Вы находитесь на странице: 1из 81

Chapter 5 Solutions

Engineering and Chemical Thermodynamics

Wyatt Tenhaeff
Milo Koretsky

Department of Chemical Engineering


Oregon State University

koretsm@engr.orst.edu
5.1

(a)
Following the example given by Equation 5.5a in the text

 u   u 
du    dT    dP
 T P  P T

(b)
 u   u 
du    dT    ds
 T  s  s T

(c)
 u   u 
du    dh    ds
 h  s  s  h

2
5.2.
The internal energy can be written as follows

 u   u 
du    dT    dv
 T v  v T

Substituting Equations 5.38 and 5.40

 u   u    P  
   cv and    T    P
 T v  v T   T  v 

into the above expression yields

  P  
du  cv dT  T    P  dv
  T  v 

From the ideal gas law, we have

 P  R
  
 T v v

Therefore,

 RT 
du  cv dT    P  dv
 v 

RT
which upon noting that P  for an ideal gas, becomes
v

du  cv dT

3
5.3
The heat capacity at constant pressure can be defined mathematically as follows

 h   u    Pv    u   v 
cP          P 
 T  P  T  P  T  v  T  P

For an ideal gas:

 v  R
  
 T  P P

Therefore,

 u 
cP    R
 T  v

One mathematical definition of du is

 u   u 
du    dT    dP
 T P  P T

 u 
We can now rewrite  T  :
 v

 u   u   T   u   P 
           cv
 T  v  T  P  T  v  P T  T  v

For an ideal gas:

 u 
  0
 P  T

so

 u 
cv   
 T  P

Substituting this result into our expression for c P gives

c P  cv  R

4
5.4
In terms of P, v, and T, the cyclic equation is

 P   T   v 
1       
 T  v  v  P  P T

For the ideal gas law:

Pv  RT

so the derivatives become:

 P  R
  
 T v v
 T  P
  
 v  P R
 v   RT v
   
 P T P 2 P

Therefore,

 P   T   v   R  P   v 
            1
 T  v  v  P  P  T  v  R  P 

The ideal gas law follows the cyclic rule.

5
5.5
For a pure species two independent, intensive properties constrains the state of the system. If we
specify these variables, all other properties are fixed. Thus, if we hold T and P constant h cannot
change, i.e.,

 h 
  0
 v  T , P

6
5.6
Expansion of the enthalpy term in the numerator results in

 h   Ts  vP 
   
  T s  T s
 h   P 
   v 
 T  s  T  s

Using a Maxwell relation

 h   s 
   v 
  T s  v  P
 h   s   T 
   v   
 T  s  T  P  v P

We can show that

 s  c
   P (use thermodynamic web)
 T P T
 T  1  RT 2a 2ab 
       (differentiate van der Waals EOS)
 v  P R  v  b v 2 v3 

Therefore,

 h  vc  RT 2a 2ab   v 2a  b  
   P      c P   1   
 T  s RT vb v 2 3
v   v  b vRT  v 

7
5.7

 h 
  :
 P T

 h   Ts  vP   s 
     T  v
 P T  P T  P T

 s 

 v
  
 P T  T
 R

   1  B' P  C ' P
P P
2

 h 
  
 P T
RT
P
 
1  B ' P  C ' P 2  v  v  v

 h 
  0
 P T

 h 
  :
 P  s

 h   Ts  vP 
    v
 P s  P s
 h 

 P
 
s
RT
P

1  B' P  C ' P 2 

 h 
  :
 T  P

 h 
   cP (Definition of cP)
 T  P

 h 
  :
 T s

 h   Ts  vP   P 
     v 
 T s  T s  T  s

 P   s   P  c  T  c P
  P   P
   
 T  s  T
 
 P  s T T  v 
 P T R 1  B' P  C ' P 2 
 h  Pv 1 
1  B' P  C ' P 2 
   cP
 T  s 
RT 1  B ' P  C ' P 2
 cP
 
1  B' P  C ' P 2 

8
 h 
   cP
 T  s

9
5.8

(a)
A sketch of the process is provided below

well m
insulated

T1 s = 0

P1 T2
P2

The diagram shows an infinitesimal amount of mass being placed on top of the piston of a
piston-cylinder assembly. The increase in mass causes the gas in the piston to be compressed.
Because the mass increases infinitesimally and the piston is well insulated, the compression is
reversible and adiabatic. For a reversible, adiabatic process the change in entropy is zero.
Therefore, the compression changes the internal energy of the gas at constant entropy as the
pressure increases.

(b)
To determine the sign of the relation, consider an energy balance on the piston. Neglecting
potential and kinetic energy changes, we obtain

U  Q  W

Since the process is adiabatic, the energy balance reduces to

U  W

As the pressure increases on the piston, the piston compresses. Positive work is done on the
system; hence, the change in internal energy is positive. We have justified the statement

 u 
  0
 P  s

10
5.9
(a)
By definition:

1  v 
  
v  T  P
and

1  v 
   
v  P  T

Dividing, we get:

 v 
 
  T  P  v   P 
     
  v   T  P  v  T
 
 P  T

where derivative inversion was used. Applying the cyclic rule:

 v   P   T 
1       
 T  P  v  T  P  v

Hence,

  P 
 
  T  v

(b)
If we write T = T(v,P), we get:

 T   T 
dT    dv    dP (1)
 v P  P  v

From Equations 5.33 and 5.36

cv  P  c  v 
ds  dT    dv  P dT    dP
T  T  v T  T  P

We can solve for dT to get:

T  P  T  v 
dT    dv    dP (2)
c P  cv  T  v c P  cv  T  P

11
For Equations 1 and 2 to be equal, each term on the left hand side must be equal. Hence,

 T  T  P 
    
 v  P c P  cv  T  v

or

 P   v    v 
c P  cv  T     T  
 T v  T P   T  P

where the result from part a was used. Applying the definition of the thermal expansion
coefficient:

 P   v  Tv 2
c P  cv  T     
 T  v  T  P 

12
5.10
We need data for acetone, benzene, and copper. A table of values for the molar volume, thermal
expansion coefficient and isothermal compressibility are taken from Table 4.4:

   
 m3 
Species v  10 6     10 3 K -1   1010 Pa -1
 mol 
Acetone 73.33 1.49 12.7
Benzene 86.89 1.24 9.4
Copper 7.11 0.0486 0.091

We can calculate the difference in heat capacity use the result from Problem 5.9b:

vT 2
c P  cv 

or


 73.33  10  6

 m3  



  293 K  1.49  10
3
K -1 
2

 mol    J 
c p  cv  
12.7  10 10
Pa -1
 37.6 
 mol  K 

 J   J 
Species c p  cv  cp  % difference
 mol  K   mol  K 
Acetone 37.6 125.6 30%
Benzene 41.6 135.6 31%
Copper 0.5 22.6 2%

We can compare values to that of the heat capacity given in Appendix A2.2. While we often
assume that cP and cv are equal for condensed phases, this may not be the case.

