Download as pdf or txt
Download as pdf or txt
You are on page 1of 62

FOCUS 4

THERMODYNAMICS

4A.2 (a) The internal energy of an open system could be increased by (1)
adding matter to the system, (2) doing work on the system, and (3) adding
heat to the system. (b) Matter cannot be added to a closed system, so
only (2) doing work on the system and (3) adding heat to the system could
be used in increase the internal energy. (c) The internal energy of an
isolated system cannot be changed.

4A.4 (a) Given four cylinders operating at 60 cycles per minute and displacing
3.6 L each under a constant pressure of 1800 bar (where 1 bar ≈ 1 atm), the
work done is given by:
w   PV   (4)(60)(1800 atm)(3.6 L )(101.325 J L1  atm1 )  1.58 108 J

(b) The expanding gas in the cylinder does work on the pistons (and to the
load connected to the pistons by the crankshaft of the engine). Therefore,
work done by the gas is negative.

(c) The internal energy of a system may be changed by doing work, and in
the absence of other changes, U  w ; change in internal energy is then

1.58 × 108 J.

4A.6 (a) Irreversible expansion:


w  PV
w  (2.00 atm)(1.00 L)
w  203 J
(b) Reversible expansion:
184 Focus 4 Thermodynamics

V2
w  nRT ln
V1
25C  298.15 K
 2.50 L 
w  (0.250 mol)(8.3145 J K 1 mol1 )(298.15 K ) ln 
 1.50 L 
w  317 J
Reversible expansion does more work.

4A.8 (a) The heat change will be made up of two terms: one term to raise the
temperature of the stainless steel and the other to raise the temperature of
the water:
q  (300.0g)(4.18J  (C) 1  g 1 )(100.0C  20.0C)
 (400.0g)(0.51J  (C) 1  g 1 )(100.0C  20.0C)
 1.00  105 J  1.63  10 4 J  1.17  10 5 J  1.17  10 2 kJ = 1.2  10 2 kJ
(b) The percentage of heat attributable to raising the temperature of the
 1.00  10 5 J
water will be  (100)  86%
 1.17  10
5
J 

(c) The use of the copper kettle is more efficient, as a larger percentage of
the heat goes into heating the water and not the container holding it.

4A.10 Heat lost by metal =  heat gained by water.


(18.0g )(24.8C  100C)(Cs )  (50.2g )(4.18J(C)1g 1 )(24.8C  22.0C)
Cs  0.434 J (C)1g 1

4A.12
1.685 g benzoic acid
( 1
)( 3228 kJmol1 benzoic acid)
122.12 gmol benzoic acid
Ccal =
2.821 °C
= 15.79 kJ (°C)1
Focus 4 Thermodynamics 185

4A.14
q
C
T
0.90 kJ
Ccalorimeter 
2.85K
Ccalorimeter  0.32 kJ K 1

qreaction  qcalorimeter  0
qreaction  qcalorimeter
q reaction  CT
q reaction  (0.32 kJ K 1 )(1.31K )
q reaction  0.42 kJ

4B.2 (a) The change in internal energy U is given simply by summing the two
energy terms involved in this process. We must be careful, however, that
the signs on the energy changes are appropriate. In this case, internal
energy will be added to the gas sample by heating, but the expansion will
remove internal energy from the sample:
U  242 kJ  171 kJ  71 kJ
(b) If the heat added had exactly matched the amount of energy lost to the
work done by the gas (i.e., if q had been 171 kJ), then U would be 0 and
the temperature of the gas would not have changed. Because more heat
was added, the temperature of the gas would have increased, and the
pressure of the gas would be higher at the end.

4B.4 To calculate this, we use the relationship ∆U = q + w, which arises from


the first law of thermodynamics. Because the system absorbs heat, q will
be a positive number, as will U , because it is an increase in internal
energy:
∆U = q + w
152 kJ = 164 kJ + w
186 Focus 4 Thermodynamics

w = – 12 kJ
Because work is negative, the system is doing work on the surroundings.

4B.6 If the heater operates as rated, then the total amount of heat transferred to
the cylinder will be:
(100. J·s1)(20.0 min)(60.0 s·min1) = 1.20 × 105 J or 120. kJ

Work will be given by w   Pext V in this case because it is an expansion


against a constant opposing pressure:
w = –(0.876 atm)(6.00L - 1.00 L) = – 4.38 L·atm

In order to combine the two terms to get the internal energy change, we
must first convert the units on the energy terms to the same values. To get
the conversion factor for L  atm to J or vice versa one can use the
equivalence of the ideal gas constant R in terms of L·atm and J:
. .
w=
.

w = – 444 J or – 0.444 kJ
The internal energy change is then the sum of these two numbers:
∆U = q + w = 120. kJ + (– 0.444 kJ) = 120. kJ
The energy change due to the work term turns out to be negligible in this
problem.

4B.8 a) Energy lost by animal doing work is:

w = mgh  (0.275 kg)(9.81m s2 )(1.01 m)  2.7247 kg m2 s2  2.72 J


Since the animal is doing work, w = -2.72 J
Animal also lost 7.8 J of energy as heat.
U = q + w  ( 7.8 J )  ( 2.7247 J )  10.5 J
b) If the experiment took place in a space station, no work would be done
because of zero gravity. The change in internal energy would be less, as it
would be due solely to the energy lost as heat.
Focus 4 Thermodynamics 187

∆U = q + w = (–7.8 J) + (0 J) = –7.8 J

4B.10 (a) True only if q = 0; (b) always true, volume is fixed and no work can
be done; (c) never true; (d) always true; (e) true if q = 0

4B.12 (a) A solid is dissolving in solution, and as indicated, there is no


volume change resulting in w = 0. However, the thermometer shows a
temperature increase, which means ΔE is positive and q must be positive.
(b) During this reaction in a pistonlike cylinder, the system expands
and w must be negative. Since the temperature drops, q is either negative
or zero.

4B.14 (a) Because the process is isothermal, U  0 and q   w. For a


reversible process:
V2
w  nRT ln
V1
n is obtained from the ideal gas law:
PV (2.57 atm)(3.42 L)
n   0.359 mol
RT (0.082 06 L  atm  K 1  mol1 )(298 K)
7.39
w  (0.359 mol)(8.314 J  K 1  mol1 )(298 K) ln  685 J
3.42
q   685 J
(b) For step 1, because the volume is constant, w  0 and U  q.
In step 2, there is an irreversible expansion against a constant opposing
pressure, which is calculated from ∆ .
The constant opposing pressure is given, and V can be obtained from

Vfinal  Vinitial  7.39 L  3.42 L


w  (1.19 atm)(7.39 L  3.42 L)
 101.325 J 
 4.72 L  atm  (  4.72 L  atm)    479 J
 1 L  atm 
The total work for part B is 0 J + (− 479 J) = − 479 J.
188 Focus 4 Thermodynamics

4B.16 The estimate considers only translational and rotational contributions to


the molar internal energy, which are 3/2 RT and RT, respectively, for a
linear molecule.

3/2 RT + RT = 5/2 RT = 6.20 kJ·mol-1

4C.2 Molecules have a higher heat capacity than monatomic gases because they
store energy in the form of rotations and vibrations, in addition to simple
translational kinetic energy, which is the only energy storage mode
available to monatomic gases. The heat capacity of C2H6 is larger than that
of CH4 because it is a more complicated molecule that has more
possibilities for the molecule to rotate and for more bonds to vibrate than
does CH4.

4C.4 (a) qp = nCp∆T = n(5/2 R)∆T


1150 J = (0.640 mol)(5/2 × 8.314 J·mol-1·K-1)(Tf – 298 K)

Tf = 384 K

At constant pressure, qp = ∆H = 1150 J

(b) qv = nCv∆T = n(3/2 R) ∆T

1150 J = (0.640 mol)(3/2 × 8.314 J·mol-1·K-1)(Tf – 298 K)

Tf = 442 K

4C.6 (a) NO is a linear molecule. Translational contribution is 3/2 RT and


rotational contribution is RT. Total contribution from molecular motions
will be 5/2 R.
(b) NH3 is a polyatomic, nonlinear molecule. Translational and rotational
contributions are both 3/2 RT. Total contribution from molecular motions
will be 3R.
Focus 4 Thermodynamics 189

(c) Hypochlorous acid is a polyatomic, nonlinear molecule. Translational


and rotational contributions are both 3/2 RT. Total contribution from
molecular motions will be 3R.
(d) SO2 is a polyatomic, nonlinear molecule. Translational and rotational
contributions are both 3/2 RT. Total contribution from molecular motions
will be 3R.

4C.8 (a) H freezing  H fus

(4.01 kJ)
H fus    5.09 kJ  mol1
 25.23 g 
 1 
 32.04 g  mol 
37.5 kJ
(b) H vap   31 kJ  mol1
 95 g 
 1 
 78.11 g  mol 

4C.10 (a) The heat change will be made up of two terms: one term to raise the
temperature of the stainless steel and the other to raise the temperature of
the water.
q  (400.0 g)(4.18 J  (C) 1  g 1 )(100.0C  22.0C)
 (500.0 g)(0.51 J  (C) 1  g 1 )(100.0C  22.0C)
 1.30  10 5 J  1.99  10 4 J  1.50  10 5 J  1.5  10 2 kJ
(b) The percentage of heat attributable to raising the temperature of the
 1.30  10 5 J
water will be  (100)  87% .
 1.50  10
5
J 

(c) The use of the copper kettle is more efficient, as a larger percentage of
the heat goes into heating the water and not the container holding it.
190 Focus 4 Thermodynamics

4C.12 This process is composed of two steps: raising the temperature of the
liquid water from 25 C to 100. C and then converting the liquid water to
steam at 100. C.
Step 1: ∆H = (276g)(4.18J (C) 1 g 1 )(100.C  25.C)  86.5 kJ
276g
Step 2: ∆H  ( 1
)(40.7kJ mol 1 )  623 kJ
18.02g mol
Total heat required = 86.5 kJ + 623 kJ = 710 kJ

4C.14 The heat lost by the metal must equal the heat gained by the water. Also,
the final temperature of the metal must be the same as that of the water.
We can set up the following relationship and solve for the specific heat
capacity of the metal: qmetal  qwater

(25.0 g)(Specific heat) metal (29.8C  90.0C)

 (50.0 g)(4.184 J  (C)1  g 1 )(29.8C  25.0C)


(Specific heat) metal  0.665 J  (C) 1  g 1
Focus 4 Thermodynamics 191

4C.16 Heating curves are usually plots of temperature versus heating time.
However it is instructive to plot the x-axis showing the energy transferred.
Notice the small plateau at 58.8 oC; it indicates phase changes, in this case
from liquid to vapor. Also notice the steeper slope after this phase
transition due to the molar heat capacity of bromine vapor being lower
than the liquid phase.

