You are on page 1of 12

Physical Chemistry 2nd Edition Ball Solutions Manual

Physical Chemistry 2nd Edition Ball Solutions


Manual

To download the complete and accurate content document, go to:


https://testbankbell.com/download/physical-chemistry-2nd-edition-ball-solutions-manu
al/

Visit TestBankBell.com to get complete for all chapters


Chapter 8
Electrochemistry and Ionic Solutions
11 (5.97  10 24 kg)(1.984  10 30 kg)
8.2. (a) F  (6.672  10 N  m /kg )
2 2

(1.494  10 8 km) 2 (1000 m/km) 2

q2
F = 3.541022 N (b) 3.54 10 22 N 
4 (8.854 10 12 C 2 /J  m)(1.494 1011 m)2
Solve for q: q = 2.971017 C. At 1.60210−19 C per electron, this works out to
1.851036 electrons, or 3.081012 moles. The mass of that many electrons is
9.109  10 31 kg
1.85  10 36 e -  -
 1.69  10 6 kg , or approximately 1700 metric tons −
1e
18 orders of magnitude less than the mass of the earth.

8.4. Since 1 dyne equals 110−5 newton, the easiest way to determine how many statcoulombs
in a coulomb is to determine what charge is needed to have a force of 110−5 N at a
q2
distance of 1 cm (0.01 m): 1  10 5 N  and solve for
4 (8.854  10 12 C 2 /J  m)(0.01 m) 2
q: q = 3.33610−10 C. Therefore, if 1 statcoulomb = 3.33610−10 C, then there are
2.998109 statcoulombs in 1 coulomb.

 1 
 
C
8.6. F  e  N A  1.602173310 19 C  6.0221367 10 23   96485 which is Faraday’s
 mol  mol
constant.

8.8. “Electromotive force” is not a force in that it is not a mass times an acceleration, or even
the equivalent. Rather, “electromotive force” is a difference between two electric
potentials, which have units of J/C and not newtons.

8.10. (a) Co  Co 2   2e  E o  0.28V

F2  2e   2F  E 0  2.866 V
Co  F2  Co 2  2F  E o  3.15V
 C 
G o   nFE  ( 2mole e  ) 96485 3.15V   607084 J
 mol 

56
Chapter 8

(b) Zn  Zn 2  2e  E o  0.7618V

Fe 2   2e   Fe E 0  .447 V
Zn  Fe 2  Zn 2  Fe E o  0.3148V
 C 
G o   nFE  ( 2mole e  ) 96485 0.3148V   60747 J
 mol 

 
(c) Zn  Zn 2   2e   3 E o  0.7618V

Fe 3
 3e   Fe  2 E o  0.037 V
3Zn  2Fe 3  3Zn 2  2Fe E o  .7248V
 C 
G o   nFE  (6mole e  ) 96485 0.7248V   419594 J
 mol 

(d) 2Hg 2   2e   Hg 22 E o  0.920V

2Hg  Hg 22   2e  E 0  0.7973V
2Hg 2   2Hg  2Hg 22  E o  0.1227 V
 C 
G o   nFE  ( 2mole e  ) 96485 0.1227 V   23677 J
 mol 
 G° for the original reaction in (d) is one−half of the G° calculated since the reaction
formed from the half−reactions is double the original.

8.12. The intensive and extensive variables are related through n, the number of electrons
transferred in the process. It turns the intensive variable E into the extensive variable
Gº.

8.14. In order for the electrochemical process to provide that much work, the G of the process
must be greater than 5.00102 kJ. Therefore, we need to determine the E for each process
and from that calculate G. (a) For this reaction, E = 1.1037 V and G = −213 kJ, so this
reaction wouldn’t provide enough energy to perform the process. (b) For this reaction, E
= 2.868 V and G = −553 kJ, so this reaction would provide enough energy to perform
the process. (c) For this reaction, E = 2.212 V and G = −427 kJ, so this reaction
wouldn’t provide enough energy to perform the process. (d) For this reaction, E = 1.096
V and G = −211 kJ, so this reaction wouldn’t provide enough energy to perform the
process.