13
5.11
We know from Equations 4.71 and 4.72

1  v  1  v 
    and    
v  T  P v  P T

Maxwell relation:

 s   P 
   
 T  T  v
v

Employing the cyclic rule gives

 P   P   v 
      
 T  v  v T  T  P

which can be rewritten as


1  v 
 
 s   P  v  T  P
    
 v T  T  v  1  P 
 
v  v T

Therefore,

 s  
  
 v T 

Maxwell Relation:

 s   v 
    
 P  T  T P

From Equation 4.71:

 v 
   v
 T P

Therefore,

 s 
    v
 P T

14
5.12

(a)
An isochor on a Mollier diagram can be represented mathematically as

 h 
 
 s  v

This can be rewritten:

 h   Ts  vP   P 
    T  
s
 v  s v  s  v

Employing the appropriate Maxwell relation and cyclic rule results in

 h   T   s 
   T  v   
s
 v  s  v  v  T

We know

 T  T  s   P 
   and    
 s  v cv  v  T  T  v

For an ideal gas:

 s   P  R
    
 v  T  T  v v

Therefore,

 h  T R  R
  T  v  T 1  
 s  v cv v  c v 

(b)
In Part (a), we found

 h  T  P 
  T  v  
 s  v c v  T  v

For a van der Waals gas:

 P  R
  
 T  v v  b

15
Therefore,

 h  RT  v 
  T   
 s  v cv  v  b 

16
5.13

(a)
The cyclic rule can be employed to give

 T   T   s 
      
 P  s  s  P  P T

Substitution of Equations 5.19 and 5.31 yields

 T  T  v 
    
 P  s c P  T  P

For an ideal gas:

 v  R
  
 T P P

Therefore,

 T  RT 1 v
   
 P  s P cP cP

(b)
Separation of variables provides

T R P

T cP P

Integration provides

R
T  P  cP
ln  2   ln  2 
 T1   P1 

which can be rewritten as

R
T2  P2  cP
 
T1  P1 

The ideal gas law is now employed

R
P2 v 2  P2  cP
  
P1v1  P1 

17
 R   R 
1  1 
cP  cP
P2 v
2  P1
v
1

where

R c P  R cv 1
1   
cP cP cP k

If we raise both sides of the equation by a power of k, we find

P2 v 2k  P1v1k
 Pv k  const.

(c)
In Part (a), we found

 T  T  v 
    
 P s c P  T  P

Using the derivative inversion rule, we find for the van der Waals equation

 v  Rv 3  v  b 
  
 T  P RTv 3  2a v  b  2

Therefore,

 T  1 RTv 3  v  b 
  
 P  s c P RTv 3  2a  v  b  2

18
5.14
The development of Equation 5.48 is analogous to the development of Equation E5.3D. We
want to know how the heat capacity changes with pressure, so consider

 c P 
 
 P  T

which can be rewritten as

 c P     h      h  
         
 P T  P  T  P  T  T  P T  P

 h 
Consider the  P  term:
 T

 h   Ts  vP   s   v 
     T   v  T   v
 P T  P T  P T  T  P

 c P 
Substitution of this expression back into the equation for   results in
 P T
 c P      v  
      T    v 
 P T  T   T P  P
 c P T  v   2v 


  

  T     v 

 P T  T 
 P  T
T P 2 P
 T
 c P    2v 
   T  
 P T  T 2 
 P

Therefore,

c Preal Preal    2v  
 dc P    T    dP
  T 2  
P 
c Pideal Pideal 

and

Preal 
  2v  
real
cP  c ideal
P   T    dP
  2
Pideal   T
 
P 

19
5.15
In order to solve this problem we need to relate the change in entropy from 10 to 12 bar to the
change in molar volume (for which we have complete data). First, we can rewrite the change in
entropy as

12 bar
 s 
s  s 2  s1     dP
 P T
10 bar

Applying a Maxwell relation, we can relate the above equation to the change in molar volume:

12 bar 12 bar
 s   v 
s 2  s1    P T dP  s1     T  dP
P
10 bar 10 bar

As 10 bar:

 v   v  4
 m3 
     5.60  10  
 T  P  T P  kg  K 

At 12 bar:

 v   v  4
 m3 
     4.80  10  
 T  P  T  P  kg  K 

 v 
To integrate the above entropy equation, we need an expression that relates  T  to pressure.
 P
Thus, we will fit a line to the data. We obtain

 v    m3   m3 
    4.0  10   P  9.6  10 
10 4
  
 T  P  
 kg  K  Pa    kg  K 

Now integrate the equation to find the entropy:

 4.0 10 P  9.6 10 4 dP 5.4960  0.104  kgkJ K   5.392  kgkJ K 
1.210 6 Pa
10
s 2  s1 
1.010 6 Pa

20
5.16
A schematic of the process follows:

Propane in

v1 = 600 cm3/mol Turbine ws


T1= 350 oC
Propane P = 1 atm
2
out

We also know the ideal gas heat capacity from Table A.2.1:

cP
 1.213  28.785  10  3 T  8.824 10  6 T 2
R

Since this process is isentropic (s=0), we can construct a path such that the sum of s is zero.

(a) T, v as independent variables


Choosing T and v as the independent variables, (and changing T under ideal gas conditions), we
get:

Ideal
step 2 Gas
v2 ,T2
 s2
volume

s1
step 1

 s=0

v1 ,T 1

Temperature

or in mathematical terms:

However, From Equation 5.33:

c  P 
ds  v dT  dv
T T v

21
To get s1

RT a  P  R
P  2 so 
vb v T v v  b
and

v 2 P v
s1   ds    dv  2 R dv  R ln v2  b 
 
 
v1 T v v1 v  b v1  b 


or, using the ideal gas law, we can put s1 in terms of T2:

RT 2  b 
P
s1  R ln  2 
 v1  b 
 

For step 2

T2 T2
c 0.213  28.785  10  3 T  8.824  10  6 T 2
s 2   v dT  R  dT
T T
T1 623.15 K

Now add both steps

s  s1  s 2  0
 RT2 
 P  b
 ln  2   0.213 ln
 v1  b  
T2 
623.15
  28.785  10  3  T2  623.15 K  

8.824  10  6 2
2

T2   623.15 K  2 
 
 
Substitute

T1  623.15 K

v1  600 cm 3 /mol 
P2  1 atm
 cm 3  atm 
R  82.06  
 mol  K 

and solve for T2:

T2  448.3  K 

(b) T, P as independent variables

22
Choosing T and P as the independent variables, (and changing T under ideal gas conditions), we
get:

P1,T1
s=0

Pressure
s1

step 1
Ideal
step 2
Gas
P2,T2 s2

Temperature

Mathematically, the entropy is defined as follows

 s   s 
ds    dT    dP  0
 T P  P T

Using the appropriate relationships, the expression can be rewritten as

c  v 
ds  P dT    dP  0
T  T  P

For the van der Waals equation

 R 
  v  b 
 v   
  
 T P  RT 2a 
  
  v  b  2 v 3 
Therefore,

 R 
T2 P2   v  b 
cP  
s   T
dT    RT 2a 
dP  0
T1 P1
  
  v  b  2 v 3 

We can’t integrate the second term of the expression as it is, so we need to rewrite dP in terms
of the other variables. For the van der Waals equation at constant temperature:

23
 2a RT 
dP     dv
 v
3
 v  b  2 

Substituting this into the entropy expression, we get

T2 v2
1.213  28.785 10  3 T  8.824  10  6 T 2 R
s   T
dT    v  b  dv  0
T1 v1

Upon substituting

T1  623.15 K
v1  600 cm 3
RT2
v2  (gas acts ideally at 1 atm)
P2
 cm 3 
b  91  
 mol 
 cm 3  atm 
R  82.06  
 mol  K 

we obtain one equation for one unknown. Solving, we get

T2  448.3 K

24
5.17

(a)
Attractive forces dominate. If we examine the expression for z, we see that at any absolute
temperature and pressure, z  1. The intermolecular attractions cause the molar volume to
deviate negatively from ideality and are stronger than the repulsive interactions.

(b)
Energy balance:

h2  h1  q

Alternative 1: path through ideal gas state


Because the gas is not ideal under these conditions, we have to create a hypothetical path that
connects the initial and final states through three steps. One hypothetical path is shown below:

P [bar]

P,T 1 P,T 2
q = h
50
step 1

step 3

h1 h3

h2 step 2
0
ideal gas

300 500 T [K]

Choosing T and P as the independent properties:

 h   h 
dh    dT    dP
 T  P  P T

or using Equation 5.46

  v  
dh  c P dT   T    v  dP
  T P 

The given EOS can be rewritten as

1 
v  R  aT 1 / 2 
P 

Taking the derivative gives:

25
 v  R  0.5
    0.5aRT
 T  P P

so


dh  c P dT  0.5aRT 0.5 dP 
For step 1

 0.5aRT1 dP  0.5aRT10.5 P  252  mol


0
0. 5 J 
h1  
50 bar

For step 2

 3.58  3.02 10 


500 K
3  J 
h2  R T  0.875T  0.5 dT  7961 
300 K
 mol 

For step 3:

 0.5aRT2 dP  0.5aRT20.5 P  323  mol


50 bar
0.5 J 
h3  
0

Finally summing up the three terms, we get,

 J 
q  h1  h2  h3  7888 
 mol 

Alternative 2: real heat capacity


For a real gas

real
h  c P

From Equation 5.48:

P real   2v 
c Preal ideal
 cP   T  dP
 2 
P ideak  T P
For the given EOS

1 
v  R  aT 1 / 2 
 P 
Therefore,

26
  2v 
   0.25aRT 1.5
 T 2 
 P
and

 
P real 50 bar

P real   2v 
  T 2
T  dP 
  0.25aRT 0.5 dP  0.875 K 1/2 RT 0.5

P ideak
 P P ideak 0 bar

real
We can combine this result with the expression for cP and find the enthalpy change.