Heating Curve for Bromine Liq → Gas

80

70

60

50

Temp. (Celsius)
40

30

20

10

0
0 1000 2000 3000 4000 5000 6000

-10

-20
Energy (J)

. .
4D.2 nSO2 = = = 0.375 mol SO2
. /
.
nO2 = = 0.3125 mol O2
. /

The SO2, the limiting reactant, can react with 0.188 mol of O2:

heat evolved = (0.375 mol SO2) = –37 kJ

4D.4 (a) C5H10O5  5O2  5CO2(g)  5H2O(l)

(b) Benzoic acid C6 H 5COOH , 122.12 g.mol1


192 Focus 4 Thermodynamics

0.825g
 6.755 103 mol
122.12 g. mol1
Heat released on combustion of 0.825g:
q  (6.755  10 3 mol)(3251 kJ·mol 1 )
 21.96 kJ
Heat capacity of bomb calorimeter:
q
C
T
21.96 kJ

1.94C
 11.32 kJ·C 1
d-Ribose, C5H10O5 , 150.13 g·mol–1

0.727g
 4.842  10 3 mol
150.13g·mol 1
Bomb calorimeter is a fixed volume.  U = q + w = q
Heat released by reaction, q(released) = heat gained by calorimeter, q(gained)
 q (released)  q (gained)  0
q ( released)  q ( gained)
 CT
 (11.32 kJ·C 1 )(22.72C  21.81C)
 10.30 kJ

On combustion of 0.727g (or 4.842  10 3 mol ) of D-ribose


U  q ( released )  10.30 kJ
 10.30kJ
Per mole of D-ribose, U 
4.842 10  3 mol
= –2127.5 kJ·mol-1
Rounding off at the end of the calculation using three significant figures:
∆U = −2.13  10-3 kJ·mol-1
Focus 4 Thermodynamics 193

4D.6 ∆U = ∆H + w
−68 kJ = ∆H + 25 kJ
∆H = −93 kJ

4D.8
P V  n RT
PV  ( 2 mol)(8.314 J K 1 mol 1 )( 298 K )
PV  4.96 kJ
H   U  P V
 684 kJ  U  4.96 kJ
U  679 kJ

4D.10 (a) Since the burn rate is 7.6 mol per minute, the total moles burned in the
natural gas mixture must add up to 7.6 mol per minute.
Dividing each of the given amounts by 1000 gives:
5.4 mol CH4 + 1.8 mol C2H6 + 0.23 mol C3H8 + 0.17 mol C4H10
= 7.6 mol (which is the moles of natural gas mixture burned per minute).
To calculate the mass of CO2 produced, we need to realize that there will
be one mole of CO2 produced per mole of carbon present in the starting
compound. So there will be one mole of CO2 produced per mole of CH4,
two per mole of C2H6, three per mole of C3H8, and four per mole of C4H10.
Total moles of CO2 produced per minute:
= (1 mol CO2/1 mol CH4) x 5.4 mol CH4 + (2 mol CO2/1 mol C2H6) x 1.8
mol C2H6 + (3 mol CO2/1 mol C3H8) x 0.23 mol C3H8 + (4 mol CO2/1 mol
C4H10) x 0.17 mol C4H10 = 10.4 mol.min–1

The mass of CO2 produced per minute:


10.4 mol·min-1 x 44.01 g·mol–1 = 4.6 x 102 g·min–1

(b) From the appendix we can find the enthalpies of combustion of the
gases involved:
Compound Enthalpy of combustion (kJ  mol 1 )
194 Focus 4 Thermodynamics

CH4(g), methane 890


C2H6(g), ethane 1560
C3H8(g), propane 2220
C4H10(g), butane 2878
The heat released per minute will be given by the enthalpies of
combustion multiplied by the number of moles of each gas combusted in
that time period:
Heat released = (5.4 mol CH4)(–890 kJ·mol–1 CH4) +
(1.8 mol C2H6)(–1560 kJ·mol–1 C2H6) +
(0.23 mol C3H8)(–2220 kJ·mol–1 C3H8) +
(0.17 mol C4H10)( –2878 kJ·mol–1 C4H10)
heat released = –8.6 × 103 kJ·min–1

3
 30.48 cm 
4D.12 (a) (12 ft  12 ft  8 ft)  7
  3.26  10 cm
3

 1 ft 
The heat capacity of air is 1.01 J·oC.g–1, and the average molar mass of air
is 28.97 g  mol1 (see Table 4.1). The density of air can be calculated
from the ideal gas law PV = nRT:
P( molar mass)
d
RT
(1.00 atm)(28.97 g·mol1 )
 1 1
 1.27 g·L1  1.27 x 10  3 g·cm 3
(0.08206 L·atm·K ·mol )(277.6 K )

40F  4.44C, 78F  25.55C


T  25.55C  4.44C  21.1C

The heat required is:


q = (mass)(specific heat capacity)(change in temperature)

= ( 1.27 103 g·cm3 )(3.26 ×107 cm3)(1.01 J·oC·g-1)(21.1 oC)


= 883.5 kJ
= 8.8 × 102 kJ
Focus 4 Thermodynamics 195

The mass of octane required to produce this much heat will be given by:
 8.8 102 kJ
1
(114.22 g.mol1 )  1.8 101 g
 5471 kJ·mol

(b)
 (1.0 gal) (3.785  103 mL  gal 1 ) (0.70 g  mL1 ) 
H   
 114.22 g  mol 1 
  10 942 kJ 
  
 2 mol octane 
  1.3  105 kJ

4D.14 The first reaction is reversed and added to the second:


2 NO(g) 
 N 2 (g)  O 2 (g) H   180.5 kJ

N 2 (g)  2 O 2 (g) 
 2 NO 2 (g) H   66.4 kJ

2 NO(g)  O 2 (g) 
 2 NO 2 (g)

H   180.5 kJ  66.4 kJ  114.1 kJ

4D.16 First, write the balanced equations for the known reactions:
CH 4 (g)  2 O 2 (g) 
 CO 2 (g)  2 H 2 O(l) H   890 kJ

CO(g)  1
2 O 2 (g) 
 CO 2 (g) H   283.0 kJ

The desired reaction can be obtained by multiplying the first reaction by 2


and adding it to the reverse of the second reaction, also multiplied by 2.
2 CH 4 (g)  4 O 2 (g) 
 2 CO 2 (g)  4 H 2 O(l) H   1780 kJ

2 CO 2 (g) 
 2 CO(g)  O 2 (g) H   566.0 kJ

2 CH 4 (g)  3 O 2 (g) 
 2 CO(g)  4 H 2 O(l)

H   1780 kJ  566.0 kJ  1214 kJ

4D.18 ∆Ho = [∆Hof (MgCl2, aq) + ∆Hof (H2O, l) + ∆Hof (CO2, g)]
– [∆Hof (MgCO3, s) + ∆Hof (HCl, aq)]
∆Ho = (-641.8 kJ  mol-1) + (-285.83 kJ  mol-1) + (-393.51 kJ  mol-1)
196 Focus 4 Thermodynamics

– [(-1095.8 kJ  mol-1) + 2(-167.16 kJ mol-1)]


∆Ho = 109.0 kJ  mol-1

4D.20 2 Al(s) + 6 HBr(aq) → 2 AlBr3(aq) + 3 H2(g) ΔHo = –1061 kJ


6 HBr(g) → 6 HBr(aq) ΔHo = 6(–81.15 kJ)
3 H2(g) + 3 Br2(g) → 6 HBr(g) ΔHo = 3(–72.80 kJ)
2 AlBr3(aq) → 2 AlBr3(s) ΔHo = 2(–368 kJ)
2 Al(s) + 3 Br2(g) → 2 AlBr3(s) ΔHo = –2502 kJ

4D.22 (a)
H   [3( H f (S, s))  2( H f ( H 2O,l ))]  [2( H f ( H 2S, g )) ( H f (SO2 ,g ))]
 [(3 mol)(0 kJ mol1 )  (2 mol)(285.83 kJ mol1 )] 
[(2 mol)(20.63 kJ mol1 )  (1 mol)(296.83 kJ mol1 )]
 233.57 kJ

(b)
H   [4(H f ( NO, g))  6(H f (H 2 O,g))]  [4(H f ( NH 3 , g))  5(H f (O 2 ,g))]
 [(4 mol)(90.25 kJ mol1 )  (6 mol)(241.82 kJ mol 1 )] 
[(4 mol)(46.11 kJ mol 1 )  (5 mol)(0 kJ mol1 )]
 905.48 kJ
(c)
H   4(H  f (H 3 PO 4 , aq ))  [(H  f ( P4 O 6 , s)) 6( H f ( H 2 O,l))]
(4 mol)(964.8 kJ mol 1 )  [(1 mol)(1640.0 kJ mol 1 )  (6 mol)(296.83 kJ mol 1 )]
 438.2 kJ

4D.24 From Appendix 2A,  H f (O 3 )  142.7 kJ  mol 1

NO 2 (g)  O 2 (g) 
 NO(g)  O 3 (g) H   200 kJ

O 3 (g) 
 3
2 O 2 (g) H   142.7 kJ
1
2 O 2 (g) 
 O(g) H   249.2 kJ

NO 2 (g) 
 NO(g)  O(g) H   306 kJ
Focus 4 Thermodynamics 197

4D.26 First, calculate the differences in heat capacities:


C p  [(1 mol)CP, m (C5 H10 N 2O3 , aq)  (1 mol)CP, m (H 2O, l)] 
[(1 mol)CP, m (C5 H8 NO 4 , aq)  (1 mol)CP, m (NH 4 + , g)
 [(1 mol)(187.0 J K 1 mol1 )  (1 mol)(75.29 J K 1 mol1 )] 
[(1 mol)(177.0 J K 1 mol1 )  (1 mol)(79.9 J K 1 mol1 )]
 5.39 J K 1

Evaluate T2  T1 .
T2  T1  (80.0  273 K )  298 K  55 K
Calculate the enthalpy change at the final temperature from
H (T2 )  H (T1 )  (T2  T1 )CP

H (55 K)  +21.8 kJ  (55 K)  (5.39 J K 1 )


 +21.8 kJ  0.30 kJ
 22 kJ

4D.28 (a) The enthalpy of vaporization is the heat required for the conversion
 CH 3 OH(g) at constant pressure. The value at 298.2 K
CH 3 OH(l) 

will be given by:


H vap at 298 K  H f (CH 3 OH, g)  H f (CH 3 OH, 1)
 200.66 kJ  mol1  (238.86 kJ  mol1 )
 38.20 kJ  mol1
(b) To take into account the difference in temperature, we need to use the
heat capacities of the reactants and products to raise the temperature of the
system to 64.5C. We can rewrite the reactions as follows, to emphasize
temperature:
CH3 OH(l)at 298 K 
 CH3 OH(g)at 298 K H   38.20 kJ

CH3 OH(l)at 298 K 


 CH3 OH(l)at 337.6 K H   (1 mol)
 (337.8 K  298.2 K)
 (81.6 J  mol1  K 1 )
 3.23 kJ
198 Focus 4 Thermodynamics

 CH3 OH(g)at 337.6 K H   (1 mol)


CH3 OH(g)at 298 K 
 (337.8 K  298.2 K)
 (43.89 J  mol1  K 1 )
 1.74 kJ
To add these together to get the overall equation at 337.6 K, we must
reverse the second equation:
CH 3 OH(l) at 298 K 
 CH 3 OH(g) at 298 K H   38.20 kJ

CH 3 OH(l) at 337.8 K 
 CH 3 OH(l) at 298 K H   3.23 kJ

CH 3 OH(g) at 298 K 
 CH 3 OH(g) at 337.8 K H   1.74 kJ

CH 3 OH(1) at 337.8 K 
 CH 3 OH(g) at 337.8 K H   36.71 kJ

(c) The value in the table is 35.3 kJ  mol 1 for the enthalpy of
vaporization of methanol. The value is close to that calculated as corrected
by heat capacities. At least part of the error can be attributed to the fact
that heat capacities are not strictly independent of temperature.

4D.30 The change in the reaction enthalpy is given by:


H r (T2 )  H r (T1 )  (T2  T1 ) C p where
C p   n  C p (prod.)   n  C p (prod.)

