Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

21 Reaction dynamics

21A Collision theory


Answers to discussion questions
21A.2 To the extent that real gases deviate from perfect gas behavior, they do so because of
intermolecular interactions. Interactions tend to be more important at high pressures, when
the size of the molecules themselves is not negligible compared to the average
intermolecular distance (mean free path). Attractive interactions, might enhance a reaction
rate compared to the predictions of collision theory, particularly if the parts of the molecules
that are attracted to each other are the reactive sites. (In that case, the both the collision
frequency and the steric factor might be enhanced.) Similarly, repulsive interactions might
reduce the frequency of collisions compared to what would be predicted for perfect gases.
In supercritical fluids, densities can be comparable to those of liquids, so the considerations
explored in the next topic (Diffusion-controlled reactions, Topic 21B) for reactions in
solution might be more relevant than those of a perfect gas.
21A.4 The RRK theory proposes a P-factor that is more related to statistical energetic
considerations than to geometric (“steric”) ones. The P-factor in RRK theory is [21A.10a]:
s−1
 E *
P =  1−
 E 
where E* is the energy required to break a bond (leading to reaction), E the energy of the
collision, and s the number of modes over which the energy can be dissipated. Like more
geometric interpretations of the P-factor, the RRK theory assigns smaller P factors to
complex molecules than to simple ones, but for different reasons. A more geometric theory
would say that an active site is only a small fraction of the “surface area” of a complex
molecule, whereas RRK theory says that complex molecules are much more effective than
simple ones at dispersing energy away from the reactive site.
Solutions to exercises
21A.1(b) The collision frequency is [1B.11a]
z = σvrelN
1/2
 16kT  p
where vrel =  [1B.10a & 1B.9], σ = πd2 = 4πR2, and N =
 π m  kT
1/2 1/2
 16kT   π 
Therefore, z = σ N  = 16 pR 2 
 π m   mkT 
= 16 × (120 × 103 Pa) × (180 × 10−12 m)2
1/2
 π 
× −27 −1 −23 −1 
 28.01 mu × 1.661 × 10 kg mu × 1.381 × 10 J K × 303 K 
= 7.90 × 109 s −1
The collision density is [Justification 21A.1]
zN A z  p  7.90 × 109 s −1  120 × 103 Pa 
Z= =  =  
2 2  kT  2 −23 −1
 1.381 × 10 J K × 303 K 
= 1.13 × 1035 s −1 m −3
For the percentage increase at constant volume, note that N is constant at constant volume,
so the only constant-volume temperature dependence on z (and on Z) is in the speed factor.
1  ∂z  1 1  ∂Z  1
z ∝ T1/2 so = and =
z  ∂T  V 2T Z  ∂T  V 2T

1
Therefore δ z = δ Z ≈ δ T = 1  10 K  = 0.017
z Z 2T 2  303 K 
so both z and Z increase by about 1.7% .
21A.2(b) The fraction of collisions having at least Ea along the line of flight may be inferred by
dividing out of the collision-theory rate constant (eqn. 21A.9) those factors that can be
− E /RT
identified as belonging to the steric factor or collision rate: f = e a

 −15 × 103 J mol


−1
 −3
(i) (1) f = exp  −1 −1  = 2.4 × 10
 (8.3145 J K mol ) × (300 K) 
 −15 × 103 J mol
−1

(2) f = exp  −1 −1  = 0.105
 (8.3145 J K mol ) × (800 K) 
 −150 × 103 J mol
−1
 −27
(ii) (1) f = exp  −1 −1  = 7.7 × 10
 (8.3145 J K mol ) × (300 K) 
 −150 × 103 J mol
−1
 −10
(2) f = exp  −1 −1  = 1.6 × 10
 (8.3145 J K mol ) × (800 K) 
− E /RT
21A.3(b) A straightforward approach would be to compute f = e a at the new temperature and
compare it to that at the old temperature. An approximate approach would be to note that f
− E /RT  − Ea 
changes from f0 = e a to exp   , where x is the fractional increase in the
 RT (1 + x) 
temperature. If x is small, the exponent changes from –Ea/RT to approximately –Ea(1–x)/RT
and f changes from f0 to
( )
−x
f ≈ e − Ea (1=
− x )/ RT
e − Ea / RT e − Ea /=
RT
f 0 f 0− x
Thus the new fraction is the old one times a factor of f0–x . The increase in f expressed as a
percentage is
f − f0 f f − x − f0
× 100% = 0 0 × 100% = ( f0 − x − 1) × 100%
f0 f0
(1) (1) f0− x = (2.4 × 10−3 )−10/300 = 1.2 and the percentage change is 20% .
(2) f0− x = (0.105)−10/800 = 1.03 and the percentage change is 3% .
(ii) (1) f0− x = (7.7 × 10−27 )−10/300 = 7.4 and the percentage change is 640% .
(2) f0− x = (1.6 1 × 10−10 )−10/800 = 1.33 and the percentage change is 33% .
1/2
 8kT 
− Ea /RT
21A.4(b) kr = Pσ   N Ae [21A.9]
 πµ 

We take P = 1, so
1/2
 8(1.381 × 10−23 J K −1) × (450 K) 
−9
kr = [0.30 × (10 m) ] ×  2

 π (3.930 m ) × (1.661 × 10−27 kg m −1) 
u u

 −200 × 103 J mol


−1

×(6.022 × 1023 mol−1 ) × exp  −1 −1 
 (8.3145 J K mol ) × (450 K) 
= 1.7 × 10−15 m 3 mol−1 s −1 = 1.7 × 10−12 dm 3 mol−1s −1 .

21A.5(b) The steric factor, P, is


σ*
P= [Topic 21A.1(c)]
σ
The mean collision cross-section is σ = πd2 with d = (dA + dB)/2
2
Get the diameters from the collision cross-sections:
dA = (σA/π)1/2 and dB = (σB/π)1/2 ,
(σ + σ B1/ 2 ) {(0.88 nm ) + (0.40 nm 2 )1/ 2 }
2 2 2
π  σ A   σ  
1/ 2 1/ 2 1/ 2 2 1/ 2

σ= + B   =
A
 =
4  π 
so
 π   4 4

= 0.62 nm 2 .
8.7 × 10−22 m 2
Therefore, P = = 1.41 × 10−3
0.62 × (10−9 m)2
21A.6(b) According to RRK theory, the steric P-factor is given by eqn. 21A.10a
s−1
 E *
P = 1 −
 E 
where s is the number of vibrational modes in the reacting molecule. For a non-linear
molecule composed of N atoms, the number of modes is [Topic 12E.1]
s = 3N – 6 = 3×4 – 6 = 6.
Rearranging eqn. 21A.10a yields
1 1
E*
= 1 − P s−1 = 1 − (0.025) 5 = 0.52
E
21A.7(b) According to RRK theory, the steric P-factor is given by eqn. 21A.10a
12−1
 300 kJ mol−1 
s−1
 E *
P = 1 − = 1 −  = 4.2 × 10−5
 E   500 kJ mol−1 

Solutions to problems
21A.2 Draw up the following table as the basis of an Arrhenius plot:

T/K 600 700 800 1000


3
10 K / T 1.67 1.43 1.25 1.00
kr / (cm3 mol–1 s–1) 4.6×102 9.7×103 1.3×105 3.1×106
ln (kr / cm3 mol–1 s–1) 6.13 9.18 11.8 14.9

The points are plotted in Figure 21A.1.