8.16. For the standard hydrogen electrode, the spontaneous process would be
2 Li + 2 H+  2 Li+ + H2
The voltage of this process is 3.04 V. For the standard calomel electrode, the process is
2 Li + Hg2Cl2  2 Li+ + 2 Hg + 2 Cl−

57
Instructor’s Manual

and the voltage is 3.31 V. Therefore, when the calomel half reaction is used as the
reduction reaction, the voltage shifts up by 0.2682 V. However, consider the other half
reaction with silver. With the hydrogen electrode, the spontaneous process is
2 Ag+ + H22 Ag + 2 H+
and the voltage is 0.7996 V. With the calomel electrode, the spontaneous reaction is
2 Ag+ + 2 Hg + 2 Cl− 2 Ag + Hg2Cl2
and the voltage is 0.5314 V, so the voltage shifts down by 0.2682 V. Therefore, the
direction of the shift depends on how the electrode reaction is used.

8.18. (a) The overall reaction can be broken down into two half reactions:
Au3+ + 3 e−  Au E = 1.498 V
Au  Au+ + e− E = −1.692 V
In order to determine the voltage of the overall reaction, we need to convert these
voltages into Gs and sum the Gs. For the first reaction: G = −(3 mol)(96,485
C/mol)(1.498 V) = −433.6 kJ. For the second reaction: G = −(1 mol)(96,485
C/mol)(−1.692 V) = +163.3 kJ. Therefore, the sum of the two reactions has a G of
(−433.6 + 163.3) = −270.3 kJ. Converting this back into an E for the two−electron
overall process: −270,300 J = −(2 mol)(96,485 C/mol)(E) E = 1.401 V.
(b) The overall reaction can be broken down into two half reactions:
Sn4+ + 2 e−  Sn2+ E = 0.151 V
Sn2+ + 2 e−  Sn E = −0.1375 V
In order to determine the voltage of the overall reaction, we need to convert these
voltages into Gs and sum the Gs. For the first reaction: G = −(2 mol)(96,485
C/mol)(0.151 V) = −29.1 kJ. For the second reaction: G = −(2 mol)(96,485
C/mol)(−0.1375V) = +26.5 kJ. Therefore, the sum of the two reactions has a G of
(−29.1 + 26.5) = −2.6 kJ. Converting this back into an E for the two−electron overall
process: −2600 J = −(4 mol)(96,485 C/mol)(E) E = 0.0067 V.

8.20. To determine if the process will work, we need to see if the following (unbalanced)
reaction is spontaneous: Al 3  Zn (s)  Al(s)  Zn 2  .
Breaking it into its half−reactions:
Zn  2e   Zn 2 E o  0.7618V
Al3  3e   Al E o  1.662V
E° for the overall reaction is −0.9002V. Since this is negative the process is
non−spontaneous and the plan would not work.

58
Chapter 8

8.22. We need to separate the reaction into its half−reactions:


Au  Au   e  E o  1.692V
4e   4H   O 2  2H 2 O E o  1.229V
E° for the overall reaction is −0.463V. Since this is negative the original reaction is
non−spontaneous and gold is resistant to corrosion.

8.24. Because these half−reactions typically occur in an aqueous solvent, the interactions
between the ionic species and the solvent molecules has an impact on the overall energy
change (in terms of E and G) of the process. Although all alkali metal ions have a +1
charge, the smaller, higher−charge−density lithium ion interacts more strongly with water
molecules, increasing the energy of the process.

