’s possible. Slight positive deviations merely mean that the physical interaction be-
tween SO2 and C4H8 makes a larger contribution to the excess Gibbs energy than does the
chemical interaction.

b)
E E
gSO ! gSO
2  isobutene 2 n butene  2


because the tendency to complex (which tends to make gE negative) is stronger with n-butene-2.
Steric hindrance in isobutene is larger than in n-butene-2.
Solutions Manual 91

10. The suggested procedure is to integrate numerically a suitable form of the Gibbs-Duhem
equation.
At low pressures, we may write the Gibbs-Duhem equation:

FG IJ  x2 FG wJ 2 IJ
x1 wJ 1
H K T J 2 H wx 2 K T
J 1 wx1
0

By assuming ideal-gas behavior,

y1P P1
J1
x1P1s x1P1s

wJ 1 1 FG wP1 IJ  P1
wx1 x1P1s H wx1 K x12 P1s
Similarly,

wJ 2 1 FG wP2 IJ  P2
wx1

x2 P2s H wx2 K x22 P2s
Substituting we find

x1 wP1 x2 wP2
P1 wx1 P2 wx 2

Because P P1  P2 , dP dP1  dP2 , then

wP2 wP
[ 1x P ]
wx2 wx2
1 2 1
x1P2

In different form:

'P2 1 'P
'x 2 x 2 P1 'x 2
1
x1P2

For P-x data, we choose a small 'x2 (say 0.05) and integrate to find 'P2 and thus P2 . We
obtain P1 by difference: P1 P  P2 .
This method is described by Boissanas, quoted in Prigogine and Defay, Chemical Thermo-
dynamics, page 346. It gives good agreement with experimental partial-pressure data for this par-
ticular system.

11. For a binary system, the Wilson equation gives


Solutions Manual 92

ª /12 / 21 º
ln J1  ln( x1  x2 /12 )  x2 «  » (1)
¬ x1  /12 x2 x2  x1/ 21 ¼

ª /12 / 21 º
ln J 2  ln( x2  x1/ 21 )  x1 «  » (2)
¬ x1  x2 /12 x2  x1/ 21 ¼

At infinite dilution these equations become

ln J 1f  ln /12  (1  / 21 ) (3)

ln J 2f  ln / 21  (/12  1) (4)

For J 1f 12.0, J 2f 3.89, solve Eqs. (3) and (4) to find,

/ 21 0.6185

/ 12 0.1220

Assuming ideal-gas behavior and neglecting Poynting correction, we may write:

y1P x1J 1P1s (5)

y2 P x2 J 2 P2s (6)

P x1J 1P1s  x 2 J 2 P2s (7)

From Perry’s, the saturation pressures at 45qC are:

P1s 0.188 bar

P2s 0.0958 bar

To construct the P-x-y diagram:


1. Choose x1 (or x2)
2. Calculate y1 (or y2) from Eq. (5) and using Eqs. (1) and (2)
3. Calculate P from Eq. (7).

12. The solution procedure would be:

1. Find P1s and P2s at each T.


2. At this low pressure, assume ideal-gas behavior and neglect Poynting correc-
tion:
Solutions Manual 93

y1P x1J1P1s

y2 P x2 J 2 P2s

For J’s use the Wilson equation with two parameters: /12 and / 21 . Assume that
(O 11  O 12 ) and (O 22  O 12 ) are independent of temperature.
/12 and / 21 are, however, temperature-dependent as given by Eqs. (6-107) and (6-108).

3. Assume value of (O 11  O 12 ) and (O 22  O 12 ) and calculate the total pressure:

Pcalc x1J 1P1s  x2 J 2 P2s

4. Repeat; assuming new values. Keep repeating until Pcalc is very close to 0.5
bar for every point; that is until

n
¦ (Pcalc  P)2 is a minimum
i 1

where n is the number of data points.

13.
a) 2-Butanone:
H H H

H C C C C H

H O H H

Cyclohexane:
(CH2)6

6 groups CH2: R = 0.6744; Q = 0.540

Molecule Group Number R Q


2-Butanone CH3CO 1 1.6724 1.488
CH3 1 0.9011 0.848
CH2 1 0.6744 0.540
Cyclohexane CH2 6 0.6744 0.540
b) We use UNIFAC activity coefficient equations to calculate J1 and J2 for the equimolar mix-
ture at 75ºC (for a detailed example of a similar UNIFAC calculation see Chapter 8 of The Prop-
Solutions Manual 94

erties of Gases and Liquids by R.C. Reid, J.M. Prausnitz, B. E Poling (4th. Ed., McGraw-Hill,
1988).
We obtain:

ln J 1 ln J 1comb  ln J 1res 0.01228  0.2595 0.27238 Ÿ J1 1.31

ln J 2 ln J 2comb  ln J 2res 0.01415  0.3420 0.35615 Ÿ J2 1.43

c) Using UNIFAC we can calculate the activity coefficients as a function of composition at


75ºC.
Total pressure is calculated from

P x1J 1P1s  x 2 J 2 P2s

and the vapor-phase composition from

x1J 1P1s
y1
P
Using Antoine vapor pressure equations at 75ºC, we obtain for 2-butanone
P1s 0.8695 bar and for cyclohexane P2s 0.8651 bar .
The following figures show the calculated results in the form of P-x1-y1 and y1-x1 dia-
grams.

1.3
L
1.2

1.1
L+V L+V
Pressure/bar

1.0

V
0.9

0.8

0.7

0.6
0.0 0.2 0.4 0.6 0.8 1.0

2-Butanone Mole Fraction


Solutions Manual 95

Vapor Mole Fraction 2-Butanone

0
0.0 0.2 0.4 0.6 0.8 1.0

Liquid Mole Fracton 2-Butanone

The table below shows the calculated activity coefficients from UNIFAC, vapor composi-
tion and total pressure.

x1 J1 J2 y1 P/bar

0 4.38 1.00 0 0.8511


0.2 2.27 1.07 0.351 1.124
0.4 1.51 1.27 0.447 1.175
0.5 1.32 1.42 0.485 1.178
0.6 1.18 1.62 0.528 1.169
0.8 1.04 2.19 0.660 1.096
1.0 1.00 3.10 1.0 0.8695

Comparison of the calculated J’s in this table with those given in the data, indicate that the
latter are actually UNIFAC predictions and not experimental data. In the tables, at x1 = 0 and x1
=1 the activity coefficients listed are, respectively, J 1f and J 2f . UNIFAC predicts J 1f 4.38 ,
which compares well with the experimental ebulliometry data at 77.6ºC, J 1f 3.70.

14. The UNIQUAC equation is


gE E
gcombinatorial E
 gresidual
Solutions Manual 96

E
gcomb ) ) z T T
x1 ln 1  x2 ln 2  ( x1q1 ln 1  x2 q2 ln 2 )
RT x1 x2 2 )1 )2

E
gres
 x1q1 ln(T1  T2 W21 )  x2 q2 ln(T2  T1W12 )
RT

x1r1 x1q1
)1 T1
x1r1  x2 r2 x1q1  x2 q2

a a
W12 exp( 12 ) W21 exp( 21 )
T T

The condition for instability of a binary liquid mixture is

F w 2 ' mix g I  0
GH wx 2 JK (1)
P,T

where ' mix g is the molar change in Gibbs energy upon mixing, or

§ w 2 gE · § 1 1 ·
¨ ¸  RT ¨  ¸  0
¨ wx 2 ¸ © x1 x2 ¹
© 1 ¹ P,T

Incipient instability occurs at

§ w 2 ' mix g ·
¨ ¸ 0
¨ wx 2 ¸
© ¹ P,T

and

§ w 3' mix g ·
¨ ¸ 0
¨ wx 3 ¸
© ¹ P,T

Given x1 , x 2 and all parameters, we could determine if Eq. (1) is satisfied. However, the
procedure is long and tedious. It is easier to graph ' mix g over the composition range and to
look for inflection points.
For the data given, phase separation occurs at -40qC.
Solutions Manual 97

xn-Hexane
0 0.2 0.4 0.6 0.8
0

-100

-200

-300
' mix g (J mol-1)

-400

-500
223.15 K
233.15 K
243.15 K
253.15 K
263.15 K

15. If the two curves cross ' mix g / RT is zero because

' mix g ' mix h  T ' mix s

This is not possible, because 'mixg must always be negative for two liquids to be miscible.
Solutions Manual 98

16. To relate J if to Hi, j :

fi
Ji
xi fi0

then,

§ f · 1 § f · Hi, j
J if lim ¨ i ¸¸ lim ¨ i ¸
xi o0 ¨ x f 0 fi0
© xi ¹ fi0
© i i ¹

Assuming fi0 # Pis ,

Hi, j
J if
Pis

then,
H1,2 2
J 1f 1869
.
P1s 107
.

H 2,1 16.
J 2f 1203
.
P2s 133
.

Using the van Laar equations,

Ac
ln J1
2
§ Ac x1 ·
¨1  c ¸
© B x2 ¹

Bc
ln J 2
2
§ B c x2 ·
¨1  c ¸
© A x1 ¹

we get

ln J 1f Ac 0.625

ln J 2f Bc 0.185

To solve for vapor composition (assuming ideal vapor and neglecting Poynting corrections),

y1P x1J1P1s

y2 P x2 J 2 P2s

P y1P  y2 P
Solutions Manual 99

At x1 x2 0.5,

0.625
ln J1 0.033 Ÿ J1 1.033
2
§ 0.625 ·
¨ 1  0.185 ¸
© ¹

0.185
ln J 2 0.110 Ÿ J2 1.116
2
§ 0.185 ·
¨ 1  0.625 ¸
© ¹

Therefore,

y1P (0.5) u (1.033) u (1.07) 0.55 bar

y2 P (0.5) u (1.116) u (1.33) 0.74 bar

P 0.55  0.74 1.29 bar

y1 0.55 / 1.29 0.426

y2 0.574

17. To estimate the vapor-phase composition, assume ideal vapor:

yi P xi J i Pis (i=1, 2, 3)

Then,

P y1P  y2 P  y3 P

To find the activity coefficients, assume that g E / RT is given by a sum of Margules terms
[Eq. (6-149)]. Then,

ln J 1 c x 22  A13
A12 c x32  ( A12
c  A13
c  A23
c ) x2 x3 (1)

ln J 2 c x12  A23
A12 c x32  ( A12
c  A23
c  A13
c ) x1 x 3 (2)

ln J 3 c x12  A23
A13 c x 22  ( A13
c  A23
c  A12
c ) x1 x2 (3)

c , A13
We can find A12 c , A23
c from binary data.
Solutions Manual 100

From (1-2) binary:

A
ln J 1f ln(13
. ) 0.262
RT

c
A12 0.262 at 320K

§ 320 ·
c
At 300K, A12 (0.262) u ¨ ¸ 0.280 (assuming regular solution).
© 300 ¹

From (1-3) binary:

At azeotrope x1 y1 , x3 y3

y1P x1J1P1s

y3 P x3 J 3 P3s

J1 J3 P / P1s 1.126

A 3
ln J1 x2
RT

A
At x3 0.5, c
A13 0.475.
RT

From (2-3) binary:

A 2
ln J 2 x
RT 3
At incipient instability,

A
2
RT c
or
A
2T c
R

A § A ·§ Tc · § 270 ·
¨ ¸ ¨¨ ¸ (2) u ¨ ¸ 1.80
RT © RT c ¹© T ¸ © 300 ¹
¹

A23 1.80

With x1 x2 x3 , from Eqs. (1), (2) and (3):


Solutions Manual 101

ln J 1 0.0322 Ÿ J1 0.968

ln J 2 0.409 Ÿ J1 1506
.

ln J 3 0.475 Ÿ J1 1607
.

y1P (1 / 3) u (0.968) u (0.533) 0.172 bar

y2 P (1 / 3) u (1506
. ) u (0.400) 0.201 bar

y3 P (1 / 3) u (1607
. ) u (0.533) 0.286 bar

P = 0.172 + 0.201 + 0.286 = 0.659 bar


Then,

y1 0.261

y2 0.305

y3 0.434

18. Using the 3-suffix Margules equation,

g E / RT x1 x2 [ A  B( x1  x2 )]

we obtain

ln J 1 ( A  3B) x 22  4 Bx23

ln J 2 ( A  3B) x12  4 Bx13

At infinite dilution,

ln J 1f A B

ln J 2f A B

which gives
A = 1.89
Solutions Manual 102

B = 0.34
For instability to occur,

F w2g I FG 1  1 IJ  0
GH wx 2 JK
E

1 P,T
 RT
H x1 x2 K
Rewriting g E as

gE RT [( A  B) x1  ( A  3B) x12  2Bx13 ]

wg E
RT [( A  B)  2( A  3B) x1  6 Bx12 ]
wx1

w2g E
RT [2( A  3B)  12Bx1 ]
wx12

Thus, the condition for instability (at constant T) is:

ª 1 1 º
RT « 2 A  6 B  12 Bx1   »0
¬ x1 1  x1 ¼

Finding the zeros of the function in brackets,

x1 0.421 and x1 0.352 in the range 0 < x1 < 1.

Thus, instability at T is in the range


0.352 < x1 < 0.421

19.
a) At the azeotrope

FG wP IJ
H wx A K T 0

With g E / RT of the form g E / RT Ax A x B ,

ln J A Ax B2

ln J B Ax 2A
Solutions Manual 103

Assuming an ideal vapor phase,

yA P x A J A PAs

yB P x B J B PBs

P x A J A PAs  x B J B PBs

P x A exp( Ax B2 )PAs  x B exp( Ax 2A )PBs

§ wP ·
¨ ¸ PAs exp( Ax B2 )(1  2 x A x B A)  PBs exp( Ax 2A )(1  2 x B x A A) 0
© wx A ¹

PAs exp( Ax B2 ) PBs exp( Ax 2A )

Ax B2 ln( PBs / PAs )  Ax 2A

At 30qC,

PAs 0.235 bar; PBs 0.658 bar; A = 0.415

Then, x A 0.30.

