You are on page 1of 87

Fuel Cell Fundamentals

Solutions

Timothy P. Holme
Ryan OHayre
Suk Won Cha
Whitney Colella
Fritz B. Prinz

Solution companion to Fuel Cell Fundamentals.


Suggested grading schemes are given after the problem numbers.

Chapter 1 solutions

Problem 1.1 (10 points) Possible answers include:


FC advantages over other power conversion devices:
1. Potentially higher efficiency.
2. Solid state components have no moving parts, giving higher reliability
and lower maintenance costs
3. silent operation
4. low emissions
5. fuel cells refuel rather than recharge, which could be faster.
Disadvantages:
1. cost
2. low power density
3. problems of hydrogen storage, production, transport, lower energy
density
4. temperature problems: PEMs cant start in the cold, high temperatures of SOFCs create materials, thermal cycling, and sealing problems
5. water management issues in PEMs.
Applications:
1. portable applications (such as laptops, cellphones, etc.) where their
fast refueling, silent operation, and independent scaling of fuel reservoir and power make fuel cells an attractive option.
2. Transportation applications where their low emissions and high efficiency makes fuel cells an attractive option.
3. Power generation applications where their silent operation, low emissions, and high efficiency make fuel cells amenable to siting in cities for
distributed generation (DG) applications, reducing the cost of power
distribution and possibly making process heat available for combined
power and heating applications.
2

Problem 1.2 (5 points) Fuel cells have lower power density than engines or batteries, but can have large fuel reservoirs, so they are much better
suited to the high capacity/long runtime applications.
Problem 1.3 (10 points) You can easily tell which reactions are reduction and which are oxidation by finding which side of the reaction the
electrons appear on.
1. Cu Cu2+ + 2e Electrons are liberated, so this is an Oxidation
reaction
2. 2H + + 2e H2 Reduction
3. O2 21 O2 + 2e Oxidation
4. CH4 + 4O2 CO2 + 2H2 O + 8e Oxidation
5. O2 + CO CO2 + 2e Oxidation
6.

1
2 O2

+ H2 O + 2e 2(OH) Reduction

7. H2 + 2(OH) 2H2 O + 2e Oxidation


Problem 1.4 (15 points) You can write full-cell reactions and then
split them into the half-cell reactions. You dont need to be an expert
chemist to do this, just use the half-cell reactions given and make sure your
equations balance with species number (ie. that O is conserved) and charge
(ie. electrons are conserved).
1. CO+ 21 O2 CO2 is a full-cell reaction common in SOFCs; the half-cell
reactions would be O2 + CO CO2 + 2e , which is an Oxidation,
or anode, reaction, and 21 O2 + 2e O2 which is a reduction, or
cathode reaction.
2.

1
2 O2 + H2

H2 O is a full cell reaction in SOFCs or PEMs, depending


on the circulating ion (O2 or H + , respectively). The half-cell reactions would be: 12 O2 + 2e + 2H + H2 O at the cathode of a PEM
and H2 2H + + 2e at the anode of a PEM; or 21 O2 + 2e O2
at the cathode of an SOFC and O2 + H2 H2 O at the anode of an
SOFC.

3. another full cell reaction in an SOFC could be CH4 + 2O2 CO2 +


2H2 O with the half cell reactions 8e + 2O2 4O2 as the reducing,
or cathode reaction, and CH4 + 4O2 CO2 + 2H2 O + 8e as the
oxidizing, or anode reaction.
3

4. to use the circulating ion (OH) , you may construct the full cell reaction 21 O2 + H2 + H2 O 2H2 O from the half-cell reactions H2 +
2(OH) 2H2 O + 2e as the oxidizing, or anode reaction, and
1

2 O2 + H2 O + 2e 2(OH) as the cathode (reducing) reaction.


Problem 1.5 (10 points) From Figure 1.6, H2 (l) has a higher volumetric
energy density but lower gravimetric energy density than H2 (g) at 7500 PSI.
On a big bus, the gravimetric energy density is probably a greater concern,
so choose H2 (g). Other considerations that could affect the choice include
safety, the amount of hydrogen lost to boil-off, and the cost of liquefaction
versus compression.
Problem 1.6 (5 points)
1. Reactant transportat high current density, there is a depletion effect.
Reactants cannot reach active sites quickly enough. The voltage loss
results from a lower concentration of reactants.(conc )
2. Electrochemical reactionvoltage loss from the sluggishness of the electrochemical reaction (act )
3. Ionic conductionresistance to ion flow in the electrolyte (ohmic )
4. Product removalin a PEM, water flooding blocks active reaction sites
(conc )
Problem 1.7 (20 points) First, multiply the whole reaction by 2 because you cant formally describe a bond in 12 O2 , but we will later divide
by 2 at the end to find the energy of the reaction given. For the reaction
2H2 + O2 2H2 O, H2 O has 2 O-H bonds, so E2H2 O = 2 2 EOH =
4 460 kJ/mol
EO2 = 494 kJ/mol
E2H2 = 2 432 kJ/mol
The energy released by the reaction is
1
1
(E2H2 O EO2 E2H2 ) = (4 460 494 2 432)
2
2

(1)

(in kJ/mol).
The energy is 241 kJ/mol . The energy is negative because an energy
input to the system is required to break bonds.
4

Problem 1.8 (20 points) You can see the benefits in a fuel cell of the
independent scaling of the fuel reservoir and the fuel cell stack. Therefore,
for this problem, you may compute the volume of the stack and the reservoir
independently.
For the stack, you need to supply 30 kW with a fuel cell that supplies
power at 1 kW/L and 500 W/kg. Therefore, the volume needed is 30 kW
1 kg
1 L
1 kW = 30 L and the weight is 30 kW 500 W = 60 kg.
For the fuel tank, you need to hold a quantity of fuel equal to
30 kJ/s

1 hr
3600 s

300 miles = 540 M J


60 miles
1 hr

(2)

note that 1 W = 1 J/s. Taking into account the 40% efficiency, you need
to hold an excess quantity of fuel, 540 M J/0.40 = 1350 M J. The hydrogen
is compressed to supply 4 M J/L and 8 M J/kg, so the fuel tank must be
1350 M J 41MLJ = 337.5 L and 1350 M J 81 MkgJ = 168.75 kg.
The entire system must occupy a volume Vsystem = Vtank + Vcell =
337.5 L+30 L = 367.5 L and weighs Wsystem = Wtank +Wcell = 168.75 kg+
60 kg = 228.75 kg
Problem 1.9 (5 points) P = V I. See the figure.

Figure 1: Sketch of Voltage and Power as a function of current density for


the fuel cell described in problem 1.9.

Chapter 2 Solutions

Problem 2.1 (6 points) When a gas undergoes a volume constriction,


possible configurations of the gas are removed. Entropy is a measure of
disorder, ie. the number of possible configurations a system can assume, so
the entropy of a gas in a smaller volume is lower (given that the temperature of the gas remains constantentropy is also a function of temperature).
Therefore, the entropy change is negative.
Problem 2.2 (6 points) G = H T S, so for an isothermal reaction
(T = 0), G = H T S.
(a) if H < 0 and S > 0 then G < 0 and the reaction is spontaneous.
(b) in this case, you cannot determine the sign of G unless you are
given the temperature and the size of the changes in entropy and enthalpy.
(c) H > 0 and S < 0, so in this case, G > 0 and the reaction is

non-spontaneous.
(d) again, you cannot make a determination from the information given
Problem 2.3 (6 points) The reaction rate is determined by the activation barrier, and not by the overall energy change of the reaction. You
may not determine which reaction proceeds faster.
Problem 2.4 (6 points) While the current scales with the amount of
reactants, the voltage does not. The thermodynamic potential comes from
the energy drop going from products to reactants, which does not scale with
reactant amount. Since E = G/nF , you can think of the scaling in n
cancelling the scaling in G.
Problem 2.5 (6 points) The Nernst equation
i
aprod
RT
E = ET
ln
i
nF
areact

(3)

shows that increasing the activity of the reactants decreases the argument
of the ln, which raises the reversible cell voltage (E) because the ln term is
negative. This, in essence, is LeChatliers principle.
Problem 2.6 (Not graded) When the reaction is in equilibrium, the
electrochemical potential of the system is zero. Components on the products side see the electrical potential P and on the reactants side see the
potential R . Note that different species do not experience different electrical potentials. Then, if is the difference in electrical potential from one
side to the other,
X
X
X
0=
i dni =
oi dni +
RT ln ai dni + nF
(4)
First, note that:
Q i
n
aprod
am
a
RT ln ai dni =
= RT ln M bN = RT ln Q i
areact
aA aB
(5)
Rearranging (4) and inserting the above result, you get
Q i
P o
aprod
i dni RT
=

ln Q i
(6)
areact
nF
nF

n
1
b
RT (ln am
M +ln aM ln aA ln aB )

G
From the thermodynamic definition of chemical potential, i n
so that
i
o
o
o
i dni = G where the in this case denotes reference concentration. We

o
o refers only to
relate this term to E by G
nF = E , but noting that the
reference concentration so the term may still depend on temperature, we
rename the quantity ET . Identifying E as the electrical potential across the
cell , we arrive at the Nernst equation (3).

Problem 2.7 (15 points) Yes, you can have a thermodynamic efficiency
greater than 1. We chose the metric of fuel cell efficiency to be G/H,
but it is in some sense an arbitrary choice.
As an example in defining efficiencies, consider the efficiency of an electrolyzer
which is a machine that makes hydrogen gas from water using electricity (this
is the exact reverse of a fuel cell, and may be used to generate hydrogen for
some fuel cell applications). An electrolyzer has an efficiency defined to be
the H of reaction (the output is the useful heat energy of hydrogen), divided by the energy input G. Therefore, the efficiency of the electrolyzer
is the inverse of fuel cell efficiencyso the fuel cell at STP with an efficiency
of 0.83, if ran in reverse as an electrolyzer, would have an efficiency of 1/0.83
which is greater than 1.
For a fuel cell, consider the following example:  G/H. For an
S
isothermal reaction, G = H T S, so  = 1 T H
. If H is going to
be negative, then you need to make S positive to get an efficiency greater
than 1. The trick is to use a solid or liquid reactant to make S positive
because solids and liquids have very low entropy compared to gases. For the
fuel cell reaction C(s) + 21 O2 CO at 298 K and 1 bar,
1
S = SCO SO2 SC = 197.70.5205.15.7 (in J/molK) = 89.45 J/molK
2
(7)
1
H = HCO HO2 HC = 110.50.500 (in kJ/mol) = 110.5 kJ/mol
2
(8)
then
S
 = 1 T H
= 1 (negative number) > 1
Problem 2.8 (15 points) Assuming constant specific heats, H(T ) =
H o + cp (T T o ) and S(T ) = S o + cp ln(T /T o ).
Find the temperature that satisfies:
X
G(T ) = 0 = H(T )T S(T ) =
[Hio + cpi (T T o ) T (Sio + cpi ln(T /T o ))]
(9)
o
o
taking out the H and S ,
X
o
o
0 = Hrxn
T Srxn
+
[cpi (T T o T ln(T /T o ))]
(10)
8

or
o
o
0 = Hrxn
T Srxn
+ (T T o T ln(T /T o ))

cpi

(11)

Substituting numbers,
0 = 41.13 kJ/molT (42.00 J/molK)+(T T o T ln(T /T o ))(3.2 J/molK)
(12)
A numerical solution (MATLAB or Excel or your graphing calculator work
fine) gives T 1020 K 747 C . The error of neglecting the dependence
of S and H on temperature led to an answer that was off by about 40
degrees in this case. A more sophisticated solution would include the variance of cp with temperature, requiring an iterative solution or an expansion
for cp in terms of T .
Problem 2.9 (a) (10 points) Remember that the effect of temperature
enters into the first term of the Nernst equation. From the Nernst equation,
if the reactants and products are ideal,
Q
Q i 



x
(P/Po )P P xi i
RT
RT

Q i = ET
E = ET
ln
ln (P/Po ) P R QP i i

nF
(P/Po ) R R xi
nF
R xi
(13)
In a reaction, the change in number of moles nG = P R . The temperature dependent term is
ET = E o +

S
(T To )
nF

(14)

For the voltages to be equal


E(T1 , P1 ) = E(T2 , P2 )
S
(T1 To )
E +
nF
S
Eo +
(T2 To ) +
nF
o

(15)

Q i 

RT1
nG QP xi
ln (P1 /Po )
i =
nF
R xi
Q i 

RT2
nG QP xi
ln (P2 /Po )
i
nF
R xi

Cancelling constant terms that appear on both sides of the equation and
solving for T2 ,
h
Q
i i
S
R
nG QP xi

ln
(P
/P
)

1
o
i
nF
nF
R xi
h
T1
(16)
Q
i i = T 2
S
R
nG QP xi

ln
(P
/P
)
2
o
nF
nF
x i
R

Simplifying for the H2 /O2 fuel cell with liquid water as product and for pure
components (xH2 = xO2 = 1) and dropping the Po which is understood to
be 1 atm the expression simplifies to
T1

S + 1.5R ln P1
= T2
S + 1.5R ln P2

(17)

(b) (5 points) At STP, using the data from Appendix B,


Srxn = SH2 O(l) SH2 0.5SO2 = 69.95130.860.5228.3 = 175.06 J/molK
(18)
If P2 = P1 /10 then from the above expression,
T2 = T1

S + 1.5R ln P1
175.06 + 1.5 8.314 ln 1
= 298
(19)
S + 1.5R ln(P1 /10)
175.06 + 1.5 8.314 ln(1/10)

Then T2 = 256 K . Note that at this temperature of 17 C, there will be


no H2 O(g) (so we shouldnt use the S for water vapor), but there will also
not be liquid waterthe product will be ice! A full solution would use the
S for liquid water down to 0 C and then the S for ice below that.
To interpret the result, since S is negative, the Nernst voltage decreases
with an increase in temperature. The decrease in pressure decreases the
Nernst voltage, so we must compensate by raising the Nernst voltage by
decreasing temperature, therefore T2 < T1 .
Problem 2.10 (15 points) Two ways of attacking this problem yield
the same result. First, you could imagine a box with water, air, and hydrogen, and the reaction H2 + 21 O2 *
) H2 O(l) is in equilibrium. To find
how much hydrogen is consumed by oxygen, find the equilibrium quantity
of hydrogen when oxygen is present. That is to say, find at what PH2 does
Grxn = 0. From the vant Hoff isotherm
o

G = G + RT ln

i
aprod
i
areact

(20)

assuming air at the cathode (xO2 = 0.21)


G = Go +RT ln
Solving for xH2 ,

a1H2 O
1
o
1.5
0.5
= Go +RT ln
1
0.5
0.5 = G RT ln(1) (xH2 )(0.21)
1.5
aH2 aO2
(P/Po ) xH2 xO
2
(21)


Go
= ln (0.21)0.5 xH2
RT
10

(22)





1
Go
(237 kJ/mol)
xH2 =
exp
= (2.18) exp
= 5.411042
0.210.5
RT
(8.314 J/mol K)(298 K)
(23)
42
atm
so PH2 = 5.41 10
Note that the partial pressure is very low, because it is very energetically
favorable for hydrogen to react with oxygen to form water.
The alternate way to solve the problem is to solve a concentration cell
where a voltage develops (1.23 V because it is a hydrogen/air system) but
E o = 0 because the concentration cell reaction is H2 + O2 H2 + O2 . In
this formulation, you solve the Nernst equation where the reactants have
activity 1 because they are pure
RT
PH2 /Po (0.21)0.5
ln
nF
1

(24)



1.23 nF
xH2 (0.21)0.5 = exp
RT

(25)

1.23 = 0
This gives the same equation

Problem 2.11 (5 points) The efficiencies of each part multiply to give


the total efficiency of the cell  = thermo V oltage f uel . For pure H2 /O2 at
STP, thermo = 0.83. We are given V , so V oltage = VE = 0.75
1.23 . We are given
1
1
, so f uel = = 1.1 . Therefore,
 = 0.83

0.75 1
1.23 1.1

 = 46%

11

(26)

Chapter 3 Solutions
Problem 3.1 A. (5 points) Reducing the potential raises the energy
of electrons in the electrode. To reduce their energy, electrons leave the
electrode, so the reaction proceeds faster in the forward direction.
B. (5 points) Increasing the potential lowers the energy of electrons in
the electrode, so the reaction is biased in the forward direction.
C. (5 points) We want to increase both reaction rates in the forward
direction. At the anode (H2 *
) H + +2e ) you want to draw electrons to the
electrode, so increase the potential. At the cathode (2H + + 2e + 21 O2 *
)
H2 O) you want electrons to leave the electrode, so reduce this potential.
The overall voltage output falls from both effects.