Figure 21A.1

The least-squares intercept is at 28.3, which implies that


A / (cm3 mol–1 s–1) = e28.3 = 2.0×1012
3
1/2
 8kT 
But comparison of eqn. 21A.9 to the Arrhenius equation tells us that A = N A Pσ 
 πµ 
1/2
A  πµ 
so P=
N Aσ  8kT 
The reduced mass is
µ = m(NO2)/2 = 46 mu × (1.661×10–27 kg mu–1) / 2 = 3.8×10–26 kg
so, evaluating P in the center of the range of temperatures spanned by the data,
1/2
2.0 × 1012 × (10−2 m)3 mol−1 s −1  π × 3.8 × 10−26 kg 
P= ×
(6.022 × 10 mol ) × 0.60 × (10 m)  8 × 1.381 × 10 J K × 800 K 
23 −1 −9 2  −23 −1

= 6.5 × 10−3 .
σ* = Pσ = (6.5×10–3) × (0.60 nm2) = 3.9×10–3 nm2 = 3.9×10–21 m2
21A.4 Example 21A.1 estimates a steric factor within the harpoon mechanism:
2
σ*  e2 
≈
σ  4πε 0 d(I − Eea ) 
Taking σ = πd2 gives
2
 e2  6.5 nm 2
σ* ≈ π   =
 4πε 0 [I(M) − Eea (X 2 )]  {(I − Eea ) / eV}2
Thus, σ* is predicted to increase as I – Eea decreases. We construct the following table from
the data:

σ* /nm2 Cl2 Br2 I2


Na 0.45 0.42 0.56
K 0.72 0.68 0.97
Rb 0.77 0.72 1.05
Cs 0.97 0.90 1.34

All values of σ* in the table are smaller than the experimental ones, but they do show the
correct trends down the columns. The variation with Eea across the table is not so good.
21A.6 Collision theory gives for a rate constant with no energy barrier
1/2 1/2
 8kT  k  πµ 
kr = Pσ  N A [21A.9] so P = r 
 πµ  σ N A  8kT 
kr / (dm 3 mol−1 s −1 ) × (10−3 m 3 dm −3 )
P=
(σ / nm 2 ) × (10−9 m)2 × (6.022 × 1023 mol−1 )
1/2
 π × ( µ / u) × (1.66 × 10−27 kg) 
× −23 −1 
 8 × (1.381 × 10 J K ) × (298 K) 
(6.61 × 10−13 )kr / (dm 3 mol−1 s −1 )
=
(σ / nm 2 ) × ( µ / mu )1/2
The collision cross-section is
1 σ 1/2 + σ 1/2 (σ 1/2 + σ 1/2 )2
σ AB = π dAB
2
where dAB = (dA + d B ) = A B
so σ = A B
2 2π 1/2 AB
4
The collision cross-section for O2 is listed in the Data Section. We would not be far wrong if
we took that of the ethyl radical to equal that of ethene; similarly, we will take that of
cyclohexyl to equal that of benzene. For O2 with ethyl
(0.401/2 + 0.641/2 )2
σ= nm 2 = 0.51nm 2
4

4
mO met (32.0 mu ) × (29.1mu )
µ= = = 15.2 mu
mO + met (32.0 + 29.1) mu
(6.61 × 10−13 ) × (4.7 × 109 )
so P= = 1.6 × 10−3
(0.51) × (15.2)1/2
For O2 with cyclohexyl
(0.401/2 + 0.881/2 )2
σ= nm 2 = 0.62 nm 2
4
m m (32.0 mu ) × (77.1mu )
µ= O C = = 22.6 mu
mO + mC (32.0 + 77.1) mu
(6.61 × 10−13 ) × (8.4 × 109 )
so P= = 1.8 × 10−3
(0.62) × (22.6)1/2

21B Diffusion-controlled reactions


Answer to discussion question
D21B.2 In the cage effect, a pair of molecules may be held in close proximity for an extended period
of time (extended on the microscopic scale, mind you) by the presence of other neighboring
molecules, typically solvent molecules. Such a pair is called an encounter pair, and their
time near each other an “encounter” as opposed to a simple collision. An encounter may
include a series of collisions. Furthermore, an encounter pair may pick up enough energy to
react from collisions with neighboring molecules, even though the pair may not have had
enough energy at the time of its initial collision.

Solutions to exercises
21B.1(b) The rate constant for a diffusion-controlled bimolecular reaction is
kd = 4πR*DNA [21B.3]
where D = DA + DB = 2×(5.2×10–9 m2 s–1) = 1.04×10–8 m2 s–1
kd = 4π × (0.4×10–9 m) × (1.04×10–8 m2 s–1) × (6.022×1023 mol–1)
kd = 3. 1 × 107 m 3 mol−1 s −1 = 3. 1 × 1010 dm 3 mol−1 s −1

21B.2(b) The rate constant for a diffusion-controlled bimolecular reaction is


8 × (8.3145 J K −1 mol−1 ) × (298 K) 6.61× 103 J mol−1
kd = 8RT [21A.4] = =
3η 3η η
(i) For decylbenzene, η = 3.36 cP = 3.36×10–3 kg m–1 s–1
6.61 × 103 J mol−1
kd = = 1.97 × 106 m 3 mol−1 s −1 = 1.97 × 109 dm 3 mol−1 s −1
3.36 × 10−3 kg m −1 s −1
(ii) In concentrated sulfuric acid, η = 27 cP = 27×10–3 kg m–1 s–1
6.61 × 103 J mol−1
kd = = 2.4 × 105 m 3 mol−1 s −1 = 2.4 × 108 dm 3 mol−1 s −1
27 × 10−3 kg m −1 s −1
21B.3(b) The rate constant for a diffusion-controlled bimolecular reaction is [21B.4]
8 × (8.3145 J K −1 mol−1 ) × (320 K)
kd = 8RT =
3η 3× (0.601× 10−3 kg m −1 s −1 )
= 1.18 × 107 m 3 mol−1 s −1 = 1.18 × 1010 dm 3 mol−1 s −1
Since this reaction is elementary bimolecular it is second-order; hence

5
t1/2 = 1 [Table 20B.3, with kr = 2kd because 2 atoms are consumed]
2kd [A]0

so t1/2 = 1 = 2.1 × 10−8 s


−1 −1 −3 −3
2 × (1.18 × 10 dm mol s ) × (2.0 × 10 mol dm )
10 3

21B.4(b) Since the reaction is diffusion-controlled, the rate-limiting step is bimolecular and therefore
second-order; hence
d[P]
= kd [A][B]
dt
where kd = 4π R∗ DN A [88.3] = 4π N A R∗ (DA + DB )
kT  1 1 2RT  1 1
= 4π N A × (RA + RB ) ×  +  [19B.19b] = (RA + RB ) ×  + 
6πη  RA RB  3η  RA RB 
2 × (8.3145 J K −1 mol−1 ) × (293K)  1 1 
kd = × (421 + 945) ×  +
−3
3 × (1.35 × 10 kg m s )−1 −1
 421 945 
= 5.64 × 106 m 3 mol−1 s −1 = 5.64 × 109 dm 3 mol−1 s −1 .
Therefore, the initial rate is
d[P]
= (5.64 × 109 dm 3 mol−1 s −1 ) × (0.155mol dm −3 ) × (0.195mol dm −3 )
dt
= 1.71 × 108 mol dm −3 s −1
Comment. If the approximation of eqn. 21B.4 is used, kd = 4.81×109 dm3 mol–1 s–1 . In this
case the approximation results in a difference of about 15% compared to the expression
used above.