At 50qC,

PAs 0.539 bar; PBs 0.658 bar; A = 0.415

Then x A 0.26.

At 70qC,

PAs 1.119 bar; PBs 1.367 bar; A = 0.330

Then x A 0.20.

b) Assuming ideal vapor,

JA y A P / x A PAs

JB y A P / x B PBs

At azeotrope, x A y A , xB yB . Then
Solutions Manual 104

JA P / PAs

JB P / PBs

Taking the ratio

JA / JB PBs / PAs

ln( J A / J B ) ln PBs  ln PAs

4050 4050
12.12   11.92 
T T

0.20

Because

ln J A Ax B2 and ln J B Ax 2A ,

ln( J A / J B ) A( x B2  x 2A ) 0.2

1
A because 0 < x A < 1
5  10 x A

If |A| > 0.2 there is an azeotrope.


The pure component boiling points are:

tbA 67qC

tbB 61qC

In the range 61qC < t < 67qC, A is always larger than 0.2. Therefore, the azeotrope exists.

c) The enthalpy of mixing equation cannot be totally consistent since the expression for g E is
quadratic in mole fraction and the expression for ' mix h is cubic. However, they may be close.
To check this, we use the Gibbs-Helmholtz equation:
wg E / RT h E
wT RT 2

gE
A(T ) x A x B
RT

wg E / RT wA
x A xB # x A x B (0.00425)
wT wT
Solutions Manual 105

Because h E ' mix h,

' mix h (1)


(323) u (0.00425) x A x B . x A xB
1373
RT
The other data indicate

' mix h (2)


(1020
.  0.112 x A ) x A x B
RT
Looking at selected values:

xA 'mixh(1)/RT 'mixh(2)/RT

0.1 0.123 0.093


0.2 0.220 0.167
0.3 0.288 0.221
0.4 0.330 0.256
0.5 0.343 0.269
0.6 0.330 0.261
0.7 0.288 0.231
0.8 0.220 0.178
0.9 0.123 0.101

The above shows the degree of inconsistency of the two sets of data.
S O L U T I O N S T O P R O B L E M S

C H A P T E R 7

1. Using regular-solution theory and data for A in CS2 we find the solubility parameter for A.
Then, we predict vapor-liquid equilibria for the A/toluene system.
Let B refer to toluene and C refer to CS2. From regular-solution theory,

RT ln J A b
v A ) 2C G A  G C g 2
(for A in CS2)

Further, assuming ideal vapor phase, we have

yA P PA xA J A PAs

8
JA 1203
.
(0.5) u (13.3)
or

ln J A 0.185

Then,
1/ 2
ª RT ln J A º 1
G A  GC r« »
¬ vA ¼ )C

with

vA 200 cm3 mol-1 and )C 0.234

GA  GC r 6.30 (J cm 3 )1/ 2

107
Solutions Manual 108

For liquid hydrocarbons, G is approximately 12-18 (J cm-3)1/2. Therefore, we take the


smaller value.
For A in toluene,

RT ln J A 2
vA)B b
GA  GB g 2

or
JA 118
.

RT ln J B b
v B) 2A G A  G B g 2

or
JB 137
.

For ideal vapor,

P PA  PB x A J A PAs  xB J B PBs 25.3 kPa

Hence,

yA 0.31

yB 0.69

2. Excess properties ( h E , s E ) are defined in reference to an ideal (in the sense of Raoult’s law)
mixture of pure components.
The partial molar quantities h E and s E are the contributions to these excess properties per
differential amount added to the solution.
The “pure” acetic acid is highly dimerized, so as the first bits go into solution thse dimers
must be broken up. This will require energy ( h1E ! 0 ) and will increase the entropy more than is
accounted for by the ideal mixing term ( s1E ! 0 ).
As x1 gets larger, some dimer will begin to exist in the solution, so these effects will dimin-
ish. Therefore, at small x1, the curves should look something like this:
Solutions Manual 109

h1 s1

0 0
0.5 x1 0.5 x1

3. The K factors for hexane (1) and benzene (2) (neglecting Poynting corrections and assuming
ideal-vapor phase) are:

J 1P1s
K1
P

J 2 P2s
K2
P
where

P x1J 1P1s  x 2 J 2 P2s

Using regular-solution theory,

RT ln J 1 b
v1) 22 G1  G 2 g 2

v ) bG G g
2 2
RT ln J 2 2 1 1 2

The volume fractions are

)1 0.389 )2 0.611

From the above equations,

J1 132
. J2 1.08

P = (0.3)u(1.32)u(0.533) + (0.7)u(1.08)u(0.380) = 0.498 bar

(1.32) u (0.533)
K1 KC6H14 1.41
0.498

(1.08) u (0.38)
K2 KC6H6 0.82
0.498
Solutions Manual 110

4. Ether and pentachloroethane hydrogen bond with each other (but not with themselves).
Then, g E  0 .

0 x 1
0

gE

5. The relative volatility of A and B is

( yA / x A )
D A,B
( yB / x B )

Assuming ideal vapor phase and neglecting Poynting corrections,

yA P x A J A PAs

yB P x B J B PBs

At the azeotrope, x A yA and

J A PAs
D A,B 1
J B PBs

From regular-solution theory,

RT ln J A 2
vA)B b
GA  GB g 2

RT ln J B v B) 2A bG A  GB g
2
Solutions Manual 111

Because v A v B , J A J B , then PAs PBs .


For the ternary mixture,

RT ln J i v i ( G i  G )2

G ¦ )i Gi
i

As ) A )B 0.2 and ) C 0.6 , G 17.2 (J cm 3 )1/ 2 .


Then

(100) u (14.3  17.2)2


ln J A Ÿ JA 1.40
(8.31451) u (300)

(100) u (16.4  17.2)2


ln J B Ÿ JB 1.03
(8.31451) u (300)

J A PAs JA 1.40
D A,B 1.36
J B PBs JB 1.03

6. Assuming ideal vapor phase and neglecting Poynting corrections,

y1P x1J 1P1s

y2 P x2 J 2 P2s

P x1J 1P1s  x 2 J 2 P2s (1)

Using regular-solution theory,

RT ln J1 v1) 22 (G1  G2 )2

RT ln J 2 v 2 )12 (G1  G2 )2

Because v1 v 2 , we can rewrite [Eqs. (7-25) and (7-26)]:

RT ln J1 v1x22 (G1  G2 )2

RT ln J 2 v 2 x12 (G1  G2 )2

As v1 v2 160 cm3 mol-1,


Solutions Manual 112

J1 exp(0.616x22 )

J2 exp(0.616 x12 )

Substitution in Eq. (1) gives

P 0.533exp(0.616 x22 )  0.800 exp(0.616 x12 )

At the azeotrope,

FG wP IJ
H wx1 K T 0

Thus, after differentiation,

0.616 x 22 0.40547  0.16 x12

Solving for x1,

x1 0.171

7. Neglecting vapor-phase non-idealities and Poynting corrections, the total pressure, P, is

P x A J A PAs  x B J B PBs

Because the two fluids are similar in size, simple and nonpolar, we can assume that J’s are
given by two-suffix Margules equations:

A 2
ln J A x
RT B

A 2
ln J B x
RT A

As xA xB 0.5,

LM FG A IJ OP u b0.427  0.493g
N H 4RT K Q
P 0.667 0.5 u exp

A = 4696 J mol-1
From Eq. (6-144) of the text,

A
Tc
2R
Solutions Manual 113

Tc 282 K

If one considers the effect of non-randomness (based on the quasichemical approximation),


Eq. (7-110) gives

A
Tc 253 K
2.23R

A w
(assuming that constant and that the temperature dependence is given by ln J propor-
2R 2k
tional to 1/T).
Thus, random mixing predicts a value higher than that given by quasichemical theory. The
observed consolute temperature is likely to be lower than both.

8.
V
yi = ?
Let:

1 = benzene
z1= z2= 0.5
2 = n-butane

xi = ?
L

There are three unknowns: x1 , y1, and V / F .


To solve for them, we use two equilibrium equations and one mass balance.
Assuming ideal vapor and neglecting Poynting corrections:

y1P x1J 1P1s

(1  y1 )P (1  x1 )J 2 P2s

FG V IJ y1  FG1  V IJ x1
z1
H FK H FK
Using regular-solution theory,

RT ln Ji vi )2j (Gi  G j )2
Solutions Manual 114

with v1 = 92 cm3 mol-1 and v 2 = 106 cm3 mol-1.


Then,

J1 exp(0.828) 22 )

J2 exp(0.950)12 )

Substitution gives

y1 0.368x1 exp(0.828) 22 ) (1)

(1  y1 ) 4.76(1  x1 ) exp(0.950)12 ) (2)

FG V IJ y1  FG1  V IJ x1
0.5
H FK H FK (3)

) 1 and ) 2 are related to x1 and x2 by Eqs. (7-25) and (7-26).


To solve Eqs. (1), (2), and (3) for x1, y1, and V / F, assume first that J i 1. This gives,

V
x1 0.856 y1 0.315 0.658
F
A second approximation ( J i z 1 ) gives

x1 xC6H6 0.94

y1 0.35

V
0.741
F

9. As derived in Sec. 7.2 of the text, the regular-solution equations can be written in the van
Laar form

Ax1 x 2
gE (1)
A
x1  x 2
B

where parameters A and B are related to pure-component liquid molar volume and solubility pa-
rameters as follows [Eqs. (7-38) and (7-39)]:

A v A (G A  G B ) 2
(2)
B v B (G A  G B ) 2
Solutions Manual 115

Substituting the given liquid molar volumes and solubility parameters, we obtain

A (120 cm3 mol 1 ) u [(18  12) 2 J cm 3 ] 4320 J mol 1


(3)
B (180 cm3 mol 1 ) u [(18  12) 2 J cm 3 ] 6480 J mol 1

As discussed in Section 6-12, the temperature and composition at the consolute point are
found from solving:

FG w ln aA IJ FG w2 ln aA IJ
H w xA K T ,P H w xA2 K T ,P 0 (4)

Upon substitution of Eq. (1) into Eq. (4), the results are given in Eq. (6-146) in the text:

c [( A / B) 2  1  ( A / B)]1/ 2  ( A / B)
xA
1  ( A / B)
(5)
c c
2 xA (1  x A )( A2 / B)
Tc
c c 3
R[( A / B) xA  (1  x A )]

where superscript c denotes consolute.


Substituting Eq. (3) into Eq. (5), we finally obtain
c
xA 0.646

Tc 328 K

10. For each phase we choose the standard-state fugacity for cyclohexane as its pure subcooled
liquid at 25ºC. The equation of equilibrium is

x 3(1) J (31) x 3(2) J (32) (1)

where subscript 3 denotes cyclohexane and superscripts (1) and (2) denote, respectively, carbon
disulfide phase and perfluoro-n-heptane phase.
Rearrangement of Eq. (1) gives

x (1) J (32)
K{ 3 (2)
x 3(2) J (31)
Solutions Manual 116

Because phase (1) contains only carbon disulfide and a trace amount of cyclohexane,
whereas phase (2) contains only perfluoro-n-heptane and a trace amount of cylohexane, J (31) and
J (32) are essentially activity coefficients at the infinite-dilution limit of cyclohexane.
Hence, we can write

x3(1) LM J (32) OP f
K{
x 3(2) MN J (31) PQ (3)

where superscript f denotes the infinite-dilution limit of cyclohexane.


f f
From the regular-solution theory [Eq. (7-37)], J (31) and J (32) are given by:

f v3
ln ª J (1) º (G1  G3 )2
¬ 3 ¼ RT
(4)
f v3
ln ª J (2) º ( G 2  G 3 )2
¬ 3 ¼ RT
Substituting the pure-component liquid molar volumes and solubility parameters, we obtain

f (109 cm3 mol-1 )


ln ª J (1) º u ª(20.5  16.8)2 J cm-3 º 0.602
¬ 3 ¼ (8.314 J mol-1 K-1 ) u (298 K) ¬ ¼
(10)
f 3 -1
(109 cm mol )
ln ª J (2) º u ª(12.3  16.8)2 J cm-3 º 0.891
¬ 3 ¼ (8.314 J mol-1 K-1 ) u (298 K) ¬ ¼

Substituting Eq. (5) into Eq. (3), we have


K 1.34

11. From the definition of the solubility parameter, G,

'ucomplete vaporization
G2
vL

[Complete vaporization means going from saturated liquid to ideal gas at constant T.]
Then,
Solutions Manual 117

F h0  h L I RT  RT (1  z L )
GH RTc JK c
G2
z L RT
Ps

LM h0  h L OP
G2
MM RTL c  1  1L PPPRs
Pc
MM z TR z
PP
N Q
Because

h0  h L
f (TR , Z )
RTc

and

zV f (TR , Z) zL f (TR , Z) PRs f (TR , Z)

for TR d 1,
G2
Pc
b g b g b
f (0) TR  Z f (1) TR  higher terms g
[Reference: Lyckman et al., 1965, Chem. Eng. Sci., 20: 703].

12.
a) Pure methanol is hydrogen-bonded to dimers, trimers, etc. In dilute solution (in iso-octante),
methanol is a monomer. For an order-of-magnitude estimate, we can assume that, to make a
monomer, approximately one hydrogen bond must be broken. Thus h E = 12 kJ mol-1.

b) From solubility parameters we get (roughly) an endothermic heat of 263 J mol-1. The molar
specific heat is (roughly) 125 J K-1 mol-1 . Thus, 't | 2D C.

c) We want a Lewis acid that can hook on to the double bond in hexene.

Good Solvents are:


Dimethyl sulfoxide U|
Sulfur Dioxide V| Strong Lewis acids

Acetonitrile
W
Solutions Manual 118

Poor Solvents are:


Ammonia UV
W
Weak Lewis acids
Aniline

It is also important that the solvent should not be something that prefers self-interaction
than that with hexene molecules. Strongly hydrogen-bonded liquids (e.g. water and most alco-
hols) would therefore be poor solvents.