Given that N2 and H2 are linear molecules, their heat capacities are both
approximately 52 R, while that of NH3 is 3R. Therefore, for this reaction
the change in the reaction enthalpy is:
H r (T2 )  H r (T1 )  (500 K  300 K) 2(3R)  3( 52 R)  52 R 
 6.65 kJ  mol1

4E.2 For the reaction AlBr3(s)  Al+3(g) + 3Br-(g)


∆HL = ∆Hof (Al, g) + 3∆Hof (Br, g) + I1 (Al) + I2 (Al) + I3 (Al) - Eea (Br)
- ∆Hof (AlBr3, s)
∆HL = 326 kJ.mol-1 + 3(111.88 kJ.mol-1) + 577 kJ.mol-1 + 1817 kJ.mol-1
+ 2744 kJ.mol-1 – 325 kJ.mol-1 – (-527.2 kJ.mol-1)
∆HL = 6002 kJ.mol-1
Focus 4 Thermodynamics 199

4E.4 The process can be broken down into the following steps:
H , kJ  mol1

Na  (g)  2 Cl  (g) 
 NaCl 2 (s) 2524 *

Na(s) 
 Na(g) 107.32

 Na  (g)  e 
Na(g)  +494

 Na 2 (g)  e
Na  (g)  +4562

Cl 2 (g) 
 2 Cl(g) +242

2 (Cl(g)  e 
 Cl , g) 2( 349)

Na(s)  2 Cl 2 (g) 
 NaCl 2 (s) H  f = +2183 kJ·mol-1
For comparison, the enthalpy of formation of NaCl(s) is −411.15 kJ·mol-1.
Because the enthalpy of formation of NaCl2 is a large positive number,

NaCl2 would be considerably unstable. The disproportionation reaction

2 NaCl 2 (s) 
 2 NaCl(s)  Cl 2 (g)

would have an energy of


-1
 Na(s)  2 Cl 2 , g) 2(-2183 kJ·mol )
2(NaCl 2 (s) 
1
2(Na(s)  1
2  NaCl(s) 2(411.15 kJ  mol )
Cl 2 (g) 
-1
 2 NaCl(s)  Cl 2 (g) - 3005 kJ·mol
2NaCl 2 (s) 

This would be very likely to take place, so NaCl2 would not be isolated.

4E.6 (a) break: 1 mol H—Cl bonds 431 kJ  mol1

1 mol F—F bonds 158 kJ  mol1

form: 1 mol H—F bonds 565 kJ  mol1

1 mol Cl—F bonds 256 kJ  mol1

Total 232 kJ  mol1

*From the assumption that H L is the same as that of MgCl2.


200 Focus 4 Thermodynamics

(b) break: 1 mol H—Cl bonds 431 kJ  mol1

1 mol C=C bonds 612 kJ  mol1

form: 1 mol C—C bonds 348 kJ  mol1

1 mol C—H bonds 412 kJ  mol1

1 mol C—Cl bonds 338 kJ  mol1

Total 55 kJ  mol 1

(c) break: 1 mol H—H bonds 436 kJ  mol1

1 mol C=C bonds 612 kJ  mol1

form: 1 mol C—C bonds 348 kJ  mol1

2 mol C—H bonds 2(412) kJ  mol1

Total 124 kJ  mol1

4E.8 (a) The Lewis structures for NO and NO2 show that the bond order in NO
is a double bond and that in NO2 it is 1.5 on average.

N O N N
NO NO2 O O O O
The bond enthalpy of NO at 632 kJ  mol1 is close to the N=O value listed

in the table, whereas the N—O bond enthalpy in NO2 of 469 kJ  mol1 is
slightly greater than the average of a N—O single and N=O double bond:
1
2 (630 kJ  210 kJ)  420 kJ
The extra stability is due to resonance stabilization.
(b) The bond energies are the same for the two bonds, because resonance
makes the bonds equivalent.

4E.10 Bonds broken in reactants: 612 kJ·mol-1 (C=C) + 436 kJ·mol-1 (H-H)
Bonds formed in products: -824 kJ·mol-1 (2 × C-H) + -348 kJ·mol-1 (C-C)
ΔHoreaction = − 124 kJ·mol-1
Focus 4 Thermodynamics 201


4F.2 (a) Rate of entropy =



.
= = 1.2 J·K−1·s−1

Sday  (1.2 J K1  s1 )(60 s min1 )(60 min hr1 )(24 hr day1 )
(b)
 1.0  105 J K1  day1
(c) Greater, because in the above equation if T is smaller, the rate of
entropy generation is larger.


4F.4 (a) ∆S = = = 1.19 J.K−1
.

(b) ∆S = = = 0.876 J.K−1
.

(c) The entropy change is smaller at higher temperatures because the


matter is already more chaotic. The same amount of heat has a greater
effect on entropy changes when transferred at lower temperatures.

4F.6 Because the process is isothermal and reversible, the relationship ∆S =

can be used. Because the process is isothermal, ∆U = 0 and hence q = −w,


where w = −PdV. Making this substitution, we obtain:

∆S = = dV = dV

∆S = nR ln
3.90 L
∆S = (0.720 mol)(8.314 J.K−1.mol−1) ln = −11.0 J.K−1
24.32 L

dq
4F.8 (a) The relationship to use is dS  . At constant pressure, we can
T
nC p dT
substitute dq  nC p dT to obtain dS  .
T
202 Focus 4 Thermodynamics

T2
Upon integration, this gives S  nC p ln . The answer is calculated by
T1
entering the known quantities. Remember that for an ideal monatomic gas
5
Cp  R
2
5  260.65
S  ( 4.10mol)  8.314J K 1 mol 1  ln  55.3J K 1
 2  498 . 86
T2
(b) A similar analysis using Cv gives S  nCv ln , where Cv for a
T1
3
monatomic ideal gas is R.
2
3  260.65
S  ( 4.10mol)  8.314JK 1 mol 1 ln  33.2JK 1
2  498.86

4F.10 Methane, CH4, 16.042 g. mol-1

70.9g
 4.42 mol
16.042 g· mol

Since P is the inverse of V for an ideal gas, we can write:


V2 P
S  nR ln  nR ln 1
V1 P2

7.00 kPa
 (4.42 mol)(8.314 J·K -1·mol 1 ) ln
350.0 kPa
 144 J·K -1

4F.12 Helium, He, 4.003g. mol−1


5.75g
 1.436 mol
4.003 g· mol -1
V
S = nRln 2
V1

Since P is the inverse of V for ideal gases we can write:


Focus 4 Thermodynamics 203

P1
S = nRln
P2
320.0 kPa
 (1.436 mol)(8.314 J·K 1·mol1 ) ln
40.0 kPa
 24.8 J·K 1

This is the entropy increase resulting from the volume increase (pressure
decrease).

T2
S = nRln
T1
273 K
 (1.436 mol)(8.314 J.·K 1·mol1 ) ln
423 K
 5.23 J·K 1
This is the entropy decrease resulting from the temperature decrease.
Net change in entropy, Snet = 24.8J·K−1 − (−5.23 J·K−1) = 30.0 J·K−1

4F.14 (a) The entropy for the vaporization per mole of water at the boiling
point under a constant pressure of 1.00 atm is found by:
q H 40700 J  mol1
Per mole, S     109 J  mol1  K 1
T T 373.2 K
∆S = 2.40 mol (109 J.mol-1.K-1) = 262 J.K-1
(b) The molar change in entropy is:
q H 4600 J  mol1
S     29.0 J  mol1  K 1
T T 158.7 K
The change in entropy for freezing 4.50g of C2H5OH (46.069 g.mol−1) is:
(0.09768 mol)( −29.0 J·mol−1·K−1) = 2.83 J·K−1

4F.16 (a) Trouton’s rule indicates that the entropy of vaporization for a number
of organic liquids is approximately 85 J.K−1.mol−1. We can then calculate
H  vap 25.76  103 J  mol 1
that TB    303 K . (b) The experimental
S  vap 85 J  K 1  mol 1
204 Focus 4 Thermodynamics

boiling point of methylamine, 267 K, is significantly different. The value


85 J·K-1·mol-1 is an average value for the entropy of vaporization of
organic liquids, and therefore deviations are expected when using this
average value for individual organic liquids. However, in the case of
methylamine the significant difference indicates that in its pure form it
must have intermolecular interactions that many organic liquids do not
have. On inspection of methylamine's structure one sees that
intermolecular H-bonds are possible between the nitrogen atom lone pair
of one molecule and a hydrogen atom of the H-N bond on another
molecule.

4F.18
Svap (210 K )  S (heating liquid 210 K to 239.7 K ) 
Svap (239.7 K )  S (cooling vapor 239.7 K to 210 K )

T2
S (heating liquid 210 K to 239.7 K )  C ln
T1
239.7 K
 (80.8 J K 1 mol1 )(ln )  10.7 J K 1 mol1
210.0K

H vap 5.65 kJ mol1


S vap (239.7 K )  
Tbp 239.7 K
 0.0236 kJ K 1 mol 1  23.6 J K 1 mol1

T2
S (cooling vapor 239.7 K to 210 K)  C ln
T1
210.0K
 (35.1 J K 1 mol1 )(ln )  4.64 J K 1 mol1
239.7K
S vap (210 K)  10.7 J K 1 mol1  23.6 J K 1 mol1  4.64 J K 1 mol1
 29.7 J K 1 mol1
Focus 4 Thermodynamics 205

4G.2 (a)
S  k B ln W
W  116  1
S 0
(b)
S  k B ln W
W  316  4.3046721  10 7
S  (1.3806  10  23 J K 1 )ln(4.3046 721  10 7 )
S  2.43  10  22 J K 1

4G.4 We would expect NO and N2O to be the most likely to have a residual
entropy at 0 K. This is because the structures are set up so that the O and
N atoms, which are of similar size, could be oriented in one of two ways
without perturbing the lattice of the solid, as shown below. Because CO2
and Cl2 are symmetrical, switching ends of the molecule does not result in
increased disorder.
N= O N= O N= O NN—O NN—O NN—O
N= O N= O N= O NN—O NN—O O—NN
N= O O= N N= O NN—O NN—O NN—O
N= O N= O N= O

4G.6
(a)

(b) 1,2-Difluorobenzene has six possible orientations:


206 Focus 4 Thermodynamics

F H H

H C F H C F H C H
C C C C C C

C C C C C C
H C H H C F H C F

H H F
H H F

H C H F C H F C H
C C C C C C

C C C C C C
F C H F C HH C H

F H H

1,3-Difluorobenzene also has six possible orientations:

F H H

H C H H C F H C H
C C C C C C

C C C C C C
H C F H C H F C F

H F H
H F H

F C H H C H F C F
C C C C C C

C C C C C C
H C H F C HH C H

F H H

1,4-Difluorobenzene has only three possible orientations:


F H H

H C H H C F F C H
C C C C C C

C C C C C C
H C H F C HH C F

F H H

The least residual molar entropy will be exhibited for the compound
that has the fewest possible orientations: 1,4-difluorobenzene
Focus 4 Thermodynamics 207

4G.8 There are three orientations that PH2F could adopt in the solid state.
P P P
F H H
H H F
H F H
The Boltzmann entropy calculation then becomes:
23 23
S  k ln 36.02  10  (1.38  1023 J  K 1 ) ln 36.02  10
S  9.13 J  K 1

4G.10 (a) W = 6 (b) W = 28 (c) Part b

4H.2 (a) O3(g), because it has more atoms and contains more elementary
particles
(b) CH4(g), because there are not as many attractive forces as there are in
CH2Br2(g)
(c) CaI2(aq), because molecules in liquids are more randomly oriented
than molecules in solids
(d) O2(g) at 278 K and 1.00 atm, because it occupies a larger volume.

4H.4 Gases will have a higher entropy than liquids, and liquids will have a
higher entropy than solids, so we expect H2O(s) to have the lowest molar
entropy. The gases will increase in entropy in the order HF(g) < NH3(g) <
NH2OH(g). Larger molecular substances will have more complexity.
NH2OH will have a higher entropy than NH3 because it is more complex,
has a larger mass, and has more fundamental particles. NH3 has a higher
entropy than HF because it is more complex and has more fundamental
particles. The final order is H2O(s) < HF(g) < NH3(g) < NH2OH(g).