 3.58  3.02 10 


500 K
3
h  R T  0.875T  0.5 dT
300 K
 J 
q  h  7888 
 mol 

The answers is equivalent to that calculated in alternative 1

27
5.18

(a)
Calculate the temperature of the gas using the van der Waals equation. The van der Waals
equation is given by:

RT a
P  2
vb v

First, we need to find the molar volume and pressure of state 1.

V
v1  1 
Al

  
0 .1 m 2  0 .4  m    m3 
 0.00016  
n n 250  mol  mol 

10000  kg   9.81  m2  


mg  s  
P1   Patm   1.01325 105  Pa   1.08 10 6  Pa 
A  
0.1 m 2

Substituting these equations into the van der Waals equation above gives

  J  m3 
 J  0 .5  
 8.314  T
 1
  mol  K    mol 
1.08  10  Pa  
6

 m3   3  2
5 m  3 
0.00016    4  10    0.00016  m  
 mol   mol   
 mol  
T1  297.5 K

Since the process is isothermal, the following path can be used to calculate internal energy:

v2 ,T2 = T1
v

u
s

v1 ,T1

Thus, we can write the change in internal energy as:

28
 u   u   u 
du    dT    dv    dv
 T v  v T  v T

Using Equation 5.40

v2
  P  
u   T  T  v  P  dv
v1

For the van der Waals EOS:

RT a
P  2
vb v
so

 P  R
  
 T  v v  b

Therefore,

v2
a
u   v 2 dv
v1

We can assume the gas in state 2 is an ideal gas since the final pressure is atmospheric.
Therefore, we calculate v 2 ,

RT2  m3 
v2   0.0244  
P2  mol 

and

 J  m3 
0.0244
0 .5  
 mol   J 
u   v 2
dv  3104.5 
 mol 
0.00016
or

U  776.1  kJ 

(b)
From the definition of entropy:

29
suniv  s sys  s surr

First, let’s solve for s sys using the thermodynamic web.

 s   s 
ds sys    dT    dv
 T v  v T

Since the process is isothermal,

 s 
ds sys    dv
 v T
v2
 P 
 s sys    T  v dv
v1

Again, for the van der Waals equation,

 P  R
  
 T  v v  b

Substitution of this expression into the equation for entropy yields

v2
R
s sys   v  b dv
v1
 J 
8.314 
 mol  K  dv  44.17  J 
0.0244
s sys    mol  K 
 3
0.00016 v  4  10  5 m
 
 mol 
J
S sys  11042 .5  
K 

The change in entropy of the surroundings will be calculated as follows

Q
s surr  surr
Tsurr

where

Qsurr  Q (Q is the heat transfer for the system)

Application of the first law provides

Q  U  W

30
We know the change in internal energy from part a, so let’s calculate W using

v2
W   n  Pdv
v1

Since the external pressure is constant,

 

  m3 
W   250  mol  1.01325  105  Pa   0.0244 
 m3  
  0.00016  
 
 mol 
 
 mol  
W  614030  J 

Now calculate heat transfer.

Q  776100  J      614030  J   1.39  10 6  J 

Therefore,

 1.39  10 6  J  J
S surr   4672  
297.5  K  K 

and the entropy change of the universe is:

J J J


S univ  11042.5    4672    6370.5  
K K  K 

31
5.19
First, calculate the initial and final pressure of the system.

Pi  10  10 5  Pa  
 20000  kg   9.81 m/s 2   4.92  10 6  Pa 
 
0.05 m 2

 Pa    30000  kg  9.81 m/s   6.89  10


2
P f  10  105 6
 Pa 
0.05 m  2

To find the final temperature, we can perform an energy balance. Since the system is well-
insulated, all of the work done by adding the third block is converted into internal energy. The
energy balance is

u  w

To find the work, we need the initial and final molar volumes, which we can obtain from the
given EOS:

vi  8.37  10 4 m 3 /mol  
vf 
8.314T f

 3.2  10 - 5 m 3 /mol 
 

6.89  10 6 1 

25 
T f 

Now, calculate the work


 
 
 
  
6
w   P f v f  vi   6.89  10 Pa 
8 .314 T f
 - 5
 3.2  10  8.37  10  4 



6 . 89  10

6 

1 
25 
T



 


  f  
We also need to find an expression for the change in internal energy with only one variable: Tf.
To find the change in internal energy, we can create a hypothetical path shown below:
step 2
v ideal gas
step 1

step 3

vi ,Ti
u = -Pf (vf -vi)

vf ,Tf

500 Tf T

For step 1, we calculate the change in internal energy as follows

32
v  RT / Plow v  RT / Plow
 u    P  
u1     dv 
 v T
 T    P dv
  T  v 
vi vi
v  RT / Plow  2  2
 aRTi  dv  aRTi ln RTi / Plow  b 
u1     T  a  2   v  b   T  a  2  vi  b 

vi  i  i

Similarly, for step 3:

vf v
 u 
f
  P  
u3    v T  dv   T  T  v  P dv
v  RT / Plow v  RT / Plow
vf  aRT 2  dv
 f  aRT f 2  vf b 
u3    T  a 2   v  b  T  a 2  RT f / Plow  b 


ln 
  
v  RT / Plow  f  f  

Insert the expression for the final molar volume into the equation for u 3 :

aRT f 2  8.314T f 
ln 
u3 
  
T f  a 2  6.89  106 1  25 / T f RT f / Plow  b    

Since the pressure is low (molar volume is big) during the second step, we can use the ideal heat
capacity to calculate the change in internal energy.

Tf Tf
u 2  
cv dT    20  R  0.05T  dT
Ti  500 K Ti  500 K


u 2  11 .686 T f  500  0.025 T f2  500 2   
If we set the sum of the three steps in the internal energy calculation equal to the work and
choose an arbitrary value for Plow, 100 Pa for example, we obtain one equation with one
unknown:

aRTi 2
 Ti  a  2
 RTi / Plow  b 
ln
vi  b
  11.686 T f  500  0.025 T f2  500 2     
 
aRT f 2  8.314T f 
ln 
  
T f  a 2  6.89  10 6 1  25 / T f RT f / Plow  b    
 
 
 

 6.89  10 6 Pa 
8 .314T f
  3.2  10 - 5  8.37  10  4 
 6 
 6.89  10 1 

25 
Tf 
  


  

33
Solving for Tf we get

T f  536.2 K

The piston-cylinder assembly is well-insulated, so

suniv  s sys

Since the gas in the cylinder is not ideal, we must construct a hypothetical path, such as one
shown below, to calculate the change in entropy during this process.