13. Let (HA) be the acid. In ionized form,

HA H+ + A-

Hexane

(HA)
H

(HA) A- + H+
W

Water

Equilibrium constants are defined as

bHAg
bHAg
W
K1
H

(H )(H ) (A )2


K2
(HA)W (HA)W

In hexane: CH = (HA)H

In water: CW = (HA)W + (A-) = (HA)W + K 2 HAb g W

b g
K1 HA H  K1K 2 HA Hb g
CW K1C H  K1K 2C H

Thus,
Solutions Manual 119

CH 1 FG1  K CH IJ
CW K1K2 H 1
CW K
with
1
a
K1K2
and
K1 b

14.

Benzene

AB 1/3 AT

AW

Water

Assume constant distribution coefficient and “reaction” equilibrium.

AB
K1
AW

AT
b AB g3
K2

In water:

CW AW

In benzene:

CB AB  3 AT

CB K1C W  3K 2 K13C W
3

CB
K1  3K 2 K13C W
2
CW
Solutions Manual 120

CB 2 .
Thus, plot as a function of CW
CW
Slope is 3K 2 K13 and intercept is K1.

15.
a) Pure CH2Cl2 and acetone do not hydrogen bond themselves but some hydrogen bonding is
likely to occur between dissimilar pairs, which explains the negative deviations from Raoult’s
law observed for this system.
Pure methanol is highly hydrogen bonded. However, in dilute solutions of CH2Cl2, metha-
nol exists primarily as monomer. Hydrogen bonding between methanol and CH2Cl2 is likely to
be weak. (Note that at infinite dilution, activity coefficient J 1f indicates the effect on a molecule
1 when surrounded by molecules of the other component).

b) Nitroethane has a large dipole moment. Both n-hexane and benzene are non-polar but, due
to S electrons, benzene is more polarizable. Therefore, we expect nitroethane/benzene interac-
tions to be stronger than those for nitroethane/n-hexane. Thus Jnitroethane in n-hexane is larger than
that in benzene.

c) Both CHCl3 and methanol are polar and slightly acidic. Although methanol has a slightly
higher dipole moment, CHCl3 is likely to solvate the coal tar more easily because methanol tends
to form strong hydrogen bonds with itself.
S O L U T I O N S T O P R O B L E M S

C H A P T E R 8

1. Given

FG1  1IJ  F
ln *1f
H rK
For r >>1,

ln *1f 1 F

If F = 0.44,

ln *1f 1.44 Ÿ *1f 4.22

As defined,

a1 a1
*1f 4.22 Ÿ a1 4.22 u 10 4
)1 10 4
Because
P1
a1
P1s

P1 (4.22 u 10 4 ) u (4.49) 0.0019 bar

For a non-volatile polymer, P2 0. Therefore,

P # P1 0.0019 bar

121
Solutions Manual 122

2. We can use the data for the Henry’s-law region to evaluate the Flory interaction parameter, F,
and then predict results at higher concentration.

Let:
1 = solvent
2 = polymer
wi = weight fraction
) i = volume fraction

Then,
w1
v1 U1
)1
v1  v 2 w1 w2

U1 U2

In the Henry’s law region,

w1 o 0, w2 o 1

a1 f1 w1H1,2
)1 )1 f10 )1 f10

But, as w1 o 0,

)1 w1U 2 / U1

)2 1

Then,
a1 U1H1,2
)1 U2 f10

If f10 P1s ,

a1 f1 U1H1,2 (0.783) u (18.3)


2.94
)1 )1 f10 U2 P1s . ) u (3340 / 760)
(111

From Flory-Huggins theory (r is large),

a
ln 1 ) 2  F) 22
)1

If )1 o 1 ,

ln(2.94) 1  F Ÿ F 0.078
Solutions Manual 123

At higher concentrations (w1 = 0.5), assume F z F(w1 ) :

(0.5) u (1/1.11)
)2 0.414 ()1 0.586)
(0.5) u (1/1.11)  (0.5) u (1/ 0.783)

Then,

a
ln 1 0.414  0.078 u (0.414) 2
)1

a1
153
. or a1 0.898
)1

P | f1 a1 f10 | a1P1s (0.898) u (3340) 3000 torr

P 3000 torr | 3.9 bar

3.
a) The generalized van der Waals partition is given by [Eq. (8-39)]
N N
1 § Vf · ª Eo · º
Q T , V , N ¨ ¸ >qext (V )@N >qint (T )@N «exp §¨  ¸» (1)
N ! ©¨ / 3 ¹¸ ¬ © 2kT ¹ ¼

Following the discussion on pages 442 and 443 of the textbook, we further have
rc
Vf § Vf ·
qext ¨¨ 3 ¸¸
/3 ©/ ¹

Eo rsK
2 2v
(2)
3
ª§ v ·1/ 3 º
Vf Wr v* «¨ ¸  1»
«© v* ¹ »
¬ ¼

V
v
Nr

Substituting Eq. (2) into Eq. (1) yields


Solutions Manual 124

ln Q  ln N ! Nrc ln / 3  Nrc ln(W r v* )


(3)
ª§ v ·1/ 3 º NrsK
 3Nrc ln «¨ ¸  1»  N ln qint 
«© v ¹
* » 2vkT
¬ ¼

Because the first, second, third and fifth terms on the right-hand side of Eq. (3) are only
functions of temperature, the equation of state is given by

P § w ln Q · 1 § w ln Q ·
¨ wV ¸ ¨ ¸
kT © ¹ N ,T Nr © w v ¹ N ,T
(4)

1 ª 3Nrc § 1 v 2 / 3 · NrsK § 1 ·º
« ¨ ¸ ¨ ¸»
Nr « (v 1/ 3  1) ¨ 3 v*1/ 3 ¸ 2kT © v 2 ¹ »¼
¬ © ¹

We can rewrite Eq. (4) as

P v v 1/ 3 1
 (5)
T v 1/ 3 1 
T v

where the reduced properties are defined by

T 2v*ckT
T
T* sK

2
P 2v * P (6)
P
* sK
P

v
v
v*

Equation (5) is the Flory equation of state [Eq. (8-45) of the textbook].

b) The configurational partition function [Eqs. (8-82) and (8-83]

§ E ·
Q QC exp ¨  ¸
© kT ¹
(7)
(V / v* ) N 1
QC (constant) N
N 0 ! N ! (V / v* ) N (r 1)

where

V
N0  r N
v*
Solutions Manual 125

z ª rN º
E =  H Nr « » (8)
2 «¬ (V / v* ) »¼

Combining Eqs. (7) and (8) gives

V §V · V
ln Q ln(constant)  ln ¨ ¸ 
v* © v * ¹ v*

§V ·
 ln N 0 !  ln N !  N (r  1) ln ¨ ¸ (9)
© v* ¹

§V ·
 H* (v*r N )2 ¨ ¸ V 2
© v* ¹

where

z H
H*
2 kT
(10)
ª§ V · º
ln N 0 ! ln «¨ ¸  r N » !
¬© v* ¹ ¼

ª§ V · º ª§ V · º ª§ V · º
«¨ * ¸  r N » ln «¨ * ¸  r N »  «¨ * ¸  r N »
¬© v ¹ ¼ ¬© v ¹ ¼ ¬© v ¹ ¼

The equation of state is thus given by

P § w ln Q ·
kT ¨ wV ¸
© ¹ N ,T

1 § V · § V · v* (1/ v* ) 1
ln ¨ ¸  ¨ ¸ 
v* © v * ¹ © v* ¹ V v*
(11)
1 §V · §V · 1 1/ v*
 ln ¨ rN ¸¨ rN ¸ 
v* ©v ¹ ©v V
rN v
* * ¹ *
v *

1/ v* § V 2 ·
 N (r  1)  H* (v*r N )2 ¨  ¸
V / v* ¨ v* ¸
© ¹

We can rewrite
Solutions Manual 126

§ PV · § 1 · P v
¨ kT ¸ ¨ r N ¸ T
© ¹© ¹
§ V ·
§V · §V · ¨ ¸
v ln ¨ ¸  v ln ¨  r N ¸  (v  1) ¨ v *
¸
© v* ¹ © v* ¹ ¨ V rN ¸
¨ * ¸
©v ¹ (12)
1 * rN
 v  1   H
r V / v*

1 ª § 1 ·º 1
 1  v ln ¨ 1  ¸ » 
r «¬ © v ¹ ¼ T v
where
T T
T
T* z H / 2k

P P
P (13)
P* z H / 2v*

v
v =
v*

Equation (12) is the Sanchez-Lacombe lattice-fluid equation of state [Eq. (8-84) of the text-
book].

4. The Flory-Huggins equation for the activity of the solvent [Eq. (8-11)] is

§ 1· 2
ln a1 ln(1  )*2 )  ¨ 1  ¸ )*2  F)*2 (1)
© r ¹

Conditions for incipient instability give [analogous to Eqs. (6-141) and (6-142)]:

§ w ln a1 ·
¨ ¸ 0
¨ w)* ¸
© 1 ¹ P,T

§ 2 ·
¨ w ln a1 ¸ 0
¨¨ *2 ¸¸
© w)1 ¹ P,T
Equivalently, we have
Solutions Manual 127

§ w ln a1 ·
¨ ¸ 0
¨ w )* ¸
© 2 ¹ P,T
(2)
§ 2 ·
¨ w ln a1 ¸ 0
¨¨ * 2 ¸¸
© w) 2 ¹ P,T

because

)1*  )*2 1

Substituting Eq. (1) into Eq. (2), we obtain

1 § 1· c
 ¨ 1  ¸  2Fc )*2 0
c © r¹
1  )*2

2
§ ·
¨ 1 ¸  2F c 0
¨¨ *c ¸¸
© 1  )2 ¹

where superscript c stands for critical.


Hence, we obtain
c 1
)*2
1  r1/ 2

2
1§ 1 ·
Fc ¨ 1  1/ 2 ¸
2© r ¹

5. The Flory-Huggins equation for the activity coefficient of HMDS (1) [Eq. (8-12)] with F 0
is

ª § 1· º § 1·
ln J1 ln «1  ¨ 1  ¸ )*2 »  ¨ 1  ¸ )*2
¬ © r¹ ¼ © r¹

Using data in Table 8-5 of the text, calculated molecular characteristic volumes V * , ratios
of molecular segments r and activity coefficients of HMDS (at )*2 0.8 ) are given in the fol-
lowing table:
Solutions Manual 128

Substance V* v*sp Mn r V2* / V1* ln J1 ()*2 0.8)


(cm3 mol-1)
HMDS 162.30  
PDMS 3 521.53 3.21 -0.249
PDMS 10 832.89 5.13 -0.389
PDMS 20 1313.8 8.09 -0.528
PDMS 100 3507.8 21.61 -0.677
PDMS 350 5530.1 34.07 -0.722
PDMS 1000 6604.8 40.70 -0.735
PDMS f f f -0.809

As we increase the molecular weight of PDMS, ln J1 becomes more negative. J1 is smaller


than unity and increasingly deviates from unity as the molecular weight of PDMS is increased.
This example illustrates the effect of differences in molecular sizes of HMDS and various
PDMS with F 0 (Fig. 8-3).

6. The flux of gas i through the membrane is given by, Eq. (8-118)

Di G
Ji
GM
G G
SiF PiF  SiP PiP (1)

Because solubility coefficients for both O2 and N2 in the feed and permeate are assumed to
be equal and the permeate pressure is vacuum, Eq. (1) reduces to

Di SiG
Ji PiF
GM
(2)
Di SiG
yi PF
GM

where yi and PF ( PF 2 u 105 Pa) denote, respectively, the mole fraction of component i and the
total pressure of the feed.
For the feed mixture (air) we have

yO2 # 0.21
(3)
yN2 # 0.79

Substituting Eq. (3) and the given data for membrane thickness, solubility and diffusion
coefficients into Eq. (2), the corresponding fluxes of O2 and N2 are
Solutions Manual 129

J O2 0.148 u 10 5 m3 m-2 s-1

J N2 0.197 u 10 5 m3 m-2 s-1

The separation factor is defined by [Eq. (8-121)]


G
DO2 SO
2
D O2 / N 2
G
DN2 SN
2

2.79

Although DO2 / N2 ! 1 , the net flux of N2 is larger than that of O2 due to the difference in
partial pressures in the feed.

7. If the feed pressure were low, we could use Eq. (8-112) to calculate J1 , the flux of carbon diox-
ide, and J2 , the flux of methane. Equation (8-113) then gives the composition (y) of the perme-
ate.
However, because the pressure of the feed is high, we must allow for the effect of pressure
on nonideality of the gas phase.
Equation (8-111) is

D1 M
J1
GM

c1F  c1MP (1)

where c1M is the concentration of carbon dioxide in the membrane; subscripts F and P refer to
feed and permeate.
To find c1MF , we use the equilibrium relation

ª M § v1P · º
( y1PM1 )F « H1c1 exp ¨ ¸» (2)
¬ © RT ¹ ¼ F

where H1 and v1 are Henry’s constant and partial molar volume for carbon dioxide in the mem-
brane, both at 300 K and 100 bar.
Fugacity coefficient M1 is given by the virial equation of state, truncated after the second
virial coefficient [Eq. (5-33)]:
Solutions Manual 130

§ 2 · P
ln M1 ¨2
¨ ¦yi B1i  Bmixt ¸
¸ RT
© i 1 ¹
(3)
2 2
Bmixt ¦¦ yi y j Bij
i 1j 1

Substituting the given temperature, pressure and second virial coefficients into Eq. (3), we
obtain

M1 0.729

Substituting Eq. (4) and all other given data into Eq. (2) yields

c1MF 0.314 mol L-1 (4)

To find c1MP we use the equilibrium relation

( y1PM1 )P ( H1c1M )P (5)

where PP 1 bar and M1P 1.


Hence, Eq. (5) reduces to
y1P
c1MP mol L1 (6)
19

The quantity y1P is an unknown in this problem.