8.26. If we reverse the bottom reaction and add the two together, we get:
CH 3 COCOO  (aq )  2H  (aq )  2e   CH 3 CHOHCOO  E '  0.166 V
HCOO   CO 2 (aq )  H  (aq )  2e  E '  0.414V
CH 3 COCOO  (aq )  HCOO   H  (aq )  CH 3 CHOHCOO   CO 2 (aq ) E '  0.248V

RT [ Zn 2 ]
8.28. EE  
ln Eº for this reaction is 1.1037 V, so we have
nF [Cu 2 ]

(8.314 J/K)(298.15 K) [ Zn 2 ]
1.000 V  1.1037 V  ln This rearranges to
(2 mol)(96,485 C/mol) [Cu 2 ]

[ Zn 2 ] [ Zn 2 ]
ln 2
 8.07279... 2
 3.21  10 3 . Unfortunately, we can’t mathematically
[Cu ] [Cu ]
determine specific concentrations without more information; we can only specify the
ratio.

8.30. First we need to determine E° for standard conditions:



6e   2MnO 4  8H   2MnO 2  4H 2 O E o  1.679V
6H 2 O  3H 2 O 2  6H   6e  E o  1.776V

2MnO 4  2H   2H 2 O(l)  2MnO 2 (s)  3H 2 O 2 E o  0.097V
Now, we can use the Nernst Equation to solve for E.
 J 
 8.314 298.15K 
E  Eo 
RT
ln
H 2 O 2 
3

 0.097 V   Kmol 
ln
0.0705M 3
nF  
2
H  MnO 4

2

  


6mol e  96485   
C  1.22  10  4 M 2.66M 2
2

 mol e  
 0.132V

59
Instructor’s Manual

 S  
8.32. According to equations 8.27 and 8.28, E  E    T . To determine E at 1700ºC,
 nF 
we need to calculate the S for the reaction. Using data in the appendix: S = [50.92 +
2(27.3)] – [2(28.30) + 87.4] = −38.5 J/K. Six moles of electrons are transferred in the
reaction. Substituting:
- 38.5 J/K
E  1.625 V   1675 K  1.625  0.111  1.514 V .
(6 mol)(96,485 C/mol)

 J 
 8.314 298.15K 
RT  Kmol  0.0077M
8.34. E 0 ln Q   ln  0.0194V
  C  0.035M
nF
 
2mol e  96485 
 mol e  

8.36. (a) Since the hydrogen half reaction has a voltage of 0.000 V, the Eº for the reaction is
RT
–0.044 V. Using the equation E   ln K :
nF
(8.314 J/mol  K)(298.15 K)
 0.044 V  ln K . Solving for K: K = 3.2510−2. (b)
(2 mol)(96,485 C/mol)
According to the equilibrium constant, D+ prefers to be in solution over H+.

 S 
8.38. In order to use the equation E  E 
  T , we will need the entropy change of the
 nF 
reaction. The entropy of H+ (aq) is defined as zero, and we are assuming that S(D+, aq) is
zero also. Therefore, the entropy change of the reaction is (using data from the appendix)
144.96 – 130.68 = 14.28 J/K. The number of electrons transferred is 2 so we have:
 14.28 J/K 
0  0.044 V   T Solving for T: T = 595 K. Therefore, if
 (2 mol)(96,485 C/mol) 
we raise the temperature from 298 K to (298 + 595) = ~893 K, the voltage of the reaction
should be about 0.

60
Chapter 8

 S 
8.40. The temperature coefficient is equal to:   . We need to calculate S°:
 nF 
   
S o  2  S o H 2 O(l)   S o CO 2 (g )   S o CH 4 (g )   2  S o O 2 (g )  
 J  J J J J
2   69.91   213.79  188.66  2  205.14  245
 mol  K  mol  K mol  K mol  K K
J
 245
S o
K V
  3.18  10  4
nF
 

8mol e   96485
C 

mol 
K

The 8 comes from an analysis of the half−reactions involved in this overall reaction.