Solutions to problems
21B.2 See Brief illustration 21B.3 for a sample scenario. In the graphs shown here, the same
parameters are used, except for the value of the rate constant. That is, n0 = 3.9 mmol of I2, A =
5.0 cm2, and D = 4.1×10–9 m2 s–1. Using these parameters, we will plot the spatial variation in
concentration at 102 s, 103 s, and 104 s. In Figure 21B.1(a), the concentration is plotted against
position in the absence of reaction. That is, kr = 0. The concentration profile spreads with time.
That is, the maximum concentration (which stays at the origin throughout) decreases, but the
distance over which there is appreciable concentration increases. Introducing a first-order
reaction with a rate constant kr = 4.0×10–5 s–1, as in the Brief illustration, has a barely
noticeable effect on the concentration profiles (plotted in Figure 21B.1(b)); the longer-time
profiles are very slightly depressed. Making the rate constant a factor of 10 greater suppresses
the 104-s profile: by that time the material has practically all reacted away (Figure 21B.1(c)).
Speeding up the reaction by a further factor of 10 (Figure 21B.1(d)) suppresses the 103-s
profile practically completely as well, and even the 102-s is visibly lower.

6
Figure 21B.1(a)

Figure 21B.1(b)

Figure 21B.1(c)

Figure 21B.1(d)

Examination of eqn. 21B.9 suggests that the spatial variation is not significantly changed by
− kr t
reaction; it is merely scaled equally at all positions by the factor e .
7
21B.4 (a) The rate constant of a diffusion-limited reaction is
8RT 8 × (8.3145 J K −1 mol−1 ) × (298 K)
kd = [21B.4] =
3η 3 × (1.06 × 10−3 kg m −1 s −1 )
= 6.23 × 106 m 3 mol−1 s −1 = 6.23 × 109 dm 3 mol−1s −1
(b) The rate constant is related to the diffusion constants and reaction distance by
kd = 4πR*DNA [21B.3]
kd (2.77 × 109 dm 3 mol−1 s −1 ) × (10−3 m 3 dm −3 )
so R*= =
4π DN A 4π × (1 × 10−9 m 2 s −1 ) × (6.022 × 1023 mol−1 )
= 4 × 10−10 m = 0.4 nm

21C Transition-state theory


Answers to discussion questions
21C.2 See Topic 21C.1(e) for detailed examples of how femtosecond spectroscopy has been used
to detect activated complexes and transition states of reactions. Because the activated
complexes are not even local minima on potential energy surfaces, they are extremely
transitory, and laser pulses of duration less than 1 ps are needed in order to detect them.
Typically one very short pulse will initiate the reaction under investigation, thereby creating
a dissociative activated complex, and a second pulse will detect a reaction product. Not only
must the pulses themselves be very short, but the delay between the creation of the complex
and the detection of its effects must also be short. By such techniques, investigators have
been able to determine just how stretched the bond in ICN must get before it breaks
(yielding free CN). Also, decay of the ion pair Na+I– has been studies in detail, revealing the
existence of two potential energy surfaces, one largely ionic and one corresponding to a
covalent NaI.
21C.4 The primary isotope effect is the change in rate constant of a reaction in which the breaking
of a bond involving the isotope occurs. The reaction coordinate in a C–H bond-breaking
process corresponds to the stretching of that bond. The vibrational energy of the stretching
depends upon the effective mass of the C and H atoms (µCH). Upon deuteration, the zero
point energy of the bond is lowered due to the greater mass of the deuterium atom.
However, the height of the energy barrier is not much changed because the relevant
vibration in the activated complex has a very low force constant (bonding in the complex is
very weak), so there is little zero point energy associated with the complex and little change
in its zero point energy upon deuteration. The net effect is an increase in the activation
energy of the reaction. We then expect that the rate constant for the reaction will be lowered
in the deuterated molecule and that is what is observed. See the derivation leading to eqns
21C.19 and 21C.20 for a quantitative description of the effect.
Sometimes the rate of reaction is lowered upon deuteration to an extent even greater than
can be accounted for by these equations. In such cases, quantum-mechanical tunneling
(Topic 8A) may be part of the reaction mechanism. The probability of tunneling is highly
sensitive to mass, so it is much less likely (and therefore much slower) for deuterium than
for 1H.
If the rate of a reaction is altered by isotopic substitution it implies that the substituted site
plays an important role in the mechanism of the reaction. For example, an observed effect
on the rate can identify bond breaking events in the rate determining step of the mechanism.
On the other hand, if no isotope effect is observed, the site of the isotopic substitution may
play no critical role in the mechanism of the reaction.

Solutions to exercises
21C.1(b) The enthalpy of activation for a bimolecular solution reaction is [Topic 21C.2(a) footnote]
∆‡H = Ea – RT = 8.3145 J K–1 mol–1 × (5925 K – 298 K) = 46.9 kJ mol–1
8
 kT   RT  kRT 2
kr = B e ∆ S / R e −∆ H / RT , B =  × O =
‡ ‡

 h  p
O
 hp
e ∆ S / R e − Ea / RT e A e − Ea / RT
= B=

 A 
Therefore, A = e B e ∆ S / R , implying that ∆= S R  ln − 1

 B 
(1.381 × 10−23 J K −1 ) × (8.3145J K −1 mol−1 ) × (298 K)2
B=
6.626 × 10−34 J s × 105 Pa
= 1.54 × 1011 m 3 mol−1 s −1 = 1.54 × 1014 dm 3 mol−1 s −1
  6.92 × 1012 dm 3 mol−1 s −1  
and hence ∆‡ S = R  ln  −1 −1 
− 1
  1.54 × 10 dm mol s  
14 3

= 8.3145 J K −1 mol−1 × (−4.10) = −34.1 J K −1 mol−1

21C.2(b) The enthalpy of activation for a bimolecular solution reaction is [Topic 21C.2(a) footnote]
∆‡H = Ea – RT = 8.3145 J mol–1 K–1 × (4972 K – 298 K) = 38.9 kJ mol–1
The entropy of activation is [Exercise 21C.1(b)]
 A 
∆=‡
S R  ln − 1
 B 
kRT 2
with B= = 1.59 × 1014 dm 3 mol−1 s −1
hp O
  4.98 × 1013  
Therefore, ∆‡ S =8.3145 J K −1 mol−1 × ln  14 
− 1 =−17.7 J K −1 mol−1
  1 .54 × 10  
Hence, ∆‡G = ∆‡H – T∆‡S = {38.9 – (298) × (–17.7×10–3)} kJ mol–1 = +44.1 kJ mol–1
21C.3(b) Use eqn. 21C.15(a) to relate a bimolecular gas-phase rate constant to activation energy and
entropy:
kr = e 2 B e ∆ S / R e − Ea / RT

 kT   RT 
where B =   ×  O  [21C.14]
 h  p 
(1.381× 10−23 J K −1 ) × (338 K)2 × (8.3145 J mol−1 K −1 )
=
(6.626 × 10−34 J s) × (105 Pa)
= 1.98 × 1011 m 3 mol−1 s −1
Solve for the entropy of activation:
 k  E
=
∆‡ S R  ln r − 2  + a
 B  T
 0.35 m3 mol−1 s −1  39.7 × 103 J mol−1
Hence ∆‡ S 8.3145 J K −1 mol−1 ×  ln
= −1 −1
− 2 +
 1.98 × 10 m mol s
11 3
 338 K
= −124 J K −1 mol−1
21C.4(b) For a bimolecular gas-phase reaction [Exercise 21C.3(b)],
 k  Ea  A E  Ea  A 
∆=