Substituting Eqs. (4) and (6) into Eq. (1) gives

§ mol · 5 u 106 § y ·
J1 ¨ ¸ ¨ 0.314 u 103  1P u 10 3 ¸ (7)
¨ cm2 s ¸ 0.1 © 19 ¹
© ¹

Applying the same procedure for methane (2), we obtain

§ mol · 50 u 10 6 § y ·
J2 ¨ ¸ ¨ 0.155 u 10 3  2 P u 10 3 ¸ (8)
¨ cm2 s ¸ 0.1 © 50 ¹
© ¹

The (steady-state) mole fraction of carbon dioxide in the permeate is given by [Eq. (8-113)]
J1
y1P (9)
J1  J2

Further, the mass conservation gives

y1P  y2 P 1 (10)

Substituting Eqs. (7), (8) and (10) into Eq. (9) gives y1 P = 0.168 for carbon dioxide in the
permeate.
Therefore, for methane in the permeate, y2 P = 0.832.
Solutions Manual 131

8.
a) Flux of water through the membrane is given by Eq. (8-128):

§ permeability · ­ L L ª v w ( PF  PP ) º ½
Jw ¨ thickness ¸ u ® xwF  xwP exp «  RT »¾ (1)
© ¹ ¯ ¬ ¼¿

L
where xwP 1 (pure water in the permeate); PP Pws 0.0312 atm.
To calculate concentration of water in the feed, we use

( Pw )F ( xwL Pws )F (2)

with

Pws 0.0312 atm

PwF (1  0.0184) u (0.0312) atm

Therefore, we obtain
L
xwF 0.9816

Because the permeate is pure water, we obtain

18.015 g mol 1
vw | vw 18.069 cm 3 mol 1
0.997 g cm 3

The feed pressure is thus given by

§ 2.6 u 10 5 g cm cm 2 s1 ·
7.2 u 10 4 g cm 2 s1 ¨ ¸
¨ 10 u 10 4 cm ¸
© ¹

°­ ª (18.069 cm3 mol 1 ) u ( PF  0.0312 atm) º °½


u ®0.9816  (1) u exp «  »¾
1 1
°¯ ¬« (82.06 atm L K mol ) u (298.15 K) ¼» ¿°

Therefore, the feed pressure is

PF = 63.93 atm

b)

1u 106 gallons/day 3785.4 m3 /day 0.0438 m3 /s 43.67 kg/s


Solutions Manual 132

43.67 u 103 g s1


Flux 7.2 u 10 4 g cm 2 s1 (3)
( A cm 2 )

where A is the membrane area needed.


Solving Eq. (3) for the area, we get
A 6.07 u 10 7 cm 2 6.07 u 10 3 m 2 6.5 u 10 4 ft 2
S O L U T I O N S T O P R O B L E M S

C H A P T E R 9

1. The solubility product is the equilibrium constant for the reaction

AgCl Ag+ + Cl-


Solid Aqueous solution
defined as

K SP (a  )(a  )
Ag Cl

being the standard states the pure solid AgCl and the ideal dilute 1-molal aqueous solution for
each ion.

mol AgCl
a) Let S be the solubility of AgCl in pure water, in (a molality).
kg water

a  J S a  J S
Ag Ag Cl Cl

K SP (a  )(a  ) (J )(J  ) S 2 (J r S ) 2 (1)


Ag Cl Ag Cl

Because the solution is very dilute, J r | 1 and S | K SP and is of the order of 10-5 molal
. u 10 5 molal. Therefore we may apply
and therefore the ionic strength is also very low: I | 131
the Debye-Hückel limiting law.

1
I [m u (1) 2  m u (1) 2 ] m S
2
Using Eq. (9-50a) of the text,

133
Solutions Manual 134

log J (rm) 0.510 u (1) u (1) I 1/ 2 0.510 S 1/ 2 (2)

We now replace J r given by Eq. (2) into Eq. (1) and solve for the solubility S:
1/ 2
[(10 0.510 S )S ]2 . u 10 10
172

S = 1.31 ×10-5 mol kg-1 ( J r | 10


. )

b) With the addition of NaCl the ionic strength increases and we need to evaluate J r because
now the solution is not very dilute and therefore we don't have J r | 10
. .
Let S be the new solubility of AgCl in this aqueous solution that contains NaCl. The mola-
lities are

m  S m  S  0.01
Ag Cl

The total ionic strength (due almost exclusively to NaCl because S is small) is

1
I [S u (1) 2  S u (1) 2  0.01 u (1) 2  0.01 u (1) 2 ] | 0.01 mol kg 1
2

We use in this case the extended limiting law [Eq. (9-52)] with AJ 1174
. kg1/2 mol 1/2 :

1174
. u (0.01)1/ 2
ln J r  Ÿ Jr 0.90
1  (0.01)1/ 2

As in a),

K SP (a )(aCl  ) (J )( J Cl  )(m )(mCl )


Ag Ag Ag

( J r )2 (S )(S  0.01) 1.72 u 10 10

or substituting J r 0.90 ,

S (S  0.01) . u 10 10
212

Because S is small and S << 0.01, we obtain S | 2.12 ×10-8 mol kg-1.
The addition of NaCl reduces the AgCl solubility from 1.31u10-5 mol kg-1 [as calculated in
a) for pure water] to 2.12u10-8 mol kg-1 (in a 0.01 molal NaCl aqueous solution). This is the
common ion (“salting out”) effect.

c) Similarly, let S be the new AgCl solubility.

Molalities are:
m S m  S
Ag Cl

Again, the ionic strength is almost exclusively due to NaNO3:


Solutions Manual 135

I | 0.01 mol kg 1 and Jr 0.90

K SP (a )(aCl  ) (J )( J Cl  )(m )(mCl )


Ag Ag Ag

( J r )2 ( S )2 1.72 u 10 10

S = 1.46 ×10-5 mol kg-1

Compared to the solubility of AgCl in pure water, the solubility of AgCl in a 0.01 molal
NaNO3 aqueous solution increases by roughly 10%, because the higher ionic strength reduces
the activity of Ag+ and Cl- ions and causes more AgCl to dissolve (“salting in” effect).

2. The solubility product is the equilibrium constant for the reaction (PbI2 is a 1-2 electrolyte)

PbI2 Pb2+ + 2I-


Solid Aqueous solution
defined as

K SP (a 2 )(a  ) 2 (J r ) 3 S 3
Pb I

being the standard states the pure solid PbI2 and the ideal dilute 1-molal aqueous solution for
mol PbI 2
each ion. In the above equation, S is the solubility of PbI2 in pure water, in (a mo-
kg water
lality). Because the solution is very dilute, J r | 1 and

K SP (S ) 3 (1.66 u 10 3 ) 3 4.57 u 10 9

For the solution with KI the ionic strength increases and we need to evaluate J r because
now the solution is not very dilute and therefore we may not have J r | 10
. .
Let S be the new solubility of PbI2 in this aqueous solution that contains KI. The molalities
are

m 2 S m 2S  0.01 m  0.01
Pb I K

The total ionic strength is

1
I [2S u (1) 2  S u (2) 2  0.01 u (1) 2  0.01 u (1) 2 ] (3S  0.01) mol kg 1 (1)
2

We use in this case the extended limiting law [Eq. (9-52)] with AJ 1174
. kg1/2 mol 1/2 :
Solutions Manual 136

1174
. u (3S  0.01)1/ 2
ln J r  (2)
1  (3S  0.01)1/ 2

Simultaneous solution of Eqs. (1) and (2) gives

Jr 0.88

S 1.89 u 10 3 mol kg 1

3. The solubility product for PbI2 in an aqueous solution is

log K SP 3log mr  3log J r (1)

where

m mPb2  m

m mI 2m (2)

1/ 3
mr ª m(2m)2 º 1.587m
¬ ¼

1 2
With I (2 m  2m) 3m Eq. (9-50a) gives
2

log J r 0.510 u (2) u (3m)1/ 2 (3)

With m 1.66 u 10 3 mol kg-1 we obtain

log K SP 7.953 (4)

For the solutions containing sodium chloride or potassium iodide saturated with PbI2, we
have

log K SP 7.953 3 u log mr  0.510 u (2) u I 1/ 2 (5)

For the NaCl solution, we write


Solutions Manual 137

m Pb2 m m I 2m mNa m Cl 0 .0 1

1/3
mr ª m u (2 m )2 º 1 .5 8 7 m
¬ ¼

1 ª 2
I 2 m  2 m  2 u ( 0 . 0 1) º 3 m  0 .0 1
2 ¬ ¼

Substitution in Eq. (1) yields

m 1.89 mol kg -1

For the KI solution, we write

m Pb 2  m m I 2 m  0.0 1 mK  0.0 1

1/ 3
mr ª m u (2 m  0.0 1) 2 º
¬ ¼

1ª 2
I (2 ) u m  2 m  2 u ( 0.0 1) º 3 m  0 .0 1
2¬ ¼

Substitution in Eq. (1) yields

m 0.24 mol kg  1

For both systems, calculated and experimental values are in good agreement. The large de-
crease in PbI2 solubility in the KI solution follows because all iodide ions are included in mr .
This reduction in solubility is called the common-ion effect.

4. The dissociation of acetic acid is represented by

CH3COOH H+ + CH3COOH-

or schematically

AH H+ + A-

The dissociation constant is


Solutions Manual 138

(aH )(aH )
K 1.758 u 10 5
aAH

(m) Q (mH )(mA ) ( J (m) 2


(mQ  )(mQ  ) ( J r ) r )
(m)
mAH J (m)
AH
(m)
mAH J (m)
AH

Designating by D the extent of ionization, and by m the stoichiometric molality of acetic


acid, in dilute solutions we may assume a  a  D and aAH m  D. Further, the activity
H A
of undissociated acetic acid approaches its molality at infinite dilution.
Assuming that the activity coefficients are unity (very dilute solutions) we have

(D )(D ) D2
K 1758
. u 10 5
mD mD
For a m = 10-3 molal aqueous solution, the equation above gives D 1.24 u 10 4 .
The fraction of acetic acid ionized is:

D . u 10 4
124
0.124
m 1 u 10 3

5. In SI units, the Debye length is defined by Eq. (9-47) of the text:

F RTH o H I 1/2
N 1 GH 2d N 2 e2 I JK
s
r

where ds is the density of the solvent in kg m-3. For water at 25ºC, ds = 997 kg m-3.
The ionic strength is I | 0.001 mol kg-1 for the 0.001 M solution and I | 0. 1 mol kg-1 for
the 0.1 M solution.
Substitution of values gives the values for the Debye length N-1 (in nm) presented in the
following table:

Solution Water Methanol


0.001 M 9.6 6.1
0.1 M 0.96 0.61

We see that N-1 decreases ten times with a hundredfold increase in concentration. For the so-
lutions at higher concentrations, shielding effects are more important and N-1 is low. Further, the
Debye length increases with increasing dielectric constant: when Hr is large (as in water), the
ionic atmosphere is weak and the coulombic interactions are strongly reduced.
Solutions Manual 139

6.
a) The molality of NaCl in seawater is

§ 3.5 ·
¨ ¸ mol NaCl
mNaCl © 58.5 ¹
(100  3.5) g water

§ 3.5 ·
¨ 58.5 ¸
© ¹ u 1000 mol
(100  3.5) kg

0.620 mol/kg water

The ionic strength is given by

1
I (m z2  m  z2 ) m 0.620 mol / kg water
2 Na  Na  Cl Cl 

which is a relatively high value.


Also,

Q Q Na  QCl 2

The molar volume of water at 25ºC is

Mw 18 g / mol
vw 18.05 cm 3 mol -1
dw 0.997 g / cm 3

To obtain the molal osmotic coefficient, I, we need an expression for J r in terms of the
ionic strength I (= m).
Since solution is not dilute, we use Bromley’s model:

AJ I 1/ 2 (0.138  138
. B)I
ln J r    2.303BI
1  I 1/ 2 . I )2
(1  15

with, for NaCl at 25ºC, AJ 1.174 kg1/2 mol 1/2 and B 0.0574 kg1/2 mol-1/2 .
From Eq. (9-11) of the text (reminding that in this case I = m),
Solutions Manual 140

1 I
I 1
³
I 0
I d ln J r

A I I 1/ 2 I
1
³
I 0 2(1  I 1/ 2 )2
dI  (0.138  1.38B)
(1  1.5I )2

(0.138  1.38 B) I I B I

I ³
0 (1  1.5I )
2
dI 
2.303 0
dI
³
Performing the integrations, we obtain:

Aª 1 º
I 1 1  I 1/ 2  2 ln(1  I 1/ 2 ) 
I «¬ »
1  I 1/ 2 ¼

(0.138  1.38 B) ª 1  3I 1 º IB
 «  ln(1  1.5I ) »  2.303
1.5 2 1.5 I 2
¬« (1  1.5I ) ¼»

Substituting AJ = 1.174, B = 0.0574, and I = 0.62 we obtain I = 0.924.

b) From the expression that relates the osmotic pressure, S, to the molal osmotic coefficient, I,
we obtain

Q RTMw
S Im
1000 v w

(2) u (8.314 J K 1 mol 1 ) u (298.15 K) u (18 g mol 1 )


u (0.924) u (0.620 mol kg1 )
(1000) u (18.05 u 10 6 )

2.83 u 106 Pascal =28.3 bar | 28 atm

Linear interpolation from osmotic pressure data of aqueous NaCl listed in Perry gives for m
= 0.62 mol kg-1, 28.0 atm, in good agreement with our calculated osmotic pressure.