P
Pf ,T f
6.9

4.9 Pi ,T i

step 3
step 1

ideal gas
Plow
step 2

500 536 T

For steps 1 and 3

Plow P
 s 
low
  v  
s1     dP       dP
Pi
 P T Pi 
 T P 
Pf P
 s 
f
  v  
s3     dP       dP
Plow
 P T Plow 
 T P 

We can differentiate the given EOS as required:

Plow
  RT  2a  T    RTi  2a  Ti   Plow 
   a  T  2 P  dP   a  T  2 ln Pi 
s1  i i

Pi  i  i
Pf

  RT 2a  T
f f  dP   RT f  2a  T f  ln Pf 
s3    a  T 2 P
   a  T f  2 P 
 low 
Plow  f 

For step 2

34
Tf Tf Tf
 s  c  20 
s 2    T  dT   P dT   
P T  T
 0.05  dT

Ti Ti Ti

Tf 
s2  20 ln 
  0.05 T f  Ti
 
 Ti 

Sum all of the steps to obtain the change in entropy for the entire process

suniv  s sys  s1  s2  s3

suniv 
 RTi  2a  Ti  P  Tf
ln low   20 ln

 
  0.05 T f  Ti 

RT f 2a  T fln
 Pf 
 
 a  Ti  2  Pi   Ti

 a Tf 2 P 
 low 
Arbitrarily choose Plow (try 100 Pa), substitute numerical values, and evaluate:

 J 
suniv  0.388 
 mol  K 
  J  J
S univ   2 mol   0.388     0.766  
  mol  K   K 

35
5.20
A schematic of the process is given by:
well
insulated

V=1L V=2L
V=1L
T = 500 K T=?
Vacuum
nCO =1 mole
nCO=1 mole

State i State f

(a)
The following equation was developed in Chapter 5:

v   2  
 P
cvreal  cvideal   T  dv
  T 2  
v ideal   v 

For the van der Waals EOS

 2P 
 0
 T 2 
 

Therefore,

cvreal  cvideal

From Appendix A.2:

 3100   J 
cvreal  R 3.376  5.57  10  4  500    R  22.0 
2  mol  K 
 500 

(b)
As the diaphragm ruptures, the total internal energy of the system remains constant. Because the
volume available to the molecules increases, the average distance between molecules also
increases. Due to the increase in intermolecular distances, the potential energies increase. Since
the total internal energy does not change, the kinetic energy must compensate by decreasing.
Therefore, the temperature, which is a manifestation of molecular kinetic energy, decreases.

(c)

36
Because the heat capacity is ideal under these circumstances we can create a two-step
hypothetical path to connect the initial and final states. One hypothetical path is shown below:

v vf ,Tf

u2

step 2
s
2
step 1
vi ,Ti
u , s
1 1

Tf 500 T

For the first section of the path, we have

Tf Tf
real ideal
u1   cv dT   cv dT
Ti Ti

Tf
u1  R


2.376  5.57  10
4
T
3100 
T2 
 dT 
500 K i 

 
u1  2.32  10  3 T f2  19.75T f 
25773.4
Tf
 10507.4

For the second step, we can use the following equation

vf
 u 
u 2    v T dv
vi

If we apply Equation 5.40, we can rewrite the above equation as

vf
  P  
u 2   T  T  v  P  dv
vi

RT a
For the van der Waals EOS, P   2 ,
v b v

 P  R
  
 T  v v  b

Therefore,

37
v f  0.002 v f  0.002
 RT  a  J 
u1    v  b  P dv   v 2 dv  73.7  mol 
v i  0.001 v i  0.001

Now set the sum of the two internal energies equal to zero and solve for Tf:

 
u1  u 2  2.32  10  3 T f2  19.75T f 
25773.4
Tf
 10507.4  73.7  0

T f  497 K

(d)
Since the system is well-insulated

suniv  s sys

To solve for the change in entropy use the following development:

 s   s 
ds sys    dT    dv
 T v  v T

Using the thermodynamic web, the following relationships can be proven

 s  cv
  
 T  v T
 s   P 
   
 T  T  v
v

For the van der Waals EOS

 P  R
  
 T  v  v  b 

Now we can combine everything and calculate the change in entropy

497 K 0.002
cv R
s sys   T
dT    v  b  dv
500 K 0.001

497 K  2.376
  
0.002
3100  dv
s sys  R     5.57  10  4    
 dT  
500 K  T 3 5 
T  0.001 v  3.95  10
 J 
suniv  s sys  5.80 
 mol  K 

38
5.21
A schematic of the process is given by:
well
insulated

V = 0.1 m3 V = 0.1 m3 V = 0.2 m3


T = 300 K T=?
Vacuum
nA =400 moles
nA =400 moles

State i State f

Energy balance:

u  0

Because the gas is not ideal under these conditions, we have to create a hypothetical path that
connects the initial and final states through three steps. One hypothetical path is shown below:

step 2
v [m 3/mol] ideal gas
step 3

step 1

0.2/400 vf ,Tf

u

0.1/400 vi ,Ti

Tf 300 T [K]

For the first section of the path, we have

v 
 u 
u1     dv
 v T
vi

If we apply Equation 5.40, we can rewrite the above equation as

v 
  P  
u1   T    P  dv
  T  v 
vi

39
For the van der Waals EOS

 P  R a
   
 T  v v  b T v 2
2

Therefore,

v  v 
 RT a  2a  J 
u1   4  v  b  Tv 2  P  dv   4 T v 2 dv  1120  mol 
v i  2.510 v i  2.510 i

Similarly for step 3:

v f  510 4 v f  510 4
 RT a  2a  168000  J 
u3   v  b 
 Tv 2
 P  dv 

 Tf v 2
dv 
Tf 
 mol  K 
v  v 

For step 2, the molar volume is infinite, so we can use the ideal heat capacity given in the
problem statement to calculate the change in internal energy:

u 2 
3
2

R T f  300 K 
If we set sum of the changes in internal energy for each step, we obtain one equation for one
unknown:

 J  3
u1  u 2  u3  1120  
 J 
  8.314    T f  300 K  
 168000
0 
 mol  2   mol  K   Tf

Solve for Tf:

T f  261.6 K

40
5.22
A schematic of the process is shown below:

Ethane 3 MPa; 500K

T surr = 293 K
Initially:
vacuum

(a)
Consider the tank as the system. Since kinetic and potential energy effects are negligible, the
open system, unsteady-state energy balance (Equation 2.47) is

 dU 
    n in hin   n out hout  Q  W s
 dt  sys in out

The process is adiabatic and no shaft work is done. Furthermore, there is one inlet stream and no
outlet stream. The energy balance reduces to

 dU 
   n in hin
 dt  sys

Integration must now be performed

U2 t

 dU   n
U1 0
in hin dt

t
n2 u 2  n1u1  hin  n in dt  nin hin  (n2  n1 )hin
0

Since the tank is initially a vacuum, n1=0, and the relation reduces to:

u 2  hin

41
As is typical for problems involving the thermodynamic web, this problem can be solved in
several possible ways. To illustrate we present two alternatives below:

Alternative 1: path through ideal gas state


Substituting the definition of enthalpy:

u 2  u in  Pin vin

or

u 2  at 3 MPa, T   uin  at 3 MPa, 500 K   Pin vin (1)

From the equation of state:

  J 
Pin vin  RT 1  B' P    8.314    
 552 K  1  2.8  10 8  3  10 6 Pa   3,800 
 J 
  mol  K    mol 
(2)

The change in internal energy can be found from the following path:

3 MPa
step 1

step 3

u1 u3

u2
Plow
step 2 ideal gas

500 K T2 T

For steps 1 and 3, we need to determine how the internal energy changes with pressure at
constant temperature: From the fundamental property relation and the appropriate Maxwell
relation:

 u   s   v   v   v 
   T   P   T    P 
 P  T  P  T  P  T  T  P  P  T

From the equation of state

 u 
  
RT
1  B'P   P  RT2    B' RT
 P T P  P 

42
So for step 1:

0 0
 u 
u1  Pin  P T dP   Pin B RTdP B RT Pin  349 [J/mol]
' '
(3)

and for step 3:

P2 P2
 u 
u 3  0  P T dP   0 B RTdP   B RT P2  0.7T
' '
(4)

For step 2

T
 u 
T
 h   Pv  
T
u 2     dT       dT    c P  R dT
500 
T  P 500 
T  P  T  P  500

or

T2

 0.131  19.225 10 


3
u 2  R T  5.56110 6 T 2 dT (5)
T1 500 K

Substituting Equations 2, 3, 4, and 5 into 1 and solving for T gives:

T2  552 K

Alternative 2: real heat capacity


Starting with:

u 2  hin

The above equation is equivalent to

h2  P2 v2  hin
 h2  hin  P2 v2

To calculate the enthalpy difference, we can use the real heat capacity

P    2v  
real
cP ideal
 cP   T   dP
  T 2  
P 
Pideal 

For the truncated viral equation,

43
  2v 
  0
 T 2 
 P

Therefore,

real ideal
cP  cP

Now, we can calculate the change in enthalpy and equate it to the flow work term.