 E o 
8.42. Using equation 8.30: H o  nF E o  T  , we can determine the temperature
 T 
E o
coefficient, . First, we need to find E° for the reaction:
T
H 2  2H   2e  E o  0.0V
E° for the overall reaction is equal to 0.5355V.
2e   I 2  2I  E o  0.5355V

 
 
 53  10 3 J   2mol e   96485
C
0.5355  298.15x 
 mol e  
V
x  2.72  10 3
K

 H 
8.44. Since heat capacity (at constant pressure) is defined as   , we can take the
 T  p
derivative of equation 8.30 with respect to temperature:
  (H)   E   E   2E     2E 
    nF   T  , which simplifies to C   nF T  .
 T  p  T  T 2  p
  T
2
  T   

8.46. First we start with the two half−reactions:


2H 2 O  2e   H 2  2OH  E o  0.8277V
H 2  2 H   2e  E o  0 .0 V
2H 2 O  2H   2OH  E o  0.8277V

61
Instructor’s Manual

Using Eqn. 8.32 and assuming molar concentrations,


 J 
 8.314 298.15K 
RT  Kmol 
E 
o
ln K  ln K  0.8277 V
  C 
nF

2mol e  96485 
 mol e  
K  1.0  10  28
This value makes sense since it is the square of Kw=1.0x10−14. The sum of the two
electrochemical half−reactions is double the equation used to find Kw.

8.48. Following Example 8.7: the reaction can be written in terms of two half reactions:
AgCl (s) + e−  Ag (s) + Cl− (aq) E = 0.22233 V
+ −
Ag (s)  Ag + e E = −0.7996 V
The overall voltage of the combination of the two reactions is –0.5773 V. Using equation
8.32:
(8.314 J/K)(298.15 K)
 0.5773 V  ln K sp Solve for Ksp: Ksp = 1.7410−10.
(1 mol)(96,485 C/mol)

8.50. Following Example 8.7: the reaction can be written in terms of two half reactions:
Hg2Cl2 (s) + 2 e−  2 Hg (s) + 2 Cl− (aq) E = 0.2682 V
+2 −
2 Hg (s)  Hg2 + 2 e E = −0.7973 V
The overall voltage of the combination of the two reactions is –0.5291 V. Using equation
8.32:
(8.314 J/K)(298.15 K)
 0.5291 V  ln K sp Solve for Ksp: Ksp = 1.2910−18.
(2 mol)(96,485 C/mol)

8.52. Following Example 8.8, E   0.05916  pH   E o (of other half  reaction )


 0.400V   0.05916  pH   0.3419V
pH  0.982

8.54. The Ksp for Hg2Cl2 was determined in exercise 8.28 and is 1.2910−18. If x moles per liter
of Hg2Cl2 dissociates, one gets x M Hg22+ and 2x M Cl−. Therefore, we have:
1.2910−18 = (x)(2x)2 1.2910−18 = 4x3 x = 6.8610−7 Since the equilibrium concentration
of Cl− is twice this, [Cl−] = 1.3810−6 M.

62
Chapter 8

RT
8.56. For this we need equation: E  E o  ln Q . The two differences between chemical
nF
standard states and biochemical standard states are the temperature (37°C for biochem)
and concentration of H+ (1x10−7M for biochem).

E'  0.105V 
(8.314 J/K)(310.15 K)
ln
1M   0.320V.

(2 mol)(96,485 C/mol) (1M ) 1  10 7 M 
8.58. Start with the definition that a   (  m  ) n , which is the definition of mean activity from
equation 8.46 but without the standard molality in the denominator. First, distribute the
exponent: a   n  m n  . Now we focus on the molality term. First, let us substitute for

the mean ionic molality from equation 8.43: a   n m m  n


  n
n  1 /( n   n  ) 
Since n = n+ +
n−, the outermost exponent cancels with the 1/(n+ + n−) exponent inside the parentheses,
leaving a   n  m n  m n  . For a salt having formula An+Bn−, for a certain molality m the
molality of the positive ion, m+, is n+m and the molality of the negative ion, m−, is n−m.
Substituting into the expression for a: a   n (n m) n  (n m) n  , which can be rearranged
by distributing the exponents: a   n  n n  m n  n n  m n  . Bringing the two molality terms
together: a   n  m n   n  n n  n n  , and now realizing again that n = n+ + n−, we do one
final substitution for the exponent on m: a   n  m n  n n  n n  , which is the expression
we’re looking for.