7. For K 2 SO4 , which is a 2-1 electrolyte, we have


Solutions Manual 141

m mK  2m ; z 1

m mSO2  m ; z 2
4

1
I (2m  22 m) 3m ; Q  2, Q  1, Q 3
2

ª 1/ 2 º
AJ z z I 1/ 2 1 « (m) AJ z z (3m) »
ln J (m)
r  bI Ÿ b ln J 
1  BaI 1/ 2 3m « r 1  Ba(3m)(3m)1/ 2 »
«¬ »¼
With

J (m)
r 0.4 for m 0.12 mol kg-1; AJ 1.174 kg1/2 mol-1/2 ; B 0.33 kg1/2 mol1/ 2 Å-1; a 4Å

we obtain

1 ª (1.174) u (2) u (0.36)1/ 2 º 1


b « ln 0.4  » kg mol
0.36 «¬ 1  (0.33) u (4) u (0.36)1/ 2 »¼

0.362 kg mol 1

For m 0.33 mol kg1 , we obtain J r 0.25 (the experimental value is 0.275).
To use Eq. (9-25) for calculating the activity of water, we first need to calculate the osmotic
coefficient, I .

1.174 0.3615 u (3 u 0.33)


I 1 (2) u (3 u 0.33)1/ 2 V( y) 
3 2

where

y (0.33) u 4 u (3 u 0.33)1/ 2 1.32

and
3ª 1 º
V( y 1.32) «1  1.32  2 ln(1  1.32)  1  1.32 » 0.257
3
(1.32) ¬ ¼
ª 1000 (g/kg) º
I 0.618 « » u ln aw
¬ 3 u 0.33 (m ol/kg) u 18 (g/mol) ¼

ln aw 0.011 Ÿ aw 0.989

Pw / Pwsat 0.989 Ÿ Pw 0.0314 bar


The vapor pressure has not changed much; it is only about 1% lower than that of pure wa-
ter because m is still small.
Solutions Manual 142

8.

S 2
cBSA  cCI  BcBSA
RT

c concentration in mol L-1

CI counter ion

Because the charge on BSA is –20, there are 20 counter ions (protons) for each molecule of
BSA.
B is the osmotic second virial coefficient characterizing the BSA-BSA interaction in a 1 M
aqueous NaCl medium. We neglect contributions from proton-proton and proton-BSA interac-
tions, and also contributions from interactions with NaCl.
Because the concentration (1 M) of NaCl is the same in both sides and because 1 M is much
larger than the concentration of counter ions, we neglect any (tiny) charges that might occur in
NaCl concentration due to interactions of Na+ and Cl- with counter ions or with BSA.

224
S 17.2 mmHg; T 298 K; R 62.36 mmHg L mol 1 K 1
13

44.6
cBSA 6.76 u 10 4 mol L1; cCI (20) u (6.76 u 10 4 ) mol L1
66, 000

Substitution gives

B = -29040 L mol-1

9.
a) To calculate the activity coefficients of water we use the Gibbs-Duhem equation:

x
d ln J w  s d ln J r (1)
xw

where subscripts w and s refer to water and salt, respectively.


Integration of Eq. (1) between mole fractions xs = 0 (or xw = 1, for which Jw = 1) and xs,
gives
Solutions Manual 143

ln J w ( x s )
z 0
xs

1
 xs
 xs
d ln J r

z
or
xs
FG
 xs d ln J r IJ
ln J w ( xs )
0 1  xs H
d xs
dxs
K (2)

The derivative in Eq. (2) can be obtained from the truncated Pitzer equation given:

8.766 x 1/ 2  124.598 x 3/2


ln J r (3)
1  9x 1/ 2

We made I = xs, because being NaBr a 1-1 electrolyte, the solution ionic strength is

1 1
Ix ¦ xi zi2
2 i 2
( xs  xs ) xs (4)

Differentiation of ln Jr in Eq. (3) in order of x and substitution in Eq. (2) gives

ln J w ( xs )
z
0 1
 xs 4.383xs 1/ 2  186.897 xs 1/ 2  1121382
xs

 x s
LM
MN 1/ 2 2
(1  9 xs )
. xs
dx s
OP
PQ (5)

The salt mole fractions are easily calculated from the given molalities (between m = 0 and m
= 5 mol kg-1) from:

nNaBr mNaBr mNaBr


xs (6)
nw  nNaBr 1000 55.51  mNaBr
 mNaBr
18.015

The following figure shows the activity coefficients of water in different NaBr aqueous so-
lutions at 25ºC, calculated using the Sympson rule to evaluate the integral in Eq. (5). As the fig-
ure shows, Jw | 1 until about m = 1.5 mol kg-1 and becomes less than one after that concentra-
tion. For example, Jw = 0.92 for m = 8 mol kg-1 and Jw = 0.88 for m =10 mol kg-1.
Solutions Manual 144

1.00

0.95
Jw

0.90

0.01 0.10 1.00 10.00

Molality NaBr

b) The mean ionic activity coefficients for NaBr aqueous solutions at 25ºC are calculated from
Debye-Hückel equation, ln J r  Ax I x , and from the Pitzer equation as given in this problem.
The following figure compares both predictions. As expected, they agree only at very low salt
concentrations.
1.2

1.0
Mean Ionic Actvity Coefficient

0.8
Pitzer

0.6

0.4 Debye-Huckel

0.2
0.001 0.01 0.1 1 10

Molality NaBr

c) Equation (4-44) gives the Van’t Hoff equation for the osmotic pressure: (valid for ideal, di-
lute solutions):

SV ns RT (7)

where V is the total volume and ns the number of moles of the salt.
Taking into consideration the nonideality of the liquid phase we write [Eq. (4-41)]:
Solutions Manual 145

S vw
 ln aw  ln(J w xw ) (8)
RT
where v w is the molar volume of the pure solvent (water).
We use Eq. (7) to calculate the osmotic pressure for the simplest case (Van’t Hoff equation)
and Eq. (8) for the more correct calculation that takes into account the solution nonideality, with
Jw obtained from Eq. (5).
Assuming that NaBr is completely dissociated into Na+ and Br- in water, we rewrite Eq. (7)
as

ns
S 2cs RT with cs mol L1 (9)
V
We obtain the salt concentrations (molarities) from the given molalities (m2) from

d ms
cs (mol L1 )
1  0.001Ms ms

where d is the mass density (in g cm-3) of the solution and Ms is the molar mass of NaBr (in g
mol-1).
To use Eq. (8) we take the molar volume of pure water as v w 18.015 cm3 mol 1 ,
xw 1  2 xs with Jw given by Eq. (5).
The following figures compares the results obtained from the Van’t Hoff equation [Eq. (9)]
with the equation that takes into account the solution nonideality [Eq. (8)].

700

600 Eq. (8)

500
Osmotic Pressue (atm)

400
Van't Hoff
300

200

100

0
0 2 4 6 8 10 12

Molality NaBr

As the figure shows, the solution behaves as an ideal solution (i.e. Van’t Hoff equation is
valid) up to a concentration of about mNaBr = 2 mol kg-1. However, for more concentrated solu-
tions (mNaBr > 2 mol kg-1), the effect of thermodynamic nonideality can not be neglected any-
more.
Solutions Manual 146

10. For the dissociation reaction

AB A+ + B-
the equilibrium constant is

(aA )(aB )
K 5 u 103 mol kg1
aAB

The equilibrium equation is


L V
fAB fAB (1)

where f denotes fugacity.


Equation (1) is equivalent to (Henry’s constant H AB,w 30 bar kg mol 1 , yAB 1)

ª P vf
AB dP
º
mAB HAB,w exp «
«¬ ³ Pwsat RT
»
»¼
MAB P (2)

Component AB in the vapor phase is in equilibrium with the undissociated AB dissolved in


water. The total molality of AB (solubility) in water is:

mT mAB  mA

If D is the fraction dissociated, we obtain

mA D mT and mAB (1  D )mT

The equilibrium constant then is

J 2r D 2 mT
K 5 u 10 3 mol kg1 (3)
1 D

and Eq. (2) becomes

ª v f ( P  Pwsat ) º
(1  D)mT H AB,w exp « AB » MAB P (2a)
«¬ RT »¼

where P  Pwsat # P because P 50 bar !! Pwsat .


The fugacity coefficient at 50 bar is

ª PBAB,AB º
MAB exp « »
¬ RT ¼

ª (50 bar) u (200 cm3mol 1 ) º


exp « » 0.668
3 1 1
¬« (83.14 cm bar K mol ) u (298 K) ¼»

1
With I (m   mB ) mA D u mT , the mean ionic activity coefficient is [Eq. (9-52)]
2 A
Solutions Manual 147

AJ D mT
ln J r  (4)
1  D mT

where AJ 1.174 kg1/2 mol 1 / 2 .


We have now three equations [Eqs. (2a), (3), (4)] and three unknowns mT , D, and J r .
Solving these gives:

mT 1.038 mol kg1

D 0.0871

Jr 0.762

Iteration procedure is:


Start with D 0 in Eq. (2a) and calculate mT .
Then calculate a first approximation for D with J r 1 using Eq. (3).
Use mT and D in Eq. (4) to obtain better J r , etc.

11. In his theory of absolute reaction rates, Eyring states that reactants A and B form an activa-
ted complex (AB) as an intermediate state in the reaction

A + B (AB) o Products (1)


By assumption, reactants A and B are in equilibrium with the activated complex (AB), so
that

a(AB)
K (2)
aA aB

The reaction rate is proportional to the concentration of (AB), i.e., constant u c(AB) .
Replacing the activities by the products of concentrations and activity coefficients in Eq. (2)
we obtain

J J
c(AB) KcA cB A B (3)
J (AB)

Hence,

J J
Rate of reaction (constant)KcA cB A B (4)
J (AB)
Solutions Manual 148

The rate of reaction can also be expressed in the usual manner by kcA cB , where k is the ob-
served specific rate. Hence, from Eq. (4), we can write

J J J J
k (constant)K A B ko A B (5)
J (AB) J (AB)

Equivalently, we can write

J J
log k log ko  log A B (6)
J (AB)

To calculate the second term on the right side of Eq. (6), we use Eq. (9-50a)

J J 2
log A B 0.510 I u ( zA  zB2  z2(AB) ) (7)
J (AB)

Substituting z(AB) zA  zB gives

J J
log A B 0.510 I u (2 zA zB ) (8)
J (AB)

Combining Eqs. (6) and (8) gives

log k log ko  1.02zA zB I (9)

Therefore, a plot of log k versus I is a straight line with slope 1.02zA zB .

Reaction zAzB Change of k

I -2 Decreases with increasing I

II 0 Constant

When the inert salt NaCl is added, I changes.


For mNaCl 0.01 and negligible molalities of reactants, we obtain

m mNa  0.01 z 1
(10)
m mCl- 0.01 z 1

and

1
I (0.01  0.01) 0.01 M (11)
2
Combining Eqs. (9) and (11) gives
Solutions Manual 149

Reaction zAzB k

I -2 -0.204

II 0 0

12. Some helpful relations for single-electrolyte solutions (i.e., one cation M and one anion X)

QM QX
Q QM  QX or 1  (1)
Q Q

Q M zM  Q X zX 0 or Q M zM Q X zX (2)

mQ mM  mX mQ M  mQ X

¦ mi Qm
i

¦ mi zi m(Q M zM  Q X zX )
i

1 1
I ¦ mi zi2
2 i 2
2
m(Q M zM 2
 Q X zX )

a) Eq. (9-59) from Eq. (I-13):

For a single electrolyte: c = 1 = M and a = 1 = X


All )’s, <’s, and O’s are zero. Equations (I-14), (I-15), (9-61) and Eq. (I-13) give
Solutions Manual 150

ln J r zM zX f J  zM zX mM mX BMX
'

QM § ·
 mX ¨ 2 BMX  ¦ mi zi cMX ¸
Q ¨ ¸
© i ¹

QX § ·
 mM ¨ 2 BMX  ¦ mi zi cMX ¸
Q ¨ ¸
© i ¹

mM mX
 2Q M zM cMX
Q

'
where BMX and BMX are given by Eq. (I-19a) and Eq. (I-19b), respectively, and CMX is given
by Eq. (I-18).
Equations (I-16), (I-19), and (9-62) – (9-65) give

' J
2BMX  I BM BMX

and

2 J 1/ 2
CMX C /(2 zM zX )
3 MX

'
From Eq. (9-45), the terms with BMX or BMX can be summarized as:

' § Q m  Q X mM ·
zM zX mM mX BMX  2¨ M X ¸ BMX
© Q ¹

2I 2 ' 1
m Q MQ X BMX  (Q MQ X m  Q X Q M m)2 BMX
mQ Q

2Q M Q X '
m( IBMX  2 BMX )
Q

§ 2Q Q · J
m2 ¨ M X ¸ BMX
© Q ¹

Summarizing the CMX terms gives:


Solutions Manual 151

ª§ Q M QX · mM mX º2 1 J
«¨ Q mX  Q mM ¸ (mM zM  mX zX )  Q 2Q M zM » 3 CM
¬© ¹ ¼ 2 zM zX 1/ 2

J
ª§ Q M Q X Q Q · m2 Q M Q X º CMX
«¨ m  X M m ¸ (Q M zM  Q X zX )m  2Q M zM »
«¬© Q Q ¹ Q »¼ 3 zM zX 1/ 2

ª 2(Q Q )3/ 2 º J
m2 « M X » CMX
¬« Q ¼»

The 3 terms above marked with are identical to the 3 terms in Eq. (9-59).

b) Eq. (9-60) from Eq. (I-10):

As before c = 1 = M and a = 1 = X and all )’s, <’s, and O’s are zero.
Equations (9-64) and (I-18) and Eq. (I-10) give

I 1 2I / ¦ mi f I


(1)

2 I
 mM mX BMX (1)
¦

mi

(2)

2 m m I
 ¦ mi zi M X1/ 2 CMX
¦ i
m 2 zM zX


(3)

Term (1):
1 2 2
2 I / ¦ mi 2 m(Q M zM  Q X zX )
2
2 2
Q M zM  Q X zX
Q

With Q M zM Q X z X , it follows
Solutions Manual 152

1§ QX Q ·
Term (1) ¨ Q M zM zX  Q X zX M zM ¸
Q© QM QX ¹

§Q Q ·
zM zX ¨ X  M ¸
© Q Q ¹

§Q Q ·
zM zX ¨ X  M ¸
© Q Q ¹

zM zX

Term (2):

2 2m2 Q M Q X § 2Q Q ·
mM mX m¨ M X ¸
¦ mi Qm © Q ¹

Term (3):

2 mM mX Q M Q X m2 m(Q M zM  Q X zX )
¦ mi zi
¦ mi 2 zM zX
1/ 2 Qm z z
1/ 2
M X

Again using Q M zM Q X zX gives


1/ 2
1/ 2 § QX ·
Q M zM zX Q X zX zX Ÿ zM zX ¨ ¸ zX
© QM ¹

and

Q M zM  Q X zX 2Q X zX

Therefore,

Q M Q X 2 2Q X zX ª 2(Q Q )3/ 2 º
Term (3) m m2 « M X »
2 1/ 2
Q Q1/
X QM zX «
¬
Q »
¼

Comparison of the results for the terms (1), (2), and (3) with those in Eq. (9-60) shows that
they are identical.
S O L U T I O N S T O P R O B L E M S

C H A P T E R 1 0

1. Let 1 = methane, 2 = benzene, 3 = m-xylene, and 4 = hexane.


Neglecting Poynting corrections and vapor non-idealities,

L
y1 J 1 f pure 1
K1
x1 M 1P

L
From Fig. 10-13 of the text we obtain f pure 1
at 366 K and 13.8 bar.
Because y1 # 1 we can use Lewis’ fugacity rule to obtain M1 by writing M1 ( f / P) pure 1 .
We find J 1 from

v1 (G1  G)2
ln J1
RT

In the first iteration, find G using x1 0.