T2
ideal
 cP dT  P2 v 2
T1  500 K
T2

 1.131  19.225  10 
T  5.561 10 6 T 2 dT  P2 v2  RT2 1  B' P2 
3
R
T1 500 K

Integrate and solve for T2:

T2  552 K

(b)
In order to solve the problem, we will need to find the final pressure. To do so, first we need to
calculate the molar volume. Using the information from Part (a) and the truncated virial
equation to do this

  J 
 8.314   552 K 
v
RT
P
1  B ' P     mol  K  
6
3  10 Pa
 
1  2.8  10  8 3  10 6 Pa 
 m3 
v  0.0014  
 mol 

This quantity will not change as the tank cools, so now we can calculate the final pressure.

  m3  
P2  0.0014  
 


 J
 mol

 
 
 1  2.8  10  8 P2
 8.314   293 K 
  mol  K  

Solve for P2 :

P2  1.66  10 6 Pa

44
The entropy change of the universe can be expressed as follows:

S univ  S sys  S surr

To solve for the change in entropy of the system start with the following relationship:

 s   s 
ds sys    dT    dP
 T P  P T

Alternative 1: path through ideal gas state


Using the proper relationships, the above equation can be rewritten as

cP  v 
ds sys  dT    dP
T  T  P

We can then use the following solution path:

3 MPa

s
step 3

1.66 MPa 3
step 1

s1
s
2
Plow
step 2 ideal gas

500 K T2 T

Choosing a value of 1 Pa for Plow, for step 1:

1 Pa 1 Pa
 v 
 dP    1  B P  dP
R
s1  
'

1.66 MPa 
T P 1.66 MPa
P

For step 3,

3 MPa 3 MPa
 v 
s1     dP   
R
1  B ' P  dP
1 Pa  T  P 1 Pa
P

For step 2:

45
293 K 293 K
cP 1.131 
s 2   dT  R    19.225  10 3  5.561 10 6 T  dT
552 K
T 552 K 
T 

Adding together steps 1, 2 and 3:

 J 
s sys  46.9 
 mol  K 
______________________________________________________________________________

Alternative 2: real heat capacity


Using the proper relationships, the above equation can be rewritten as

c real  v 
ds sys  P dT    dP
T  T  P

For the truncated virial equation

 v  1 
   R  B ' 
 T P P 

Now, substitute the proper values into the expression for entropy and integrate:

293 K 1.6610 6 Pa
1.131 3 6  1 
s sys  R   T  19.225  10  5.561 10 T  dT  R ^  P  B'dP
552 K 310 Pa
 J 
s sys  46.9 
 mol  K 
______________________________________________________________________________

In order to calculate the change in entropy of the surroundings, first perform an energy balance.

u  q

Rewrite the above equation as follows

h   Pv   q

Since the real heat capacity is equal to ideal heat capacity and the molar volume does not change,
we obtain the following equation

Tf

 cP
ideal

dT  v P f  Pi  q 
Ti

46
T f  293K
R  1.131 19.225 10 3
 
 mol

 m3  
T  5.561 10  6 T 2 dT   0.0014 

6 6

  1.66  10 Pa - 3  10 Pa  q
Ti  552 K    
 J 
q  15845 
 mol 

Therefore,

 J 
q surr  15845 
 mol 

and

 J 
s surr  54.08 
 mol  K 

Before combining the two entropies to obtain the entropy change of the universe, find the
number of moles in the tank.

n
 
0.05 m 3
 75.7 mol
 m3 
0.0014  
 mol 

Now, calculate the entropy change of the universe.

  J   J 
S univ   75.7 mol  54.08    46.9  
  mol  K   mol  K  
J
S univ  544  
K

47
5.23
First, focus on the numerator of the second term of the expression given in the problem
statement. We can rewrite the numerator as follows:

uTr , v r  uTideal
r ,vr

 uTr , v r  uTideal
r ,vr  
 
 uTideal
r ,vr
 uTideal
r ,vr 

For an ideal gas, we know

uTideal  uTideal 0
r ,vr r ,vr 

Therefore,

uTr , v r  uTideal
,vr
 uTr , v r  uTideal
,vr  
r r

Substitute this relationship into the expression given in the problem statement:

uTdep uTr , v r  uTideal uTr , v r  uTideal


,v r ,vr r ,vr  
r r
 
RTc RTc RTc

ideal
Now, we need to find an expression for uT , v  uT , v   . Note that the temperature is constant.
r r r r

Equation 5.41 reduces to the following at constant temperature:

  P  
duT  T    P  dv
  T v 

The pressure can be written as

zRT
P
v

and substituted into the expression for the differential internal energy

  RT  z  RT  zRT   RT 2  z  
duT  T      
  dv      dv
  v  T  v v 
v
v 
 
 v  T  v 

Applying the Principle of Corresponding States

duTrT 2  z  
 r    dvr
RTc  vr
  Tr vr 

If we integrate the above expression, we obtain

48
v duTr uTr , v r  uTideal
v 
vr  2
T  z  
  r, v
   r    dv r
v 
RTc RTc  vr
v    Tr vr 

Therefore,

uTdep
,v
uTr , v r  uTideal
, vr
uTr , v r  uTideal
, vr  
vr  2
T  z  
r r
 r
 r
   r    dvr
RTc RTc RTc v
v  r  Tr vr 

49
5.24
We write enthalpy in terms of the independent variables T and v:

 h   h 
dh    dT    dv
 T  v  v T

using the fundamental property relation:

dh  Tds  vdP

At constant temperature, we get:

  P   P  
dhT  T    v   dv
  T  v  v T 

For the Redlich-Kwong EOS

 P  R 1 a
   
 T  v v  b 2 T 3 / 2
v v  b 

 P  RT a a
    
 v T  v  b  2 T 1 / 2 v 2  v  b  T 1 / 2 v v  b  2

Therefore,

 RT RTv 3 a a 
dhT       dv
 v  b  v  b  2 2 T 1 / 2 v v  b  T 1 / 2  v  b  2 

To find the enthalpy departure function, we can integrate as follows

v v  RT 
RTv 3 a a
h dep   dhT        dv
v  v  b  v  b
v  
2 2T 1 / 2
v v  b  T 1 / 2
 v  b  
2

Since temperature is constant, we obtain

RTb 3a  v  a
h dep   ln   1/ 2
v  b 2bT 1 / 2 vb T  v  b

To calculate the entropy departure we need to be careful. From Equation 5.64, we have:

sT , P  sTideal
,P
gas

 sT , P  sTideal  
gas ideal gas
, P  0  sT , P  sTideal gas
, P 0 
However, since we have a P explicit equation of state, we want to put this equation in terms of v.
Let’s look at converting each state. The first two states are straight -forward

50
sT , P  sT , v

and

sTideal gas ideal gas


, P  0  sT , v  

For the third state, however, we must realize that the ideal gas volume v’ at the T and P of the
system is different from the volume of the system, v. In order to see this we can compare the
equation of state for an ideal gas at T and P

RT
P
v'

to a real gas at T and P

RT a
P 
v b T v v  b 

The volume calculated by the ideal gas equation, v’, is clearly different from the volume, v,
calculated by the Redlich-Kwong equation. Hence:

gas  ideal gas gas 


sTideal
,P
gas
 s ideal gas
 sTideal
,v s '  sTideal
,v 
T ,v '  T ,v 

Thus,

sT , P  sTideal
,P
gas

 sT , v  sTideal gas
,v   
 sTideal
,v
gas
 sTideal gas
,v   
  s ideal
 T ,v
'
gas
 sTideal
,v
gas 