Figure 2: Schematic of activation voltage losses for problem 3.1


Problem 3.2 (5 points) Yes, it is possible to have a negative Galvani
Potential at one electrode. It means that one half-cell reaction requires
energy input, and the other results in energy output. So long as the total
potential adds up to the measured full-cell potential, it is impossible to know
what each half-cell potential is!
Problem 3.3 (10 points) Alpha is the charge transfer coefficient, it
describes whether the center of the reaction, or peak of the reaction activation barrier, falls nearer to one side of the reaction or the other. In
this figure, note that alpha does not change the final electrochemical energy
12

Figure 3: Schematic of different Galvani potentials problem 3.2


change, only the height of the peak in electrochemical energy.
Problem 3.4 (5 points) The exchange current density is the current
density of the forward and reverse reactions at equilibrium (at open circuit).
Problem 3.5 (5 points) (a) The Tafel equation, which holds in the
exponential regime, reads:
act = a + b log j

(27)

and in the exponential regime, the Butler-Volmer equation simplifies to


act =

RT
RT
ln jo +
ln j
nF
nF

(28)

Note: to convert between log and ln, use the conversion ln x = 2.3 log x.
The terms that go as a logarithm with current density are equal:
b log j =

RT
RT
ln j b/2.3 =
nF
nF
RT
b = 2.3 nF

(b) Identifying the constant terms, we get


RT
a = nF
ln jo

13

(29)

Figure 4: Effect of on the electrochemical energy pathway for problem


3.3.
Problem 3.6 (5 points) The full-cell reaction is CO + 21 O2 CO2 .
The half-cell reaction at the anode is O2 + CO CO2 + 2e and at the
cathode is O2 + 4e 2O2
Problem 3.7 (5 points) The main job of a fuel cell catalyst is to be
able to form intermediate strength bonds with reactants and products, ie.
to yield a low Gact . Also, it should have a long lifetime, which means
that it is resistant to poisoning, and does not migrate or agglomerate on
the membrane. The requirements for an effective fuel cell catalyst-electrode
structure are: porosity, a high degree of interconnection between the pores, a
high effective catalyst area, high TPB density, high electronic conductivity,
14

and high exchange current density. It also must have a long lifetime, meaning
high mechanical strength and resistance to corrosion. Ideally, it would also
be cheap and easy to manufacture.
Problem 3.8 (5 points) For reaction A, the net reaction rate in mol/s
2.5A/cm2
A
is nF52cm
= 1.30 105 mol/s cm2 . For reaction B, the net
2 =
2F

cm2

3A/cm
A
reaction rate is nF155cm
= 1.04 105 mol/s cm2 . Therefore,
2 =
3F
reaction A has a higher reaction rate.

Problem 3.9 (10 points) At equilibrium, j = 0. The Butler-Volmer


equation is


CR
CP
nF
(1 )nF
o
j = jo
] =0
(30)
o exp [ RT ] C o exp [
CR
RT
P
After canceling the joo term and rearranging, the equation reads

/C o
]
exp [ (1)nF
CR
( 1)nF nF
nF
R
RT
= exp [

] = exp [
]
=

o
nF

CP /CP
RT
RT
RT
exp [ RT ]
(31)
Solving for ,
o
C /CR
RT
=
ln [ R
]
(32)
nF
CP /CPo

Moving the negative sign to the right and inverting the argument of the
logarithm,
C /C o
RT
=
ln [ P Po ]
(33)
nF
CR /CR
The overvoltage, , is the amount by which the actual voltage is less than
the standard state voltage
ET E =
(34)
For an ideal gas, the concentration is the activity, so we have
aP
RT
ln [ ]
nF
aR

(35)

ai
RT
ln [ Pi ]
nF
aR

(36)

E = ET
which is almost the Nernst equation
E = ET

The difference lies in the ln term, which does not account for a multiplicity
of products and reactants. This arises from the fact that the Butler-Volmer
15

equation only accounts for the concentration of the limiting products and
reactants. If you inspect the derivation of the BV equation in the text, youll
find that it is only for one electrode reaction. Since one fuel cell reaction
is usually orders of magnitude slower than the other reactions, we typically
neglect the slugishness of fast reactions and only employ the BV equation
to find the loss at one electrode. A more rigorous form of the BV equation
would include all of the products and reactants (and youd have to write one
equation for the anode and one for the cathode) and would reduce exactly
to the Nernst equation.
Problem 3.10 A (2 points) : P = IV = (1 A)(2.5 V ) = 2.5 W
B (2 points) : By putting 5 cells together in series, you get an overall
voltage of 5(0.5 V ) = 2.5 V , the necessary voltage.
C (3 points) : Note that when you stack cells together in series, the
current does not change. However, the amount of hydrogen required does
increase. One way to think of it is that when you are required to supply
2.5 W of power, you must supply 2.5 J/s of hydrogen, not just the 0.5 J/s
per cell. What is going on is that the first fuel cell uses 0.5 J/s of hydrogen
to raise the voltage of its cathode up to 0.5 V . When this cathode connected
to the anode of the next cell it is raised to a higher voltage, and this cell
uses 0.5 J/s of hydrogen to raise its output voltage, and so on...
The amount of hydrogen per cell is determined by the current by N H2 =
s
i/nF . The electric charge needed is 1 A 100 hrs. 3600
1 hr = 360, 000 C. Converting to moles of hydrogen, 360, 000 C/2F = 1.866 mol H2 . Converting
g
to grams, 1.866 mol 1 2mol
= 3.73 g H2 per cell. So the fuel cell stack uses
five times that, 18.7 g H2
D (3 points) : From the ideal gas law, V = nRT /P = (1.866
5 mol)(0.08205 L atm/mol K)(298K)/(500 atm) = 0.457 L.
V = 457 cm3
For storage in a metal hydride, the 5 wt.% hydrogen gives a storage density of 10 g/cm3 0.05 = 0.5 g/cm3 . To store 18.7 grams, you need a volume
of (18.7 g)/(0.5 g/cm3 ) = 37.3 cm3 . This is a significant improvement over
storage at 500 atm!
Problem 3.11 (10 points) The exchange current density is given by
jo = nF

G
CR
e RT
o
CR

16

(37)

Grouping the preexponential terms into a constant that is temperature independent,


jo (T ) = Ce

G
RT

(38)

We may multiply and divide the equation by jo (To )


jo (T ) =

G 1
1
jo (To ) G
Ce RT = jo (To )e R ( To T )
jo (To )

(39)

This is our equation for jo (T ). To find G , first rearrange the general


expression
1
1
jo (T )
= G ( )
(40)
R ln
jo (To )
To T
Now, plugging in the numbers given,
G =

( T1o

1
T)

ln

jo (T )
= (8.314 J/molK)/(1/300 K1/600 K) ln [104 /108 ]
jo (To )
(41)
G = 45.9 kJ/mol

Problem 3.12 (10 points) The general expression for current density
as a function of concentration is
jo = nF

CR

f1 eG /RT
o
CR

(42)

oo ), the current density at a reference


Multiplying and dividing by jo (CR
concentration,

oo )
CR
jo (CR

jo =
nF
f1 eG /RT
(43)
oo
o
jo (CR )
CR

Canceling, we obtain the general expression


oo
jo = jo (CR
)

CR
oo
CR

(44)

) = j(T, 10C )
We must find the temperature T 0 for which j(T 0 , CR
R

nF

10CR
CR

0
G /RT
f
e
=
nF
f1 eG /RT
1
o
o
CR
CR

(45)

Canceling,
/RT

10eG

/RT 0

= eG

17

(46)

ln 10 =

G
(1/T 1/T 0 )
R

(47)

Solving for T 0 ,
1/T 0 = 1/T

1
R
ln 10 T 0 =
R

G
1/T G
ln 10

(48)

Note, the problem should have read: G = 20 kJ/mol. Using this value,
T 0 = 421 K

(49)

So the temperature change is T = 121 K .


Problem 3.13 (not graded) The Butler-Volmer equation is


G
CP
CR
nF
(1 )nF
RT
j = Ce
]
o exp [ RT ] C o exp [
CR
RT
P

(50)

Doubling the temperature or halving the activation barrier will have the
same effect on the first term in the equation. Doubling the temperature will
also effect the terms in parenthesis. Since the fuel cell is producing current
(there is a fixed overvoltage), the first term in parenthesis is larger than the
second, so changing T will have a larger impact on that term, the forward
current term. So increasing T decreases the contribution from the term in
parenthesis, reducing the current. At a fixed overvoltage, there is a larger
increase in current density from halving the activation barrier than from
doubling temperature.
Problem 3.14 (not graded) As stated in the text, the bond strength
of oxygen is about 8.8 eV , which corresponds to a temperature of T =
8.8 eV /kB = 1.0105 K = 1.0105 C. Breaking the bond on a Pt catalyst
requires 2.3 eV of energy input, which corresponds to a temperature of T =
2.3/kB = 2.7 104 K = 2.6 104 C. While the Pt catalyst helps greatly,
this is still a temperature far too large to allow thermal decomposition of
oxygen. This is why we must supply an overvoltage to the cathode, to
provide an extra driving force for oxygen dissociation.

18

Chapter 4 Solutions

Problem 4.1 (5 points) There must be a voltage gradient (an electric


field) to induce charge to flow. Another way to think about it is the intrinsic
resistance of real materials will cause flowing charge to drop in voltage.
Problem 4.2 (5 points) For a given current density, the voltage drop
depends on the ASR, ohmic = ASR j. The cell with the lower ASR will
be the cell with the lowest voltage drop. The second cell has an ASR0 =
A0 R0 = 10A R/9, or an ASR 10
9 higher than the first cell, so the losses
10
increase by a factor of 9 over the first cell.
Problem 4.3 (5 points) Conductivity is a function of carrier concentration and carrier mobility = nF cu
Problem 4.4 (5 points) Ion transport in materials is harder because
ions are much more massive than electrons, so ion < electron . Also, the
concentration of free electrons in metals is much higher than typical ion
concentrations that can be achieved in electrolytes, celectron > cion . The
two effects combine to make ionic conductivity much lower than electronic
conductivity.
Problem 4.5 (5 points) Fuel cell electrolytes must have high ionic
conductivity, low electronic conductivity, high stability under corroding environments, mechanical strength, low fuel crossover rates. Ideally, they are
easy to manufacture at low cost. The stability requirement is the hardest
to fulfill because the membrane must be stable in both the reducing environment at the anode and oxidizing environment at the cathode.
Problem 4.6 (10 points) There are two differences from the PEM
case: first, the charge carrier has the opposite sign, and second, it flows in the
opposite direction, from the cathode to the anode. These two effects combine
to make the diagram exactly the same. The ohmic loss still decreases the
output voltage of the fuel cell, and the oxygen ion is negative it flows up a
potential gradient. Note that the graph is exactly the same as the figure in
the text!
Problem 4.7 (5 points) See the figure that shows ohmic and activation
losses.

19

Figure 5: Voltage losses for problem 4.6.

Figure 6: Problem 4.7: the left panel shows only ohmic losses in the electrolyte, the right panel shows activation losses at the anode and cathode, as
well as ohmic losses at the electrolyte.

Problem 4.8 (5 points) Find the thickness of a membrane that has


an electric field of 108 V /m with a voltage of 1 V . This can be done by
dimensional analysis, or knowing that, with no net charge distribution, the
electric field will be dropped linearly across the membrane. L = V /E =
(1 V )/(108 V /m).
L = 10 nm
Problem 4.9 (10 points) We must minimize the voltage loss as a
function of electrolyte thickness. The ohmic loss is given by the current
20

L
density times the ASR (ASR = R A = A
A = L ), so ohmic = L j.
The leakage loss is given to be of the form leak = A ln jleak = A ln[B/L].
Therefore, the total loss is

T = A ln[B/L] + j

(51)

To find the optimal thickness, set the derivative with respect to L equal to
zero:
j
dT
A
=0= +
(52)
dL
L

L = A/j
To check that the result makes sense, note that as A increases, the leakage
losses increase, so the optimal thickness must increase to offset the increased
leakage; therefore L is proportional to A. As increases, ohmic losses decrease, so the membrane may be thicker to offset more leakage loss; therefore
L is proportional to . Finally, as current density is increased, ohmic losses
increase, so the optimal membrane thickness should decrease to minimize
ohmic loss, so L 1/j.
Problem 4.10 (5 points) The two resistances simply add, since they
are resistors in series. Therefore, the ohmic loss is
ohmic = j(ASRelec + ASRelectrolyte )

(53)

ohmic = (0.5 A/cm2 )[(0.01 )(5 cm2 ) + (100 m)/(0.1 ( cm)1 )] (54)
ohmic = 75 mV
Problem 4.11 Starting with the equation for conductivity,
=

c(nF )2 D
RT

(55)

insert the equation for diffusivity,


D = Do eGact /RT

(56)

and you get


=

c(nF )2 Do eGact /RT


RT
21

(57)

Multiply each side by T and group the temperature independent preexponential factor into a constant o , and the equation looks very close to the
final result:
T =

c(nF )2 Do Gact /RT


e
= o eGact /RT
R

(58)

The last step requires a little concentration to keep your units straight.
If Ea is expressed in units of eV /mole, and Gact is in units of J/mole, and
as stated in the text, Ea = Gact /F , then we can convert by noticing some
relations between the fundamental constants: R = kNA and F = qNA .
Gact = Ea F = Ea (qNA ) = Ea (q

R
)
k

(59)

The argument of the exponential in equation 58 becomes:


Ea (q R
Ea q
Gact
k)
=
=
RT
RT
kT

(60)

The definition of an electron volt is the energy given to an electron as it is


raised through a potential of one volt, so 1 eV = q Joules. The factor of q
in equation 60 is the conversion from electron volts to Joules. So with the
understanding that Ea is to be expressed in eV /mol, equation 58 becomes
T = o eEa /kT

(61)

Problem 4.12 Well need the saturation water pressure at 70 C and


90
The empirical fit for psat is
C.

log10 psat = 2.1794 + 0.02953T 9.1837 105 T 2 + 1.4454 107 T 3 (62)


Plugging in T = 70 C,
psat (70 C) = 102.1794+0.02953(70)9.183710

5 (70)2 +1.4454107 (70)3

= 100.5127 = 0.307 bar


(63)

and for T = 90 C,
psat (90 C) = 102.1794+0.02953(90)9.183710

5 (90)2 +1.4454107 (90)3

= 100.1602 = 0.692 bar


(64)
To find the humidity of the exit stream, we need to track the water inflows
and outflows, given some information about the proton flow and inlet humidity. First, it is simple to find the hydrogen fluxes. The inlet hydrogen is
22

Figure 7: Problem 4.12: Atom fluxes into the gas channel, into the GDL,
and out of the gas channel must sum so that there is no net accumulation
of any species in the gas channel.