For a second estimate, first calculate y1 from y1 1  y1  y3  y4 , where

x2 J 2 f20
y2 , etc.
P
Then,
y1
x1 (other xi from relative amounts)
K1

Recalculate G for second estimate of J 1 to find

153
Solutions Manual 154

K1 KCH 4 34

2. Let 1 = argyle acetate and 2 = helium.


Because we have two data points, we can use the Krichevsky-Kasarnovsky equation to evaluate
the two parameters H 2,1 and v2f .

FG f2 IJ (Ps ) v2f (P  P1s )


ln
H x2 K ln H 2,11 
RT

Assume:

1. J *2 1

2. P1s  P

Then, for helium,


f2 y2M 2 P

From data given in App. C for helium,

B22 (293 K) 12.1 cm3 mol 1

Using the virial equation and assume y2 1,

B22 P
ln M 2
RT
Find f 2 at different pressures:

P (bar) M2 f2 (bar)

25 1.012 25.3
75 1.038 77.8
150 1.077 161.5

At 25 bar,

f2
ln ln(25.3 u 10 4 )
x2
(25) u ( v2f )
ln H 2,1 
. ) u (293)
(8314

(75) u (v2f )
ln(27 u 10 4 ) ln H 2,1 
. ) u (293)
(8314
Solutions Manual 155

These equations give

vf
2 32.4 cm3 mol 1

and

ln H 2,1 12.4

At 150 bar,

ln x2 ln f2  12.62

x2 x He 5.37 u 10 4

3. Over a small range of values for G, Hildebrand has shown that log x 2 is linear (approximately)
in G1. A plot of log x 2 vs. G solvent gives x2 in liquid air.

log x 2

G1

Because
1/ 2 1/ 2
§ ' vap ui · § ' vap hi  RT ·
Gi ¨¨ v ¸¸ | ¨¨ ¸¸
© i ¹ © vi ¹

we can calculate G’s from data given:

G CH 4 15.0 (J cm 3 )1/ 2

and
G CO 1.4 (J cm 3 )1/2

For air, assume a mixture of N2 and O2:

G air 11.4 (J cm 3 )1/2


Solutions Manual 156

From plot of log x 2 vs. G1 we find for G 11.4 (J cm 3 )1/2 ,

x2 (in air) xH2 2.63 u 10 3

Note that in this case we hit one of our known points. In general, we must interpolate or ex-
trapolate.

4.
a) The equilibrium equation between the gaseous phase containing oxygen and the liquid phase
saturated in dissolved oxygen is:

fOG fOL (1)


2 2

At the low pressures of interest here, we assume the gas phase as an ideal gas mixture:

fOG yO P (2)
2 2

Moreover, because oxygen is sparingly soluble, i.e., oxygen is present at very low concen-
trations in the liquid phase, Henry's law holds:

fOL x O 2 kO 2 (3)
2

From Eqs. (1), (2) and (3),

yO P x O kO (4)
2 2 2

Equation (4) is the condition for phase equilibrium that characterizes the dissolution of a
sparingly soluble gas. Under the given pressure and gaseous composition, Henry's law constant
kO2 can be determined, once the solubility xO2 is known. This solubility is given here by the
Bunsen coefficient D.
Substituing t = 20ºC in the equation for D we obtain,

Ncm3 (O2 )
D = 31.01u10-3
cm3 (H2 O)

where Ncm3 stands for normal cubic centimeters, i.e., cubic centimeters of gas measured at 0ºC
and 1 atm.
We convert normal cubic centimeters of gas to moles using the ideal gas law:
Solutions Manual 157

RT (8.31451) u (293.15)
v O2
P 101325

0.022414 m3 mol 1

22414 cm3 mol 1

The molar density of water at 20ºC and 1 atm is

1 U H 2O 0.9982
0.0554 mol cm 3
v H 2O M H2O 18.015

Under the conditions of the Bunsen experiment, we have:

D . u 10 3
3101 mol O 2
Dissolved oxygen 13835
. u 10 6
v O2 22414 cm3 H 2 O

The liquid phase is made exclusively of H2O and O2. Then the mole fraction of O2 in the
liquid is

FDI
GH vO2 JK .
13835 u 10 6
xO2
F D I F 1 I u 10 6  0.0554
GH vO2 JK GH v H2O JK
.
13835

.
13835 u 10 6 mol O 2
| 2.497 u 10 5
0.0554 mol O 2  mol H 2 O

By the definition of D, yO2 10


. and P = 1 atm = 1.01325 bar.
From Eq. (4),

yO 2 P . ) u (101325
(10 . )
kO 2 4.058 u 10 4 bar
xO2 (2.497 u 10 5 )

b) Here we want to determine xO2 given yO2 , P and kO2 , using again Eq. (4).
The atmospheric air above the water phase is a mixture of water vapor, oxygen and other
atmospheric gases (predominantly nitrogen).
Assuming that the air is saturated in water vapor,

PHs O 17.5
2
yH2O 0.0230
P 760
The mole fraction of oxygen in the vapor phase is then
Solutions Manual 158

PO2 (0.2095) u [(760)  (17.5)]


yO2 0.02047
P (760)

For the ambient pressure we obtain form Eq. (4),

yO2 P (0.2047) u (101325


. ) mol O 2
xO2 5112
. u 10 6
kO 2 (4.057 u 10 4 ) mol liquid

This solubility can now be expressed as mass of gas per volume of liquid, making the sim-
plifying assumption that the liquid phase is pratically pure water:

(5.112 u 10 6 mol O2 / mol liquid) u (32.0 g O2 / mol O2 )


Solubility=
§ 1 ·
(18.015 g H2 O/mol H2 O) u ¨ ¸ u (1 mol H2 O / mol liquid)
¨ 0.9982 g H O/cm3 H O ¸
© 2 2 ¹

9.06 u 10 6 g O2 /cm3 H2 O = 9.06 mg dm 3 9.06 ppm

5. The number of moles for each component is

180 420 28
n1 10.0 mol n2 5.0 mol n3 10
. mol
18.015 84.16 28.012
The volume available for the vapor phase is

FG 420  180 IJ u 10 6
V 3.0 
H 0.774 0.997 K 2.28 u 10 3 m 3

Therefore the pressure inside the vessel is (assuming vapor  formed almost exclusively by
nitrogen whose second virial coefficient at 25ºC is zero  as ideal),

n3 RT (1) u (8.314) u (29815


. )
P
V (2.28 u 10 6 )

10.88 u 10 5 Pa 10.88 bar

Because the solubilities are very small, we use Henry's law to describe the fugacity of
cyclohexane in gas phase:

f3 | p3 y3 P x3 H3,i .

Further, because we neglect mutual solubility of water and cyclohexane, we obtain for the
solubility of nitrogen in water,
Solutions Manual 159

FG H3,1 IJ 1
FG 86,000 IJ 1
x3,w
H p3 K H 10.88 K . u 10 4
126

and for the solubility of nitrogen in cyclohexane,

FG H3,2 IJ 1
FG 1,300 IJ 1
x3,c
H p3 K H 10.88 K 8.37 u 10 3

6. Let 1 = N2 and 2 = H2.


From Orentlicher’s correlation,

f2 ( Ps ) A 2 v f ( P  P1s )
ln ln H2,11  ( x1  1)  2
x2 RT RT

Assuming that the vapor is pure H2:

f2 f2V y2 M 2 P 88 bar (with y2 1)

Because,

P1s 1 bar

A 7.1 L bar mol 1

H2,1 467 bar

v 2f 31.3 cm3 mol 1

by trial and error we find

x2 0.17

7. From Fig. 10-11 with G1 14.9 (J cm 3 )1/ 2 , at 25oC and at 1.01325 bar partial pressure,

log x2 3.1 Ÿ x2 8 u 10 4

For t = 0oC, use Eq. (10-26):


Solutions Manual 160

x (T ) 's T
ln 2 2 # 2 ln 2
x 2 (T1 ) R T1

But at 25oC and 1.01325 bar,

 R ln x2 59.29 J mol 1 K 1

and from Fig. 10-7,

s2L  s2G # 17 J mol 1 K 1

Then,

's 2 s2L  s2G 17 J mol 1 K 1

x2 17 273
ln ln
8 u 10 4 8.31451 298

x2 6.7 u 10 4 (at 1 bar)

Assuming

x2 (P2 ) ( y2 P) 2
x 2 (P1 ) ( y2 P)1

we obtain at 0oC and 2 bar partial pressure,

x2 = 1.34 ×10-3

8. Let 1 = ethylene oxide and 2 = CH4.


Then,

f (Ps ) v2f (P  P1s )


ln 2 ln H 2,11 
x2 RT

From Tables 10-2 and 10-3 at 100C,

H 2,1 621 bar

v2f 45 cm3 mol -1

At 10oC, P1s | 1 bar. Then,


Solutions Manual 161

f
ln 2 6.5
x2

f2
665 bar
x2

But

f2 M 2 y2 P

Assuming y2 | 1,

B22 P
ln M 2
RT

with B22 49 cm3 mol 1 (Table 10-3)

M2 0.949

f2 (0.949) u (25) 23.7 bar

f2
xCH 4 x2 0.036
665

9.
a) Henry’s constant H2,1 is calculated from [Eq. (10-21)]:

(Ps )
H2,11 P1s M 2L,f

where M 2L ,f is the fugacity coefficient of solute 2 in the liquid phase at infinite dilution (x2 = 0);
M2 may be obtained from Eq. (12-64) of the text with

bmixture | b1

v mixture | v1s

P | P1s

x2 | 0

Then, H2,1 is given by


Solutions Manual 162

RT b2 a1b2 v1
ln H 2,1 ln  
v1  b1 v1  b1 RTb1[v1 (v1  b1 )  b1 (v1  b1 )]

a1b2 (v1  2.414b1 ) a12 (v  2.414b1 )


 ln  ln 1
2 2 RTb12 (v1  0.414b1 ) RT 2b1 (v1  0.414b1 )

where

a12 (a1a2 )1/ 2 (1  k12 )

k12 0.0867

Constants a1 and a2 are obtained from Eqs. (12-61) to (12-63) and constants b1 and b2 from
Eq. (12-60). Substitution gives

H 2,1 360 bar


b) From Eq. (10-22),

LM F wP I OP
M M GH wn JK2 T ,V ,n PP
MM FGH wwVP IJK
v2f 1
PP
N T ,n1,n2
Q n2 0

Using Peng-Robinson equation of state [Eq. (12-59)], we find

RT (b2  v1s  b1 ) 2a12 2a1b2 (v1s  b1 )


 s s 
(v1s  b1 ) 2 v1 (v1  b1 )  b1 (v1s  b1 ) [v1s (v1s  b1 )  b1 (v1s  b1 )]2
v2f
RT 2a1 (v1s  b1 )

(v1s  b1 ) 2 [v1s (v1s  b1 )  b1 (v1s  b1 )]2

Substitution gives

v 2f 69.5 cm 3 mol 1

c) Margules parameter A can be found from Eq. (10-23):

F L I
A 
2
GH
RT w ln M 2
wx 2
JK
P P1s ,T , x2 0

The result is:

R|
A  S|
RT b2 (b2  b1 )  v' b2

a1b2 v'2b2 v1s (a12  a1 )

v' (b2  b1 )
2
T(v1s  b1 ) 2 RTb1[v1s (v1s  b1 )  b1 (v1s  b1 )] v1s  b1
Solutions Manual 163


{ } { }
a1b2v1s (b2  b1 ) v1s (v1s  b1 )  b1 (v1s  b1 )  2b1 v' (v1s  b1 )  (v1s  b1 )(b2  b1 )

{ }
2
RT b1 v1s (v1s  b1 )  b1 (v1s  b1 )

FG
2b1 (a12  a1 )  a1 (b2  b1 ) 2a12 b2 IJ
v s  2.414b1

2 2 RTb12 Ha1

b1
ln 1
K
v1s  0.414b1

a1 LM a1 (a2  a12 )  2a12 (a12  a1 )  b2 (b2  b1) OP ln v1s  2.414b1


2 RTb1 MN PQ v1s  0.414b1

a12 2b12

FG IJ U|

a1 2a12 b2

H
v1s (b2  b1 )  b1v'
K V|
RTb1 a1 b1 (v1s  2.414b1 )(v1s  0.414b1 )
W
with

FG wv IJ
v'
H wx2 K x2 0 v2f  v1s

Substitution gives

A 13,900 bar cm3 mol1


S O L U T I O N S T O P R O B L E M S

C H A P T E R 1 1

1. Assuming

fi s s
f pure i

and

fiL L
xi f pure i (i.e. J i 1)

with Tt Tm , 'c p 0,

L
f pure FG 1 IJ ' fus hi LM1  T OP
s i
i
ln
f pure
ln
H xi K RT N Tm Q
Rearranging,

ª º
« »
§ ' fus hi · « 1 »
T ¨ R ¸«
© ¹« § 1 · ' fus hi »»
ln ¨ ¸
«¬ © xi ¹ RTm,i »¼

For i = benzene, ' fus hi 9843 J mol-1, xi = 0.95, Tm,i 278.7 K , we obtain

T 275 K

For i = naphthalene, ' fus hi 19008 J mol-1, xi = 0.05, Tm,i 353.4 K , we obtain

T 241 K

165
Solutions Manual 166

We choose the higher temperature (i.e., benzene precipitates first).