Using a Maxwell relation:

 ds   P 
   
 T  T  v
dv

Therefore,

 P 
dsT    dv
 T  v

For the Redlich-Kwong EOS

 P  R 1 a
   
 T  v v  b 2 T 3 / 2
v v  b 

so

51
s   Rv 
1 a
T ,v  sTideal gas
,v       dv
v  b 2 T
v  
3 / 2
v v  b  

For an ideal gas

 P  R
  
 T  v v

so

s 
v
ideal gas R
T ,v  sTideal gas
,v     v  dv
v 

Finally:

v'
dv v' RT
s ideal gas
 sTideal
,v
gas
 R  R ln  R ln
T ,v ' v v Pv
v

Integrating and adding together the three terms gives:

s dep  R ln
 v  b  a  v  RT
ln   R ln
v 2bT 3/ 2 vb Pv

52
5.25
Calculate the reduced temperature and pressure:

Tc  647.3  K 
Pc  220.48  bar  (Table A.1.2)
w  0.344

300  bar 
Pr   1.36
220.48  bar 
673.15 K
Tr   1.04
647.3 K

By double interpolation of data from Tables C.3 and C.4


( 0) (1)
 h dep   h dep 
 Tr , Pr   2.921  Tr , Pr   1.459
 RTc   RTc 

 
 
 

From Tables C.5 and C.6:


( 0) (1)
 s dep   s dep 
 Tr , Pr   2.292  Tr , Pr   1.405
 R   R 

 
 
 

Now we can calculate the departure functions

 dep 
( 0)
 h dep 
(1) 
  hTr , Pr  
 w 
dep Tr , Pr
h  RTc  
   RTc 
  RTc 
     
  J   J 
h dep   8.314    647.3    2.921  0.344  1.459   18421  
  mol  K    mol 

 dep 
( 0)
 s dep 
(1) 
 sTr , Pr 
 R    w 
dep Tr , Pr
s 
 R   R 
    
  
  J   J 
s dep   8.314     2.292  0.344  1.405   23.07 
  mol  K    mol  K 

To use the steam tables for calculating the departure functions, we can use the following
relationships.
dep
h  hT , P  hTideal
,P

s dep  sT , P  sTideal
,P

53
From the steam tables

 kJ   kJ 
hT , P  2151.0   and sT , P  4.4728  
 kg   kg  K 

We need to calculate the ideal enthalpies and entropies using the steam tables’ reference state.

673.15 K
hTideal
, P  h
vap
 0 .01º C   c ideal
p dT
273.16 K

vap  kJ 
We can get h  45.1  from the steam tables and heat capacity data from Table A.2.2.
 mol 
Using this information, we obtain

673.15 K 
 kJ    kJ   3 0.121  10 5 
hTideal
, P  45.1  
  0.008314 
 mol  K    3.47  1 .45  10 T   dT
 mol     273.16 K  T2 
 kJ 
, P  59.14 
hTideal
 mol 

Now, calculate the ideal entropy.

673.15 K c ideal
p P 
sTideal
, P  s
vap
 0 .01º C   dT  R ln 2 
T  P1 
273.16 K

From the steam tables:

 kJ 
s vap  0.01 º C   0.165 
 mol  K 

Substitute values into the entropy expression:

 kJ    
673.15 K 
 3.47 3 0.121 105   30
sTideal
,P  0.165  
 0 .008314 
 mol  K       1.45  10   dT  ln 
    273.16 K  T T3   0.000613
 J 
sTideal
, P  107  mol  K 
 

Now, calculate the departure functions:

 kJ   kJ   kJ 
h dep  2151.0    0.0180148  kg/mol   59.14    20.4 
 kg   mol   mol 
 kJ 
s dep  4.4728    0.0180148  kg/mol    0.107  kJ   0.0264  kJ 
 kg  K   mol  K   mol  K 

54
Table of Results
Generalized Percent Difference
Steam Tables
Tables (Based on steam tables)
 kJ 
h dep  -18.62 -20.4 9.9
 mol 
 kJ 
s dep  -0.0231 -0.0264 12.5
 mol  K 

55
5.26
State 1 is at 300 K and 30 bar. State 2 is at 400 K and 50 bar. The reduced temperature and
pressures are

30  bar  50  bar 
P1, r   0.616 P2,r   1.026
48.74  bar  48.74  bar 
300 K 400 K
T1, r   0.982 T2,r   1.31
305.4 K 305.4 K

and

  0.099

By double interpolation of data in Tables C.3 and C.4


(0) (1)
 h dep   h dep 
 T1, r , P1, r   T1, r , P1, r 
 RT   0.825    0.799
c RTc

 
 
 

(0) (1)
 h dep   h dep 
 T2 , r , P2 , r   T2 , r , P2 , r 
   0.711    0.196
RTc RTc

 
 
 

Therefore,

 h dep 
 T1, r , P1, r 
 RT   0.825  0.099  0.799   0.904
 c 
 
 h dep 
 T2 , r , P2 , r   0.711  0.099  0.196   0.730
 RTc 

 

The ideal enthalpy change from 300 K to 400 K can be calculated using ideal cP data from Table
A.2.1.

400 K
3
hTideal
T
R 1.131  19.225 10 T  5.561 10  6T 2 dT  717.39 R
1 2
300 K

The total entropy change is

h   hTdep, P  hTideal
T2
 hTdep, P
1, r 1, r 1 2,r 2, r

h  R     0.904TC  717.39  0.730TC 


  J 
h   8.314   0.904 305.4 K   717.39 K  0.730 305.4 K  
  mol  K  

56
 J 
h  6406.2 
 mol 

Using the data in Table C.5 and C.6


 s dep 
 T1, r , P1, r   0.601  0.099  0.756  0.676
 R 

 

 s dep 
 T2 , r , P2 , r 
 R   0.394  0.099  0.224  0.416

 

Substituting heat capacity data into Equation 3.62, we get

 400 K 1.131  19.225  10 3 T  5.561  10 6 T 2  50 bar 


s ideal  R   dT  ln 
300 K T  30 bar 

s ideal  1.542 R

Therefore,

s  sTdep, P  s ideal  sTdep, P  R 0.676  1.542  0.416 


1, r 1, r 2, r 2, r

 J 
s  14.98  
 mol  K

57
5.27
The turbine is isentropic. Therefore, we know the following

s   sTdep, P  s ideal  sTdep, P  0


1, r 1, r 2,r 2,r

Using the van der Waals EOS, we can find P1,r, which leaves one unknown in the above
equation: T2.

  3   atm  cm 3 
 82.06  cm  atm   623.15 K  91  10 5  
  mol  K    mol 2 
P1   
  3  3  2
 600  cm   91  cm     3 
 600  cm  
  mol   mol    
   mol  
P1  75.19  atm  76.19  bar 

Calculate reduced temperature and pressures using data from Table A.1.1

76.19  bar  1.013  bar 


P1, r   1.8 P2,r   0.024
42.44  bar  42.44  bar 
623.15 K
T1, r   1.68
370.0 K

Also,

  0.152

From Tables C.5 and C.6:

 s dep 
 T1, r , P1, r   0.327  0.152  0.102  0.343
 R 

 

Substituting heat capacity data into Equation 3.62, we get

 T2 1.213  28.785  10  3 T  8.824  10  6 T 2 


ideal  1 atm 
s  R  dT  ln 
 T  75.19 atm 
623.15 K 

Therefore,

 T2  s dep 
 1.213  28.785  10 3 T  8.824  10 6 T 2  T2 , r , P2 , r    s
R 0.343   T
dT  4.32   R 
 623.15 K 
 

 
We can solve this using a guess-and-check method

T2  600 K : T2,r  1.62

58
 J 
s  33.84  
 mol  K 

T2  450 K : T2,r  1.22


 J 
s  0.77  
 mol  K 

T2  446.6 K : T2,r  1.21


 J 
s  0 
 mol  K 

Therefore,

T2  446.6 K

59
5.28
A reversible process requires the minimum amount of work. Since the process is reversible and
adiabatic

s  0

which can be rewritten as


dep dep
s   sT , P  s ideal  sT , P  0
1, r 1, r 2,r 2,r

Calculate reduced temperature and pressures using data from Table A.1.1

1  bar  10  bar 
P1, r   0.0217 P2,r   0.217
46.0  bar  46.0  bar 
300 K
T1, r   1.57
190.6 K