H2 ,in = I/nF = (8 A)/(2 96485 C/mol) = 4.15 105 mol/s. The fuel cell
current is I = jA = (8 cm2 )(0.8 A/cm2 ) = 6.4 A. The number of hydrogen
molecules (that split into protons to cross the membrane) leaving the gas
channel towards the membrane is
H2 ,memb = I/(nF ) = (6.4 A)/(2 96485 C/mol) = 3.32 105 mol/s (65)
The amount of hydrogen that flows out of the gas channel is 8 A 6.4 A =
1.6 A or in mol/s, H2 ,out = I/nF = 8.29 106 mol/s.
To find the water fluxes is slightly more complicated. From the drag
coefficient, the amount of water crossing the membrane is H2 O,memb =
H2 ,memb = 0.8 3.32 105 mol/s = 2.65 105 mol/s. The inlet
w
humidity tells us that the influx of water is ppsat
= 0.80. Assuming that the
inlet pressure is 1 bar, the mole fraction of water at the inlet is
yw = 0.8

psat
0.692 bar
= 0.8
= 0.554
Po
1 bar

(66)

by knowing the inlet flow of hydrogen, we can determine the inlet flow of
23

water

0.554
= 5.14 105 mol/s
(67)
1 0.554
Knowing the inlet water and the amount that crosses through the membrane,
we can find the outlet water flow
H2 O,in = H2 ,in

H2 O,out = H2 O,in H2 O,memb = (5.14105 mol/s)(2.65105 mol/s) = 2.49105 mol/s


(68)
Finally, we can find the activity of water in the exhaust
aw =

H2 O,out
yw Pout
Pout
pw

=
=
psat
psat (Tout )
H2 O,out + H2 ,out psat (70 C)

(69)

Assuming that the pressure drop along the gas channel is negligible,

aw =

2.49 105 mol/s


1 bar

= 2.44
5
6
2.49 10 mol/s + 8.29 10 mol/s 0.307 bar

(70)

Since the activity is greater than 1, we have liquid water in the exhaust.
Problem 4.13 This problem is quite similar to example 4.4. Following
the same method, we first find the water content of the membrane by solving
an ODE subject to boundary conditions, then find the conductivity of the
membrane as a function of temperature and water content, and finally find
the overpotential by Ohms law.
(a) First, the boundary conditions on are, from equation 4.34:
(aw = 1) = 14; (aw = 0.5) = 3.45

(71)

Assuming the current of water through the membrane to be of the form


j
JH2 O = 2F
where is an unknown constant, we can rearrange equation
4.44 to find an equation for the variation of water content through the
membrane



d
j
Mm
=
2ndrag

(72)
dz
2F dry D ()
22
To find D (), we again assume that, since D is a slowly varying function
of , it is a constant over the range we consider. Now, we may analytically
solve the differential equation
2ndrag
d
= k(
)
dz
22
24

(73)

Where k

jMm
2F dry D .

The solution is

(z) = C exp(k

ndrag
11
z) +
= Cez + /
11
kndrag

(74)

Where kndrag /11. We determine C from the boundary conditions. If


the anode is at z = 0, then for part (a),

= 14 C = 14

tm
(tm ) = Ce
+ = 3.45

1
tm
+ = 3.45 [1 etm ] = 3.45 14etm
[14 ]e

3.45 14etm
=
1 etm
(0) = C +

(75)
(76)
(77)
(78)

For lambda, we find


(z) = (14

3.45 14etm z 3.45 14etm


)e +
1 etm
1 etm

(79)

Then the conductivity of the membrane is found from


(T, ) = (0.005193 0.00326) exp [1268 (1/303 1/T )]

(80)

and the resistivity of the membrane is


Z tm
Z tm
dz
dz
[1268(1/3031/T )]
=e
R=
z + /) 0.00326)

(0.005193(Ce
0
0
(81)
The integral works out to
Z
dz
z
= ln(b + aez )/b
(82)
z
ae + b
b
so the membrane resistance is


e[1268(1/3031/T )]
1 0.005193 0.00326 + 0.005193Cetm
tm ln
R(T ) =
0.005193 0.00326

0.005193 0.00326 + 0.005193C


(83)
To estimate our parameters, we calculate D with = 10 to be
D = e2416(1/3031/T ) (2.5630.33+0.02642 0.0006713 )106 = 3.81106 cm2 /s
(84)
25


To find we also need ndrag = 2.5 22
= 1.14, dry = 1970 kg/m3 = 1.97
3
3
10 kg/cm , Mm = 1 kg/mol

ndrag jMm
11
1.14
=
= 71.3 cm1
11 2F dry D
11 2 96485 1.97 103 3.81 106
(85)
Then we can find alpha to be

3.45 14etm
= 1.52 103 cm1
1 etm

(86)

so that C is
C = 14 / = 7.33

(87)

R(T ) = e[1268(1/3031/T )] .319

(88)

Finally, R(T ) is
R(80 C) = 0.176 cm2 so the ohmic loss is
R = 0.176 V
(b) In this case, the boundary conditions on change

3.45 = C +
14 = Cetm

(89)
(90)

These equations may be solved for alpha


=

14 3.45etm
= 277 cm1
1 etm

(91)

using our previous value for . This gives C = 7.33, so that R(T) is
R(T ) = e[1268(1/3031/T )] .398

(92)

R(80 C) = 0.220 cm2 so the ohmic loss is


R = 0.220 V
This ohmic loss is higher than in part (a), so we can see that humidifying
the anode is more effective under these conditions.
Problem 4.14 (15 points) (a) The equation for diffusivity is
Gact
1
D = vo (x)2 e RT
2

26

(93)

I didnt subtract points for assuming that the hop distance was the lattice constant, 5
A. A more sophisticated analysis takes the crystal structure of cubic ZrO2 into account. Zirconia, in the cubic form, has the fluorite structure. The oxygen atoms occupy tetrahedral sites in an FCC
lattice. Those sites are 1/4 of the way along the cube body diagonal,
or have positions 14 (1, 1, 1). The vector between nearest neighbor sites is
1
1
1
4 (3, 1, 1) 4 (1, 1, 1) = 4 (2, 0, 0). Therefore, the distance between oxygen
nearest neighbor sites is a/2 = 2.5
A.
1

D = (1013 s1 )(2.5 108 cm)2 e (8.314


2

100 kJ/mol
J/molK)(1273 K)

(94)

D = 2.46 107 cm2 /s


(b) The vacancy fraction is
xV ' eHv /2kT = 0.0105

(95)

the density of oxygen sites is


cO =

8
= 6.4 1022 atoms/cm3 = 0.106 mol/cm3
(5
A)3

(96)

The vacancy concentration is therefore


cV = 1.113 mol/cm3
(c) Note: n = 2 since the charge of a vacancy has the opposite charge of
the ion (O2 ). The intrinsic conductivity is
=

(nF )2 cD
(2 96485 C/mol)2 (1.11 103 mol/cm3 )(2.46 107 cm2 /s)
=
RT
(8.314 J/mol K)(1273 K)
(97)
= 9.65 104 ( cm)1

Problem 4.15 (20 points) First, calculate the carrier concentration,


which requires a bit of crystal stoichiometry. The crystal is (ZrO2 ).92 (Y2 O3 ).08 .
If we let the symbol stand for a generic cation atom, we have 1.08 O2.08 . If
this crystal were pure zirconia, with a ratio of one Zr to two O, it would be
Zr1.08 O2.16 . Therefore, the 8% doping introduced an absence of 0.08 oxygens
0.08
for every 2.16 oxygen sites, or the fractional vacancy level of 2.16
= 3.70%.
27

This extrinsic doping changes the energetics of intrinsic vacancies, so we


must neglect intrinsic vacancies. At reasonable temperatures, the extrinsic
doping level will dominate. The concentration of oxygen sites is
cO =

8
= 6.4 1022 atoms/cm3 = 0.106 mol/cm3
(5
A)3

(98)

So the concentration of vacancies is


cV = (0.0370)(0.106 mol/cm3 ) = 3.93 103 mol/cm3

(99)

Now we may set up two equations with the two unknowns of Do and Gact :
(T ) =

(nF )2 cDo eGact /RT


L
=
AR(T )
RT

(100)

Solving for Do , and inserting values for T = 700 K


Do =

L
R(700 K)(e+Gact /(8.314
A(47.7 )
(nF )2 c

J/molK)(700 K) )

(101)

Inserting this equation for Do when T = 1000 K, many things cancel and
we are left with:
(700 K)(e+Gact /(8.314 J/molK)(700
(47.7 )

K) )

(1000 K)(e+Gact /(8.314 J/molK)(1000


(0.680 )
(102)

K) )

Solving for Gact ,


Gact =

1
1000

1
700

ln

700 K 0.680
1000 K 47.7

(103)

Gact = 89.4 kJ/mol


Now, we may solve for Do from equation 101:
Do =

100 m
(8.314 K/mol K)(700 K)(e+(89.4 kJ/mol)/(8.314 J/molK)(700
(1 cm2 )(47.7 )
(2 96485 C/mol)2 (3.93 103 mol/cm3 )
(104)
Do = 3.90 102 cm2 /s

To check this result, you may check the diffusivity at a normal operating
temperature and see that it compares to standard values: at T = 1000 K,
we get D = Do eG/RT 8 107 cm2 /s, which is in the right ballpark for
YSZ.
28

K) )

Chapter 5 Solutions
Problem 5.1 (20 points) The use of a different gas in the mixture
will change the effective diffusion constant of oxygen, which changes the
limiting current density jL . Intuitively, since helium is much lighter than
nitrogen (MHe  MN2 ), then DO2 He  DO2 N2 . Therefore, jL,He > jL,N2
so conc,He < conc,N2 . More rigorously:
jL = nF Def f

coR

(105)

To have a higher jL , a higher Def f is desired. The effective binary diffusion


constant is related to gas properties by:
s
s
(pci pcj )1/3 (Tci Tcj )5/12
1
1
1
1
+
= (pci pcj )1/3 (Tci Tcj )5/12b/2
+
Dij
b/2
Mi Mj
Mi Mj
(Tci Tcj )
(106)
The relevant properties are Tc,N2 = 126.2 K, Tc,He = 5.2 K, pc,N2 = 33.5 atm, pc,He =
2.24 atm, MN2 = 28, MHe = 4, MO2 = 32, b 1.823. Therefore, to compare
diffusion coefficients,
q
1
1
1/3 (T
5/12b/2
(p
)
)
c,N
c,N
2
2
MN2 + MO2
DO2 N2
q
=
(107)
1
DO2 He
(pc,He )1/3 (Tc,He )5/12b/2
+ 1
MHe

DO2 N2
DO2 He

M O2

q
1
1
+ 32
(33.5)1/3 (126.2)5/121.823/2 28
0.076
q

=
0.307
1
1
(2.24)1/3 (5.2)5/121.823/2 4 + 32

(108)

Therefore, the effective binary diffusion constant of oxygen in helium is


larger, so the limiting current will be larger in synthetic air. Therefore, the
concentration losses are larger in real air than in synthetic air.
Problem 5.2 (10 points) Most importantly, SOFCs are not plagued
with the problems of water removal that PEMs are. SOFCs operate at high
enough temperature that liquid water is not formed and does not have to
be removed. Second, diffusion is faster at higher temperatures (D T b ) so
temperature aids mass transport and not as much engineering must go into
the flow structures.

29

Problem 5.3 (15 points) The limiting current density is set by the
diffusion constant, the diffusion layer thickness, and the bulk concentration.
jL = nF Def f

coR

(109)

You may ensure high coR by designing flow structures that evenly distribute
reactants. You may decrease the diffusion layer thickness by optimizing
electrode structure (though reaction kinetics places other constraints on the
MEA). Finally, you may increase Def f by optimizing fuel cell operating
conditions such as temperature and inlet gas mole fractions (again, there
will be other constraints at work).
Problem 5.4 (25 points) To find the limiting current density, we need
to calculate the diffusion constant and the bulk concentration of oxygen.
First, the diffusion constant is
!b
0.5

T
1
a
1
ef f
1.5
1/3
5/12
p
+
DO2 N2 = 
(pc,O2 pc,N2 ) (Tc,O2 Tc,N2 )
p
MN 2
MO 2
Tc,O2 Tc,N2
(110)
The relevant properties are a = 2.745104 , b = 1.823, Tc,O2 = 154.4 K, Tc,N2 =
126.2 K, pc,O2 = 49.7 atm, pc,N2 = 33.5 atm, MO2 = 32, MN2 = 28. Plugging
in,
!1.823
4
298 K
ef f
1.5 2.745 10
p
DO2 N2 =0.4
1 atm
(154.4 K)(126.2 K)


1 0.5
1
1/3
5/12
+
((49.7 atm)(33.5 atm)) ((154.4 K)(126.2 K))
28 32
So Def f = 0.0520 cm2 /s. Finding the bulk concentration of oxygen is easier.
Neglecting the consumption of oxygen as air travels down the flow path, we
can say that the concentration of oxygen in air is 21%. From the ideal gas
P
law, we find the concentration in bulk is coR = 0.21 N
V = 0.21 RT =
1 atm
3 mol/L = 8.59 106 mol/cm3 .
0.21 (0.08205 Latm/molK)(298
K) = 8.59 10
Note that n = 4 because the species we are considering is oxygen, which
transfers 4 electrons per mole. Finally, the limiting current density is
jL = (4 96485 C/mol)(0.0520 cm2 /s)

8.59 106 mol/cm3


500 104 cm

jL = 3.44 A/cm2
30

(111)

Problem 5.5 (25 points) Using equation 5.23,


conc = c ln

jL
jL j

(112)

the following plot was generated.