4H.6 Cyclohexane has more atoms, is more flexible, and has the greater
standard molar entropy.
208 Focus 4 Thermodynamics

4H.8 (a) Entropy should increase, because CO2 is more dynamic as a gas than as
a solid.
(b) Entropy should decrease, because the number of moles of gas is fewer
on the product side of the reaction.
(c) Entropy should increase, because liquid water is more dynamic than
ice.

4H.10 The total change in entropy is given by:


T2 T
S  S Ne  SF2  nNe  CV,Ne  ln  nF2  CV,F2  ln 2
T1 T1
where the heat capacities of the monatomic gas Ne and diatomic gas
F2 are 3/2 R and 5/2 R, respectively. This equation can be reduced to one
unknown by the fact that the sum nNe  nF2 is equal to the total number of

moles of gas present, ntot calculated using the ideal gas law:

P V 3.32 atm  2.5 L


ntot    0.370 mol .
R  T  0.0820574 L  atm  K 1  mol1   273.2 K

Therefore,
T2 T
S  0.345 J  K 1  nNe  CV,Ne  ln   0.370 mol  nNe   CV,F2  ln 2
T1 T1
3 288.15 K
0.345 J  K 1  nNe  R  ln
2 273.15 K
5 288.15 K
  0.370 mol  nNe  R  ln
2 273.15 K
1 1
0.345  nNe  0.667 J  K  mol
  0.370 mol  nNe 1.111 J  K 1  mol1
nNe  0.150 mol
and rounding off at the end to two decimal places, one obtains
nF2 = 0.370 mol - 0.150 mol = 0.220 mol = 0.22 mol
Focus 4 Thermodynamics 209

4H.12 (a) 4 Al(s) + 3 MnO2(s) → 3 Mn(s) + 2 Al2O3(s)


∆So = [So (Mn, s) + So (Al2O3, s)] - [So (Al, s) + So (MnO2, s)]
∆So = [3(32.01 J·mol-1·K-1) + 2(50.92 J·mol-1·K-1)] – [4(28.33 J·mol-1·K-1)
+ 3(53.05 J·mol-1·K-1)]
∆So = -74.60 J·K-1
A small decrease in entropy for this solid state reaction.
(b) 7 H2O2(l) + N2H4(l) → 2 HNO3(aq) + 8 H2O(l)
∆So = [So (HNO3, aq) + So (H2O, l)] - [So (H2O2, l) + So (N2H4, l)]
∆So = [2(-146.4 J·mol-1·K-1) + 8(69.91 J·mol-1·K-1)] − [7(109.6 J·mol-1·K-1)
+ (121.21 J.mol-1.K-1)]
∆So = −621.93 J·K-1
Compared to part (a) a larger decrease in entropy.
(c) SiO2(s) + 2 C(s) → Si(s) + 2 CO(g)
∆So = [So (Si, s) + So (CO, g)] - [So (SiO2, s) + So (C, s)]
∆So = [(18.83 J·mol-1·K-1) + 2(197.67 J·mol-1·K-1)] – [(41.84 J·mol-1·K-1)
+ 2(5.740 J·.mol-1·K-1)]
∆So = 360.85 J·K-1
With solid state reactants producing a gas product, there is an increase in
entropy.
° ,
(d) 4 NH3(g) + 5 O2(g) 4 NO(g) + 6 H2O(g)
∆So = [So (NO, s) + So (H2O, g)] - [So (NH3, s) + So (O2, s)]
∆So = [4(210.76 J·mol-1·K-1) + 6(188.83 J·mol-1·K-1)]
– [4(192.45 J·mol-1·K-1) + 5(205.14 J·mol-1·K-1)]
∆So = 180.52 J·K-1
An increase in entropy with more moles of gas on the product side.

4H.14 The standard molar entropy of a substance at temperature Ti may be found


by plotting Cp/T vs. T for that substance and finding the area under the
curve from T = 0 to T = Ti. If at low temperatures CpT3, the standard
molar entropy of the substance is:
210 Focus 4 Thermodynamics

T
Cp T
T3 1
S (T )  
0K
T
dT  
0K
T
dT  T 3
3

4I.2
q
S =
T
 25000 J
S 
700 K
S  35.7 J K 1

 25000 J
S 
320 K
S  78.1 J K 1
Stotal  35.7 J K 1  78.1 J K 1  42.4 J K 1

4I.4 Assume the transfer of energy as heat is reversible.


(a) The value can be estimated from:
H vap  T S vap
 (329.4 K)(85 J  mol1  K 1 )
 28.0 kJ  mol1
(b) ∆Socondensation = -85 J·mol-1·K-1
If one mole of acetone condensed, the entropy of the surroundings would
increase by +85 J.mol-1.K-1. For the condensation of 10. g:
 10g 
∆Sosurr =   (85 J·mol-1·K-1) = 15 J·K-1
-1 
 58.08g·mol 

4I.6 For this problem we can treat the two samples of ethanol as different
systems and find the total change in entropy by addition. The change in
T2
entropy for both samples of ethanol is found using S  C ln , so our
T1
first job is to determine the final equilibrium temperature after the two
samples mix:
Focus 4 Thermodynamics 211

m1 T1  m2 T2 320.0 g  291.15 K + 120.0 g  329.15 K


TF  
m1  m2 320.0 g + 120.0 g
 301.514 K
The change in entropy for the ethanol initially at 18.0oC is:
301.51 K
 
S sys  320.0 g  2.42 J  g 1  K 1 ln  291.15 K
 27.077 J  K 1  27.1 J  K 1

and the change in entropy for the ethanol initially at 56.0oC is:
301.51 K
 
S sys  120.0 g  2.42 J  g 1  K 1 ln  329.15 K
 25.471 J  K 1  25.5 J  K 1

The total change in entropy is that of the entire system; there is no change
in entropy of the surroundings, as the vessel isolates the system from the
surroundings:
∆Stot = 27.1 J.K−1 + (−25.5 J.K−1) = +1.6 J.K−1

4I.8 (a) The change in entropy will be given by:


H system 1.00 mol   5.65  103 J  mol1
Ssurr    28.9 J  K 1
T 195.4 K
H system 1.00 mol  5.65  103 J  mol1
Ssystem    28.9 J  K 1
T 195.4 K
(b)
H system (1.00 mol   3.16  103 J  mol1 )
Ssurr    18.0 J  K 1
T 175.2 K
H system 1.00 mol   3.16  103 J  mol1
Ssystem    18.0 J  K 1
T 175.2 K
(c)
H system (1.00 mol  40.7  103 J  mol1 )
Ssurr    109 J  K 1
T 373.2 K
H system 1.00 mol  40.7  103 J  mol1
Ssystem    109 J  K 1
T 373.2 K
212 Focus 4 Thermodynamics

4I.10 (a) The total entropy change is given by S tot  Ssurr  S . S for an

isothermal, reversible process is calculated from


qrev  wrev V
S    nR ln 2 . To do the calculation we need the value of
T T V1
n, which is obtained by use of the ideal gas law:
(0.6789 atm)(12.62 L) = n(0.082 06 L  atm  K 1  mol 1 )(412 K); n =
0.253 mol.
19.44 L
S  (0.253 mol)(8.314 J  K 1  mol1 ) ln  0.909 J  K 1 .
12.62 L
Because the process is reversible:
 S tot  0, so  S surr   S  0.909 J  K 1 .

(b) For the irreversible process, S also  0.909 J  K 1 . No work is


done in free expansion, so w = 0. Because ∆ U = 0, it follows that q = 0.
Therefore, no heat is transferred into the surroundings, and their entropy is
unchanged:
 Ssurr = 0. The total change in entropy is therefore  S tot  0.909 J  K 1 .

4I.12 A solid is converted to a gas, which is an entropy increase for the system.
However, the temperature of the system drops, which is an entropy
decrease for the system. However, we are told that the system undergoes a
spontaneous change and therefore the net entropy for the system must
increase. Since the temperature of the system dropped, we can assume that
the system is isolated from the surroundings. If the system were not
isolated, heat would have gone from the surroundings to the system (+qsys)
and the system would have maintained the same temperature. Since there
is no heat exchange between the system and surroundings, there is no
entropy change in the surroundings.
Focus 4 Thermodynamics 213

4J.2 Under constant temperature and pressure conditions, it is the sign of


Gr that determines whether or not a process is spontaneous. If Gr < 0,

( Gr is negative), the process is spontaneous. Gr is related by the

equation Gr  H r  T Sr to the enthalpy and entropy changes in a

reaction. If a reaction is endothermic ( H r positive), then the

reaction will be spontaneous only if T S r is larger than H r . So,


effectively, a reaction that is endothermic can be spontaneous only if the
entropy change in the system outweighs the enthalpy change’s effect on
the entropy of the surroundings.

4J.4 (a)
G  H vap  TS vap
 140.0  273.15  133.15 K
G  8.2 kJ mol 1  [(133.15 K)(73.0 J K 1 mol 1 )]
 8.2 kJ mol 1  9.72 k J mol 1
 1.5 kJ mol 1
The reaction has a negative Gibbs free energy change and is therefore
spontaneous at –140.0 °C and 1 atm.
(b)
G  H vap  TS vap
 180.0  273.15  93.15 K
G  8.2 kJ mol 1  [(93.15 K)(73.0 J K 1 mol 1 )]
 8.2 kJ mol 1  6.80 k J mol 1
 1.4 kJ mol 1
The reaction has a positive Gibbs free energy change and is therefore
not spontaneous at –180.0 °C and 1 atm.
214 Focus 4 Thermodynamics

4J.6 (a) 1
2 H 2 (g)  1
2 Cl 2 (g) 
 HCl(g)

H r  H f (HCl, g)  92.31 kJ  mol1


S r  S m (HCl, g)  [ 12 S m (H 2 , g)  1
2 S m (Cl2 , g)]
 186.91 J  K  mol  [ 12 (130.68 J  K 1  mol1 )
1 1

 12 (223.07 J  K 1  mol1 )]
 10.04 J  K 1  mol1
Gr  92.31 kJ  mol1  (298 K)(10.04 J  K 1  mol1 ) /(1000 J  kJ 1 )
 95 kJ  mol1

(b) 6 C(s), graphite  3 H 2 (g) 


 C 6 H 6 (l)

H r  H f (C6 H 6 , l)  49.0 kJ  mol1


S r  S m (C6 H 6 , l)  [6 S m (C, s)  3S m (H 2 , g)]
 173.3 J  K 1  mol 1  [6(5.740 J  K 1  mol1 )
 3 (130.68 J  K 1  mol 1 )]
 253.2 J  K 1  mol1
G r  49.0 kJ  mol1  (298 K)(253.2 J  K 1  mol1 ) /(1000 J  kJ 1 )
 124.5 kJ  mol1

S  m (C 6 H 6 , l)  173.3 J  K 1  mol 1

(c) Cu(s) + S(s) + 9/2 O2(g) + 5 H2(g) 


 CuSO4˜5 H2O
H r  H f  Cu 6SO 4  5 H 2 O, s   2279.7 kJ  mol1
S r  S m (CuSO 4  5 H 2 O, s)  [ S m (Cu, s)  S m (S, s)
 92 S m (O 2 , g)  5 S m (H 2 , g)]
 300.4 J  K 1  mol1  [(33.15 J  K 1  mol1 )
 (31.80 J  K 1  mol1 )  92 (205.14 J  K 1  mol1 )
 5 (130.68 J  K 1  mol1 )]
 1341.1 J  K 1  mol1
G r  2279.7 kJ  mol1
 (298 K)(  1341.1 J  K 1  mol1 )/(1000 J  kJ 1 )
 1880.0 kJ  mol 1

S  m (CuSO 4  5H 2 O,s)  160.7 J  K 1  mol 1

(d) Ca(s)  C(s), graphite  3


2 O 2 (g) 
 CaCO 3 (s)
Focus 4 Thermodynamics 215

H r  H f (CaCO3 , s)  1206.9 kJ  mol 1


S r  S m (CaCO3 , s)  [ S m (Ca(s)  S m (C, s)  32 S m (O 2 , g)]
 92.9 J  K 1  mol 1  [41.42 J  K 1  mol1  5.740 J  K 1  mol1
 32 (205.14 J  K 1  mol 1 )]
 262.0 J  K 1  mol1
G r  1206.9 kJ  mol1
 (298 K)(  262.0 J  K 1  mol1 )/(1000 J  kJ 1 )
 1.1288  103 kJ  mol1

S  m (CaCO 3 , s)  92.9 J  K 1  mol 1

The S f value is negative because of the reduction of the number of


moles of gas during the reaction. For all of these, the important point to
gain is that the S m value of a compound is not the same as the S f for the

formation of that compound. S f is often negative because one is


bringing together a number of elements to form that compound.