From Tables C.5 and C.6:

 s dep 
 T1, r , P1, r 
 R   0.00457  0.008  0.0028   0.0046

 

Substituting heat capacity data into Equation 3.62, we get

 T2 1.702  9.081  10 3T  2.164  106 T 2 


ideal  10 bar 
s  R  dT  ln 
 T  1 bar 
300 K 

Therefore,

 T2  s dep 
 1.213  28.785  10 3 T  8.824  10 6 T 2  T2 , r , P2 , r 
s  R 0.0046   dT  2.303   
T R
 300 K 
 

 
We can solve using a guess-and-check method

T2  400 K : T2,r  2.10


 J 
s  4.98 
 mol  K 

T2  385 K : T2,r  2.02


 J 
s  1.42 
 mol  K 

T2  379 K : T2,r  1.99

60
 J 
s  0.018  
 mol  K 

Therefore,

T2  379 K

An energy balance reveals that

h2  h1  h  ws

We can calculate the enthalpy using departure functions. From Tables C.3 and C.4:

 h dep 
 T1, r , P1, r 
 RT   0.0965  0.008  0.011  0.0966
 c 
 
 h dep 
 T2 , r , P2 , r 
 RTc   0.0614  0.0089 0.015  0.0613

 

Ideal heat capacity data can be used to determine the ideal change in enthalpy

379 K 
h ideal
 R   1.702  9.081  10  3 T  2.164  10  6 T 2 dT 
300 K 

Therefore,

  
379 K
  J
h   8.314  
  190. 6 K  0. 0966  0. 0613   1.702  9 .081  10 3
T  2. 164  10 6 2
T dT 
  mol  K    300 K

 J 
 s  h  3034.2 
w 
 mol 
and
  mol    J 
W S  1 / 30    3034.2     101.1 W
  s    mol  

61
5.29
Equation 4.71 states

1  v 
   
v  T  P
 v 
 v   
 T  P

This can be substituted into Equation 5.75 to give

v T  1
 JT 
cP

62
5.30
For an ideal gas

 v  R
  
 T P P

Therefore,

 RT 
  v
 JT  
P    v  v  0
cP cP

This result could also be reasoned from a physical argument.

63
5.31
The van der Waals equation is given by:

RT a
P  (1)
v  b v2

The thermal expansion coefficient is given by:

1  v   1   T 
        (2)
v  T  P  v   v  P

Solving Equation 1 for T:

 a  v  b 
T   P  2  
 v  R 

Differentiating by applying the chain rule,

 T   a  1  v  b  2a Pv 3  av  2ab
    P  2      (3)
 v P  v  R  R  v3 Rv 3

Substitution into Equation 2 gives

Rv 2

Pv 3  av  2ab

Substituting Equation 1 for P gives b in terms of R, T, v, a , and b:

Rv 2  v  b 

RTv 3  2a  v  b 
2

The isothermal compressibility is given by:

1  v   1   P 
         
v  P  T  v   v  T

From the van der Waals equation:

 P  2a  RTv 3  2a  v  b 
2
RT
    
 v T  v  b 2 v 3 v 3  v  b
2

so

64
v 2  v  b
2

RTv 3  2a v  b 
2

For the Joule-Thomson coefficient, we can use Equation 5.75:

Substituting the van der Waals equation into Equation 3 gives

 T  RTv 3  2a v  b  2a  v  b 
2
T
     (4)
 v  P  v  b  Rv 3
 v  b Rv 3

Thus, the second derivative becomes:

  2T  T 1  T  2 a 6a  v  b 
 2        3
 v P  v  b  v  b  v  P Rv
2
Rv 4

or simplifying using Equation 4,

  2T  2a  v  3b 
 2   (5)
 v P Rv 4

Substituting Equations 5 and 4 into Equation 5.75 gives:

 bRTv 3  2av v  b 
2

RTv 3  2a v  b 
2
 JT  Preal
 RTv 4 
cP   
ideal
dP
Pideal 
2a v  3b  

At a given temperature the integral in pressure can be rewritten in terms of volume using the van
der Waals equation to give:
 bRTv 3  2av v  b 
2

RTv 3  2a v  b 
2
 JT 
 RTv  RTv 3  2a v  b  2
vreal

c Pideal  
videal
 2a v  3b  
   v  b  2
dv

65
5.32
We can solve this problem by using the form of the Joule-Thomson coefficient given in Equation
5.75. The following approximation can be made

 vˆ   vˆ 
   
 T  P  T  P

At 300 ºC,

 vˆ  vˆ 350 º C,1MPa   vˆ 250 º C,1MPa 


  
 T P 350  250 º C
 m3   m3 
0.28247    0.23268  
 vˆ   kg 
  kg 

  
 T  P 350  250 º C
 vˆ   m3   m3 
   0.0005    0.0005  
 T  P  kg  º C   kg  K 

A similar process was followed to find cP.

 hˆ   ˆ
cˆ P      h 
  
 dT  P  T  P

At 300 ºC,

 hˆ  hˆ 350 º C,1MPa   hˆ 250 º C,1MPa 


  
 T  350  250 º C
 P
 kJ   kJ 
3157.7    2942.6  
 hˆ   kg   kg 
  
 T  350  250 º C
 P
 hˆ   ˆ    
cˆ P      h   2.15  kJ   2.15  kJ 
  T 
 dT P  P  kg  º C   kg  K 

Now,  JT can be found.

 3 
  573.15 K   0.0005  m    0.25794  m  
3
  vˆ 
T      
  T   vˆ    
 kg  K   
 kg  
 
 JT   P  
cˆ P  kJ 
2.15  
 kg  K 

 m3  K 
 JT  0.0133  
 kJ 

66
67
5.33
At the inversion line, the Joule-Thomson coefficient is zero. From Equation 5.75:

This is true when the numerator is zero, i.e.,

For the van der Waals equation, we have

RT a
P  2
vb v

Solving for T:

 a  v  b 
T   P  2  
 v  R 
so

 T   a  1  v  b  2a Pv 3  av  2ab
    P     
 v  P  v2  R  R  v3 Rv 3

Substituting for P:

 T  RTv 3  2a v  b 
2

  
 v  P  v  b  Rv 3

Hence,

 v   bRTv 3  2av v  b 
2
T  v  0 
 T  P RTv 3  2a v  b 
2

Solving for T:

2av v  b 
2
T (1)
bRv 3

Substituting this value of T back into the van der Waals equation gives

2av v  b  a a 2v  3b 
P  2  (2)
bv 3 v bv 2

68
We can solve Equations 1 and 2 by picking a value of v and solving for T and P. For N2, the
critical temperature and pressure are given by Tc = 126.2 [K] and Pc = 33.84 [bar], respectively.
Thus, we can find the van der Waals constants a and b:

Using these values in Equations (1) and (2), we get the following plot:

Joule-Thomson inversion line

1000

800

600
P [bar]

400

200

0
0 100 200 300 400
T [K]

69
5.34
We can solve this problem using departure functions, so first find the reduced temperatures and
pressures.