Figure 8: Concentration losses as a function of current density for various


values of c for problem 5.5
As you can tell from the plot, for higher fuel cell performance, a lower
value of c is desired.
Problem 5.6 (optional) Using the power law to find the viscosity of
gases,
N2 (T = 800 C) = 16.63 106 kg/m s(

1073 0.67
)
= 4.16 105 kg/m s
273

1073 0.69
)
= 4.93 105 kg/m s
273
1073 1.15
H2 O (T = 800 C) = 11.2 106 kg/m s(
)
= 4.06 105 kg/m s
350
Using the molecular weights, viscosity, and mole fractions of each gas, we
find from equation 5.34,
O2 (T = 800 C) = 19.19 106 kg/m s(

31

Species i

Species j

Mi /Mj

i /j

ij

xj ij

xj ij

N2

N2
O2
H2 O

1
0.876
1.555

1
0.844
1.02

1
0.919
1.001

0.711
0.174
0.100

0.985

N2
O2
H2 O

1.142
1
1.776

1.19
1
1.21

1.094
1
1.093

0.778
0.189
0.109

1.08

O2

H2 O

N2
0.643
0.976 0.980 0.697
0.563
0.824 0.902 0.171 0.967
O2
H2 O
1
1
1
0.1
Equation 5.33 gives the mixture viscosity mix = 4.29 105 kg/m s.
The molecular weight of the mixture is Mmix = 27.77 g/mol. The density
is obtained from the ideal gas law:
=

p
101325 P a
=
= 0.315 kg/m3
RT /Mmix
(8.314 J/mol K)(1073 K)/(0.02777 kg/mol)

In a circular pipe, laminar flow holds for Re 2000, so


Vmax =

Remix
2000 4.29 105 kg/m s
=
= 272 m/s
L
(0.315 kg/m3 )(0.001 m)

(113)

This very fast flow will not be achieved in SOFCs, so flow will always be laminar. Laminar flow allows a higher velocity in the case of higher temperature
because the density of the gas is diminished at higher temperature.
Problem 5.7 From the current required, 1 A/cm2 , you can find the
amount of reactant required: JO2 /A = j/nF , where JO2 is the flux of oxygen
and A is the area.
The limit of laminar flow found in example 5.1 is at the velocity vmax =
38.03 m/s. The mole fraction of oxygen carried in the air is xO2 = 0.168.
The amount of reactant supplied is therefore JO2 = xO2 vmax where vmax is
in units of mol/s.
j
vmax xO2
=
(114)
A
nF
To convert the given velocity from m/s to mol/s we use the factor
where is the density in mol/m3 and is the cross sectional area of the gas
32

channel. Then v = v. To find the area that can be supplied with that
amount of reactant, we get
A = xO2 vmax

nF
j

(115)

We are told that the stoichiometric number is 2, and the geometry


from Example 5.1 has circular cross sections of diameter 1 mm, so that
= 0.00785 cm2 . In addition, = 0.921 kg/m3 /(0.02669 kg/mol) =
34.51 mol/m3 . Therefore,
A = (0.168)(3803 cm/s)(0.00785 cm2 )(34.51106 mol/cm3 )

2 96485 C/mol
1 A/cm2
(116)

A = 34 cm2
For a channel of 1 mm to cover an area of 34 cm2 , it would have to be
3.4 m long! Such a long, thin channel will lead to large pumping losses, so
channels are not usually designed like this. Therefore, channels remain in
the laminar flow regime to very good approximation.
Problem 5.8 The equation developed to describe the oxygen content
as a function of channel length was:
!
j
HC
HE
X
O2 (x = X) = O2 (x = 0) MO2
+ ef f +
(117)
4F ShF DO2
uin HC
D
O2

First, the diffusion constant of O2 in H2 O (neglecting multicomponent diffusion with N2 ) is:


DO2 H2 O

a
=
p

T
p
Tc,O2 Tc,H2 O

DO2 H2 O

3.64 104
=
1

!b
1/3

(pc,O2 pc,H2 O )

5/12

(Tc,O2 Tc,H2 O )

1
+
MH2 O
M O2
(118)
1

0.5

!2.334

353

1/3

(154.4)(647.3)

((49.7)(217.5))

((154.4)(647.3))

(119)
Or DO2 ,H2 O = 0.371 cm2 /s. To modify for the effective diffusion constant,
ef f
take into account the porosity: DO
= 1.5 DO2 H2 O = 0.0940 cm2 /s. The
2 H2 O
remainder of the values are specified. We take the inlet density of oxygen to
be half of that in the example from section 5.3.3, because that example uses
33

5/12

1
1
+
18.02 32

0.5

Figure 9: Oxygen density along a fuel cell flow channel for problem 5.8
p = 2 atm (O2 (x = 0) = 1.2 kg/m3 ). Therefore, a plot of oxygen density
along the flow channel can be constructed.
The only tricky part is to make sure you use consistent units. Make sure
that distances are in cm or m, and if is in kg/m2 , then MO2 should be in
kg/mol, not g/mol or kg/kmol.
Problem 5.9 The current in the fuel cell creates water at a flux rate of
j(x)
JH2 O (x, y = C) = MH2 O
2F

(120)

The diffusion from the GDL to the gas channel is (note: water is diffusing
in the opposite direction that oxygen diffuses, the positions of the E and
C are switched in this equation compared to the equation in the book)
dif f
ef f H2 O (x, y = E) H2 O (x, y = C)
JH
(x, y = E) = DH
2O
2O
HE

(121)

The convective transport from the gas channel to the GDL is given by (you
can check that the direction is correct because a positive flux results when
the density in the channel is larger than the density at the electrode)
conv
JH
(x, y = E) = hm [H2 O (x, y = E) H2 O (X, y = channel)]
2O

34

(122)

dif f
From the steady state conditions JH
(x, y = E) = JH2 O (x, y = C) =
2O
conv (x, y = E), we get the first relation
JH
2O
ef f
DH
2O

j(x)
H2 O (x, y = E) H2 O (x, y = C)
= MH2 O
HE
2F

(123)

that we may solve for the desired quantity, the density of water at the
catalyst layer,
H2 O (x, y = C) = H2 O (x, y = E) +

HE
ef f
DH
2O

MH2 O

j(x)
2F

(124)

From the second equality in the steady state condition, we obtain the second
relation
hm [H2 O (x, y = E) H2 O (X, y = channel)] = MH2 O

j(x)
2F

(125)

We may make the substitution for hm with the Sherwood number and isolate
the density of water diffusing into the GDL:
H2 O (x, y = E) = H2 O (X, y = channel)

HC
j(x)
MH2 O
ShF DH2 O
2F

(126)

Plugging this term into equation 124,


H2 O (x, y = C) = H2 O (X, y = channel)

j(x)
j(x)
HC
HE
MH2 O
+ ef f MH2 O
ShF DH2 O
2F D
2F
H2 O
(127)

Factoring out a term,


HC
j(x) HE
[ ef f
]
2F D
ShF DH2 O
H2 O
(128)
The last step is to take care of the x dependence. Wed like the equation
to be in terms of something we can set or measure, like H2 O (0, y = channel),
not H2 O (X, y = channel). Since the flow streams will likely be humidified,
there will be water flowing into the gas channel at a speed uin . Integrating
the flux of water from the GDL to the flow channel over the length of the
gas channel is the difference between inlet and outlet water concentration
in the gas channel (note the sign reversal):
Z X
[JH2 O (x, y = E)]dx = uin HC H2 O (0, y = channel)+uout HC H2 O (X, y = channel)
H2 O (x, y = C) = H2 O (X, y = channel) + MH2 O

(129)
35

Alternatively, we can calculate the change in water concentration by integrating the water produced by current in the fuel cell.
Z
0

MH2 O
[JH2 O (x, y = E)]dx =
2F

j(x)dx

(130)

These two equations must be equal. Assuming that the velocity of the gas
does not appreciably change in the channel (uin = uout )
MH2 O
2F

j(x)dx = uin HC [
H2 O (X, y = channel) H2 O (0, y = channel)]
0

(131)
Again, we make the simplifying assumption that j(x) is a constant, the
above equation becomes
X

MH2 O j
= uin HC [
H2 O (X, y = channel) H2 O (0, y = channel)] (132)
2F

Solving for H2 O (X, y = channel) and plugging into equation 128, we get
the final result:
H2 O (x, y = C) = H2 O (0, y = channel)+

HC
X MH2 O j
j HE
+MH2 O
[ ef f
]
uin HC 2F
2F D
ShF DH2 O
H2 O
(133)

or
j HE
HC
X
[ ef f
+
]
2F D
ShF DH2 O uin HC
H2 O
(134)
To check the signs, note that water concentration increases with X, as it
should. Increasing the inflow of water by increasing uin increases the water
that reaches the catalyst. Decreasing the diffusion layer thickness (HE )
increases the amount of water swept away by the gas, so the density of
water at the catalyst goes down, as it should. As the last check, if the
amount of gas is fixed and HC , the channel size, is decreased, the density of
water increases.
H2 O (x, y = C) = H2 O (0, y = channel)+MH2 O

Problem 5.10 Mass concentration effects become important at high


current density, so we will use the Tafel approximation to the Butler Volmer
equation
cR nF /RT
j = j00 0
e
(135)
cR
36

If we assume that voltage, and therefore overpotential is constant along the


flow channel, the only part that varies in x is the concentration of oxygen. Via a control volume analysis, you can find that the ODE governing
concentration along the flow channel is
dc(x)
j 0 enF /RT
= 0 0
c(x)
dx
cR

(136)

That is, the amount of oxygen flowing into the catalyst layer from the flow
channel is proportional to the concentration in the flow channel. The solution is a decaying exponential
c(x) = Aeax
where
a=

(137)

j00 enF /RT


c0
R

(138)

Applying the boundary condition at the inlet gives the constant A


c(0) = cin = A

(139)

Plugging this into equation 5.62, we get


O2 |x=X,y=C

MO2
= O2 |x=0,y=channel
4F

= O2 |x=0,y=channel

MO2
4F

a
j(X) HE j(X)
+
+
ef
f
hm
uin HC
DO2
j(X) HE j(X)
a
+
+
ef
f
hm
uin HC
DO2

Z
0

!
cO2 (x)dx

(140)
!
Z X
cin eax dx
0

(141)
MO2
= O2 |x=0,y=channel
4F

= O2 |x=0,y=channel

MO2
4F

!
j(X) HE j(X)
acin
eax X
+
+
(
| )
ef f
hm
uin HC
a 0
DO
2
(142)
!
j(X) HE j(X)
cin
+
+
(1 eaX )
hm
uin HC
Def f
O2

(143)
MO2
M O2
= O2 |x=0,y=channel [1
(1 eaX )]
4F uin HC
4F

37

j(X) HE j(X)
+
ef f
hm
DO
2
(144)

This says that the oxygen falloff in the channel is exponential rather than
linear. The falloff is faster for larger overpotential or higher exchange current
(faster kinetics meaning faster use of reactant).

38

Chapter 6 Solutions
Problem 6.1 (4 points each) 1. High resistance will lead to a steep
linear section of the curve, this is exhibited in (e)
2. A leakage current will shift the curve left, as in (b)
3. Poor kinetics appear as large activation losses, (c)
4. Low ohmic resistance will show up as a near-horizontal portion of the
linear section of the curve, (a)
5. Reactant starvation leads to large concentration losses (d)
Problem 6.2 (10 points) We intended for students to only consider
voltage efficiency. In that case, while both curves show the same VOC and
the same short circuit current, the power (and efficiency) of the SOFC is
higher because the curve lies at higher voltage for every current level.
Another key difference between the SOFC and PEM is that the SOFC
will be operated at much higher temperature. If you consider overall efficiency, then lets find the effect of operating at higher temperature:  =
thermo voltage f uel and assume that fuel efficiency is roughly the same for
each case, lets see how thermodynamic efficiency changes:
G V
nF G V
E V
V
V
P
=

= H = H =
o + T
H E
nF H E
E
(H
P R cp,i )/ nF
nF
nF
(145)
Remember that H o < 0. Well assume constant specific heats, so that
term at 298 K is
X
cp,i = cp,H2 O cp,H2 0.5xp,O2 = 33.628.8.528.9 = 9.7 cp (146)


P R

The effect of higher temperature (larger T ) is to make the denominator


larger, making efficiency lower. To get a quantitative feel for the size of the
effect, lets assume the SOFC is operating at 1000 K and the PEM at 298
K.
SOF C
VSOF C
=

P EM
VP EM
=

H o
nF
H o +T cp
nF

VSOF C
285 kJ/mol

VP EM 241 kJ/mol 9.7 J/mol K (1000 K 298 K)

VSOF C
1.15
VP EM

Remember that we use values for liquid water for the PEM and water vapor
for the SOFC. That means that HP EM > HSOF C , even though the
39

SOFC is operated at higher temperature. Then both effects combine to


make the SOFC higher efficiency.
Problem 6.3 (15 points) (a) 1. If electrode pores shrink or the electrode becomes thicker, the catalytically active area increases because there
are more possible reaction sites.
2. If the ionic resistance increases, active sites further away from the
electrode are less effective because ions will have to be transported through
thicker resistive material. The catalytically active area therefore decreases.
3. Similar to above, if the resistance increases, areas further from the
current collector become less active, so active area decreases.
4. The charge transfer resistance is a measure of how effective the catalyst is. If the entire catalyst becomes less active, the area does remains
unaffected, though performance decreases.
(b) At high operating temperatures, SOFCs are generally less affected
by mass transport and charge transfer, but they generally have low ionic
conductivity. Therefore, the extent to which the electrochemically active
area extends into the electrode is limited by the ionic conductivity of the
electrode.
(c) PEMs are more commonly limited by charge transfer resistance, while
ionic conductivity is generally high. Therefore increasing the thickness of
catalyst layer increases the electrochemically active area.
Problem 6.4 (10 points) The leakage current is larger than the exchange current density, so we should first assume Tafel kinetics, then check
the validity of the assumption. Neglecting ohmic and concentration losses
at this low current, the voltage loss from leakage due to activation losses is
=

RT
j
ln
= 59.1 mV
nF
j0

(147)

Using this value for in the full Butler Volmer equation gives a 1% error in
the current, so this is sufficiently accurate.
Problem 6.5 1. (5 points) Starting from the Nernst equation
aPi
S
RT
E=E +
(T To )
ln
nF
nF
aRi
o

(148)

The problem does not state whether liquid water or water vapor is the
product. This solution is worked for liquid water, but may be easily solved