8.60. Using the definition of ionic strength in equation 8.47:

(a) I 
1
2
 
(0.0055m)(1) 2  (0.0055m)( 1) 2  0.0055m

(b) I 
1
2
 
(0.075m)( 1) 2  (0.075m)( 1) 2  0.075m

(c) I 
1
2
 
(0.0250 m)(2) 2  (0.0500 m)( 1) 2  0.0750 m

(d) I 
1
2
 
(0.0250 m)( 3) 2  (0.0750 m)( 1) 2  0.150 m

8.62. We will assume that the concentrations are so low that we can assume that the molarity is
equal to the molality.
(a)
I int ra 
1
2
0.012m  12  0.139m  12  0.004m  12  0.012m  12  0.0835m 
I extra 
1
2
0.140m  12  0.005m  12  0.105m  12  0.024m  12  0.137 m 
(b) Speculation is left to the student. Hint: osmoregulation

63
Instructor’s Manual

8.64. In the equation H2 (g) + I2 (s)  2 H+ (aq) + 2 I− (aq), the overall enthalpy of reaction is –
110.38 kJ. Using the concept of products−minus−reactants to determine H, we need the
heats of formation of the products (one of which is the object of this calculation) and the
heats of formation of the reactants. The two reactants are elements, so their fHs are zero.
By convention, the fH of H+ (aq) is also zero, so the only non−zero fH is that for I−.
Therefore, we have
−110.38 kJ = (2 mol)(fH[I−]) fH [I−] = −55.19 kJ/mol.

2
8.66. (a) The real equation of interest is: Ca 2   CO 3  CaCO 3 (s, arag ) .

  
H o   f H o CaCO 3 , (s)    f H o Ca 2   f H o CO 3
2

kJ  kJ   kJ 
 1207.1    542.83     413.8   250.5kJ
mol  mol   mol 
  
G o   f G o CaCO 3 , (s)    f G o Ca 2   f G o CO 3
2

kJ  kJ   kJ 
 1127.8    553.54     386.0   188.3kJ
mol  mol   mol 

(b) The real equation of interest is: Ba 2   SO 4 2   BaSO 4 (s )

  
H o   f H o BaSO 4 , (s)    f H o Ba 2   f H o SO 4
2

kJ  kJ   kJ 
 1473.19    537.64     909.3   26.25kJ
mol  mol   mol 
  
G o   f G o BaSO 4 , (s)    f G o Ba 2   f G o SO 4
2

kJ  kJ   kJ 
 1362.3    560.77     744.6   56.93kJ
mol  mol   mol 

8.68. H o   f H o (Ca 2 )  2   f H o (Cl  )   f H o (CaCl 2 (s))

 kJ   kJ 
 81.3kJ    542.83   2x    795.80 
 mol   mol 
x   f H o (Cl  )  167.1kJ
This value is very close to the value calculated in Example 8.11.

8.70. As you go from F− to I− in column 7A, the ions become increasingly large. fHo’s are
obviously more negative for the smaller ions. This large negative number represents
greater stability of the smaller ions. This makes sense because if the ion is smaller, water
molecules can approach the ion more closely. From Coulomb’s law, we know that this
increases the binding energy between the ion and the water molecules and results in
greater stability of the ion.

64
Chapter 8

8.72. The complete expression for A is worked out in the text. The student need simply verify
that the numbers and units do reduce to 1.171 molal−1/2.