50  bar 10  bar 
P1, r   0.99 P2,r   0.2
50.36  bar  50.36  bar 
273.15 K
T1, r   0.967
282.4 K

Since the ethylene is in two-phase equilibrium when it leaves the throttling device, the
temperature is constrained. From the vapor-liquid dome in Figure 5.5:

T2,r  0.76
 T2  214.6

The process is isenthalpic, so the following expression holds

h   hTdep, P  hTideal
T2
 hTdep, P  0
1, r 1, r 1 2, r 2,r

Therefore,
dep dep
hT , P  hT , P  hTideal
2, r 2,r 1, r 1, r 1 T2

From Table A.2.1:

 1.424  14.394  10 
214.6 K
3
hTideal
T
R T  4.392  10  6 T 2 dT
1 2
273.15 K

From Tables C.3 and C.4    0.085 :

 h dep 
 T1, r , P1, r   3.678  0.085  3.51  3.976
 RTc 

 

Now we can solve for the enthalpy departure at state 2.

hTdep, P  214.6 K 
1
2,r 2,r
   3.976   1.424  14.394  10  3 T  4.392  10  6 T 2 dT 
RTc  282.4 K 
 273.15 K 
hTdep, P
2,r 2,r
 3.01
RTc

We can calculate the quality of the water using the following relation

70
hTdep, P hTdep,, liq
P
hTdep,, vap
P
 1  x 
2,r 2, r 2, r 2,r 2 ,r 2,r
x
RTc RTc RTc

where x represents the quality. From Figures 5.5 and 5.6:

hTdep,, liq
P
2,r 2 ,r
 4.6  0.085  5.5  5.068
RTc
hTdep,, vap
P
2,r 2 ,r
 0.4  0.0859  0.75  0.464
RTc

Thus,

x  0.447

55.3% of the inlet stream is liquefied.

71
5.35
Density is calculated from molar volume as follows:

MW

v

2
Substitute the above into the expression for Vsound :

 
 
P   1  P 
2
Vsound   
  MW   MW   1 / v   s
  
  v  s

The following can be shown using differentials:

1 v
   
v v2

Therefore,

2  P  v 2  P 
Vsound       
   s MW  v  s

72
5.36
From Problem 5.35:

2  P  v 2  P 
Vsound       
   s MW  v  s

The thermodynamic web gives:

 P   P   s   T   P   T   s   P   s 
                   
 v  s  s  v  v  P  v  s  T  s  s  v  v T  s T  T  P
 P   T  s   P   c  c  P   T 
         P    P    
 v  s  cv  v T  s T  T   cv  T  v  v  P

If we treat air as an ideal gas consisting of diatomic molecules only

cP 7  P  R  T  P
      
cv 5  T v v  v P R

Therefore,

 P   7  P 
     
 v  s  5  v 
and

v 2  7  P  7  RT 
Vsound       
MW  5  v  5  MW 

Vsound  343  m/s

The lightening bolt is 1360 m away.

73
5.37
From Problem 5.35:

2  P  v 2  P 
Vsound       
   s MW  v  s

The thermodynamic web gives:

 P   P   s   T   P   T   s   P   s 
                   
 v  s  s  v  v  P  v  s  T  s  s  v  v T  s T  T P

so
 P   T  P   T   c 
         P 
 v  s  cv  T  v  v  P  T 

For liquids c P  cv water at 20 ºC, so

 P   P   T 
      
 v  s  T  v  v P

However, the cyclic rule gives:

 P   T   v 
1       
 T v  v  P  P T

So

 P   P 
   
 v  s  v T

From the steam tables, for saturated water at 20 oC:

 m3 
P = 2.34 kPa and vˆ  .001002  
 kg 

For subcooled water at 20 oC:

 m3 
P = 5 MPa and vˆ  .0009995  
 kg 
So

74
 P   P   P  5000  2.34  kg kPa 
       
 vˆ s  vˆ T  vˆ  .0009995  .001002  m3 

and
 P 
Vsound  vˆ 2    1414  m/s
 vˆ  s

75
5.38

(a)
The fundamental property relation for internal energy is

dU  Qrev  Wrev

Substituting the proper relationships for work and heat, we obtain

dU  TdS  Fdz

The fundamental property relation for the Helmholtz energy is

dA  dU  d  TS   dU  TdS  SdT

Substitute the expression for the internal energy differential:

dA  Fdz  SdT

(b)
First, relate the entropy differential to temperature and length.

 S   S 
dS    dT    dz
 T z  Z T

Now we need to find expressions for the partial derivatives.

 u   TS  Fz   S 
nc z  n     T 
 T z  T z  T z

Therefore,

 S  nc a 
   z  n  b 
 T  z T T 

The following statement is true mathematically (order of differentiation does not matter):

  A     A  
      
Z  T  z  T T  Z T  z

Furthermore,

  A    S 
     
Z  T  z  T  Z T

76
  A    F 
    
T  Z T  z  T  z

 S   F 
      k  z  z0 
 Z T  T  z

Substituting the expressions for the partial derivatives into the expression for the entropy
differential, we obtain

a 
dS  n  b dT  k  z  z 0  dz
T 

(c)
First, start with an expression for the internal energy differential:

 U   U 
dU    dT    dz
 T z  Z T

From information given in the problem statement:

 U 
   n a  bT 
 T  z

Using the expression for internal energy developed in Part (a) and information from Part (b)

 U   S 
   T   F  T   k  z  z 0    kT  z  z 0   0
 Z T  z T

Therefore,

dU   n a  bT   dT  0dz   n a  bT   dT

(d)
We showed in Part (c) that

 U 
FU    0
 z T

 S 
Using the expression for  z  developed in Part B, we obtain
 T

 S 
FS  T    kT  z  z0 
 z T

77
(e)
First, perform an energy balance for the adiabatic process.

dU  W

Substitute expressions for internal energy and work.

 n a  bT   dT  Fdz  kT  z  z 0  dz
Rearrangement gives

dT kT  z  z 0 

dz  n a  bT  

The right-hand side of the above equation is always positive, so the temperature increases as the
rubber is stretched.

78
5.39
The second law states that for a process to be possible,

suniv  0

To see if this condition is satisfied, we must add the entropy change of the system to the entropy
change of the surroundings. For this isothermal process, the entropy change can be written

cv  P   P 
ds  dT    dv    dv
T  T v  T  v

Applying the van der Waals equation:

R
ds  dv
v b

Integrating

v2  J 
s sys  R ln  11 .5 
v1  mol K 

For the entropy change of the surroundings, we use the value of heat given in Example 5.2:

 J 
q   q surr  600 
 mol 

Hence the entropy change of the surroundings is:

q surr  600  J 
s surr    1.6 
Tsurr 373  mol K 

and
 J 
suniv  s sys  s surr  9.9 
 mol K 

Since the entropy change of the universe is positive we say this process is possible and that it is
irreversible.

Under these conditions propane exhibits attractive intermolecular forces (dispersion). The closer
they are together, on average, the lower the energy. That we need to put work into this system
says that the work needed to separate the propane molecules is greater than the work we get out
during the irreversible expansion.

79
5.40
A schematic of the process is given by:

Gas A in

Pi = 100 bar Turbine ws


Ti = 600 K
Gas A Pf = 20 bar
out
Tf = 445 K

The energy balance for this process is provided below:

h  w S

Because the gas is not ideal under these conditions, we have to create a hypothetical path that
connects the initial and final states through three steps. One hypothetical path is shown below:

100 Pi ,Ti
ideal gas

P [bar]

step 1

h

Pf ,Tf
20
step 3

step 2
Plow

445 600 T [K]

For the first section of the path, we have

P0
 h 
h1     dP
 P T
P1

If we apply Equation 5.45 we can rewrite the above equation as

P0
  v  
h1    Ti 
 T
  v  dP
P
Pi  

For the given EOS:

80
 v  R aP
    2
 T  P P Ti

Therefore,

P 0 v 0
 aP   2aP   J 
h1   5   P 
Ti
 v 

dP   
T
 b dP  2467  
 mol 
Pi 10010 Pi 10010 5  i 

Similarly for step 3

Pf  20 10 5
 2aP   J 
  b dP  250 
h3    Tf   mol 
P 0  

For step 2, the pressure is zero, so we can use the ideal heat capacity given in the problem
statement to calculate the enthalpy change.

Tf 445 K
 J 
h2   c P dT    30  0.02T dT  6270 
 mol 
Ti 600 K

Now sum each part to find the total change in enthalpy:

 J 
h  h1  h2  h3  8487 
 mol 
 J 
ws  8487 
 mol 

In other words, for every mole of gas that flows through the turbine, 8487 joules of work are
produced.

81

Вам также может понравиться