40

for water vapor product using aH2 O(g) 6= 1. The thermodynamic voltage is
#
"
1
1
o
(149)
E=E +
S(T To ) RT ln[ 1 0.5 ]
nF
pH2 pO2


1
0.5 P 1.5
E=E +
S(T To ) + RT ln[xH2 xO2 ( ) ]
nF
Po
o

(150)

To find S, we need the entropy of liquid water at 330 K, which can be


found from the appendix as SH2 O(l) = 77.57 J/mol K. Then the entropy
change of reaction is
1
Srxn = SH2 O(l) SO2 SH2 = 77.570.5209.02133.60 = 160.04 J/molK
2
(151)
Inserting values,


1
0.5 1 1.5
160(330 298) + 8.314 330 ln[1 0.21 ( ) ]
E = 1.23 +
2 96485
1
(152)
E = 1.19 V
2. (5 points) In problem set 3 we found the Tafel constants in the Tafel
equation
act = ac + bc ln j
(153)
The constants are
ac =

RT
8.314 330
ln jo =
ln 103
nF
0.5 2 96485

(154)

ac = 0.196 V
We may use this constant without modification since the conversion between
ln and log was not used. The other constant did use the conversion factor
of 2.3, which may be dropped so that
bc =

RT
8.314 330
=
nF
0.5 2 96485

(155)

bc = 0.0284 V
3. (5 points) The ASR is given by
ASR = L/ = 0.01/0.1
41

(156)

ASR = 0.1 cm2


4. (5 points) First, the diffusion coefficient is
!b

0.5
1
1
p
DO2 N2
(pc,O2 pc,N2 ) (Tc,O2 Tc,N2 )
+
MN 2
MO 2
Tc,O2 Tc,N2
(157)
4
The relevant properties are a = 2.74510 , b = 1.823, Tc,O2 = 154.4 K, Tc,N2 =
126.2 K, pc,O2 = 49.7 atm, pc,N2 = 33.5 atm, MO2 = 32, MN2 = 28. Plugging
in,
a
=
p

DO2 N2

2.745 104
=
1 atm

1/3

5/12

!1.823

330 K
p

(154.4 K)(126.2 K)
1/3

((49.7 atm)(33.5 atm))

5/12

((154.4 K)(126.2 K))

1
1
+
28 32

0.5

The diffusion coefficient is DO2 N2 = 0.248 cm2 /s. The effective diffusion
coefficient is then
ef f
DO
= 0.201.5 0.248
(158)
2 N2
ef f
DO
= 0.0222 cm2 /s
2 N2

5. (5 points) The limiting current density is


jL = nF Def f

c0R

(159)

The bulk concentration of oxygen is


N
P
0.21 atm
=
=
= 7.75 106 mol/cm3
V
RT
0.08205 L atm/mol K 330 K
(160)
Note that n = 4 since we are calculating losses at the cathode (oxygen is
the species of interest) so that we find
c0R =

jL = 4 96485 0.0222

7.75 106
0.05

jL = 1.33 A/cm2

42

(161)

6. (10 points) The various losses add:


V = E act ohmic conc

(162)

Neglecting anodic activation losses, we get


V = E [ac + bc ln(j + jleak )] j ASR c ln

jL
jL (j + jleak )

(163)

with the values found above, this becomes


V = 1.19 [0.196 + 0.0284 ln(j + 0.005)] 0.1j 0.1 ln

1.33
1.33 (j + 0.005)
(164)

The i V and i p plots are shown in the figure.

Figure 10: Voltage and power density as a function of current density for
problem 6.5
Note that since we are using the Tafel form, the activation losses are
underestimated at low current density.
7. (10 points) Finding the maximum power density can be done analytically, but as a first guess, we can just determine from the graph that
the maximum power density is about 0.77 W/cm2 which occurs at a current
density of about 1.1 A/cm2 .
43

For an analytical solution, we need to find the maximum in the power


density curve:
dp
dV
p=Vj
=V +j
(165)
dj
dj
So the optimum occurs at
jL
dp
= 0 =E [ac + bc ln(j + jleak )] j ASR c ln
+
dj
jL (j + jleak )


bc
c
j
ASR
j + jleak
jL (j + jleak )
A numerical solution shows that our guess from the graph was very accurate.
The peak power density is pmax = 0.775 W/cm2 at a current density of
j = 1.12 A/cm2 .
8. (5 points) If f is the fuel utilization fraction, the overall fuel cell
efficiency is
V
G
V
 = thermo f =
f
(166)
E
HHHV E
Using values from the appendix to find G, we find
1
Grxn = GH2 O(l) GH2 GO2 = 309.15(42.91)0.5(75.07) = 228.59 kJ/mol
2
(167)
We use the voltage at the maximum power point (V = 0.692 V ) to find the
overall efficiency
228.59 0.692
=

0.90
(168)
286
1.19
 = 41.8%
Problem 6.6 Using the fact that the total number of moles is conserved,
dxi = dxj (which also follows from xi + xj = 1), we can find that
Ji + Jj = Dij

dxj
dxi
dxi
+ Dji
= (Dij Dji )
dz
dz
dz

(169)

In steady state, there is no net flux, so Ji + Jj = 0. In general, 6= 0 and


i
there may be concentration gradients, so dx
dz 6= 0, which means that we must
have Dij = Dji .

44

Problem
6.7 Showing that x1 + x2 + + xN = 1 is equivalent to
P
showing
dxi = 0. Proceeding that way, sum the Stefan-Maxwell equation
over all components to get
X dxi
i

dz

RT

X xi Jj xj Ji
ef f
P Dij

j6=i

RT X xi Jj xj Ji
ef f
P
Dij
i,j6=i

(170)

Since Dij = Dji , each term in the sum cancels to give 0, as we require. To
show that each term cancels, we write out the sum for two species i, j = 1, 2:

X xi Jj xj Ji
i,j6=i

ef f
Dij

X
x1 Jj xj J1 x2 Jj xj J2 x1 J2 x2 J1 x2 J1 x1 J2
+
+

=
D1j
D2j
D12
D21

j6=i |
{z
} |
{z
}
|
{z
} |
{z
}
i=1

i=1,j=2

i=2

i=2,j=1

(171)
= (x1 J2 x1 J2 x2 J1 + x2 J1 )/D12 = 0

(172)

Problem 6.8 Using the values from the text, the j-V curve is shown below. From the figure, the limiting current density is approximately 2.3 A/cm2 .

Problem 6.9 The j-V curve at T = 873 K is shown. It is clear from the
plot that the ohmic losses increase dramatically at lower temperaturesthe
maximum current density produced by the fuel cell at this temperature is
only about 0.43 A/cm2 !
Problem 6.10 The j-V curves for an electrolyte supported SOFC at
T = 1073 K and T = 873 K are shown below. Given the very high ohmic
losses, especially at lower temperatures, it is clear that it is not optimal to
support an SOFC with a thick electrolyte.
Problem 6.11 (a) The linear approximation to the Butler-Volmer equation is
RT jp0
RT
jp0
anode =
=
(173)
2F j0 p
2F j0 pA xH2
which is similar to equation 6.22 in the book. Using an analogous equation
to 6.26 for xH2 , we find
anode =

RT
2F

j

j0

pA

xH2 |a

tA

jRT
ef f
2F pA DH
,H
2

45


2O

(174)

Figure 11: IV plot of the SOFC model for problem 6.8

Figure 12: IV plot of the SOFC model at T = 873 K for problem 6.9

46

Figure 13: IV plot of the SOFC model at T = 1073 K for problem 6.10

Figure 14: IV plot of the SOFC model at T = 873 K for problem 6.10

47

where we again implicitly express pC in atm, allowing us to drop the p0 .


(b) Using anode parameters from table 6.1 in the book, and the rest of
the parameters from table 6.4, the j-V curve may be plotted. Anode losses
in this case are very small.

Figure 15: IV plot of the SOFC model for problem 6.11


Problem 6.12 The i-V curves for an anode supported SOFC according
to our model are shown in the figure below. In this case, anode losses
remain small, but are not negligiblelosses are dominated by ohmic loss.
The limiting current density for the cathode is approximately 14.3 A/cm2 ,
and for the anode around 18 A/cm2 . In this simple model, the anode losses
do not blow up as the cathode losses do, because we have taken the linear
form for the losses, which breaks down long before the limiting current
density.
Problem 6.13 At each value of current density, we must solve for the
constants and C to find (z) in order to find the resistance. Following
the equations in the text, it is possible to find the complete j-V curve, as
is shown below. Important steps along the way are the following equations.

48

Figure 16: IV plot of the anode supported SOFC model for problem 6.12
The conductivity of the membrane is
h
i
1
1
(z) = 0.005193(4.4 + Ce0.000598jz/D ) 0.00326 e1268( 303 T ) = s1 +s2 es3 z
(175)
Where several constants are introduced to simplify notation. Then the resistance of the membrane is


Z tm
Z tm
dz
dz
z
ln(s1 + s2 es3 z ) tm
Rm =
=
=

(176)
(z)
s1 + s2 es3 z
s1
s1 s3
0
0
0
Rm =


1
s3 tm ln(s1 + s2 es3 tm ) + ln(s1 + s2 )
s1 s3

(177)

Note that the ohmic losses are not linear with current, as at higher
current, the membrane is dried and the resistance increases. As current
increases, the first and third term in the above equation increase, and R
increases.
Problem 6.14 (a) The j-V curve for = 1.2 is shown below, as are the
power density curves for = 1.2 and = 2.0. The maximum power density
in the first case is approximately 0.25 W/cm2 , so if the pump consumes
49

Figure 17: IV plot of the PEMFC model for problem 6.13


10% of this power, the power delivered is approximately 0.23 W/cm2 . If
= 2.0, the max power density is approximately 0.56 W/cm2 , so if the
pump consumes 20% of this power, the power delivered is approximately
0.45 W/cm2 , still more than in the first case.

50

Figure 18: IV plot of the SOFC model for = 1.2 problem 6.14
Chapter 7 Solutions
Problem 7.1 (5 points) First, to quantitatively determine how good
a fuel cell is. Second, to determine why it is good or bad.
Problem 7.2 (10 points) There isnt just one right answer. As long
as you give a good argument behind your answer, we accept the following
answers. Major operational variables include:
1. Temperature: The most important temperature dependence is in the
equation for the exchange current density.
/RT

j0 = AeG

(178)

You can see that the exchange current increases exponentially with
temperature. From the Butler-Volmer equation in the low overpotential regime, act 1/jo , so that the overpotential depends exponentially on temperature. In the high overpotential regime, act ln jo ,
and the dependence is linear. The Butler-Volmer equation shows a

51

Figure 19: Power density plot of the SOFC model for problem 6.14. = 1.2
line is blue, = 2.0 line is green.
temperature dependence that is less strong:


CR
CP
nF
(1 )nF
o
j = jo
]
o exp [ RT ] C o exp [
CR
RT
P

(179)

From this equation, the overpotential is linear in temperature. Also,


from the Nernst equation
 i 
aP
RT
ln
(180)
E = ET
nF
aRi
the temperature dependence is linear. For transport, temperature
comes into play in a ceramic by the equation for conductivity
c(nF )2 Do e
=
RT

Gact
RT

(181)

So the ohmic overpotential depends exponentially on temperature.


2. Gas Pressure: both Nernst and Butler-Volmer are important here too.
We showed in class and in an example that the Nernstian pressure
52

dependence is logarithmic, that is to say weak. Also, the transport


equation shows logarithmic dependence:
conc =

c0
RT
ln R
nF
cR

(182)

3. Gas composition: again, Nernst and BV, as well as transport equation


182
Other operational variables include the compression force that presses the
fuel cell together, the flow rates of gases, etc.
Problem 7.3 (5 points) An EIS measurement gives more detailed
information, though it is harder to interpret than a current interrupt measurement. The current interrupt measurement can be done on largerhigher
powerfuel cells, can be done faster, in parallel with an i-V measurement,
and with simpler measurement equipment.
Problem 7.4 (a) (5 points) A current scan from zero current to a high
current that is too fast will show a voltage that is too high, because the fuel
cell hasnt reached a steady state. At each current step, the voltage must
drop down to its steady-state value. The faster scan rate (100 mA/s) will
show a higher voltage, ie. a better performance.
A voltage scan will have the opposite effect. Since you scan from OCV
(high voltage) to zero voltage, at a step downwards in voltage, the current
must ramp up to the new steady-state value. If you step down in voltage too
fast, the current will not ramp up to the steady-state value, and you measure
a current that is too low. In this case, the faster scan rate (100 mV /s) will
show a current that is too low, ie. a worse performance.
(b) (5 points) The slowest step in the fuel cell reaction will be the most
affected by a fast scan. As electrochemistry is quite fast in comparison, the
gas diffusion is the slowest step in the fuel cell reaction (as shown by the
low frequency of the Warburg element representing mass transport in figure
7.12). Thus, the gas diffusion, which dominates at high current density, will
be the most affected by the faster scan rate.
Problem 7.5 Intuition should get students most of the way on the
problem. The circuit diagram is shown in the first figure.
You should inspect the easy, limiting cases of high frequency and low
frequency first. (1) At high frequency, capacitors act as short circuits, so
the combination of blocking + activated electrode acts as just the resistor
53

Figure 20: One blocking and one activated electrode for problem 7.5.
Rb . The Nyquist plot of this resistor is just a point at Rb . (2) At low
frequency, capacitors act as open circuits, so we can ignore Ca . The circuit
now acts as a resistor Rb + Ra in series with a capacitor. The Nyquist
plot of this circuit is just a vertical line, with the intercept at Rb + Ra . At
intermediate frequencies, we have some of the character of the parallel RC
circuit, which is a semicircle in a Nyquist plot. The amount of the semicircle
you will see depends on the distance between the point Rb and the vertical
line at Rb + Ra . Well answer that question later.
If you dont buy that argument, the mathematical treatment is given
here. The impedance of the circuit is
ZT otal = ZRb + ZCb +

ZRa ZCa
ZRa + ZCa

(183)

Inputting the complex impedances of resistors and capacitors


ZT otal = Rb

j
j
jRa /Ca
j
jRa Ra + Ca
(184)
+
= Rb

2
Cb Ra j/Ca
Cb Ca Ra + 21C 2
a

Separating the real and imaginary components of the impedance,


ZRe = Rb +

Ra / 2 Ca2
1
= Rb + Ra
1
2
2
1
+
(R
Ra + 2 C 2
a Ca )

(185)

ZIm =

1
R2 /Ca
1
1
1
+ 2a
=
+

1
Cb Ra + 2 C 2
Cb Ca 1 + (Ra Ca )2

(186)

The low frequency asymptote = 0 shows that ZRe = Rb + Ra and ZIm =


, and the high frequency ( = ) shows that ZRe = Rb and ZIm =
0. This checks with our inspection of the circuit diagram above. Given
an intercept and asymptote, information about the relative magnitudes of
54

the RC components can be used to sketch in the intervening region. As


given in the problem, when the RC time constants are well separated, the
resolution of each segment should be possible. For the skeptics, a more
detailed treatment follows.
A key question is: is the semicircle visible not? This question is actually
much tougher to answer, and probably cant be found by inspection. If you
see part of a semicircle, that means that ZIm has a maximum. So, here is
the derivative of Zim with respect to frequency:




d(ZIm )
1
1
2(Ra Ca )2 3
1
1
+

=0=

d
Cb 2 Ca 2 1 + (Ra Ca )2
Ca
(1 + (Ra Ca )2 )2
(187)
Let = (Ca Ra )1 , and factor out 1/ 2


1
1
1
22
0= 2

+
(188)

Cb Ca (1 + 2 ) Ca (1 + 2 )2
Since we arent interested in the solution = , we discard that solution
and clear the denominators:
0=
Let C =

Cb
Ca .