8.74. 0.9% NaCl implies 0.9 g NaCl in 99.1 g water. Assuming that the volume of the solution
(0.9 g)(1 mol/58.5 g)
is 100 mL = 0.100 L = 0.100 kg:  0.154 molal . Therefore, the
0.100 kg
ionic strength is I  (0.154 molal)(1) 2  (0.154 molal)(-1) 2   0.154 molal .
1
2

8.76. From Coulomb’s law, we know that the force of interaction for higher−charged ions
qq
(q=2,3..) is much greater than that of singly charged ions: F  1 2 2 . In addition, the
r
ionic strength of a solution is much higher for higher−charged ions since the ionic
strength is proportional to the square of the charge on the ion. Since the Debye−Huckel
theory assumes solutions with very low ionic strength, solutions with many
higher−charged ions can be problematic.

8.78. (a) Identity of the counterion is necessary to determine the ionic strengths of the
solutions.
(b) If the counterion were sulfate instead of nitrate, the ionic strengths would need to be
recalculated. If sulfate were the counterion, the salt’s formula is Fe2(SO4)3 and the
sulfate concentration is 3/2 of the iron ion concentration:
I (Fe3+ soln) = ½[(0.100)(+3)2 + (0.150)(−2)2] = 0.750 molal
If sulfate were the counterion, the other salt’s formula would be CuSO4 and the
sulfate concentration would equal the copper ion concentration:
I (Cu2+ soln) = ½[(0.050)(+2)2 + (0.050)(−2)2] = 0.200 molal
Using equation 8.52 for each ion:
(1.171 molal-1/2 )(3) 2 (0.750 molal)1/2
ln  Fe3 
1  (2.32  10 9 m -1 molal-1/2 )(9.00  10 10 m)(0.750 molal)1/2
ln  Fe 3  3.250  (Fe3+) = 0.0388 Therefore, the activity of Fe3+ is (0.0388)(0.100
m) = 0.00388 m.
Similarly, for the Cu2+:
(1.171 molal1/2 )(2) 2 (0.200 molal)1/2
ln  Cu 2   
1  (2.32  10 9 m -1 molal-1/2 )(6.00  10 10 m)(0.200 molal)1/2
ln  Cu 2   1.291  (Cu2+) = 0.275 Therefore, the activity of Cu2+ is (0.275)(0.050 m)
= 0.0138 m.

65
Physical Chemistry 2nd Edition Ball Solutions Manual

Instructor’s Manual

Substituting these activities into the Nernst equation:


(8.314 J/K)(298.15K) (0.00388) 2
E  0.379 V  ln  0.379  0.0075  0.386 V
(6 mol)(95,485 C/mol) (0.0138) 3
When you compare this to the voltage from the example (0.372 V), we see how the
ionic strength of the solution, as influenced by the counterion, can have an influence
on the voltage even though the counterions don’t participate in the reaction.

N  E
8.80. Equation 8.61 is I  e 2  | z | 2  i   A  . Equation 8.5 shows that E has units of
V  6ri
N/C, e has units of C, z is unitless (it is simply the magnitude of the charge on the ion),
the fraction (Ni/V) has units of 1/m3, A has units of m2, and in SI units, viscosity has units
of kg/ms. The numbers 6 and  have no units. The radius r has units of m. Combining all
of these units:
 1  2 N/C C2  m2  N  m  s
C  3 m 
2
This can be rearranged to get . One of the
m  (kg/m  s)(m) C  m 3  m  kg
C units cancels, as do the three m units in the numerator and three of the four m units in
C N s
the denominator. This results in . If we break down the newton unit into kgm/s2,
kg  m
we have
C  kg  m  s
. The kg and m units cancel, as do one of the second units in the numerator
s 2  kg  m
and denominator. What’s left is C/s, and one coulomb per second is an ampere, the unit
of current.

8.82. For a galvanic cell, oxidation (the loss of electrons) occurs at the anode and is considered
the negative electrode. Therefore, I− is the current towards the cathode, and I+ must be the
current towards the anode. In an electrolytic cell, oxidation (the loss of electrons) still
occurs at the anode, so I− is still the current towards the cathode and I+ is the current
towards the anode. For any given cell reaction, however, the identities of the cathode and
anode are switched for galvanic and electrolytic cells.

66

Visit TestBankBell.com to get complete for all chapters

You might also like