S R  ln r − 2  + = R  ln − a − 2  + = R  ln − 2 
 B  T  B RT  T  B 
kRT 2
where B =
hp O
For two structureless particles, the rate constant is the same as that of collision theory
1/2
 8kT  −∆E0 /RT
kr = N A σ ∗  e [Example 21C.1]
 πµ 
9
The activation energy is [20D.3]
d ln kr d  1 8k 1 ∆E0 
 ln N Aσ + 2 ln πµ + 2 ln T − RT 

Ea = RT 2 = RT 2
dT dT  
 1 ∆E0  RT
= RT 2  +  = ∆E0 + 2 ,
 2T RT 2 
so the prefactor is
1/ 2 1/ 2
 8kT  −∆E0 / RT ∆E0 / RT 1/ 2  8kT  1/ 2
= A k= re N Aσ 
Ea / RT ∗
 e e = e( )
N Aσ ∗   e
 πµ   πµ 
  8kT 
1/ 2
1 kRT 2   σ ∗ p O h  8 1/ 2 3 
Hence ∆‡ S R ln N Aσ ∗ 
=  += − ln − 2  R ln 3/ 2   − .
  πµ  2 pOh   (kT )  πµ  2 
For identical particles,
µ =m/2 = (92 u)(1.661×10–27 kg u–1)/2 = 7.6×10–26 kg ,
and hence
∆‡ S = 8.3145 J K −1 mol−1
 0.45 × (10−9 m) 2 × 105 Pa × 6.626 × 10−34 J s  8 
1/ 2
3 
× ln   − 
 π × 7.6 × 10 kg 
−23 −1 −26
 (1.381× 10 J K × 450 K) 3/ 2
2 

= −79 J K −1 mol−1 .

21C.5(b) At low pressure, the reaction can be assumed to be bimolecular. (The rate constant is
certainly second-order.)
 A 
(a) ∆=‡
S R  ln − 2  [Exercise 21C.4(b)]
 B 
kRT 2 1.381× 10−23 J K −1 × 8.3145 J K −1 mol−1 × (298 K)2
where B = [Exercise 21C.4(b)] =
hp O
6.626 × 10−34 J s × 105 Pa
= 1.54 × 1011 m 3 mol−1 s −1 = 1.54 × 1014 dm 3 mol−1 s −1.
 2.3 × 1013 dm3 mol−1 s −1 
Hence ∆‡ S 8.3145 J K −1 mol−1 ×  ln
= −1 −1
− 2
 1.54 × 10 dm mol s
14 3

= −32 J K −1 mol−1
(b) The enthalpy of activation for a bimolecular gas-phase reaction is [Topic 21C.2(a)
footnote]
∆‡H = Ea – 2RT = 30.0 kJ mol–1 – 2 × 8.3145 J mol–1 K–1 × 298 K = 25.0 kJ mol–1
(c) The Gibbs energy of activation at 298 K is
∆‡G = ∆‡H – T∆‡S = 25.0 kJ mol–1 – (298 K) × (–32×10–3 kJ K–1 mol–1)
∆‡G = +34.7 kJ mol–1
21C.6(b) Use eqn. 21C.18 to examine the effect of ionic strength on a rate constant:
log kr = log kr° + 2A|zAzB|I1/2
Hence log kr° = log kr – 2A|zAzB|I1/2 = log 1.55 – 2 × 0.509 × |1×1| × (0.0241)1/2 = 0.032,
and kr° = 1.08 dm6 mol–2 min–1 .

Solutions to problems
21C.2 log kr = log kro + 2 AzA zB I 1/2 [21C.18]
This expression suggests that we should plot log kr against I1/2 and determine zB from the
slope, since we know that |zA| = 1. We draw up the following table:

I / (mol kg–1) 0.0025 0.0037 0.0045 0.0065 0.0085


{I / (mol kg–1)}1/2 0.050 0.061 0.067 0.081 0.092
log {kr / (dm3 mol–1 s–1)} 0.021 0.049 0.064 0.072 0.100
10
These points are plotted in Figure 21C.1.

Figure 21C.1

The slope of the limiting line in Figure 21C.1 is approximately 2.5. Since this slope is equal to
2 AzA zB × (mol dm −3 )1/2 = 1.018 zA zB , we have zA zB ≈ 2.5 . But |zA| = 1, and so |zB| = 2.
Furthermore, zA and zB have the same sign because zAzB > 0. (In fact, the data refer to I– and
S2 O82− .)

21C.4 Figure 21C.2 shows that log kr is proportional to the ionic strength even when one of the
reactants is a neutral molecule.

Figure 21C.2

From the graph, the intercept at I = 0 is –0.182, so the limiting value of kr is


kr° = 10–0.182 = 0.658 dm3 mol–1 min–1
Compare the equation of the best-fit line to eqn 21C.16b:
γC γI γH O

= kr log kr ° − log=
Kγ log kr ° − log = log kr ° + log

2 2
log
γI γH O

2 2
γC ‡

γI γH O

which implies that log 2 2
= 0.145I
γC ‡

If the Debye-Hückel limiting law holds (an approximation at best), the activity coefficients of
I– and the activated complex are equal, which would imply that log γ H O = 0.145I , that is,
2 2

that the activity coefficient of a neutral molecule depends on ionic strength.

11
[H + ][A − ] 2 [H ][A ]γ ±
+ − 2

21C.6 Ka = γ± ≈
[HA]γ HA [HA]
[HA]K a
Therefore, [H + ] =
[A − ]γ ±2
[HA] [HA]
and log[H + ] = log K a + log −
− 2 log γ ± = log K a + log − + 2 AI 1/2
[A ] [A ]
Write v = kr [H + ][B] .
Then log v = log(kr [B]) + log[H + ]
[HA]
= log(kr [B]) + log + 2 AI 1/2 + log K a
[A − ]
[B][HA]K a
= log v° + 2 AI 1/2 , v° = kr .
[A − ]
That is, the logarithm of the rate should depend linearly on the square root of the ionic
strength.
v 1/ 2
log = 2 AI 1/ 2 so v= v°× 102 AI

That is, the rate depends exponentially on the square root of the ionic strength.
21C.8
We use the Eyring equation (combining eqns. 21C.9 and 21C.10) to compute the bimolecular
rate constant
kT  RT  N Aq C‡  −∆E0  ( RT ) q C‡  −∆E0 
O 2 O