(1 + 2 )2 (1 + 2 ) 22

+
Cb
Ca
Ca

(189)

Now expand and group like terms to get a quadratic in 2 :

0 = (1 + 22 + 4 ) + (1 + 2 )C 22 C = 4 + 2 (2 C) + (1 + C) (190)
The solution is
2 =

C 2

p
(C 2)2 4(1 + C)
2

(191)

This will have a solution if the determinant is positive, so in order to see


most of our semicircle, we require
(C 2)2 4(1 + C) > 0 = C 2 4C + 4 4 4C > 0

(192)

Therefore, C 2 8C > 0 that is, since C > 0 by definition, C > 8. So to see


the peak in the semicircle, Cb > 8Ca , or the blocking R C must be larger
than the activated R C. Note that this requirement does not depend on
the relative values of the resistance!
(a) (10 points) When RC is much smaller for the parallel RC than for
the series RC, the entire semicircle can be seen before the = 0 asymptote
55

Figure 21: EIS plot for 7.5(a), when Ra Ca < Rb Cb

is approached. A MATLAB plot of the EIS curve for some chosen values of
R and C is shown here.
(b) (10 points) When R C for the parallel RC is much larger than
the series RC, the semicircle is only beginning before the asymptote is approached. Only the first linear part of the semicircle can be seen. A MATLAB plot of the EIS curve is shown here.
(c) (10 points) The circuit diagram is shown here.
The impedance of the circuit is

q 
j

C2 R3 + (1 j) tanh( j
jR1
D)
C
1
q
Z=
+
R
+
(193)
2
j
j
j

R1 C

+
R
+
(1

j)
tanh(
)
1
3
C2
D

1
To simplify notation, make theq
following definitions: 1
1 C1 R1 , 2
C2 R3 , and W R tanh( j
D ). Clearing out the imaginary denomi3
nators,

Z = R1 1

1 j
j[1 + 2W + 2 W + 2W 2 + j2 (1 + W )]
+
R
+
R

2
3
2
(1 + W )2 + (2 + W )2
1 + 21
(194)

56

Figure 22: EIS plot for 7.5(b), when Ra Ca > Rb Cb

Figure 23: Circuit diagram for problem 7.5


The real impedance is
ZRe = R1

21
2 (1 + W )
+ R2 + R3 2
(1 + W )2 + (2 + W )2
1 + 21

(195)

and the imaginary impedance is


ZIm =

R1 1
1 + W (2 + 2 + 2W )
+ R3 2
(1 + W )2 + (2 + W )2
1 + 21

(196)

The high frequency asymptote (1 = 2 = 0) gives


ZIm = 0

ZRe = R2 ;

57

(197)

which could be confirmed by inspection of the circuit diagram. The low


frequency (1 = 2 = ) gives
ZRe = R1 + R2 + R3 (1 + W );

ZIm = R3 W

(198)

Which makes sense because the definition of W has R3 in the denominator.


A program such as MATLAB can plot the results for given values for the
constants. A plot that uses numbers from Table 7.2 is shown.

Figure 24: EIS plot generated for problem 7.5


Problem 7.6 (10 points) Porosity refers to the amount of free space
in a structure, but for a structure to be permeable, the holes need to go the
entire way through a material. A sketch of a highly-porous material with
low permeability (in the direction in the plane of the page) is shown.
Problem 7.7 (15 points) In Example 7.1, we found the values at point
(b)
(b)
(b): ohmic = 0.10 V and act = 0.30 V .
At point (c), the current appears to be 1.75 A. From the left intercept
of the Zreal axis, the ohmic resistance is 0.10 . The ohmic overvoltage is
(c)

ohmic = iRohmic = (1.75 A)(0.10 )


58

(199)

Figure 25: The highly-porous material with low permeability for problem
7.6.

(c)

ohmic = 0.175 V
Remember that Rf is the diameter of the semicircle, not the intercept of
the semicircle. Then from the right intercept of the Zreal axis minus the left
intercept, Rf = 0.20 , so the activation overvoltage is
(c)

act = iRf = (1.75 A)(0.20 )

(200)

(c)

act = 0.350 V
We may fit the data to the equation to extract the remaining desired
values. If you plug in numbers first, since i(b) = 1, things simplify and the
solution is shorter. Or, solving the general case:
RT
i(b)
ln
nF
io

(201)

RT
i(c)
RT
i(c)
ln
=
ln
(c)
nF
io
io
nF act

(202)

(b)

act =
and
(c)

act =

Substituting into equation 201


(b)

(b)
(c) ln[i /io ]
ln[i(c) /io ]

act = act

(203)

Isolating the io term


(b)

act
(c)
act

(b)

ln i(c) ln i(b) = ln io

59

act
(c)

act

!
1

(204)

Therefore, io is

io = exp

(b)

act
(c)
act

ln i(c) ln i(b)

(b)
(c)
act /act

"

= exp

0.30
0.35

ln(1.75) ln(1)
0.30/0.35 1

#
(205)

io = 0.0348 A
Substituting this value back into equation 202 gives alpha
=

1.75
(8.314)(300)
ln
(2)(96485)(0.35) 0.0348

(206)

= 0.145
Problem 7.8 (10 points) The scan rate allows us to convert from
an i-V plot to an i-t plot. To find the total charge Qh , remember that
charge equals current times time, so the area under the i-t curve gives the
1
charge. Estimating the width of the peak to be 400 mV 10 mV
/s = 40 s
and the peak of the curve to be about 40 A, the area under the curve is
approximately
Qh (40 s)(40 A) = 1600 C
(207)
Then the active catalyst area coefficient is
Ac =

Qh
1600 C
=
Qm Ageometric
210 C/cm2 (.1 cm)2
Ac = 760

60

(208)

Chapter 8 Solutions
Problem 8.1 (5 points)(a) Ni is used in high temperature fuel cells
as a catalyst. While it is not as good a catalyst as some noble metals, it is
far cheaper. The position of Ni in the periodic table is above Pt and Pd,
both good catalysts, showing that its electronic structure is well suited to
catalysis.
(b) YSZ is mixed with Ni in SOFCs primarily to add ionic conductivity.
In addition, it adds thermal expansion compatibility, mechanical stability,
and maintains the high porosity of the electrode.
(c) Cr is added to Ni in MCFCs to maintain porosity and high surface
area.
Problem 8.2 (5 points) As long as a convincing argument is provided,
we recommend accepting either of the answers:
1) High temperature fuel cells benefit from faster reaction kinetics, and
therefore require lower noble metal loadings as catalysts. As cost is a primary barrier to widespread fuel cell adoption and noble metals add greatly
to the cost of low temperature fuel cells, the lower noble metal loading of
high temperature fuel cells is the largest advantage.
2) High temperature fuel cells benefit from fuel flexibility. Whereas low
temperature fuel cells are poisoned by CO presence in the ppm level, high
temperature fuel cells use CO as a fuel. Since generation of neat hydrogen
is currently very costly, in the foreseeable future, the only economical way
towards a hydrogen economy is via reforming of hydrocarbons. In that case,
fuel flexibility is an enabling feature, as impurities such as CO are present
in large levels in reformed fuel.
Note: alleviation of water clogging issues is also an important advantage
of high temperature fuel cells. The availability of high-quality waste heat
is a nice perk, but is not a main advantage of high temperature fuel cells.
Most of the present research on high temperature fuel cells concerns how
to lower the operating temperature, which shows that people consider the
waste heat to be of lesser importance.
Problem 8.3 (10 points) See the figure below.
Problem 8.4 (5 points)(a) The stack power is 25 kW . That power
is produced by 40 cells, each with a cell area of 6000 cm2 . The area-based

61

Figure 26: Fuel cell in problem 8.3


power density is therefore
25 kW/40 cells
6000 cm2 /cell

(209)

104 mW/cm2
(b) Each cell has a volume
120 cm 81.4 cm 0.65 cm = 6349.2 cm3

(210)

. Then the volumetric power density of the stack is


25 kW
40 cells 6349.2 cm3 /cell

(211)

98.4 mW/cm3
Problem 8.5 (5 points) Assuming the hydrogen is stored at room
temperature, the quantity of hydrogen on board the vehicle is
N=

PV
350 atm 156.6 L
=
= 2.24 kmol
Latm
RT
0.08205 molK
298 K

(212)

The energy stored in this fuel is (using the HHV of hydrogen, 286 kJ/mol)
2.24 kmol 286 kJ/mol = 641 M J
62

(213)

If this energy is used continually over a span of 4.3 hours (430 km/100 km/hr),
the power output is
641 M J
= 41.4 kW
4.3 hr 3600s/hr

(214)

If the fuel cell is 55% efficient, the fuel cell produces 22.8 kW during operation under these conditions.
Problem 8.6 (10 points)(a) Carnot efficiency is given by
Carnot = 1
The efficiency is Carnot = 1

Tc
Th

(215)

373
1073

Carnot = 65.2%
(b) If the SOFC is 55% efficient, then 45% of the input energy is rejected
as heat. This heat is available to the heat engine that has efficiency of
HE = 0.6Carnot = 39.1%. The efficiency of the combined cycle is T ot =
SOF C + (1 SOF C )HE = 0.55 + 0.45 0.391.
T ot = 72.6%

63

Chapter 9 Solutions
Problem 9.1 (5 points) Fuel flows up the center of the donut and in the
porous anode. Air flows in around the sides of the porous cathode . Sealing
prevents mixing of air and fuel. One advantage of the structure is that it
facilitates oxygen transport: usually oxygen is the limiting reactant. As it
flows through the cathode, it gets depleted, but it has access to the most
reactive area around the edges where it is least depleted. As the oxygen gets
depleted along the cathode, it accesses less area so that the depletion is not
as detrimental to performance. The figure shows a sketch of the structure,
the inset shows the gas flow.

Figure 27: One possible configuration for a donut shaped fuel cell for problem 9.1
Problem 9.2(5 points) The DOE canceled the on-board fuel reforming
project because current technologies are inadequate and there is no promising path forward. There is no interest from industry, and no reason to
suspect that a future FCV (fuel cell vehicle) with on-board reforming will
outperform future gasoline-electric hybrids. Since the great interest in fuel
cell cars lies in their zero emissions capability, we do not perceive a huge
impetus to implement FCVs that run on gasoline.
Problem 9.3(10 points) (a) The heat generation comes from ineffi64

ciency. The thermodynamic efficiency at STP is 83%, and we must find


the voltage efficiency V = 0.7/1.23 = 57%. So the fuel cell efficiency is
0.83 0.57 = 47%. Since the fuel cell generates 1000 W of electricity at
47% efficiency, the input fuel must flow at a rate of 1000 W/.47 = 2128 W ,
so it must generate 2128 W .53 = 1128 W of heat. The heat rejection is
1128 W .
(b) The parasitic power consumed by the cooling system is 1128 W/25 =
45.1 W .
Problem 9.4(10 points) The molar mass of methanol is 32.0 g/mol, so
g/cm3
3
the molar density of methanol is 0.79
32 g/mol = 0.0247 mol/cm . The molar
3

g/cm
mass of water is 18 g/mol, so the molar density of water is 1.0
18 g/mol =
0.0556 mol/cm3 . If we assume that the volume of methanol and water do
not change upon mixing, we must find x, the number of moles of water and
methanol that will fill one liter.
x
x
1
1
1L=
+
= x = (
+
)1 = 17.1 mol
55.6 mol/L 24.7 mol/L
55.6 mol 24.7 mol
(216)
Next, we find the HHV of methanol. A methanol combustion reaction is

3
CH3 OH + O2 CO2 + 2H2 O
2

(217)

So that the HHV of methanol is obtained by


3
HfO2
2
3
= 393.51 + 2(285.83) (238.4 0) = 727 kJ/mol
2
H2 O(l)

HHHV =HfCO2 + 2Hf

CH3 OH(l)

Where the value Hf

CH3 OH(l)

Hf

= 238.4 kJ/mol was found from the NIST

kJ/mol
Webbook (a good resource if you dont know it). In kW h, this is 727
3600 s/hr =
0.202 kW h/mol. Finally, we know how much fuel we have, so the fuel tank
capacity is 17.1 mol 0.202 kW h/mol = 3.45 kW h. For our 2 liter system,
we have an energy density of 1.72 kWh/L . The small difference between
this value and the quoted value may be explained if the Hf numbers were
obtained from different sources.

Problem 9.5(10 points) Effectiveness is defined in the text as


ef f ectiveness =

% conversion of carrier energy to electricity


% conversion of neat H2 to electricity
65

(218)

We can rewrite this as


ef f ectiveness =(% conversion of carrier energy to dirty 00 H2 )
% conversion of dirty 00 H2 to electricity
% conversion of neat H2 to electricity
Using the numbers given, we find
ef f ectiveness = 0.75 0.8 = 0.6

(219)

So that the effectiveness of our system is 60% .