T, K 353 K

279 K
275 K

Benzene Naphthalene

At T = 275 K, a solid phase appears.

We need activity aA at x A sat


2. xA / 2.
At saturation,

fAL fA s s
f pure A

Because

fAL L
aA f pure A

§ fL · ' fus h § T · 'c p § Tt  T · 'c p Tt


s ¸¸¹pure
sat
 ln aA ln ¨ A ¨1  ¸   ln
¨f RT © Tt ¹ R ¨© T ¸¹ R T
© A

Assuming, Tm sat
Tt , we find aA 0.118 .
sat
Because xA 0.05 , J A 2.36 .
To find the activity at another composition, assume that

RT ln J A D(1  x A ) 2

Using the above data, we find D 2357 J mol 1 .


Hence at xA sat
xA /2,

JA 2.47
Solutions Manual 167

aA xA J A 0.0618
Then
T
8.04 Ÿ T 0.89
1 T

Thus, 89% of sites are occupied.

3. Klatt’s data are really at 00C, not -700C. This is above the freezing points of toluene and xylene
P P P P

and near that of benzene.


Let HF be component 1 and the solute (2) be A (benzene), B (toluene), and C (m-xylene).
The order of increasing substitution (basicity) is A, B, C. To simplify things, ignore the solubil-
ity of 1 in 2.
Then:

f pure 2 f2, in 1 x2 J 2 f20

But, f20 f pure 2 , so this reduces to: 1 x2 J 2 . Therefore, x2 is inversely proportional to J2.
We might think that J2 depends only on the 1-2 interaction. On this basis, we expect
J C  J B  J A , and thus x C ! x B ! x A . Klatt’s data show the reverse.
There is, however, another factor: the strength of the 2-2 interactions. At 00C, U U P P

PAs | 0.036 bar , PBs | 0.009 bar and PCs | 0.002 bar . This means that pure C “holds on” to its
molecules more tightly than B which in turn has a tighter grip than A.
In other words, the more volatile solute (that has the weakest 2-2 interactions) exerts more
“pressure” to enter the solvent phase. This is discussed in a qualitative manner by Hildebrand,
1949, J. Phys. Coll. Chem., 53: 973.
It may be helpful to look at this from a lattice theory (interchange energy) perspective. Us-
ing the simplest form of this theory, we can say:

w 2
ln J 2 x
kT 1
1
w z [*12  (*11  * 22 )]
2

where w is the interchange energy.


With the attractive interaction, we expect *1A ! *1B ! *1C .
This produces a higher w (and hence a higher J2 and lower x2) for the less-substituted mole-
cule.
But, if we look at the vapor pressures we see that the less-substituted molecules have less
U U

attractive 2-2 interactions. Hence, *AA ! *BB ! *CC . This produces a higher w (and hence a
higher J2 and lower x2) for the more-substituted molecule. Sometimes, this effect is greater than
U U

that of the 1-2 interactions; that is apparently true in this case.


Solutions Manual 168

4. Let 1 = naphthalene, 2 = iso-pentane, and 3 = CCl4.


At saturation,

f1s f1L x1J 1 f1,L pure

s
Assuming, f1, pure f1 ,s
§ f1L · ' fus h § T ·
f1s ¸¹ pure
ln ¨ ¸  ln a1 ¨1  ¸  ln x1J1
¨ RT © Tt ¹
©

From the regular-solution theory, assuming x1 = 0 initially, we obtain

G 14.9 (J cm-3)1/2

and
v1 (G1  G )2
ln J 1 1.45
RT

G1 20.3 (J cm-3)1/2

ln a1 118
.

x1 0.073

x2 7
Now repeat the calculation using x1 = 0.073 and , to obtain
x3 3

G 15.3 (J cm-3)1/2

and

x1 0.093

One more iteration gives x1 xnaphthalene | 0.10 .

5. The equilibrium equation for benzene (B) is


s
Partial pressure of B yB P xB J B Pliquid B (1)
Solutions Manual 169

where xB ( xB 0.10 ) and J B denote liquid-phase mole fraction and activity coefficient of ben-
s
zene; Pliquid B is the vapor pressure of pure, subcooled liquid benzene at 260 K.
s
To find Pliquid B , we use the approximation

s
Pliquid § fL ·
s ¸¸¹pure B
B
¨ (2)
s
Psolid ¨f
B ©

s s
where Psolid B ( Psolid B 0.0125 bar at 260 K ) is the vapor pressure of pure solid benzene at 260
K; the fugacity ratio for pure benzene is calculated from Eq. (11-13) neglecting 'c p for benzene

§ fL · ' fus h § Tm
s ¸¸¹pure B
·
ln ¨  1¸ (3)
¨f
© RTm ¨© T ¹

Substituting ' fus h 30.45 cal g1 9944.07 J mol 1 , Tm 278.7 K , T 260 K and R = 8.314
J K-1 mol-1 into Eqs. (2) and (3), we obtain
s
Pliquid B
1.362 (4)
s
Psolid B

Hence,
s
Pliquid B 1.362 u (0.0125 bar) 0.0170 bar (5)

To calculate J B in Eq. (1), we use Eq. (7-55):

vB
ln J B (GB  G )2 (6)
RT
where
3
G ¦ ) i Gi
i 1
(7)
xi vi
)i
3
¦ x jv j
j

Substituting T 260 K , R 8.314 J K 1 mol 1 and the given liquid-phase mole fractions,
pure-component molar volumes and solubility parameters into Eqs. (6) and (7), we obtain

JB 1.305 (8)

Combining xB 0.10 and Eqs. (1), (5) and (8) yields

Partial pressure of B yB P 0.0022 bar


Solutions Manual 170

6. This is similar to Problem 1, but includes activity coefficients.

fi s s
fi,pure fi L xi J i fi,Lpure

Then, considering 'c p 0 and Tt Tm ,

§ fL · § ·
s
' fus hi T § 1 ·
ln ¨ i ¸ ¨¨ 1  ¸¸ ln ¨ ¸
¨f ¸ RT © Tm,i ¹ © xi J i ¹
© i ¹pure

Using the regular-solution theory,

vi (G1  G 2 )2 ) 2j
ln J i
RT
Let 1 = benzene and 2 = n-heptane

)2 0.935

. ) 2 u (0.935) 2
(89) u (18.8  151 128
ln J 1
(8.31451) u (T ) T

Then,

9843 § T · ª § 128 · º
1  ln «(0.1) u exp ¨
RT ¨© 278.7 ¸¹ ¬ © T ¹¼
¸»

Solving for T,
T 200 K

Similarly, for n-heptane,

)1 0.065

(148) u (18.8  15.1)2 u (0.065)2 1.03


ln J 2
(8.31451) u (T ) T

Then,

14067 § T · ª § 1.03 · º
1  ln «(0.9) u exp ¨
RT © 182.6 ¸¹
¨
¬ © T ¹¼
¸»

Solving for T,
T 181 K
Solutions Manual 171

As temperature decreases, benzene starts to precipitate at 200 K.

7. Plotting the data we obtain:

1245

1200
t ( C)

1000
935

805
800

100 80 60 40

Mole % Cu2O

A compound, Cu2OP2O5, is formed with a congruent melting point at 1518 K. Eutectics oc-
cur at 1208 K and 1078 K.

8.
a) According to the ideal solubility equation, Tm of the solvent has no influence on the solubil-
U U

ity. Any difference would have to come from nonideality (i.e. activity coefficients).
If we look at solubility parameters, we find that the solubility parameter of CS2 is closer to
that of benzene. Therefore, we expect greater solubility in CS2.
b) Let 1 = benzene, 2 = CS2 and 3 = n-octane.
Assuming

s
f1 s
f pure 1 f1L L
x1J 1 f pure 1

and
Solutions Manual 172

'c p 0

Tt # Tm

we have

' fus h § T ·
 ln J1x1 ¨1  ¸ (1)
RT ¨© T f ¸
¹

Using the regular-solution theory,

v1 (G1  G ) 2
ln J 1 (2)
RT
with
3
G ¦ )iGi
i 1

From Tables:

Component G (J cm-3)1/2 v (cm3 mol-1)

1 18.8 89
2 20.4 61
3 15.3 164

From Eq. (1) with x1 = 0.3,

ln J 1 0.144

Then, from Eq. (2),

v1 (G1  G )
ln J 1 0.144
RT

G 171 . (or 20.6; this value is probably meaningless since it is higher than G’s of pure com-
ponents).
3
G ¦ )iGi
i 1

x1v1
)1 , etc.
x1v1  x 2 v 2  x 3v 3
Solving for x2 and x3,
Solutions Manual 173

x1 0.30

x2 xCS2 0.32

x3 0.38

9. At x 2 0.25 ,

s
f2 s
f pure 2 f2L L
x 2 J 2 f pure 2

x2 J 2 (f s / f L )pure 2 0.56

J2 2.24

s
Because f pure 2 s
P2 ,sat
0.99 bar,

L
f pure 2 . bar
177

Assuming ln J 2 Ax12 , at x 2 0.25 , then J 2 2.24 , and

A 1.434
At x 2 0.05 ,

f2V f2L L
x 2 J 2 f pure 2

ln J 2 (1.434) u (0.95) 2

J2 3.65

P2 PCO2 (0.05) u (3.65) u (177


. ) 0.323 bar

10. At 250 K, we need a standard-state fugacity for a hypothetical liquid.


Solutions Manual 174

§ fL · ' fus h § T ·
ln ¨ A ¸
¨ fs ¸ ¨1 
RT © Tm ¹
¸
© A ¹pure

(13000) § 250 ·
u ¨1 
(8.31451) u (250) © 300 ¸¹

Hence,

fAL
fAs
2.84

Because solid is pure,

s
fA s
f pure A s
PA ,sat 35 torr

L § 35 ·
fpure A ¨ ¸ u (2.84) 0.1325 bar
© 750.06 ¹

For the A-CCl4 system,

fAV fAL L
x A J A f pure A

Neglecting vapor-phase non idealities and the Poynting correction factor,


L
yA P x A J A f pure A

(5 / 750.06)
JA 1.677
(0.03) u (0.1325)

Using the regular-solution theory,

RT ln J A v A (G A  G CCl 4 ) 2 ) 2CCl
4

Thus,

(8.31451) u (250) u (ln1.677)


(GA  GCCl4 )2
2
ª (0.97) u (97) º
(95) u « »
¬ (0.03) u (95)  (0.97) u (97) ¼

G A  G CCl 4 3.4 (J cm-3)1/2

As for G CCl 4 17.6 (J cm-3)1/2,


P P

GA . (J cm-3)1/2
210 P

or
Solutions Manual 175

GA 14.2 (J cm-3)1/2

Because A is a branched hydrocarbon, we choose G A 14.2 (J cm-3)1/2.


For the A-hexane system,
L
yA P x A J A f pure A

v A (G A  Ghex )2 2
ln J A ) hex
RT

2
ª (0.99) u (132) º
(95) u (14.2  14.9)2 u « »
¬ (0.01) u (95)  (0.99) u (132) ¼
(8.31451) u (250)

JA 102
.

Then,

yA P PA (0.01) u (102
. ) u (0.1325)

PA 0.00135 bar

11. Let

fAL L
x A f pure A (J A 1)

fBL L
x B f pure B (J B 1)

From Eq. (11-13) with T # Tm , 'c p # 0 ,

§ L ·
fpure A ' fus hA § T ·
ln ¨
¨ s A ¸¸
fpure
¨¨ 1 
RT © Tm,A
¸¸
¹
© ¹

§ L ·
fpure B ' fus hB § T ·
ln ¨ ¸
¨ s B¸
fpure
¨¨ 1 
RT © Tm,B ¸¹
¸
© ¹

Because solids A and B are mutually insoluble,

fAs s
f pure A
Solutions Manual 176

fB s s
f pure B

At equilibrium,

fAs L
x A f pure A

s
fB L
xB f pure B

Then,

§ L ·
fpure A § 1 · (8000) § T ·
ln ¨
¨ s ¸
fpure A ¸¹
ln ¨ ¸
© xA ¹
u 1
(8.31451) u (T ) ¨© 293 ¸¹
©

1 (12000) § T ·
ln u ¨1 
xB (8.31451) u (T ) © 278 ¸¹

with x B 1  x A .
Solving the above equations, we obtain x A 0.516 (or 51.6 mol % A), xB 0.484 and
T 244 K.
S O L U T I O N S T O P R O B L E M S

C H A P T E R 1 2

1.

i) T < TU ii) T = TU

P P

L+L’ L U L+L’ L’

V
L+
L
L’
V V
L+ L+V
V

n-Alkane x Water n-Alkane x Water

iii) TU < T < TCn iv) TCn < T < TCb

P P

L+L’ L’ L+G L’

V
L+

V V

n-Alkane x Water n-Alkane x Water

177
Solutions Manual 178

v) T = TCb vi) TCb < T < TCH O


2
P P
G+G’

L+
G
V

n-Alkane x x
Water n-Alkane Water

vii) T > TCH O


2

P
G+G’

n-Alkane x Water

2.

a) T = T1 b) P1 < PUCEP

P T V

L L’ L+V

L+L’ L’+
V
L+
V L+L’
L
L’
L’+V
V

1 x 2 1 x 2
Solutions Manual 179

P2 = PUCEP P3 > PUCEP

T V T V

L+V L+V

L’+
V L
L
L+L’ L’
L+L’

1 x 2 1 x 2

3. Let A stand for alcohol.


For alcohol distributed between phases ' and "
' ' " "
xA JA xA JA

Then,

F xA' I F J "A I
K
x
lim G "J
A o0 H xA K x
limG'J
A o0 H J A K

At 0qC and 1 bar,

2400
ln J 'A ' )2
(1  x A
RT

320
ln J "A " )2
(1  xA
RT

The pressure correction to JA is

J A (P2 ) J A (P1 ) exp


z P2

P1
vE
RT
dP

The temperature correction is:

J A (T2 ) J A (T1 ) exp


z T2

T1
h E
RT 2
dT

Thus, we can write,


Solutions Manual 180

2400 100 16
ln J 'A [ 
³ dP
(8.31451) u (273) 1 (8.31451) u (273)

303 4800
³273 (8.31451) u (T 2 ) dT ]u (1  xA )
' 2


' 2
0.9178u (1  xA )

320 100 10


ln J"A [ 
³ dP
(8.31451) u (273) 1 (8.31451) u (273)

303 600
³273 (8.31451) u (T 2 ) dT ]u (1  xA )
" 2


0.0712(1  x"A )2

This gives

J "A
K lim 0.429
xA o 0 J 'A

4. For pure benzene, neglecting fugacity coefficients and assuming constant density of each phase
with respect to pressure,

fBL fB s
PBs,L exp
v L (P  PBs,L )
PBs, s exp v s (P  P s ) s,
B
RT RT

with v L 87.7 cm3 mol 1 and v s 77.4 cm 3 mol 1 , as obtained from density data.