= kr κ  O O O exp   ≈ exp  
h  p  q H q D2  RT  hp q H q D2
O O O
 RT 
We are to consider a variety of activated complexes, but the reactants, (H and D2) and their
partition functions do not change. Consider them first. The partition function of H is solely
translational:
1/ 2
RT  h2  RT (2π kTmH )3/ 2
= q O
= and ΛΗ  =  so q H
O

p ΛH  2π kTmH  p O h3
H O 3

We have neglected the spin degeneracy of H, which will cancel the spin degeneracy of the
activated complex. The partition function of D2 has a rotational term as well.
RT kT RkT 2 (2π kTmD2 )3/ 2
q D2 = O 3 ×
O
=
p ΛD2 σ hcBD2 2 p O h 4 cBD2
We have neglected the vibrational partition function of D2, which is very close to unity at the
temperature in question. The symmetry number σ is 2 for a homonuclear diatomic, and the
rotational constant is 30.44 cm–1. Now, the partition function of the activated complex will
have a translational piece that is the same regardless of the model:
q CO‡ = q CTO
‡ × q CR‡ × q CV‡
RT (2π kTmHD )3/2
where q C‡ =
TO 2

p O h3
Let us aggregate the model-independent factors into a single term, F where
( RT ) 2 q CTO  −∆E0  2h3 cBD2 mHD2 3/ 2  −∆E0 
= F =

exp   exp  
hp q H q D2  RT  kT (2π mH mD2 kT )  RT 
O O O 3/ 2

1/ 2
 53   −∆E0 
= h cBD2  3
3 3 3 5 
exp  =  2.71× 104 dm3 mol−1 s −1
 RT 
3
 2mH (4) p T k 
where we have taken mHD = 5mH and mD = 4mH .
2 2

Now kr = F × q C‡ × q C‡
R V

12
The number of vibrational modes in the activated complex is 3×3–6=3 for a non-linear
complex, one more for a linear complex; however, in either case, one mode is the reaction
coordinate, and is removed from the partition function. Therefore, assuming all real vibrations
to have the same wavenumber v
q CV‡ = q mode
2
(non-linear) or q mode
3
(linear)
−1
  −hcv  
where q mode = 1 − exp  kT   = 1.028
  
if the vibrational wavenumbers are 1000 cm–1 . The rotational partition function is
3/ 2 1/ 2
kT 1  kT   π 
= q ‡ R
(linear) or        ( non-linear)
C 
σ hcB σ  hc   ABC 
where the rotational constants are related to moments of inertia by

B = where I ∑ mr
2
=
4π cI
and r is the distance from an atom to a rotational axis.
(a) The first model for the activated complex is triangular, with two equal sides of
s =1.30 × 74 pm = 96 pm
and a base of
b =1.20 × 74 pm = 89 pm

The moment of inertia about the axis of the altitude of the triangle (z axis) is

I1 = 2mD (b / 2) 2 = mH b 2 so A = = 21.2 cm −1
4π cmΗ b 2
To find the other moments of inertia, we need to find the center of mass. Clearly it is in the
plane of the molecule and on the z axis; the center of mass is the position z at which
∑ mi (zi − z) = 0 = 2(2mH )(0 − z) + mH (H − z)
i
where H is the height of the triangle,
H = [s 2 − (b / 2)2 ]1/2 = 85 pm
so the center of mass is z = H / 5 .
The moment of inertia about the axis in the plane of the triangle perpendicular to the altitude
is
I 2 = 2(2mH )(H / 5)2 + mH (4H / 5)2 = (4mH / 5)H 2

=
so B = 28.3 cm −1
4π c(4mH / 5) H 2
The distance from the center of mass to the D atoms is
rD = [(H / 5)2 + (b / 2)2 ]1/2 = 48 pm
and the moment of inertia about the axis perpendicular to the plane of the triangle is
I 3 = 2(2mH )rD 2 + mH (4H / 5)2 = 2(2mH )[(H / 5)2 + (b / 2)2 ] + mH (4H / 5)2
= (4mH / 5)(s 2 + b2 ).
so
13

=C = 12.2 cm −1
4π c(4mH / 5)( s 2 + b 2 )
The rotational partition function is
1  kT   π 
3/ 2 1/ 2

=q CR‡ =
σ  hc   ABC
   
47.7

(The symmetry number σ is 2 for this model.) The vibrational partition function is
q CV‡ = q mode
2
= 1.057
So the rate constant is:
kr = F × q CR‡ × q CV‡ = 1.37 × 106 dm 3 mol−1 s −1
(b) To compute the moment of inertia, we need the center of mass. Let the terminal D atom be
at x = 0, the central D atom at x = b, and the H atom at x = b + s. The center of mass is the
position X at which
∑ mi (xi − X ) = 0 = 2mH (0 − X ) + 2mH (b − X ) + mH (s + b − X )
i

5X = 3b + s so x = (3b + s) / 5
The moment of inertia is
I = ∑ mi (xi − X )2 = 2mH X 2 + 2mH (b − X )2 + mH (s + b − X )2
i

= 3.97 × 10−47 m kg 2

and = B = 7.06 cm −1
4π cI
The rotational partition function is
kT
= q R‡ = 39.4
C σ hcB
(The symmetry number σ is 1 for this model.) The vibrational partition function is
q CV‡ = q mode
3
= 1.09
So the rate constant is
kr = F × q CR‡ × q CV‡ = 1.16 × 106 dm 3 mol−1 s −1
(c) Both models are already pretty good, coming within a factor of 3 to 4 of the experimental
result, and neither model has much room for improvement. Consider how to try to change
either model to reduce the rate constant toward the experimental value. The factor F is model-
independent. The factor q CV‡ is nearly at its minimum possible value, 1, so stiffening the
vibrational modes will have almost no effect. Only the factor q CR‡ is amenable to lowering, and
even that not by much. It would be decreased if the rotational constants were increased, which
means decreasing the moments of inertia and the bond lengths. Reducing the lengths s and b
in the models to the equilibrium bond length of H2 would only drop kr to 6.5×105 (model a) or
6.9×105 (model b) dm3 mol–1 s–1, even with a stiffening of vibrations. Reducing the HD
distance in model (a) to 80% of the H2 bond length does produce a rate constant of 4.2×
105 dm3 mol–1 s–1 (assuming stiff vibrations of 2000 cm–1); such a model is not intermediate in
structure between reactants and products, though. It appears that the rate constant is rather
insensitive to the geometry of the complex.
21C.10 Eqn. 21C.18 may be written in the form:
1 log ( kr / kr )
o

zA =
2

2A I 1/ 2
where we have used zA = zB for the cationic protein. This equation suggests that zA can be
log ( kr / kro )
determined through analysis that uses the mean value of from several
I 1/ 2
experiments over a range of various ionic strengths.

14
1 log ( kr / kr )
o

zA =
2A I 1/ 2
We draw up a table that contains data rows needed for the computation:

I 0.0100 0.0150 0.0200 0.0250 0.0300 0.0350


kr/kro 8.10 13.30 20.50 27.80 38.10 52.00
log(kr/kro) / I1/2 9.08 9.18 9.28 9.13 9.13 9.17

log ( kr / kro )
= 9.17
I 1/ 2

1 log ( kr / kr )
o
9.16
zA = = = +3.0
2A I 1/ 2 2 ( 0.509 )
We used the positive root because the protein is cationic.