4J.8 (a) CH4(g) + H2O(g) 


 CO(g) + 3 H2(g)
S r  S m (CO, g)  32S m (H 2 , g)  [S m (CH 4 , g)  S m (H 2 O, g)]
 197.67 J  K 1  mol1  3(130.68 J  K 1  mol1 )
 [186.26 J  K 1  mol1  188.83 J  K 1  mol1 ]
  214.62 J  K 1  mol1
H r  H f (CO, g)  [H f (CH 4 , g)  H f (H 2 O, g)]
 (110.53 kJ  mol1 )  [(74.81 kJ  mol1 )  (241.82 kJ  mol1 )]
 206.10 kJ  mol1
Gr  Gf (CO, g)  [Gf (CH 4 , g)  G f (H 2 O, g)]
 (137.17 kJ  mol1 )  [(50.72 kJ  mol 1 )
 (228.57 kJ  mol1 )]
 142.12 kJ  mol1
G r can also be calculated from S r and H r using this relationship:
216 Focus 4 Thermodynamics

G r  H r  T S r  206.1 kJ  mol1
 (298 K)(  214.62 J  K 1  mol 1 )
 142.1 kJ  mol1
(b) NH4NO3(s) 
 N2O(g) + 2 H2O(g)
S r  S m (N 2 O, g)  2S m (H 2 O, g)  S m (NH 4 NO3 , s)
 219.85 J  K 1  mol1  2(188.83 J  K 1  mol1 )
 151.08 J  K 1  mol1
 446.43 J  K 1  mol1

H r  H f (N 2 O, g)  2H f (H 2 O, g)  [H f (NH 4 NO3 , s)]


 82.05 kJ  mol1  2(241.82 kJ  mol1 )  [365.56 kJ  mol1 ]
 36.03 kJ  mol1
Gr  Gf (N 2 O, g)  2Gf (H 2 O, g)  [Gf (NH 4 NO3 , s)]
 104.20 kJ  mol 1  2(228.57 kJ  mol 1 )  [ 183.87 kJ  mol1 ]
 169.07 kJ  mol1
G r can also be calculated from S r and H r using the relationship:

Gr  H r  T S r
 36.03 kJ  mol1  (298 K)(446.43 J  K 1  mol 1 ) /(1000 J  kJ 1 )
 169.07 kJ  mol

4J.10 (a)
H r  2279.7 kJ  mol1  [771.36 kJ  mol1  5(285.83 kJ  mol1 )]
 79.19 kJ  mol1

S r  300.4 J  K 1  mol1  [109 J  K 1  mol 1  5(69.91 J  K 1  mol1 )]


 158 J  K 1  mol1
Gr  1879.7 kJ  mol1  [661.8 kJ  mol1  5(237.13 kJ  mol1 )]
 32.3 kJ  mol1

Gr may also be calculated from  H r and  S r . (The numbers


calculated differ slightly from the two methods due to rounding
differences.)
Focus 4 Thermodynamics 217

Gr  H r  T S r
 79.19 kJ  mol 1  (298 K)(  158 J  K 1  mol1 )/(1000 J  kJ 1 )
 32.1 kJ  mol1
(b)
∆H° 4 1288.34 kJ ∙ mol — 2984.0 kJ ∙ mol
6 285.83 kJ ∙ mol
= -454.4 kJ ∙ mol
∆S° 4 158.2 J ∙ K ∙ mol — 228.86 J ∙ K ∙ mol
+ 6(69.91 J ∙ K ∙ mol
15.52 J ∙ K ∙ mol
∆G° -4 1142.54 kJ ∙ mol — 2697.0 kJ ∙ mol
6 237.13 kJ ∙ mol
450.4 kJ ∙ mol
or
∆G° ∆H° T∆S°
. ∙ ∙
454.4 kJ ∙ mol

449.8 kJ ∙ mol
(c)
∆H° 207.36 kJ ∙ mol — 174.10kJ ∙ mol
-33.26 kJ ∙ mol
∆S° 146.40 J ∙ K ∙ mol — 155.60J ∙ K ∙ mol
9.20J ∙ K ∙ mol
∆G° -111.25 kJ ∙ mol 80.71kJ ∙ mol
= -30.54 kJ ∙ mol
or
∆G° ∆H° T∆S°
. ∙ ∙
33.26kJ ∙ mol

30.5 kJ ∙ mol
218 Focus 4 Thermodynamics

(d) In order to rank the ability of these compounds to remove water, we


can examine the free energies for the reactions involved. The one with the
greatest driving force (most negative ∆Gºr) is the reaction of
phosphorus(V) oxide with water, followed by hydration of copper sulfate,
with the reaction of HNO3(l) with water to form aqueous nitric acid further
behind.

4J.12 Use the relationship Gr   Gf (products)  Gr (reactants):

(a) Gr  Gf (NH3 , g)  Gf (HCl, g)  [Gf (NH 4Cl, s)]

 (16.45 kJ  mol1 )  (  95.30 kJ  mol1 )  [  202.87 kJ  mol1 ]


 91.12 kJ  mol1
The reaction is not spontaneous.
(b) G r  G f (H 2 O, l)  [ G f (D 2 O, l)]
 ( 237.13 kJ  mol 1 )  [  243.44 kJ  mol1 ]
 6.31 kJ  mol1
The reaction is not spontaneous.
(c) G r  G f (N 2 O, g)  G f (NO, g)  [G f (NO 2 , g)]
 104.20 kJ  mol 1  86.55 kJ  mol 1  51.31 kJ  mol 1
 139.44 kJ  mol 1
The reaction is not spontaneous.
(d) Gr  2Gf (CO 2 , g)  4Gf (H 2 O, l)  [2Gf (CH3 OH, g)]
 2(394.36 kJ  mol1 )  4(237.13 kJ  mol1 )
 [2(161.96 kJ  mol1 )]
 1413.32 kJ  mol1
The reaction is spontaneous.

4J.14 To answer this question, we examine the standard Gibbs free energies of
formation of the compounds. These values from the Appendix:
(a) C3H6(g), cyclopropane, +104.45 kJ·mol-1
(b) CaO(s), -604.03 kJ·mol-1
Focus 4 Thermodynamics 219

(c) N2O(g), +104.20 kJ·mol-1


(d) HN3(g), +328.1 kJ·mol-1
Compounds with a negative free energy of formation are stable, whereas
those with a positive free energy of formation are unstable with respect to
the elements that compose them. Accordingly, compounds (a), (c), and (d)
are thermodynamically unstable, whereas (b) is thermodynamically stable.

4J.16 To understand what happens to G r as the temperature is raised, we use

the relationship Gr  H r  T S r . From this it is clear that the free


energy of the reaction becomes less favorable (more positive) as
temperature increases only if S r is a negative number. Therefore, we
only have to find out whether the standard entropy of formation of the
compound is a negative number. This is calculated for each compound as
follows:
(a) 3C(s), graphite + 3H2(g) → C3H6(g)
∆Sor = Som(C3H6, g) - [3Som(C, s) + 3Som(H2, g)]
= 237.4 J·K-1·mol-1 - [3(5.740 J·K-1·mol-1) + 3(130.68 J·K-1·mol-1)]
= 237.4 J·K-1·mol-1 - 17.22 J·K-1·mol-1 - 392.04 J·K-1·mol-1
= -171.86 J.K-1.mol-1 = -171.9 J.K-1.mol-1
The compound is less stable at higher temperatures.
(b) Ca(s) + 0.5 O2(g) → CaO(s)
∆Sor = Som(CaO, s) - [Som(Ca, s) + 0.5 Som(O2, g)]
= 39.75 J.K-1.mol-1 - 41.42 J.K-1.mol-1 - 0.5(205.14 J.K-1.mol-1)]
= -104.24 J·K-1·mol-1
The compound is less stable at higher temperatures.
(c) N2(g) + 0.5 O2(g) → N2O(g)
∆Sor = Som(N2O, g) - [Som(N2, g) + 0.5 Som(O2, g)]
= 219.85 J·K-1·mol-1 - 191.61 J·K-1·mol-1 - 0.5(205.14 J·K-1·mol-1)]
= -74.33 J·K-1·mol-1
The compound is less stable at higher temperatures.
(d) 1.5 N2(g) + 0.5 H2(g) → HN3(g)
220 Focus 4 Thermodynamics

∆Sor = Som(HN3, g) - [1.5 Som(N2, g) + 0.5 Som(H2, g)]


= 238.97 J·K-1·mol-1 - [1.5(191.61) + 0.5(130.68)] J·K-1·mol-1
= 238.97 J·K-1·mol-1 - 287.415 J·K-1·mol-1 - 65.34 J·K-1·mol-1
= -113.78 J·K-1·mol-1
The compound is less stable at higher temperatures.