Problem 9.6(10 points) (a) Given the constraint that the total power
is specified to be P , we say that the power density of the fuel cell is pF C =
P
xV . Now, the efficiency explicitly written as a function of x is
P
(220)
xV
Now we may write equation 9.9 in the book, for total energy of the system
as a function of x, as
(x) = A B

P
)
xV
Setting the derivative with respect to x equal to zero, we find
E = (1 x)V eF (A B

(221)

P
dE
= BV eF 2 V eF A = 0
(222)
dx
x V
By inspection of this equation, we can see that the extremum we will obtain
will be a maximum. Solving for x,
r
P
BP
B = V A = x =
(223)
2
x
AV
So for a fuel cell with higher efficiency (lower B or larger A), the volume we
devote to the fuel cell decreases. For larger power requirements, we devote
more volume to the fuel cell, and for larger volume allowance, we devote a
smaller fraction of the volume to the fuel cell.
(b) Plugging in numbers,
r
0.003 L/W 500 W
x=
(224)
.7 100 L
So x = 0.146 . We devote about 15% of the system to the fuel cell. For a
500 W
3
power density of pF C = 0.146100
L = 34 W/L = 0.034 W/cm , a reasonable
power density, wed operate a fuel cell at an efficiency of 0.7 0.003 W/L
34 W/L = 0.6, which is also reasonable.
66

Chapter 10 Solutions

Problem 10.1 The four primary sub-systems of a fuel cell system are 1)
the fuel cell sub-systems, 2) the fuel processing sub-system, 3) the thermal
management sub-system, and 4) the power electronics sub-system.
For example, a fuel reformer in the fuel processing sub-system depends
on the thermal management sub-system to provide heat from other system
components for an endothermic reaction at the reformer or to extract heat
from an exothermic reaction at the reformer.
As another example, the fuel cell sub-system depends on the power electronics sub-system to regulate the output voltage of the system. As shown
by the shape of the fuel cells polarization curve, the fuel cells voltage declines at higher currents and higher powers. Converters enable the system
to provide a constant voltage.
As another example, the heat produced by an afterburner downstream
of the fuel cell stack depends on the anode utilization of the stack. At
high anode utilizations, less anode off-gas will be available for downstream
combustion. Instead of a downstream afterburner, a downstream microturbine could be used to recapture energy in the anode off-gas.
One way fuel cell systems can be integrated is to use excess heat from
exothermic chemical reactions to preheat inlet fuel, oxidant, and water.
Problem 10.22) The Pressure Swing Absorption Unit (PSA) is a gas
purification unit that uses a physical mechanism to remove contaminants.
From a hydrogen-rich gas stream containing CO, HCs, CO2 and other
species, a PSA unit can produce a 99.99% pure hydrogen stream. One
of its adsorbent beds adsorbs all non-hydrogen species because of the relatively higher molecular weight of these species as compared with hydrogen.
The swing in PSA refers to the switching between two adsorbent beds.
When one adsorbent bed is being regenerated, the other is adsorbing nonhydrogen species. When one bed is full and the other bed is regenerated,
the hydrogen-rich stream is re-diverted (swings) to the regenerated bed.
Problem 10.3 An exothermic reaction is defined as a chemical reaction
that releases heat. An endothermic reaction is a chemical reaction that
absorbs heat.
Endothermic: steam reforming
Exothermic: oxidation of hydrogen fuel in a fuel cell, partial oxidation,
water gas shift reaction, selective methanation, selective oxidation, combus67

tion of fuel cell exhaust gases


Neither: 1) a chemical reaction takes place but heat is neither released
nor absorbed: autothermal reforming, 2) no chemical reaction takes place
but heat may be released: hydrogen separation via palladium membranes,
pressure swing adsoption, condensing water vapor to a liquid, expansion of
hydrogen gas (hydrogen heats up as a result of the Joule-Thompson Effect),
compression of natural gas (in practice, releases heat due to frictional losses
at the compressor)
Problem 10.4 A process diagram for a fuel cell system for a scooter is
included in the accompanying figure.

Figure 28: Process diagram for a fuel cell system for a fuel cell scooter.
(Problem 10.4)
Problem 10.5 A process diagram for a fuel cell system for a scooter
with metal hydride storage is shown. Metal hydride storage tanks are composed of a special granular metal that absorbs and releases hydrogen like a
sponge absorbs and releases water. Depending on the metal hydride material, it may need to be pressurized to charge it with hydrogen and heated to
discharge the hydrogen. The process of charging and discharging the metal
hydride requires a more complicated thermal management sub-system. One
68

possible design is sketched in the figure: Heat from the fuel cell stack is
used to heat the hydride to release hydrogen. Coolant air is used to cool the
hydride so that it can absorb hydrogen. Because hydrogen absorption is an
exothermic reaction, heat must be continuously removed from the alloy bed.
The rate at which a metal hydride can absorb or release hydrogen is primarily dependent upon the rate at which the alloy can release or absorb heat,
respectively. Although not shown in the figure, the thermal management
sub-system becomes more complicated than the one shown if the fuel cell
operates at a lower temperature (for example, 80 C for a PEM) than the
minimum temperature required for hydrogen release in the metal hydride
(for example, 150 C for a N aAlH4 metal hydride).

Figure 29: Process diagram for a fuel cell system for a fuel cell scooter with
metal hydride hydrogen storage. (Problem 10.5)
The absorption reaction with the metal hydride depends on the pressure
and temperature of the hydrogen gas. If the gas pressure is above a certain
equilibrium pressure at a certain temperature, the metal absorbs hydrogen.
If the hydrogen gas pressure is below the equilibrium pressure, the metal
hydride releases its hydrogen.
Problem 10.6 For partial oxidation, all of the products are CO and

69

H2 . A balanced reaction is
C8 H18 + a(O2 + 3.76N2 ) bCO + cH2 + 3.76aN2

(225)

C8 H18 + 4(O2 + 3.76N2 ) 8CO + 9H2 + 15.04N2

(226)

The hydrogen yield is then


yH2 = 9H2 /(8CO + 9H2 + 15.04N2 ) = 28%.

(227)

However, if all of the CO in the product stream was shifted to hydrogen via
the water gas shift reaction,
CO + H2 O CO2 + H2

(228)

then the hydrogen yield could be greater. The overall reaction would be
C8 H18 + 4(O2 + 3.76N2 ) + 8H2 O 8CO2 + 17H2 + 15.04N2

(229)

The hydrogen yield is then


yH2 = 17H2 /(8CO2 + 17H2 + 15.04N2 ) = 42%

(230)

The WGS reaction can increase the hydrogen yield by 14%. (This calculation
does not directly consider the addition energy input required to vaporize
liquid water for the WGS reaction.)
Problem 10.7 We do a back-of-the envelope calculation based on the
example problems to produce a reasonable estimate at STP. For steam reforming, for a 100% efficient transfer of heat between a methane burner and
a steam generator, the mass/moles/volume of methane needed for combustion is at a minimum about 31.5% of the moles/mass/volume of methane
consumed by the steam reformer (Example Problem 1.6) at STP. Assuming only 72% efficient heat transfer, then the moles of methane needed for
combustion is at a minimum about (31.5%/0.72 =) 43.8% of the moles of
methane consumed by the steam reformer at STP. Under these conditions,
the fuel reformer efficiency in terms of HHV is
R =

HHHV,H2
4 mol H2 (284 kJ/mol H2 )
=
= 89.8%
HHHV,F uel
1.438 mol CH4 (880 kJ/mol CH4 )

(231)

A more exact calculation can be conducted for reactants entering and


CH (1000 K) = 36.62 kJ/mol, h
CO (1000 K) =
products leaving at 1000 K. h
4
2

360.11 kJ/mol, hH2 O(g) (1000 K) = 215.83 kJ/mol, hO2 (1000 K) =


70

rxn,CH =
27.71 kJ/mol, so the enthalpy of methane combustion is h
4
H (1000 K) =
810.57 kJ/mol. The enthalpy of hydrogen combustion is (h
2
rxn,H = 250.37 kJ/mol. Then the fuel reformer effi20.68 kJ/mol) h
2
ciency is
F R =

4 mol H2 (250.37 kJ/mol H2 )


= 85.9%
1.438 mol CH4 (810.57 kJ/mol CH4 )

(232)

Problem 10.8 We conduct a back-of-the envelope calculation based on


the example problems to produce a reasonable estimate. We ignore the
enthalpy of formation of nitrogen because it does not chemically react. The
autothermal reforming reaction is
Cx Hy + zH2 O(l) + (x z/2)O2 xCO2 + (z + y/2)H2 CO, CO2 , H2 , H2 O
(233)
The value for the steam-to-carbon (S/C) ratio, here shown as z/x, should
be chosen such that the reaction is energy neutral, neither exothermic nor
endothermic. For autothermal reforming of propane, the only variable is z
in the reaction:
C3 H8 + zH2 O(l) + (3 z/2)O2 3CO2 + (z + 4)H2

(234)

rxn = 0. You can find the equation for z to


We must find for what z the h
be:
H h
H O + 1 h
O ) = h
C H + 3h
O 3h
CO
z(h
2
2
2
3 8
2
2
2
z=

12 kJ/mol + 3 22 kJ/mol 3 (360 kJ/mol)


= 4.6
20 kJ/mol (216 kJ/mol) + 12 22 kJ/mol

(235)
(236)

The steam to carbon ratio is then z/x = 4.6/3 = 1.5. The reformer would
produce 4.6 + 4 = 9.6 moles of H2 per mole of fuel.
Problem 10.9
Steam reforming
The heat of reaction for this steam reforming reaction at 1000 K is

hrxn = 190.89 kJ/mol. The heat of reaction for the complete combustion
rxn = 800.57 kJ/mol. Therefore, assuming
of methane at 1000 K is h
all of the reactants are preheated to 1000K, only 190/800 = 0.24 moles of
methane must be combusted to provide energy for every 1 mole of methane

71

that goes through the endothermic steam reforming process. The reaction
stoichiometry is
1.24CH4 + 2H2 O + 0.48O2 1.24CO2 + 4H2 + 0.48H2 O

(237)

These reactions are not usually combined in the same reactor. Because
combustion is usually done in a separate reactor from the steam reforming,
the products of combustion CO2 and H2 O and the gas N2 do not dilute
the hydrogen yield in the product gas. The hydrogen yield is then yH2 =
4 H2 /(1CO2 + 4H2 ) = 80%, as shown in the example problem. The ratio of
hydrogen produced per unit of methane consumed is 4/1.24 = 3.23.
Partial Oxidation
The hydrogen yield for the partial oxidation reaction is the same as
shown in the text. For partial oxidation, all of the products are CO and H2 .
A balanced reaction is
CH4 + 0.5(O2 + 3.76N2 ) CO + 2H2 + 1.88N2

(238)

The hydrogen yield is then yH2 = 2H2 /(1CO + 2H2 + 1.88N2 ) = 41% The
ratio of hydrogen produced per unit of methane consumed is 2. However, if
all of the CO in the product stream was shifted to hydrogen via the water
gas shift reaction then the hydrogen yield could be greater. The overall
reaction would be
CH4 + 0.5(O2 + 3.76N2 ) + H2 O CO2 + 3H2 + 1.88N2

(239)

The hydrogen yield is then yH2 = 3H2 /(1CO2 + 3H2 + 1.88N2 ) = 51%.
The WGS reaction can increase the hydrogen yield by 10%. The ratio of
hydrogen produced per unit of methane consumed is 3. (This calculation
does not directly consider the addition energy input required to vaporize
liquid water for the WGS reaction.) The autothermal reforming reaction for
methane is
Autothermal Reforming
We have the reaction
CH4 + zH2 O(g) + (1 z/2)O2 CO2 + (z + 2)H2

(240)

We need to find the value of z for which the enthalpy of this reaction is
zero at 1000 K (and assuming water enters as a gas at 1000 K), which gives
an equation for z:
CO + (z + 2)h
H h
CH zh
H O (1 z/2)h
O
0 = h
2
2
4
2
2
72

(241)

H O = 215.83,
CH = 36.62, h
Using the values at 1000 K (in kJ/mol) h
2
4

hO2 = 22.71, hCO2 = 360.11, hH2 = 20.68, we find z = 1.22.


The number of moles of N2 involved is 3.76 (1 1.22/2) = 1.46 moles.
The hydrogen yield is then yH2 = 3.23H2 /(1CO2 +3.23H2 +1.46N2 ) = 57%.
The ratio of hydrogen produced per unit of methane consumed is 3.23. This
is the same ratio as with steam reforming, but with steam reforming, the
hydrogen yield is higher due to the separation of the combustion reaction
and the steam reforming reaction in separate reactors.
Problem 10.10 From example 10.8, the heat available to heat the building is 11.6 kW . We need to find the volume of space that can be heated,
assuming heat is lost through conduction through the walls. Then the heat
transfer is
T
Q = kA
(242)
l
Where l is the wall thickness, which well assume to be 10 cm. The thermal
conductivity of the walls, k, we can take to be 0.1 W/mK, which is be
approximately the value for wood or some types of concrete. Then the
surface area of the heated room is
A=

Ql
(11.6 kW )(0.1 m)
=
= 504 m2
kT
(0.1 W/mK)(23 K)

(243)

If the building is cubic, the volume of the building is V = (A/6)3/2 = 770 m3 .


Alternatively, if the building is 8 f t high, standard for a one story building
in the US, the building may have a square footprint of 18.1 m per side
(assuming no heat is lost through the ground, it is only lost through the
roof and sides).
Problem 10.11 The T-H diagram looks similar to that for the condenser
in figure 10.11.
As in table 10.9, we assume that half of the waste heat from the fuel cell
exits via the cathode exhaust gas. Then the heat flow in the exhaust that
needs to be removed by the condenser is QF C /2 + Qc , where QF C is the
heat produced by the fuel cell and Qc is the heat needed to condense water,
given by the latent heat of condensation of water. Convection removes an
amount of heat
Q = hA(Th Tc )
(244)
where A is the surface area of the condenser in contact with the air and h is
a function of the velocity of the air flowing past the condenser. To removing
enough heat from the exhaust to condense the water, one must design the
73

Figure 30: T-H diagram of a condenser. (Problem 10.11)


condenser to have an area large enough that Q = QF C /2 + Qc even for low
velocity.
If we assume that the fuel cell operates at a voltage of 0.5 V and has a
range of 2 hours at maximum power, then the rate of hydrogen use is
P/V
N H2 =
= 0.010 mol/s
2F

(245)

The rate of water production is N H2 O = N H2 , and the amount of water


produced is
NH2 O = (0.010 mol/s)(2 3600 s) = 74.6 mol = 1.34 kg

(246)

The volume of this water, at STP, is 1.34 L.


Problem 10.12 An alternative heat exchanger network configuration
that increases the pinch is to use two parallel cold streams. The first absorbs
heat in series from the fuel cell, the aftercooler, the selective oxidation unit,
and the post-shift converter. The second absorbs heat from the condenser.
The following figures show the new T-H diagrams and the new pinch point
temperatures. The pinch no longer occurs in the condenser, and the pinch
in the overall system increases. The pinch is greater than ten degrees over a
range of mass flow rates of water in each cold stream. The last figure shows
this range of the ratio of mass flow rates in one cold stream (that flows
through the fuel cell, aftercooler, selective oxidation unit, and post-shift
converter) vs. the other cold stream (flowing through the condenser).
74

This configuration manages the cooling loop by placing four thermal


sources in series with each other and altogether in parallel with respect to
the condenser, a fifth thermal source. The primary advantages of this design
are 1) its ability to increase the pinch point temperature and 2) under certain
design conditions, this configuration can capture much of the waste heat.
The disadvantages of this design are 1) greater system complexity than with
respect to the first configuration examined and 2) more complex control
compared with initial configuration.