21D The dynamics of molecular collisions


Answers to discussion questions
21D.2 The saddle point on the potential energy surface corresponds to the transition state of a
reaction. The saddle-point energy is the minimum energy required for reaction; it is the
minimum energy for a path on the potential energy surface that leads from reactants to
products. Because many paths on the surface between reactants and products do not pass
through the saddle point, they necessarily pass through points of greater energy, so the
activation energy can be greater than the saddle-point energy. Thus, the saddle-point energy
is a lower limit to the activation energy.
21D.4 Attractive and repulsive potential-energy surfaces are discussed in Section 21D.4(b). An
attractive surface is one whose saddle point is closer to reactants than to products, so that
the transition state occurs early in the reaction. On such a surface, trajectories in which
excess energy is translational tend to end in products whereas trajectories in which the
reactant is vibrationally excited tend not to cross the saddle point and end in products.
Conversely, on a repulsive surface, the oscillatory motion of a trajectory that has excess
vibrational energy in the reactant enhances the likelihood that the trajectory will end in
products rather than simply reflect back to reactants.

Solutions to exercises
21D.1(b) Refer to Figure 21D.20 of the main text, which shows a repulsive potential energy surface
as well as trajectories of both a successful reaction and an unsuccessful one. The trajectories
begin in the lower right, representing reactants. The successful trajectory passes through the
transition state (marked by a circle with the symbol ‡ near it). The unsuccessful trajectory is
fairly straight from the lower right through the transition state, indicating little or no
vibrational excitation in the reactant. Therefore most of its energy is in translation. That
trajectory runs up a steep portion of the surface and rolls back down the valley representing
the reactant. Without vibrational energy, it cannot go around the corner to the transition
state. The successful trajectory, conversely, is able to turn that corner only because it has a
substantial amount of energy in vibration (which is represented by side to side motion in the
valley representing reactants). That is, the reactant is relatively high in vibrational energy.
Once that successful trajectory passes through the transition state, it rolls pretty much
straight into the valley representing products, so the product is high in translational energy
and low in vibrational energy.
15
21D.2(b) The numerator of eqn. 21D.6 is

( )
1 − E / RT E = −V
(1 − e−V / RT )
∞ V 1
∫ ∫0 e dE =
P ( E )e − E / RT dE = − E / RT
− e =
0 RT E = 0 RT
Thus, if the cumulative reaction probability were a step function that vanishes at high
temperature, then the numerator would decrease with increasing temperature. (The
exponential term increases with increasing temperature, but it diminishes the expression
because of the negative sign in front of it. The 1/T factor also decreases with increasing
temperature.) The rate constant, then, would also decrease with increasing temperature. (In
fact, the rate constant would decrease with increasing temperature even faster, because the
denominator of eqn. 21D.6 would increase with increasing temperature.)
Comment: The cumulative reaction probability is more likely to be a step function in the
opposite sense, one that vanishes for energies below a threshold.
Comment: The solution to Exercise 21D.2(a) can be obtained from this solution by taking
the limit V → ∞.

Solutions to problems
21D.2 The number density of scatterers (Ns) and the path length L are the same in the two
experiments. Because
I
I = I 0 e − σ N L [Problem 21D.1] so ln = −σ N L ,
I0
σ (CH 2 F2 ) ln 0.6
we have = = 5
σ (Ar) ln 0.9
CH2F2 is a polar molecule; Ar is not. CsCl is a polar ion pair and is scattered more strongly by
the polar CH2F2.
21D.4 Refer to Figure 21D.1.

Figure 21D.1

The scattering angle is θ = π – 2α if specular reflection occurs in the collision (angle of impact
b
equal to angle of departure from the surface). For b ≤ R1 + R2, sin α =
R1 + R2 (v)
  b 
 π − 2 arcsin   b ≤ R1 + R2 (v)
θ (v) =   R1 + R2 (v) 

 0 b > R1 + R2 (v)
where R2(v) = R2e-v/v* , R1 = R2/2 = b.
 1 
(a) θ (v) = π − 2 arcsin  − v/v* 
 1 + 2e 
(Note: The restriction b ≤ R1 + R2(v) transforms into R2/2 ≤ R2/2 + R2e-v/v* , which is valid for
all v.) This function is plotted as curve (a) in Figure 21D.2.

16
Figure 21D.2

(b) The kinetic energy of approach is E = mv2/2, so


 1 
θ (E) = π − 2 arcsin  −( E/ E*)1/ 2 
where E* = m(v*)2/2 .
 1 + 2e 
This function is plotted as curve (b) in Figure 21.6.

21E Electron transfer in homogeneous systems


Answer to discussion question
21E.2 Electron tunneling plays an important role in electron transfer. From considerations in
Topic 8A, we would expect that tunneling would be more important for electrons than any
other particles that participate in chemical reactions because they are so much lighter than
even the lightest atoms or ions. Tunneling is responsible for the exponential distance
dependence of the factor Het(d)2 (eqn. 21E.4), which is directly proportional to the electron-
transfer rate constant (eqn. 21E.5). A more thorough discussion can be found in Topic
21E.2(a).
Solutions to exercises
21E.1(b) For a donor–acceptor pair separated by a constant distance, assuming that the reorganization
energy is constant, eqn. 21E.9 holds:
(∆ G O )2 ∆ r G O
ln ket = − r − + constant
4RT ∆ER 2RT
or, using molecular units rather than molar units,
(∆ G O )2 ∆ r G O
ln ket = − r − + constant
4kT ∆ER 2kT
Two sets of rate constants and reaction Gibbs energies can be used to generate two
equations (eqn. 21E.9 applied to the two sets) in two unknowns, ∆Gr and the constant.
(∆ G O )2 ∆ G O (∆ G O )2 ∆ G O
ln ket,1 + r 1 + r 1 = constant = ln ket,2 + r 2 + r 2
4kT ∆ER 2kT 4kT ∆ER 2kT
(∆ r G1O )2 − (∆ r G2O )2 k ∆ G O − ∆ r G1O
so = ln et,2 + r 2
4kT ∆ER ket,1 2kT
(∆ r G1O ) 2 − (∆ r G2O ) 2
and ∆ER =
4kT ln et,2 + 2 ( ∆ r G2O − ∆ r G1O )
k
k et,1
17
(−0.665 eV)2 − (−0.975 eV)2
∆ER = = 1.53 eV
4(1.381 × 10−23 J K −1 )(298 K) 3.33 × 106
ln − 2(0.975 − 0.665) eV
1.602 × 10−19 J eV −1 2.02 × 105
If we knew the activation Gibbs energy, we could use eqn. 21E.5 to compute Het (d) from
either rate constant, and we can compute the activation Gibbs energy from eqn. 21E.6:
(∆ G O + ∆Er ) 2 {(−0.665 + 1.53) eV}2
∆‡G = r = =0.123 eV
4∆Er 4(1.53 eV)
1/ 2
2{H et (d )}2  π3   −∆‡G 
Now ket =   exp  
h  4kT ∆ER   kT 
1/ 2 1/ 4
 hk   4kT ∆ER   ∆‡G 
H et (d ) =  et    exp  
 2   π 
3
 2kT 
1/ 2
 (6.626 × 10−34 J s)(2.02 × 105 s −1 ) 
= 
 2 
1/ 4
 4(1.53 eV)(1.602 × 10−19 J eV −1 )(1.381× 10−23 J K −1 )(298 K) 
× 
 π3 
 (0.123 eV)(1.602 × 10 J eV ) 
−19 −1
× exp  −23 −1 
 2(1.381× 10 J K )(298 K) 
so = 9.6 × 10−24 J
21E.2(b) Equation 21E.8 applies:
ln ket = –βd + constant
The slope of a plot of ket versus d is –β. The slope of a line defined by two points is:
∆y ln ket,2 − ln ket,1 ln 4.51 × 104 − ln 2.02 × 105
slope = = = −β =
∆x d2 − d1 (1.23 − 1.11) nm
so β = 12.5 nm −1
Inserting data from either rate constant allows calculation of the constant:
constant = ln 2.02 × 105 + (12.5 nm −1 )(1.11 nm) = 26.1
Taking the exponential of eqn. 21E.8 yields:
ket = e − β d + constant s −1 = e −(12.5/nm )(1.59 nm )+26.1 s −1 = 5.0 × 102 s −1