4J.18 (a)
∆H° 2 33.18 kJ ∙ mol — 2 90.25kJ ∙ mol
-114.14kJ ∙ mol
∆S° 2 240.06 J ∙ K ∙ mol — 2 210.76J ∙ K ∙ mol
205.14J ∙ K ∙ mol
146.54 J ∙ K ∙ mol
∆G° ∆H° T∆S°
. ∙ ∙
114.14kJ ∙ mol

37.5 kJ ∙ mol
∆G° 0 to be spontaneous
0 ∆H° T∆S°
. ∙ ∙
0 114.14kJ ∙ mol

T = 778.9 K
For the reaction to be spontaneous, ∆G°r must be less than zero. To make
this statement true, the temperature must be less than 778.9 K. So T <
778.9 K.
(b)
∆H° 986.09 kJ ∙ mol — 635.09 kJ ∙ mol
285.83 kJ ∙ mol
-65.17 kJ ∙ mol
∆S° 83.39 J ∙ K ∙ mol — 39.75 J ∙ K ∙ mol
Focus 4 Thermodynamics 221

69.91 J ∙ K ∙ mol
26.27 J ∙ K ∙ mol
∆G° ∆H° T∆S°
. ∙ ∙
-65.17 kJ ∙ mol

51.4 kJ ∙ mol
∆G° 0 to be spontaneous
0 ∆H° T∆S°
. ∙ ∙
0 65.17 kJ ∙ mol

T = 2481 K
For the reaction to be spontaneous, ∆G°r must be less than zero. To make
this statement true, the temperature must be less than 2481 K. So T < 2481
K.
(c)
∆H° 1002.82 kJ ∙ mol 226.73 kJ ∙ mol
— 59.8 kJ ∙ mol 2 285.83 kJ ∙ mol
-144.6 kJ ∙ mol
∆S° 74.50 J ∙ K ∙ mol 200.94 J ∙ K ∙ mol
— 69.96 J ∙ K ∙ mol 2 69.91 J ∙ K ∙ mol
83.34 J ∙ K ∙ mol
∆G° ∆H° T∆S°
. ∙ ∙
144.6 kJ ∙ mol

101.0 kJ ∙ mol
∆G° 0 to be spontaneous
0 ∆H° T∆S°
. ∙ ∙
0 144.6kJ ∙ mol

T = 1735 K
For the reaction to be spontaneous, ∆G°r must be less than zero. To make
this statement true, the temperature must be less than 1735. K. So T <
1735 K.
222 Focus 4 Thermodynamics

4.2 (a) The amount of heat lost by the piece of stainless steel must equal the
amount of heat absorbed by the water. This problem is complicated
because the water may undergo one or more phase changes during this
process. In order to answer this question, we analyze each step of the
reaction to determine whether the steel can transfer enough heat to the
water to cause the given change.
For stainless steel, the enthalpy change will be given by:

121.3 g  0.51 J  (C)1  g 1  T  62 J  (C)1  T


For water, five separate processes that may be involved:
(1) heating solid water to 0 C, (2) converting the solid water to liquid at
0 C, (3) raising the temperature of the water from 0 C to 100 C, (4)
converting the liquid water to vapor at 100 C, (5) heating the vapor
above 100 C.
To heat the solid water from 23C to 0 C:

37.55 g  2.03 J  (C)1  g1  23C  1.75 kJ


For this process the corresponding decrease in temperature of the steel
would be: 1.75 kJ  0.062 kJ  (C)1  28C
Temperature of the steel after heating the solid water to 0C will be:
482C  28C  454C.
Because the temperature of the steel is still above that of the water, there is
adequate heat for the next transformation to occur.
To convert all the solid water to liquid water at 0C, we would require:
 37.55 g H O 
 2
1
 (6.01 kJ  mol 1 )  12.53 kJ
 18.01 g  mol H 2O 
This change would require the temperature of the steel to decrease further
by: 12.53 kJ  0.062 kJ  (C)1  202C
Focus 4 Thermodynamics 223

The temperature of the steel after melting all of the solid water to 0 C will
be 454C  202C  252C. The temperature of the steel is still above that
of the water, so more heat can be transferred to heat the liquid water.
To heat the liquid water from 0 C to 100 C:

37.55 g  4.184 J  (C)1  g 1  100C  15.7 kJ


This change would require the temperature of the steel to decrease further

by: 15.7 kJ  0.062 kJ  (C)1  253C


The temperature of the steel is not sufficient to convert all of the water to
vapor (The temperature of the steel after heating all of the liquid water to
100 C would be 252 C – 253 C = -1 C − much less than 100 C.)
The final temperature will then be between 0C and 100C with the water
in liquid form.
- 62.0 J  ( o C)-1 ( T) = (37.55 g)(4.184 J  ( o C)-1  g -1 )( T)
- 62.0 J  ( o C)-1 (TF - 252 oC) = (37.55 g)(4.184 J  ( o C)-1  g -1 )(TF - 0 oC)
TF = 71 oC
(b) The final temperatures of the two systems is Tf = 71 °C.
Thus the water is in liquid form.

4.4 (a) First, we calculate the amount of heat needed to raise the temperature
of the water. Converting the temperatures from Fahrenheit to Celsius:
C  95 (F  32), so 65F corresponds to 18C, and 108F corresponds to

42C.
H  (100 gal)(3.785 L  gal1 )(1000 cm3  L1 )(1.00 g  cm 3 )
 (4.18 J  (C) 1  g 1 )(42C  18C)
= 38 MJ
The enthalpy of combustion of methane is −890 kJ·mol.
The mass of methane required will be calculated as follows:
 38  103 kJ 
 1 
(16.04 g  mol1 CH 4 )  6.8  102 g CH 4
 890 kJ  mol 
224 Focus 4 Thermodynamics

(b) This quantity can be obtained from the ideal gas law. 6.8  102 g of
CH4 is 42 mol of CH4:
nRT (42 mol)(0.082 06 L  atm  K 1  mol1 )(298 K)
V    1030 L
P 1.00 atm

15.50 L
4.6 (a) w  (1.00 mol)(8.314 J  K 1  mol1 )(298.2 K)ln
7.00 L
 1.97 kJ
(b) w = -Pext ∆V = -(1.00 atm)(15.5 L – 7.00 L)(101.325 J.L-1.atm-1)
= -0.861 kJ
(c) The fact that the expansion occurs adiabatically − which means that
the system is isolated from its surroundings so that no heat is transferred −
is important. It tells us that q = 0, and therefore ∆U = w = -0.861 kJ.
But U will also be equal to  nCvT because U is a state function. The
heat capacity of an ideal gas is 12.5 J·K-1·mol-1; we will assume it is
constant over this temperature range. Therefore,
−861 J = (1.00 mol)(12.5 J·K-1·mol-1)(∆T)
∆T = −68.9 K
The final temperature will be 298.2 K – 68.9 K = 229.3 K.

4.8 In step 1, q = 60 J and w = 0, since it is at constant volume. So ∆U = 60 J.


In step 2, q = -12 J and ΔU 60 J, since the internal energy returns to
initial; therefore, 60 J 12 J w.
This gives w = −48 J, since w = −P∆V and P = 1.00 atm
+ 48 J (0.08206 L  atm  K -1  mol-1 )
V = ( ) = 0.48 L
1.00 atm 8.314 J  K -1  mol-1
The gas constants are used to convert J to L.atm. Because the value V is
positive, this change will indicate an expansion.
Focus 4 Thermodynamics 225

4.10 Not counting internal vibrations, a nonlinear molecule will have a heat
capacity of CV,m = 3R. The temperatures at which the three vibrations of
SO2 become available are 1968 K, 1680 K, and 768 K. If a vibration is
accessible, it will contribute a factor of R to the heat capacity. Therefore,
(a) At 1968 K the high temperature limit is reached and CV,m = 6R.
(b) CV,m = 5R
(c) CV,m = 3R

4.12
N
N N

N N
N
The reaction in question is N 6 
 3 N 2 (g)

For this to occur, we will need to break three N—N and three N=N bonds
and form three N—N triple bonds.
Energy required to break bonds plus energy released on bond formation:
3(163 kJ·mol-1)
3(409 kJ·mol-1) energy required to break bonds
−3(944 kJ·mol-1) energy released on bond formation
−1116 kJ·mol-1 net enthalpy change

The reaction to form N2(g) is very exothermic and so the reaction is likely
to occur. Although N6 might be resonance-stabilized like benzene, it is
unlikely that resonance stabilization would overcome the tendency to form
the very strong NN triple bond, so we do not expect N6 to be a stable
molecule.
226 Focus 4 Thermodynamics

4.14 The total amount of heat released by the reaction will be given by
5.270 kJ  (C) 1  1.140C  6.008 kJ
In this case, the heats of reaction for the two different outcomes will be
given simply by the H f values for SO 2 ( 296.83 kJ  mol 1 ) and

SO 3 ( 395.72 kJ  mol 1 ) . The number of moles of sulfur is

0.6192 g S(s)  32.06 kJ  mol1  0.01931 mol S(s) . We do not know the
relative amounts of SO2 and SO3 formed, but these can be determined
using the following two relationships, which are required by the
stoichiometries of the reactions.
Let x = number of moles of SO2 formed
Let y = number of moles of SO3 formed
Then x + y = 0.01931 mol
(296.83 kJ  mol 1 ) x  (395.72 kJ  mol 1 ) y  6.008 kJ
x = 0.01652 mol
y = 0.00279 mol
The ratio of SO2 to SO3 will be 0.01652 mol  0.00279 mol = 5.92:1.

4.16 In order of increasing enthalpy of vaporization:


H2 < N2 < CH4 < C6H6 < H2O < NaCl
A large enthalpy of vaporization indicates strong intermolecular
interactions in the liquid phase. The stronger these interactions, the more
energy needed to free a molecule into the gas phase. Hydrogen and
nitrogen, being homonuclear diatomic molecules, interact through other
H2 or N2 molecules in the condensed phase through very weak London
dispersion forces. CH4, being a nonpolar molecule, also interacts with
other CH4 molecules through weak London dispersion forces. However,
the electron distribution along the C−H bonds is not uniform, which leads
to stronger interactions than in N2. C6H6 has no net dipole moment, but
because of pi bonding, above and below the ring of carbons it tends to be
negative compared to the hydrogen atoms along the edge of the ring. This
Focus 4 Thermodynamics 227

gives giving rise to stronger bonding interactions than in CH4 and N2.
Water is a highly polar molecule that can bind to other water molecules
through relatively strong hydrogen bonding interactions, making its
enthalpy of vaporization higher than that of the previous molecules.
Finally, NaCl forms an ionic solid with strong ionic bonds extending
throughout a three-dimensional crystalline lattice, requiring a large
amount of energy to free a NaCl molecule from the condensed phase.

4.18 (a) T = Tinitial − c(V − Vinitial)

Ideal gas, use: PV  nRT and w    Pext dV

Reversible, therefore Pext = Pint = P


nRT T T - c (V - Vinitial )
w( non  isothermal )    dV  nR  dV   nR  initial dV
V V V
T - cV + cVinitial T cV
  nR  initial dV   nR  initial  c  initial dV
V V V
V V
  nR[c V  Tinitial ln 2  cVinitial ln 2 ]
V1 V1
V2
)(Tinitial  cVinitial )
 nRc V  ( nR ln
V1
V
(b) For an isothermal expansion, w   nRT ln 2 , which is the same as
V1

V2
the second term in the above equation,  nRTinitial ln . The remaining
V1

V2
terms in the equation: nRc V  nRcVinitial ln , which can be rewritten as
V1

V2
nRc ((V2  V1 )  V1 ln ).
V1

V2 V
In an expansion (V2  V1 )  V1 ln ; therefore, nRc ((V2  V1 )  V1 ln 2 )
V1 V1
will be a positive value.
V2
Since w(non-isothermal) =  nRTinitial ln plus a positive value,
V1
228 Focus 4 Thermodynamics

there is less energy lost in the form of work by the non-isothermal


expansion. In other words, the work done is smaller than that of the
isothermal expansion.
V2
Note: One should show that the statement (V2  V1 )  V1 ln is true.
V1

V2 V
The above statement can be rewritten as (  1)  ln 2 and also as
V1 V1

V2
(  1)
V1 V2
e  which is the same as e( x  1)  x . This last statement is
V1
always true when x > 1 which is the case, since V2 > V1 in any expansion.

4.20 (a) The average kinetic energy is obtained from the expression:
3
average kinetic energy  RT
2
(a) 3718 J·mol−1
(b) 3731 J·mol−1
(c) 3731 J·mol−1 − 3718 J·mol−1 = 13 J·mol−1

4.22 The formula relating enthalpy and internal energy is ΔH = ΔU + Δ(pV)


Where Δ(pV) takes into account the change in pressure and volume during
a reaction to give:

ΔH = ΔU + ΔpV + pΔV

Under standard conditions pressure does not change. Therefore ΔpV = 0.

The question also states constant volume. Therefore pΔV = 0.

Which gives ΔH = ΔU

a) ΔUf° = ΔHf° = -241.82 kJ·mol1 at constant volume for H2O(g)


b) ΔUf° = ΔHf° = -285.83 kJ·mol1 at constant volume for H2O(l)
Focus 4 Thermodynamics 229

The formation of liquid water is more negative than the formation of


water vapor because a phase change from water vapor to liquid water
(condensation) is exothermic, so the value for liquid water must be lower.