ª (87.7) u (200  P s,L ) º


PBs,L exp « B » PBs, s exp ª« (77.4) u (200  PBs,s ) º»
«¬ (83.1451) T »¼ «¬ (83.1451)T »¼

Temperature T can be found from the intercept of the curves obtained by representing each

s
side of the last equation as a function of temperature.
In an alternate way, we can express PBs,L and PBs, from vapor-pressure equations:
Solutions Manual 181

10 (7.96221785/T )
PBs,L
750.06

PBs, s 10 (9.846  2310 /T )


750.06
which, together with the last equation, can be solved for T:
Tm(200 bar) = 284.4 K

5. The Redlich-Kwong equation is:

RT a
P 
v  b T 1/ 2v(v  b)

with
2a 2
a ¦ ¦ zi z j aij zA AA  2zA zBaAB  zBaBB
i j

b ¦ zi bi zA bA  zBbB
i

Assuming that aAB is given by the geometric rule,

aAB (aAA aBB )1/ 2

we get for zA zB 0.5,

a 4.35 u 10 8 bar (cm3 mol-1) K1/2

b = 91.5 cm3 mol-1


Substitution in the R-K equation gives for total pressures:
P 413
. bar

Because this result is absurd, we use Henry’s constant data to find aAB .
For infinitely dilute solutions of A in B,

H A,B (PM A ) xA f
PBs M B
0

s
At infinite dilution, Ptotal # Ppure B which can be obtained from the R-K equation with the

appropriate constants [ a 4.53 u 10 8 bar (cm3 mol-1)2 K1/2 and b 82.8 cm3 mol-1].
Solutions Manual 182

This gives,

PBs 113
. bar

Therefore,

f H A,B 7.01
MA 6.195
PBs 113
.

For the R-K equation, fugacity coefficients are given by:

f v b 2aAB vb abA vb b LM


Pv OP
ln M A ln  A 
v  b v  b RT b
3/2
ln
v

3/2
RT b 2
ln
v

vb
 ln
N
RT Q
Using b # bB , a # aB and v # v B (infinite dilution of A).
Solving for aAB ,

aAB 3.963 u 10 8 bar (cm3 mol-1)2 K1/2 P

Then, for the mixture,

a 4.159 u 10 8 bar (cm3 mol-1) K1/2

b 91.5 cm3 mol-1

Calculating again the pressure we obtain,


P 4.14 bar

6. Let 1 = C2H6 and 2 = C6H6.


The K factor of component i is defined as

yi M iL (P, x )
Ki
xi M Vi (P, y)

with

M c (0)  c (1) P  c (2) zi

Thus, we need to solve for P and y1 (or y2).


At equilibrium,

y1M1V x1M1L (1)


Solutions Manual 183

(1  y1 )M V2 (1  x1 )M 2L (2)

From the given equations, we rewrite Eqs.(1) and (2) for x1 = 0.263:

(12545
.  2.458 u 10 4 P  0.4091y1 ) y1 11699
.  0.008345P

[0.74265  7.0069 u 10 3 P  0.50456(1  y1 )](1  y1 ) 0.24596  0.001874 P

The above equations can be solved (either graphically or numerically) for P and y1:

P 59 atm

y1 0.715 ( y2 0.285)

Then,
y1 0.715
K1 2.72
x1 0.263

y2 0.285
K2 0.387
x2 0.737

[From Kay’s data: K1 2.73 and K 2 0.41 ].

7. The stability criterion is [see Eq. (6-131) of text]:

F w2 g E I  RT F 1  1 I  0
GH wx 2 JK GH x1 x2 JK
1 T ,P

We need an expression for gE valid at high pressures.


Because
F wg E I
GH wP JK vE
T ,x

we write

g E (T , P, x ) g E (T , P 1 atm, x )  z1
P E
v dP

(RT ) u (1877
. ) x1 x 2  z
1
P
x1 x 2 (4.026  0.233 ln P )dP
Solutions Manual 184

Thus,

g E (T , P, x ) (RT ) u (1877
. ) x1 x 2  (P  1) u (4.026) x1 x 2

 (P  1) u (0.233) x1 x2  0.233P (ln P ) x1 x 2

(42043  4.259P  0.233P ln P) x1 x 2 Ax1 x 2

For gE of this form ( g E Ax1 x 2 , where A is a constant), the stability criterion is (see Sec.
6.12):
A
!2
RT
or
42043  4.259P  0.233P ln P ! (2) u (82.0578) u (273)

Solving for P,
P 1046 atm (or 1060 bar)

At pressures higher than 1060 bar, the system splits into two phases.
To solve for the composition at a higher pressure, we use:

x1' J 1' x1" J 1"

(1  x1' )J '2 (1  x1" )J "2

where

RT ln J i (42043  4.259P  0.233P ln P ) x 2j

At 1500 atm (or 1520 bar) and 273 K,

ln J i 2.0477 x 2j

Thus,

x1' exp[2.0477(1  x1' ) 2 ] x1" exp[2.0477(1  x1" ) 2 ]

(1  x1' ) exp[2.0477( x1' ) 2 ] (1  x1" ) exp[2.0477( x1" ) 2 ]

Solving (either graphically or numerically), we obtain

x1' 0.37 ( x2' 0.63)

x1" 0.63 ( x2" 0.37)


Solutions Manual 185

8. We want to relate hE to volumetric data. Relations given in Chapter 3 of the text may be used.
We write hE at any pressure P relative to hE at 1 bar as:

h E (P )  h E (1 bar)
z LMMN
P

1
vE  T
F wv E I OP dP
GH wT JK P
PQ

Thus, we need the above integrand as a function of pressure at 333 K.


From volumetric data, using linear regression at each pressure between 323 K and 348 K,

F wv E I F wv E I
GH wT JK 0.0186 GH wT JK 0.0154
1 bar 100 bar

F wv E I F wv E I
GH wT JK 0.01239 GH wT JK 0.00963
250 bar 500 bar

Using linear interpolation, at 333 K,

v E (1 bar) 1.091 v E (100 bar) 0.9638

v E (250 bar) 0.8284 v E (500 bar) 0.6846

If
F wv E I
F (P) vE  T GH wT JK
P
then:

P (bar) 1 100 250 500

F(P) (J bar mol-1) -0.5102 -0.4164 -0.3297 -0.2522

Using a trapezoid-rule approximation,


360
128 J mol 1
³
1
F ( P)dP

Therefore, at 333K
360
h E (360 bar) h E (1 bar) 
³1 F ( P)dP

1445  128

h E (360 bar, 333K) 1317 J mol 1


Solutions Manual 186

9. For condensation to occur,


L V
fW ! fW

To find the temperature for condensation (at constant pressure and vapor composition), we
solve the equilibrium relation
L V
fW fW

The liquid phase is assumed to be pure water. Its fugacity is given by

L
fW L
f pure exp
z s
PW
P vW
RT
dP

As a good approximation, let


s s
L # P s exp v W ( P  Pw )
fW (obtain data from Steam Tables)
W
RT

Thus, we are neglecting M sW and we assume that (liquid) water is incompressible over the
s and P (150 atm).
pressure range between PW
The vapor phase is described by an equation of state. Therefore,

V
fW yW M VW P

To obtain M VW , we use the Redlich-Kwong equation of state:

RT a
P  (1)
v  b T 1/ 2v(v  b)

from which we obtain

v b §vb·
ln MW ln  W  (2¦ y j aWj )( RT 1.5b)1 ln ¨ ¸
vb vb j © v ¹
(2)
abW § vb b · Pv
 ¨ ln v  v  b ¸  ln RT
1.5 2
RT b © ¹

where v is the molar volume of the mixture and


Solutions Manual 187

a ¦¦ yi y j aij
i j

b ¦ yi bi
i

In these equations, aW and aCO2 are given as functions of temperature; the cross-coefficient
is

aij (ai (0) a j (0) )1/ 2  0.5R 2T 2.5 K

V Use the following procedure:


With a trial-and-error procedure, we can calculate f W .

1. Guess temperature.
2. Calculate v from equation of state [Eq. (1)].
3. Use T and v (along with P and y) to calculate M W from Eq. (2).
4. Calculate the fugacity of vapor.
5. Compare f W V with saturation pressure of water at that temperature.

Typical results are:

V
T (K) MW fW (atm)

475 0.588 17.7


500 0.633 19.9
525 0.711 21.3
550 0.746 22.4

V and L as a function of temperature (see figure), we see that


Plotting f W fW

L V
fW fW

at T | 482 K. That is the temperature where condensation first occurs (dew-point temperature of
the mixture).
Solutions Manual 188

100 L
fW
Fugacity of water, atm

75

50

V
fW
25

0
400 450 500 550 600
Temperature, K

Fugacity of water in vapor phase and in liquid phase at P = 150 atm.

10.
a) For equilibrium between solid solute and solute dissolved in the supercritical fluid,

s
f2 (P, T ) f2f (P, T , y2 )
or

d ln f2s d ln f2f (1)

where subscript 2 refers to solute and superscript f to fluid phase.


Expanding Eq. (1) with respect to T, P and composition (see Sec. 12.4), we obtain (tempera-
ture is constant):

F w ln f2s I
GG wT JJ 0
H K P, y
F w ln f2f I
GG wT JJ
H K P, y
0
Solutions Manual 189

F w ln f2s I
GH wx2 JK 0 (pure solid solute)
P,T

F w ln f2s I dP F w ln f2f I F fI
GH wP JK GG wP JJ dP  GG w lnwy2f2 JJ dy2
H K T ,y H K T ,P
(2)
T

But because

F w ln f2s I s
v2
GH wP JK RT
T

F w ln f I f
v2f
GG wP JJ 2
H K T
RT

Equation (2) becomes:

s
(v 2  v2f ) F w ln f2f I F w ln f2f I
GG wy2 JJ dy2 GG w ln y2 JJ d(ln y2 )
H K T ,Pdy H K T ,P
dP (3)
RT

Finally, because

f2f y2M 2 P

Equation (3) becomes

s
v 2  v2f
FG w ln y IJ
H wP K
RT
FG IJ
2 (4)
w ln M 2
H K
T 1
w ln y2 T ,P

b) Maxima (or minima) occur when

FG w ln y IJ
H wP K
2
0
T

Because w ln M 2 / w ln y2 is always greater than –0.4, the above derivative is zero when
v 2s v2f .
f
It is necessary, then, to calculate v2 as a function of pressure.
Using Eq. (12-41),
Solutions Manual 190

FG wP IJ
H wn2 K T ,V ,n1
v2 
FG wP IJ
H wV K T , all n
and the Redlich-Kwong equation of state with the mixing rules,

a ¦¦ xi x j aij
i j

b ¦ xi bi
i

we obtain

2(¦ xi a2i )  ab2 /(v  b)


RT § b · i
1 2 
v  b ¨© v  b ¸¹ v(v  b)T 1/ 2
v2
RT a ª 2v  b º

2 1/ 2 « 2 2»
( v  b) T ¬« v  (v  b) ¼»

Assuming that the fluid phase is almost pure solvent, v, a and b are those for pure solvent 1.
Cross parameter a12 is given by:

a12 (a11a22 )1/ 2 (1  k12 )

Constants are:

a11 0.7932 u 108 bar (cm3 mol 1 )2 K1/2


a22 0.11760 u 1010 bar (cm3 mol 1 )2 K1/2
a12 0.3264 u 109 bar (cm3 mol 1 )2 K1/2
b1 40.683 cm3 mol 1 b2 140.576 cm3 mol 1

Using volumetric data for ethylene at 318 K (IUPAC Tables), and because

s
v2
128174
.
112 cm3 mol 1
1144
.

the maximum (and minimum) occurs ( v2 s


v 2 ) at (see figure below)

minimum = 19 bar

maximum = 478 bar


These values are in good agreement with results shown in Fig. 5-39 of the text.
Solutions Manual 191

500 Solubility Solubility


minimum maximum

s
v 2 =112
0
10 100 1000 10,000 P (bar)

19 bar 478 bar


v2 (cm3 mol -1)

-500

-1000

-1500

-2000

Partial molar volumes of naphthalene infinitely dilute in ethylene at 318 K calculated from
Redlich-Kwong equation of state with k12 = 0.0182.

You might also like