Solutions to problems
21E.2 Estimate the bimolecular rate constant kr for the reaction
2+
Ru(bpy)33+ + Fe(H 2 O)6 → Ru(bpy)32 + +Fe(H 2 O)63+
by using the approximate Marcus cross-relation:
kr ≈ (kDDkAAK)1/2
The standard cell potential for the reaction is
E O = Ered
O
(Ru(bpy)33+ ) − Ered
O
(Fe(H 2 O)63+ ) = (1.26 − 0.77)V = 0.49 V
so the equilibrium constant is
νF EO   (1)(96485 C mol−1 s −1 )(0.49 V) 
K = exp  = exp  (8.3145 J K −1 mol−1 )(298 K)  = 1.9 × 10
8

 R T 
The rate constant is approximately
kr ≈ {(4.0×108 dm3 mol–1 s–1)(4.2 dm3 mol–1 s–1)(1.9×108)}1/2
kr ≈ 5.7×108 dm3 mol–1 s–1
21E.4 Does eqn. 21E.8
ln ket = − β d + constant
apply to these data? Draw the following table and plot ln ket vs. d (Figure 21E.1):
18
d / nm ket / s–1 ln ket / s–1
0.48 1.58×1012 28.1
0.95 3.98×109 22.1
0.96 1.00×109 20.7
1.23 1.58×108 18.9
1.35 3.98×107 17.5
2.24 6.31×101 4.14

Figure 21E.1

The data fall on a good straight line, so the equation appears to apply. The least-squares linear
fit equation is:
ln ket / s = 34.7 − 13.4d / nm R 2 (correlation coefficient) = 0.991
so we identify β = 13.4 nm −1

21F Processes at electrodes


Answer to discussion question
21F.2 For electron transfer to occur at an electrode, several steps are necessary. A species in a
bulk solution phase must lose its solvating species and make its way through the electode-
solution interface to the electrode. Once there, its hydration sphere must be adjusted by the
electron transfer itself, and then the species must detach and reverse its steps as it were,
passing back through the interface into the bulk solution phase. Because there are energy
requirements associated with these steps, they are said to be activated. How the activation
Gibbs function depends on applied potentials and on the resemblance of transition state to
oxidized and reduced species is treated in Topic 21F.2(a).

Solutions to exercises
21F.1(b) The conditions are in the limit of large, positive overpotentials, so eqn 21F.5b applies:
ln j = ln j0 + (1 – α)fη
F 96845 C mol−1
where f = = = 38.9 V −1
RT (8.3145 J K −1 mol−1 ) × (298 K)
Subracting this equation from the same relationship between another set of currents and
overpotentials, we have

19
j′
ln = (1 − α ) f (η ′ − η )
j
which rearranges to
ln ( j ′ / j) ln(72 / 17.0)
η′ = η + = (105 × 10−3 V) + = 0.169 V
(1 − α ) f (1 − 0.42) × (38.9 V)−1
21F.2(b) Use eqn 21F.5a; then
j0 = j e–(1–α)ηf = (17.0 mA cm–2) × e–(1-0.42)×0.105 V×38.9/V = 1.59 mA cm–2 .
21F.3(b) In the high overpotential limit [21F.5a]
(1−α ) f η j (1−α ) f ( η1−η2 ) (1−α ) f ( η2 −η1 )
j = j0e so 1 = e and j2 = j1e .
j2
So the current density at 0.60 V
(1−0.50)×(0.60 V−0.50 V)×(38.9/V)
j2 = (1.22 mA cm −2 ) × e = 8.5 mA cm −2
about a 7-fold increase compared to the current at 0.50 V.
21F.4(b) (i) The Butler–Volmer equation is [21F.1]
j = j0 (e(1−α ) f η − e −α f η )
= (2.5 × 10−3 A cm −2 ) × (e(1−0.58)×(0.30 V)×(38.9/V) − e −0.58×(0.30 V)×(38.9/V) ) = 0.34 A cm −2
(ii) Eqn 21F.5a (also known as the Tafel equation) corresponds to the neglect of the second
exponential above, which is very small for an overpotential of 0.3 V. (Even when it was
kept, in part (a), it was negligible.) Hence
j = 0.34 A cm–2 .
The validity of the Tafel equation increases with higher overpotentials, but decreases at
lower overpotentials. A plot of j against η becomes linear (non-exponential) as η → 0.
The validity of the Tafel equation improves as the overpotential increases.
21F.5(b) The Butler–Volmer equation (21F.1]), with transfer coefficients from Table 21F.1, is
j = j (e( ) − e −α f η ) = j (e0.42 f η − e −0.58 f η )
1−α f η
0 0

Recall that η is the overpotential, defined as the working potential E′ minus the zero-current
potential E. The latter is given by the Nernst equation (6C.4):
RT 1 a(Fe 2+ ) 1 a(Fe 2+ )
E = EO − ln Q = E O − ln = 0.77 V − ln
vF f a(Fe3+ ) f a(Fe3+ )
1 a(Fe 2+ ) 1
Thus η = E ′ − 0.77 V + ln = E ′ − 0.77 V + ln r ,
f a(Fe3+ ) f
where r is the ratio of activities. Specializing to the condition that the ions have equal
activities yields
η = E′ – 0.77 V
and j = (2.5 mA cm −2 ) × (e0.42 fE ′−0.42 f ×0.77 V − e −0.58 fE ′+0.58 f ×0.77 V ) .
Evaluating the constant parts of the exponentials (with f = 38.9 V–1). and incorporating them
as numerical factors yields
j = (8.6 × 10−6 mA cm −2 ) × e0.42 fE ′ − (8.8 × 10−7 mA cm −2 )e −0.58 fE ′

21F.6(b) The current density of electrons is j0/e because each one carries a charge of magnitude e.
Look up j0 values in Table 21F.1, and recall that 1 A = 1 C s–1 .
For Cu | H2 | H+ j0 = 1.0×10–6 A cm–2
j0 1.0 × 10 A cm −2
−6
= = 6.2 × 1012 cm −2 s −1
e 1.602 × 10−19 C
For Pt | Ce4+, Ce3+ j0 = 4.0×10–5 mA cm–2
j0 4.0 × 10−5 A cm −2
= = 2.5 × 1014 cm −2 s −1
e 1.602 × 10−19 C