4.24 No, the gases do not have the same final temperature. The CH4 molecules
have more internal vibrations in which to store energy than do N2
molecules. As a result, it will require more energy to increase the
temperature of CH4(g), and if the same amount of energy is supplied to
both gases, the temperature of N2 will be higher than that of CH4.

4.26 Given the composition of “synthesis gas,” it is easiest to first calculate the
enthalpy of combustion of 1.000 L of the gas. In 1.000 L of synthesis gas
there is 0.40 L of CO. The moles of CO present is given by:

P V 1 atm    0.40 L 
nCO    0.0164 mol of CO
R  T  0.0820574 KLatm
mol    298 K 
The enthalpy of combustion of CO is calculated, using enthalpies of
formation and the balanced reaction 2CO(g) + O2(g)  2CO2(g), to be
−283.0 kJ/mol. Therefore, if 1.000 L of synthesis gas were to burn,
 0.0164 mol    283.0 kJ/mol   4.64 kJ of energy would be released
by the combustion of CO. The energies released due to the combustion of
H2(g) and CH4(g) when 1.000 L of synthesis gas burns is calculated in a
similar way:

P V 1 atm    0.25 L 
nH2    0.0102 mol of H 2
R  T  0.0820574 KLatm
mol    298 K 
 0.0102 mol of H2   (285.83 kJ  mol1 )  2.92 kJ
P V 1 atm    0.25 L 
nCH4    0.0102 mol of CH 4
R  T  0.0820574 KLatm
mol    298 K 
 0.0102 mol of H 2   (890 kJ  mol1 )  9.10 kJ
Altogether, when 1 L of synthesis gas burns, −16.66 kJ of energy is
released.
230 Focus 4 Thermodynamics

The amount of energy needed to heat 5.5 L of H2O(l) by 5oC is given by:
 1000 g 
   4.184 J g C    5 C   115 kJ
1 o 1 o
5.5 L  
 L 
Therefore, the volume of synthesis gas that must be burned to provide
this much energy is:
115 kJ
 6.90 L
16.66 kJ  L1

kJ hr days kJ
4.28 (a) (420 )(1 )(150 ) = 6.3 × 10 4
hr day year year
(b)
 trips  gal.  L  mL  g 
150  0.40  3.785 1000  0.702 
 year  trip  gal.   L  mL 
 1 mol   kJ  6 kJ
  5471   7.6  10
 114.23 g   mol  year

4.30 (a)
CO2 (g) + 2H2 (g) → C (s) + 2H2O (l)
2H2O (l) → 2H2 (g) + O2 (g)

CO2 (g) → C(s) + O2(g)


(b) Need to calculate moles of O2 (g) produced.
PV = nRT
n =

(0.82 atm)(32.0 L)
=
(0.0820574 L  atm  K -1  mol-1 )(300 K)
= 1.07 mol

H (reaction)   H f  (products) −  H f  (reactants)

= −(−393.51 kJ.mol-1)
= + 393.51 kJ.mol-1
Focus 4 Thermodynamics 231

Since +393.51 kJ.mol-1 reaction produces 1 mole of O2 (g) , the

enthalpy change associated with the production of 1.07 mole of O2 (g)

would be  4.21 102 kJ (on rounding off to two significant figures).


(c) Highly endothermic reaction and the reactor would have to be heated.

4.32 (a) The heat output of each sample is found using the heat capacity of the
calorimeter and the observed change in temperature of the calorimeter. For
Brand X, T was 8.8oC and the energy content of 1.00 g is:

8.8 C   600 J  K   5300 J  g


o 1 1
.

For the ABC product,T was 8.5oC, and the energy content of 1.00 g is:

8.5 C   600 J  K   5100 J  g


o 1 1

(b) 30 g of Brand X cereal contains



(30 g)(5.3 kJ·g-1) = 38 calories, while 30 g of ABC product
.

contains (30 g)(5.1 kJ·g-1) = 37 calories
.

4.34 Useful information. The enthalpy of vaporization of water at 25 oC is


44 kJ.mol-1. The specific heat capacity for air is 1.01 J.oC-1.g-1 and the
density for air is 1.25 kg.m-3.
Volume of room = (4.0 m)(5.0 m)(3.0 m) = 60 m3
Mass of air in room = (volume)(density)
= (60 m3)(1.25 kg·m-3)
= 75 kg
Now calculate how much energy is needed to cool the air in the room.
Energy needed = (mass in grams)(specific heat capacity)(change in temp)
= (75,000 g)(1.01 J· oC-1·g-1)(20 oC)
= 1515 kJ
water, Hvap = 44 kJ·mol-1
1515 kJ
= 34 mol H2O
44 kJ.mol-1
232 Focus 4 Thermodynamics

mass of water evaporated = (34 mol)(18 g.mol-1) = 612 g


Which rounds to 6.1 × 102 g.

4.36 Normally one would expect dissolution of one substance in another to


increase the entropy of the solution. However, the ability of acetic acid to
hydrogen-bond to water molecules makes the solution more ordered than
in the pure liquid in spite of the mixing. This is confirmed by the standard
molar entropies of 159.8 and 86.6 J.K-1.mol-1 for CH3COOH(l) and
CH3COOH(aq) respectively.

4.38 For C2H5OH, the molar enthalpy of vaporization is found by dividing the
amount of energy supplied to cause the phase transition by the number of
moles of C2H5O that underwent the transition:

H vap 
 500 J  s  (4.0 min)  60 s  min   43 kJ  mol
1 1
1

 400.15 g  271.15 g 
 46.07 g  mol1 
 
The associated change in entropy is then found using:
H vap 43 kJ  mol1
Svap    120 J  mol1  K 1
T 351.5 K
For C4H10 the enthalpy and entropy of vaporization are:

H vap 
 500 J  s  (4.0 min)  60 s  min   22 kJ  mol
1 1
1

 398.05 g  74.95 g 
 1 
 58.123 g  mol 
H vap 22 kJ  mol1
Svap    81 J  mol-1  K 1
T 273.2 K
For CH4O the enthalpy and entropy of vaporization are:

H vap 
 500 J  s  (4.0 min)  60 s  min   38 kJ  mol
1 1
1

 395.15 g  294.25 g 
 1 
 32.042 g  mol 
H vap 38 kJ  mol1
Svap    110 J  mol1  K 1
T 337.7 K
Focus 4 Thermodynamics 233

4.40 First, calculate the G r for both reactions:

(a) Gr  2Gf (CO, g)  [Gr (TiO 2 , s)]


 2(  200 kJ  mol1 )  [(762 kJ  mol1 )]
  362 kJ  mol1

(b) Gr  Gf (CO 2 , g)  [Gr (TiO 2 , s)]


 (  396 kJ  mol1 )  [762 kJ  mol1 ]
  366 kJ  mol1
Neither reaction is spontaneous, so TiO2 cannot be reduced by carbon at
1000 K.

4.42 (a) First, balance the equation for the combustion of 0.825 mol of C6H6(l)
to give carbon dioxide and water vapor:
0.825 C 6 H 6 (l) + 6.1875 O 2 (g)  4.95 CO 2 (g) + 2.475 H 2 O(g)

The work term will be dominated by the change in the number of moles of
gas, which in this case is 7.425 moles – 6.1875 moles = 1.2375 moles
Work will be given by:
w = −P∆V
Approximating using ideal gas equation P∆V = ∆nRT, we have:

∆V =

=w −P

=w −∆nRT
=w −(1.2375 moles)(8.314 J·mol−1·K−1)(298.15 K)
w = 3068 J = 3.07 kJ
(b) 1 bar ≈ 1 atm
. . . .
V2 = =

V2 = 181.75 L
. . . .
V1 = =

V1 = 151.46 L
∆S = ∆nR ln
234 Focus 4 Thermodynamics

∆S = (1.2375 mol)(8.314 J·mol−1·K−1) ln (181.75/151.46)


∆S = 1.876 J·K−1

4.44 (a) Gr = Gf(products) - Gf(reactants)


= 2(−16.45 kJ·mol−1) – 0
= −32.90 kJ·mol−1
(b) Approximate by using ideal gas equations.
Need to find the moles of H2(g) and N2(g).
pV  nRT
pV
n( H 2 ) 
RT
(1bar )(50.1L )

(8.314 10 L·bar·K-1·mol-1 )(298K )
2

 2.022 mol

pV
n( N 2 ) 
RT
(1bar )(15.6L )

(8.314 10 L·bar·K-1·mol-1 )(298K )
2

 0.630 mol
Balanced equation shows H2(g):N2(g) is 3:1.

.
Actual ratio is: 3.2
.

 H2(g) is in excess and the maximum yield of NH3(g) that can be


produced is limited by N2(g). Assuming all the N2(g) reacts, 1.26 mol
NH3(g) can be formed.
1.26 mol
(c) Aqueous concentration of ammonia =
2.00 L
 0.630 mol·.L-1

4.46 (a) Reactions with negative reaction free energies are thermodynamically
favored, but the thermodynamics will not tell us how fast a process takes
place. For example, some reactions with large negative free energies do
Focus 4 Thermodynamics 235

not happen unless initiated, as in the case of the reaction of hydrogen gas
with oxygen gas to produce water.
(b) This statement is false because the sample of the element must be in
its standard state. Not all forms of the element at a given temperature have
the same energy; the one chosen as the standard state will have the lowest
energy.
(c) False. For this process, H r is a negative number, and S r will be
positive because the number of moles of gas increases. According to the
relationship Gr  H r  T S r , if H r is negative and S r is

positive, G r must be negative.

4.48 The reactions are:


(a) CH 3 OH(l)  CO(g) 
 CH 3 COOH(l)

(b) C 2 H 5 OH(l)  O 2 (g) 


 CH 3 COOH(l)  H 2 O(l)

(c) CH 4 (g)  CO 2 (g) 


 CH 3 COOH(l)

To understand these three reactions completely, we must calculate


H r , S r , and Gr for each reaction.

(a) H r  H f (CH 3 COOH, l)

 [H f (CH 3 OH, l)  H f (CO, g)]


 (  484.5 kJ  mol1 )  [(238.86 kJ  mol1 )
 (110.53 kJ  mol1 )]
 135.1 kJ  mol1
S r  S m (CH 3 COOH, l)  [ S m (CH3 OH, l)  S m (CO, g)]
 159.8 J  K 1  mol1  [126.8 J  K 1  mol1
 197.67 J  K 1  mol1 ]
 164.7 J  K 1  mol1
Gr  Gf (CH3 COOH, l)  [Gf (CH3 OH, l)  Gf (CO, g)]
 (389.9 kJ  mol1 )  [(166.27 kJ  mol1 )
 (137.17 kJ  mol1 )]
 86.5 kJ  mol1
236 Focus 4 Thermodynamics

(b)
H r  H f (CH 3 COOH, l)

 H f (H 2O, l)  [ H f (C2 H 5OH, l )]


 (  484.5 kJ  mol 1 )  ( 285.83 kJ  mol 1 )
 [277.69 kJ  mol 1 ]
 492.6 kJ  mol 1
S r  S m (CH 3COOH, l)  S m (H 2 O(l))
 [ S m (C 2 H 5OH, l)  S m (O 2 , g)]
 159.8 J  K 1  mol 1  69.91 J  K 1  mol 1
 [160.7 J  K 1  mol 1  205.14 J  K 1  mol 1 ]
 136.1 J  K 1  mol 1
G r  G f (CH 3COOH, l)  G f (H 2 O, l)  [ G f (C 2 H 5OH, l)]
 ( 389.9 kJ  mol 1 )  ( 237.13 kJ  mol 1 )  [ 174.78 kJ  mol 1 ]
 452.3 kJ  mol 1
(c)
H r  H f (CH 3 COOH, l)

 [H f (CH 4 , g)  H f (CO 2 , g)]


 (  484.5 kJ  mol1 )  [(74.81 kJ  mol 1 )  (393.51 kJ  mol1 )]
 16.2 kJ  mol1
S r  S m (CH 3 COOH, l)  [ S m (CH 4 , g)  S m (CO 2 , g)]
 159.8 J  K 1  mol1  [186.26 J  K 1  mol1  213.74 J  K 1  mol1 ]
 240.2 J  K 1  mol1
G r  G f (CH 3 COOH, l)  [Gf (CH 4 , g)  G f (CO 2 , g)]
 (389.9 kJ  mol1 )  [(50.72 kJ  mol1 )  (394.36 kJ  mol1 )]
 55.2 kJ  mol1
It is clear from these numbers that the second reaction, the oxidation of
ethanol, is by far the most favorable thermodynamically. The addition of
CO to methanol is favorable but less so than the oxidation of ethanol. The
addition of carbon dioxide to methane is not thermodynamically favored.
Focus 4 Thermodynamics 237

4.50 (a)

cis-2-Butene trans-2-Butene 2-Methylpropene

(1) (2) (3)


(b) For the three reactions, the calculation of G, H , and S  are as
follows:
Gr  Gf (2)  Gf (1)
 62.97 kJ  mol1  65.86 kJ  mol1
 2.89 kJ  mol1
H r  H f (2)  H f (1)
 (11.17 kJ  mol1 )  ( 6.99 kJ  mol1 )
 4.18 kJ  mol1
Gr  H r  T S r
 2.89 kJ  mol1  4.18 kJ  mol1  (298 K) (S r )/(1000 J  kJ 1 )
S r  4.33 J  K 1  mol1

Gr  Gf (3)  Gf (1)


 58.07 kJ  mol1  65.86 kJ  mol1
 7.79 kJ  mol1
H f  H f (3)  H f (1)
 (16.90 kJ  mol1 )  (6.99 kJ  mol1 )
 9.91 kJ  mol1
Gr  H r  T S r
7.79 kJ  mol1  9.91 kJ  mol1  (298 K) (S r )/(1000 J  kJ 1 )
S r  7.11 J  K 1  mol1
238 Focus 4 Thermodynamics

Gr  Gf (3)  Gf (2)


 58.07 kJ  mol1  62.97 kJ  mol1
  4.90 kJ  mol1
H r  H f (3)  H f (2)
 (16.90 kJ  mol1 )  (11.17 kJ  mol1 )
 5.73 kJ  mol1
Gr  H r  T S r
 4.90 kJ  mol1  5.73 kJ  mol 1  (298 K) (S r )/(1000 J  kJ 1 )
S r   2.78 J  K 1  mol1
(c) The most stable of the three compounds is 2-methylpropene.
(d) Because S  is also equal to the difference in the Sm values for the
compounds, we can examine those values to place the three compounds in
order of their relative absolute entropies.
The ordering is S m (1)  S m (2)  S m (3).

4.52 For any reaction, G r  H r  T S r . If a liquid boils above room

temperature, G r for this phase change must be positive at room

temperature. S r for a vaporization reaction will generally be positive


because the increase in disorder leads to a negative −T∆S°r term.
Therefore, for G r to be positive, H r must be greater than T S r or:

 
H r  kJ  mol1    298 K  S r  J  mol1  K 1  1000
1 kJ
J 
 0.3  S r  J  mol1  K 1 

Therefore, for the liquid to have a boiling point above room temperature,
H r in kJ mol 1 must be greater than S r in J mol 1 K 1 .
Focus 4 Thermodynamics 239

4.54

4.56 By inspection of the equations in exercise 8.96, e−E/kT = 2 for there to be


twice as many electrons with up spins as with down spins.
1 1 2 2
pdown   p up  
1 2 3 1 2 3
Since E is the energy difference between the low-energy and high-energy
spin states, it is a positive value. Therefore, the only way e−E/kT can equal
2 is to have a negative value for T.

4.58 Use the molar heat capacity of any simple gas, such as N2 gas.
N2 CV,m = 20.81 J·K−1·mol−1
N2 CP,m = 29.12 J·K−1·mol−1
Also, assume CV,m and CP,m are constant over temperature range.
(a) ΔSsys step (1)
Process at constant volume; use CV,m.
T2 75.6 K
S  nCV,m ln = (1.00 mol) (20.81 J·K−1·mol−1) ln
T1 302 K

= −28.82 J.K−1
240 Focus 4 Thermodynamics

ΔSsys step (2)


Process at constant pressure; use CP,m.
T2 302 K
S  nCP,m ln = (1.00 mol) (29.12 J·K−1·mol−1) ln
T1 75.6 K

= +40.33 J.K−1
ΔSsys step (3)
Isothermal compression of an ideal gas. Need to calculate V1.
At the end of step (1):
nRT
PV  nRT Therefore, P 
V
(1.00 mol)(0.08206 L.atm.mol -1 .K -1 )(75.6 K)
P = 2.068 atm
3.00 L
Which rounds to 2.07 atm.
At the end of step (2), the pressure is the same (2.068 atm) but T = 302 K.
nRT
PV  nRT Therefore, V 
P
(1.00 mol)(0.08206 L.atm.mol -1 .K -1 )(302 K)
V = 11.98 L
2.068 atm
At the end of step (2), beginning of step (3), V1 = 11.98 L.
At the end of step (3) given V2 = 3.00 L.
We can now calculate ΔS step (3).
V2 3.00 L
S  nR ln = (1.00 mol) (8.314 J·K−1·mol−1) ln = −11.51 J·K−1
V1 11.98 L
ΔSsys for the entire cycle
∆Ssys = ∆Sstep 1 + ∆Sstep 2 + ∆Sstep 3 = (−28.82 J·K−1) + (+40.33 J·K−1) +
(−11.51 J·K−1) ΔSsys = 0

ΔSsys for the entire cycle is zero, and this confirms that entropy is a state
function.
ΔU of the system, step (1)
U  q  w
Focus 4 Thermodynamics 241

At constant volume no expansion work can be done.


w  0  U  q
q  nC V,m  T and  U  nC V,m T

∆U = (1.00 mol)(20.8 J·mol−1)(75.6 K − 302 K) = − 4711.4 J


Which rounds to − 4.71 kJ
ΔU of the system, step (2)
U  q  w Calculate w at constant pressure, using w   Pex V .

w = -(2.068atm)(11.98 L - 3.00 L) = -18.57 L·atm, which rounds to


−18.6 L·atm
w = −18.57 L·atm × 101.325 J·L−1·atm−1 = −1881.6 J, which rounds to
−1.88 kJ.
Calculate q at constant pressure, using q  nC P,m  T .

q = (1.00 mol)(29.12 J·K−1·mol−1)(302 K − 75.6 K) = + 6592.8 J


Which rounds to +6.59 kJ.
Therefore, ∆ = (+6592.8 J) + (−1881.6 J) = +4711.2 J, which
rounds to +4.71 kJ.
ΔU of the system, step (3)
This is an isothermal compression of an ideal gas.  U  0
ΔU for the entire cycle
 U   U step i   U step ii   U step iii = (−4.71 kJ) + (+4.71 kJ) + (0) = 0

ΔU for the entire cycle is zero, and this confirms that internal energy is a
state function.
(b) q and w are not state functions. Therefore, we have to calculate q and
w for each step and then sum them to get q and w for the entire cycle.
Step (1)
At constant volume w = 0.
U  q  w
U  q From part (a): U  4.71kJ  q  4.71kJ
Step (2)
From part (a): w  1.87 kJ and q = +6.59 kJ
242 Focus 4 Thermodynamics

Step (3)
Isothermal compression of an ideal gas:
Vfinal
w   nRT ln
Vinitial

3.00 L
w   (1.00 mol)(0.08206 L.atm.mol -1 .K -1 )(302 K) ln
11.98 L

w  34.31L.atm ×101.325 J.L-1 .atm -1


w = +3476.9 J, which rounds to +3.48 kJ.
Since it is an isothermal compression of an ideal gas, U  q  w = 0.
q   w = −3476.9 J, which rounds to −3.48 kJ
q and w for the entire cycle
q (entire cycle) = (−4711.4 J) + (+6592.8 J) + (−3476.9 J) = −1595.5 J
= −1.60 kJ
w (entire cycle) = (0) + (−1874.4 J) + (+3476.9 J) = 1602.5 J = +1.60 kJ
Notice that since ΔU for the entire cycle is zero, q and w for the entire
cycle must be equal and opposite (which is what our calculations have
shown).
(c) ΔStot for the entire cycle
 S tot   S sys   S surr

From part (a) we know ΔSsys = 0.


Since it is a reversible process, we know ΔStot = 0.
Therefore, ΔSsurr = 0.
(d) For a process to be reversible, it must be at equilibrium.

4.60 In order for this process to occur spontaneously, entropy cannot decrease.
However, to get as much work out of the transfer of energy as possible, we
do not want to waste energy increasing entropy unnecessarily. Therefore,
the most work will be done when ∆S for the process is infinitesimally
greater than zero, just enough to make the process spontaneous. This will
ensure that as little energy is lost to entropy as possible but enough energy
Focus 4 Thermodynamics 243

is lost to make the process spontaneous. The change in entropy due to the
extraction of 200 J of energy from the hot energy source is:
∆Shot = q/T = (−200 J)/673 K = −0.2972 J·K−1
In order for the net change in entropy to be greater than zero, a slightly
greater increase in entropy must be realized at the cold energy sink. The
minimum energy needed to affect an offsetting increase in entropy is
given by:
Shot = q/T = (q)/293 K = 0.2972 J  K -1
q = 87 J
Therefore, of the 200 J extracted from the heat source, slightly more than
87 J must be deposited in the cold energy sink to make the ∆S positive
and the transfer spontaneous. Slightly less than 113 J is therefore available
to do work. Since we set ∆Shot = ∆Scold, this is the maximum efficiency.

4.62 The balanced chemical reaction is:


2AlCl3 (g)  9
2 O 2 (g) 
 Al 2 O 3 (s)  6ClO(g)

and ∆Gor for the reaction at 2000 K is:

Gr  1034 kJ  mol1  6(75 kJ  mol1 )  2(467 kJ  mol1 )


Gr  350 kJ  mol1
Given the sign of ∆Gor, it does not appear that the reaction will be
spontaneous; therefore, this reaction does not appear to be a good
candidate for rocket propulsion.

4.64 (a) 2NaN 3 (s)  3N 2 (g) + 2Na(g)

(b)  S ro is positive because a gas is produced (from a solid).


+ 0 0
(c) Na + N 3 = N = N NN Overall the nitrogen is oxidized.

(d) Sro   Smo  products   Smo  reactants

= [3(191.61) + 2(153.71)] − [2(96.9)]


244 Focus 4 Thermodynamics

= [574.83 + 307.42] − [193.8]

= + 688.45 J·K−1·mol−1 = + 688.4 J·K−1·mol−1

(e) H r    Hf  products    Hf  reactants


o o o

= [3(0) + 2(107.32)] − [2(21.7)]

= 214.64 − 43.4 = 171.24 = 171.2 kJ·mol−1

∆ ° ∆ ° ∆ ° = (171.2×103 J·mol−1) − 298 K(688.4 J·K−1·mol−1)

= −3.39 × 104 J·mol−1


(f) yes
(g)  H ro and  S ro are both positive. Therefore, the reaction could be
nonspontaneous if the temperature is lowered.

You might also like