20
(1.0 × 10−2 m)2
There are approximately = 1.5 × 1015 atoms in each square centimeter of
(260 × 10−12 m)2
surface. The numbers of electrons per atom are therefore 4.2×10–3 s–1 and 0.17 s–1 ,
respectively.
21F.7(b) When the overpotential is small, its relation to the current density is [21F.4]
RTj j
η= =
Fj0 fj0
which implies that the current through surface area S is
I = Sj = Sj0fη .
An ohmic resistance r obeys η = Ir, and so we can identify the resistance as
η 1 1 2.57 × 10−2 Ω
r= = = = [1V = 1A Ω]
I Sj0 f 1.0 cm 2 × 38.9 V −1 × j0 ( j0 / A cm −2 )
(a) Pb | H2 | H+ j0 = 5.0×10–12 A cm–2
2.57 × 10−2 Ω
r= = 5.1 × 109 Ω = 5.1 GΩ
5.0 × 10−12
(b) Pt | Fe3+, Fe2+ j0 = 2.5×10–3 mA cm–2
2.57 × 10−2 Ω
r= = 10. Ω
2.5 × 10−3
21F.8(b) Zn can be deposited if the H+ discharge current is less than about 1 mA cm–2. The exchange
current, according to the high negative overpotential limit, is
j = j0e–αfη [21F.6a] = (0.79 mA cm–2) × e–0.5×(38.9/V)×(–0.76 V) = 2.1×106 mA cm–2
This current density is much too large to allow deposition of zinc; that is, H2 would begin
being evolved, and fast, long before zinc began to deposit.
Solutions to problems
21F.2 Deposition may occur when the potential falls to below E. (Recall that η < 0 for cathodic
processes.) E is given by the Nernst equation (6C.4):
RT
E = EO + ln a(M + )
zF
Simultaneous deposition will occur if the two potentials are the same; hence the relative
activities are given by
E O (Sn, Sn 2+ ) + RT ln a(Sn 2+ ) = E O (Pb, Pb2+ ) + RT ln a(Pb2+ )
2F 2F

( )
2+
a (Sn ) 2 F {E O (Pb, Pb 2 + ) − E O (Sn, Sn 2 + )}
=
or ln
a (Pb 2 + ) RT
= 2 × (38.9 V −1 ) × (−0.126 + 0.136) V = 0.78
That is, we require a(Sn2+) = e0.78a(Pb2+) = 2.2a(Pb2+) .
21F.4 This problem differs somewhat from the simpler one-electron transfers considered in the text.
In place of Ox + e– → Red we have here
In3+ + 3 e– → In
namely, a three-electron transfer. Therefore equations that contain the Faraday constant F (or
f, which is proportional to F) need to be modified by including the factor z (in this case 3). In
place of eqns 21F.5b and 21F.6b, we have
ln j = ln j0 + z(1 – α)fη anode
ln(–j) = ln j0 – zαfη cathode
We draw up the following table
j/(A m–2) –E/V η/V ln (j/A m–2)
0 0.388 0
0.590 0.365 0.023 –0.5276
1.438 0.350 0.038 0.3633
3.507 0.335 0.053 1.255
21
We carry out a linear regression of ln j against η with the following results (see Figure 20F.1).

Figure 20F.1

slope = z(1 – α)f = 59.42 V–1, standard deviation = 0.0154


y-intercept = ln j0 = –1.894, standard deviation = 0.0006
R = 1 almost exactly
The fit of the three data points to the Tafel equation is almost exact. Solving for α from the
slope, we obtain
59.42 V −1  59.42 V −1 
α = 1− = 1−   = 0.50
3f  3 × (38.9 V −1 ) 
which matches the usual value of α exactly.
j0 = e–1.894 A m–2 = 0.150 A m–2 .
The cathodic current density at E = –0.365 V is obtained from
ln(− jc ) = ln j0 − zα f η
= −1.894 − (3 × 0.50 × 0.023 V) × (38.9 V −1 ) = −3.26
so jc = e −3.26 = 0.038A m −2

21F.6 Start from the Butler-Volmer equation (21F.1), and expand it in powers of η:
= j j0 (e(1−α ) f η − e −α f η )
= j0 {1 + (1 − α )η f + 12 (1 − α ) 2η 2 f 2 +  − 1 + α f η − 1 α 2η 2 f 2 + }
2
= j0 {η f + 12 (η f ) 2 (1 − 2α ) + }
Average over one cycle (of period 2π/ ω):
=
j {
j0 η f + 12 (1 − 2α ) f 2 η 2 + }
ω 2 π /ω
where η = 0 , because
2π ∫ 0
cos ω t dt = 0

ω 2 π /ω
η 2 = 12 η02 , because
2π ∫ 0
cos 2 ω t dt = 1
2

Therefore, j = 1
4
(1 − 2α ) f 2 j0η02
and j = 0 when α = 12 . For the mean current,

22
I = 14 (1 − 2α ) f 2 Sj0η02
1 − 0.76
= × (1.0 cm 2 ) × (7.90 × 10−4 A cm −2 ) × (0.0389 mV −1 ) × (10 mV)2
4
= 7.2 µA

21F.8 (a) The roughly symmetrically distributed positive maximum and negative minimum suggest
a reversible one-electron transfer. Compare to Figures 21F.12 and 21F.13(b) of the main part
of the chapter as discussed in Topic 21F.3 and Example 21F.2.
(b) There are two roughly symmetrically distributed positive maxima and negative minima,
suggesting a reversible two-electron transfer brought about by sequential reversible one-
electron transfers.
(c) The shape is typical of an irreversible reduction: the positive maximum has no
corresponding negative minimum. Compare to Figure 21F.13(a) of the main part of the
chapter as discussed in Example 21F.2.
(d) Two reductions are apparent in this voltammogram, the second of which is reversible and
the first not. (The first positive maximum has no corresponding minimum; the second does.)
Compare to Figure 21F.14 of the main part of the text.

Integrated activity
21.2 Both the Marcus theory of photo-induced electron transfer (Topic 21E) and the Förster
theory of resonance energy transfer (Topic 20G) examine interactions between a molecule
excited by absorption of electromagnetic energy (the chromophore S) and another molecule
Q. They explain different mechanisms of quenching, that is, different ways that the
chromophore gets rid of extra energy after absorbing a photon through intermolecular
interactions. Another common feature of the two is that they depend on physical proximity
of S and Q: they must be close for action to be efficient.
In the Marcus theory, the rate of electron transfer depends on the reaction Gibbs energy of
electron transfer, ∆rG, and on the energy cost to S, Q, and the reaction medium of any
concomitant molecular rearrangement. The rate is enhanced when the driving force (∆rG)
and the reorganization energy are well matched.
Resonant energy transfer in the Förster mechanism is most efficient when Q can directly
absorb electromagnetic radiation from S. The oscillating dipole moment of S is induced by
the electromagnetic radiation it absorbed. It transfers the excitation energy of the radiation
to Q via a mechanism in which its oscillating dipole moment induces an oscillating dipole
moment in Q. This energy transfer can be efficient when the absorption spectrum of the
acceptor (Q) overlaps with the emission spectrum of the donor (S).

23

You might also like