Figure 31: Temperature change of the cold and hot streams and four thermal
sources, as a function of enthalpy change of the two streams. (Problem
10.12)
Analysis of Pinch Points for Configuration 3 at Full Power (6kWe)
A model for the new parallel configuration can be made to visualize
the change in enthalpy versus flow stream temperature. Figures 31 and
32 show the results. The change in enthalpy refers to the aggregate level
of enthalpy being exchanged between the hot stream and the cold stream,
and the temperatures shown at each enthalpy point are the inlet and outlet
temperatures for the thermal source. Figure 31 shows the temperature vs.
enthalpy plot for the four thermal sources in series on one side of the parallel
loop: the fuel cell, aftercooler, selox, and post shift sources. Figure 32 shows
the temperature vs. enthalpy plot for the condenser on the other side of the

75

Figure 32: Temperature change of the cold and hot streams in the condenser,
as a function of enthalpy change of the two streams. (Problem 10.12)
parallel loop. The combined total flow rate for both streams is the mass
flow rate needed to capture all the heat available so as to increase the initial
inlet temperature for the domestic cooling stream from 25 C to the desired
outlet temperature of 80 C. The flow rates chosen for each of the streams in
parallel are based on the mass flow ratio of 0.58 (first parallel stream) to 0.42
(second parallel stream), which allows each parallel stream to independently
achieve an exit temperature of 80 C.
Figures 31 and 32 show each parallel streams pinch point temperature,
the minimum temperature difference between hot and cold streams for effective heat transfer. From figure 31, one can see that the first parallel streams
pinch point temperature is approximately 3500 W cumulative heat load in
the aftercooler at a temperature 8 C. Figure 32 shows the second parallel
streams pinch point temperature to be approximately 1800 W cumulative
heat load in the condenser at a temperature of 18 C.
Keeping the combined total mass flow rate of water to the domestic cooling loop constant at the same rate, the ratios of the flows in the two parallel
branches should be varied to find the mass flow rate ratio at which the
pinch point in the system is maximized. Figure 33 shows the results of this
analysis, by plotting the minimum pinch point temperature for each thermal source with respect to the ratio of mass flow of the first parallel stream
(containing the fuel cell, aftercooler, selox, and post shift) with respect to
the total flow rate. The intersection of the aftercooler pinch point curve and
the condenser pinch point curve shows that the pinch point for the system
76

can be maximized to a temperature difference of 12.5 C by operating at a


mass flow ratio of 0.64. If the mass flow ratio range is maintained between
0.60 and 0.67, the pinch point temperature will fall at or above 10 C.

Figure 33: Pinch point analysis considering all five thermal sources as a
function of the two streams flow rates. Results are for a power of 6 kW e.
The overall pinch point follows the black line. (Problem 10.12)
To understand this result consider that at one extreme, the flow ratio is
limited to above a certain minimum by the pinch point in the aftercooler
(shown by the square symbols in figure 33) to maximize the pinch point
temperature in the aftercoolers hot and cold streams. At the other extreme,
the flow ratio is limited to below a certain maximum by the pinch point in the
condenser (shown by the starred crosses), because the condensers cooling
water must remain below 100 C to avoid vaporization. The composite
sketch of the systems overall pinch point range is shown as the Overall Pinch
Point curve (shown by a single curve), which is the composite minimum of
the aftercooler pinch point and condenser pinch point curves.
Problem 10.13 Reading off figure 10.11 the location of the pinch is the
aftercooler. To make this quantitative, we need to find first the temperature
at which the water in the first stream condensesthis is where the pinch point
77

occurs. As in example 10.10, we equate the heat transfer of the entire process
with the heat transfer to a stream of liquid water and a stream with water
vapor to solve for the condensation temperature Tc :
Q = (mc
p )1 (Th Tc ) + (mc
p )2 (Tc Tl )
Tc =

Q (mc
p )1 Th + (mc
p )2 Tl
(mc
p )1 (mc
p )2

(247)
(248)

Using the data for mc


p and the high and low temperatures from table 10.9,
we solve to find Tc = 84.78 C. This is the temperature of the hot stream at
the pinch point. Next, we need to find the temperature of the cold stream
at this point, so we need to find what H point on the curve this temperature
corresponds to:
dH = (mc
p )1 (Tc Tl ) = (276 W/ C)(84.78 C 60 C) = 6838 W (249)
We can now find the temperature of the cold stream
Tb Tbi =

dH
6838 W
=
Tb = 72.82 C
mc
p
143 W/K

Therefore, the pinch point is 84.78 72.82 12 C.

78

(250)

Chapter 11 Solutions

Problem 11.1 The primary steps of Process Chain Analysis (PCA) are
as follows:
1. Research and develop an understanding of the supply chain from raw
material production to end use.
2. Sketch a supply chain showing important processes and primary mass
and energy flows.
3. Identify the bottleneck processes, which consume the largest amounts
of energy or which produce the largest quantities of harmful emissions
(or both).
4. Analyze the energy and mass flows in the supply chain using a control
volume analysis and the principles of conservation of mass and energy.
5. Having analyzed the individual processes within the supply chain, evaluate the entire supply chain as a single control volume. Aggregate net
energy and emission flows for the chain.
6. Quantify the environmental impacts of these net flows, for example,
in terms of human health impacts, external costs, and potential for
global warming.
7. Compare the net change in energy flows, emissions, and environmental
impacts of one supply chain with another.
8. Rate the environmental performance of each supply chain against the
others.
9. Repeat analysis for an expanded, more detailed number of processes
in the supply chain.
Problem 11.2 Several gases and particles are understood to have a
warming effect on the Earth, a good summary is given in figure 11.5. Section
11.3.3 describes the following effects from gases and particles:
Anthropogenic greenhouse gases include CO2 , CH4 , H2 O, and nitrous
oxide (N2 O). Greenhouse gases selectively absorb infrared radiation, and
then re-emit this radiation partly back towards the Earths surface.
79

In addition to these gases, certain particles also have a warming effect


on the Earth, but through a different mechanism. Dark-colored particles,
such as soot, absorb sunlight, re-emit this energy as infrared radiation, and
therefore also warm the Earth. Dark-colored particles that contribute to
global warming include black carbon (BC). The warming effect of black
carbon is enhanced by Organic Matter (OM), which focuses additional light
onto black carbon. The center portion of Figure 11.5 shows the warming
mechanism of dark-colored particles. Figure 11.5 shows that these gases
and particles re-emit infrared radiation towards the Earths surface to cause
warming; they also re-emit infrared away from the Earth.
In contrast, light-colored particles reflect sunlight and have a cooling
effect. Light-colored particles that cool the Earth include sulfates (SULF)
and nitrates (NIT). SULF also attract water, which reflects light as well.
Emitted gases that have a cooling effect include sulfur oxides (SOx ), nitrogen oxides (N Ox ), and non-methane organic compounds or volatile organic
compounds (VOC). These gases react in the atmosphere and convert to particles, which are mostly light in color. SOx converts to SULF, N Ox converts
to NIT, and VOC convert to organics that are mostly light in color. The
right portion of Figure 11.5 shows the cooling mechanism of light-colored
particles.
Problem 11.3 Section 11.4 describes the six primary emissions that
create air pollution: ozone, carbon monoxide, nitrogen oxides, particulate
matter, sulfur oxides, and volatile organic compounds (VOC). VOC are nonmethane organic compounds, such as the higher hydrocarbons. Some of
these compounds are air pollutants themselves. Others react with chemicals
to produce air pollution. Effects of air pollution on human health can include
respiratory illness, pulmonary illness, damage to the central nervous system,
cancer, and increased mortality.
Four of the more important air pollutants known to affect human health
include CO, N O2 , O3 , and P M10 . CO is known to cause increased incidents of headaches, hospitalization, and mortality. N O2 is known to cause
increased irritation to the sinuses, throat and eyes. O3 increases the incidents of asthma attacks, eye irritation, lower respiratory illness, upper respiratory illness, and other conditions. P M10 is known to increase incidents
of asthma attacks, respiratory restriction, chronic illness, and morality. (See
Table 11.3)
Problem 11.4 A National Emission Inventory (NEI) is a database of a
countrys emissions from various sources, which records the type of emission,

80

its quantity, the location of the emission, and its time of release. National
emission inventories vary in the depth of material they provide. The US
EPAs inventory records emissions from vehicles, power plants, factories,
chemical processing plants, industrial facilities and other sources. Emissions
may be recorded on an annual, monthly, or even daily basis. Since the
location of stationary facilities is known, the location of these emissions is
known. The location of emissions from mobile sources (such as vehicles)
may be estimated based on the location of the sale of the fuel.
Problem 11.5 Leaked hydrogen can combust with oxygen in air if a
certain quantity of hydrogen accrues in high enough concentrations at high
enough temperatures. The self-ignition temperature of hydrogen is 858 K
and its ignition limits in air are between 4% and 75% at STP. Without
proper ventilation, a closed indoor area could accrue high concentrations of
hydrogen that could lead to ignition.
Problem 11.6 As described in section 11.3.5, the mechanisms through
which hydrogen might contribute to global warming and air pollution are
very complex and still the subject of intense research.
One mechanism through which released H2 might increase global warming is by indirectly increasing the concentration of the greenhouse gas CH4 .
In the troposphere (lower atmosphere), H2 reacts with the hydroxyl radical
(OH), according to the reaction
H2 + OH H2 O + H

(251)

If H2 did not consume OH in this reaction, OH might otherwise reduce the


presence of CH4 via the reaction
CH4 + OH CH3 + H2 O

(252)

The net effect of hydrogen then is to reduce the concentration of OH, which
increases the concentration of CH4 , a greenhouse gas. However, numerous
other chemical reactions must also be considered.
One mechanism through which released H2 might increase one type of
air pollutant is through a series of chemical reactions that enhance the concentration of O3 . In the troposphere, H2 might increase O3 by increasing
the concentration of atomic hydrogen (H). After several years in the atmosphere, molecular hydrogen decays to atomic hydrogen in the presence of
the hydroxyl radical (OH), via the reaction
H2 + OH H2 O + H
81

(253)

Atomic hydrogen (H) could then react with oxygen in air in the presence
of photon energy (h) from light to increase O3 in the presence of M, any
molecule in the air that is neither created nor destroyed during the reaction
but that absorbs energy from the reaction.
H + O2 + M HO2 + M

(254)

N O + HO2 N O2 + OH

(255)

N O2 + h N O + O

(256)

O + O2 + M O3 + M

(257)

However, other sets of reactions must also be considered, with a focus


on their net effect on air pollution. The net effect of these reactions may be
able to be determined with computer simulations of chemical reactions in the
atmosphere (atmospheric models). As we learned in PCA, these simulations
should model not the mere addition or subtraction of an individual chemical
component, but rather the net change in emissions among different scenarios
to be accurate.
Problem 11.7 One possible answer is to compare and contrast different
energy conversion scenarios using Process Chain Analysis (PCA) and data
from an NEI. For example, one may examine the biological generation of
hydrogen via different mechanisms. Some researchers have concluded that
the energy input requirements are too great to merit development of some
of these. This assertion could be examined. A research proposal abstract
should be no more than a page in length and address a technically literate
but non-expert audience.
Problem 11.8 We can perform this calculation from the following equation (11.14)

CO2,equivalent =mCO2 + 23mCH4 + 296mN2 O + (mOM,2.5 + mBC,2.5 )


(258)
[mSU LF,2.5 + mN IT,2.5 + 0.40mSOX + 0.10mN OX + 0.05mV OC ]
where m is the mass of each species emitted, with, for example, mOM,2.5
indicating the mass of organic matter 2.5 microns in diameter and less. The
coefficient can range between 95 and 191. The coefficient can range
between 19 and 39. We will perform the calculation for a low case and a

82

high case, where for the high case, = 191 and = 19. For the low case,
= 95 and = 39.
We use the data in table 11.2; please note that for mV OC you may
use the entry for total non-methane organics. Further, the problem asks
only to consider organic gas and particulate matter. Then we find for the
high case mCO2 ,equivalent = 7.35 108 metric tons/yr, and for the low case,
mCO2 ,equivalent = 3.98 108 metric tons/yr.
Problem 11.9 We first need to find how much fuel is consumed by a
fleet of fuel cell vehicles. The fuel consumption rate can be found from the
consumption of gasoline in the current fleet after adjusting for the relative
efficiencies of the vehicles.
VM T
m
H2 C =
(259)
Fh
where
Fh =

barMgvf Vc HHHV,h h
g HHHV,g g

(260)

We are given that h /g = 2, so using the numbers from example 11.2 we
find
Fh =

(17.11 mi/gal)(264 gal/m3 )(142 M J/kg)


2 = 36.16 mi/kg
(750 kg/m3 )(47.3 M J/kg)

(261)

The vehicle miles per year is VM T = 2.681012 mi/yr so we find that annual
fuel consumption is m
H2 C = 7.41 1010 kg/yr. If the hydrogen leakage is
2%, the annual hydrogen production must be 7.56 1010 kg/yr.
According to the Example 10.3, in the steam reforming reaction approximately 4 moles of H2 are produced for every one mole of CH4 . If the heat for
this endothermic steam reforming reaction is provided by the waste heat of
another process, no additional CH4 need be consumed. If CH4 is combusted
to provide heat for the reaction, some additional CH4 is consumed. In example 10.6, a back-of-the envelope calculation shows that approximately
31.5% excess CH4 must be combusted to provide sufficient heat for the reaction. Then, for every 4 moles of H2 produced, approximately 1.315 moles
of CH4 may be consumed. (The actual ratio depends on the design of the
fuel reformer, the supply of external heat to the reformer from other sources,
and the efficiency of the reformer.) Then the mass of CH4 needed for this
hydrogen production is
m
H2

1.315 16 g/mol CH4


= 1.99 1011 kg/yr
4
2 g/mol H2
83

(262)

Considering the 1% methane leakage, actual production must be 2.01


1011 kg/yr.
To compare this to current production, we must assume that natural gas
is simply CH4 , and find that the US current annual production is 23.14
1015 BT U/yr. To find this in kg, convert
(23.14 1015 BT U/yr)

1055 J
1 kg CH4

= 4.40 1011 kg/yr


1 BT U 5.55 107 J

(263)

So the amount we are considering for a hydrogen fleet is 46% of current US


production.
The CO2 equivalent of the leaked methane is calculated from equation
258. The CO2 equivalent of the leaked methane is 4.63 1012 of CO2
equivalent/yr.
The external cost of the leaked methane can be estimated from the external cost of global warming, estimated as $0.026 and $0.067 per kg of CO2
equivalent (section 11.3.7). Based on these estimates, the external cost of
the leaked methane (1% of the total) is between $1.2 and $3.1 billion/yr.
Problem 11.10 The EPA gives a Natural Gas Hydrogen Fuel Cell Vehicle (HFCV) Scenario as an alternative to the base case. In both scenarios,
they predict CO2 equivalent emissions.
The base case gives CO2 equivalent estimates (low and high) of 5.33109
and 5.86 109 tons/yr. The HFCV scenario estimates are 4.58 109 and
5.08 109 tons/yr. Therefore, reasonable estimates for the change in CO2
equivalent can range from a reduction of 12% to 15% of all anthropogenic
sources in the natural hydrogen fuel cell vehicle case (compared with the
1999 fleet). The actual reduction depends on choices pertaining to the scenario, such as the type of reformer and the energy requirements required for
storing hydrogen on the vehicle (gas or liquid). See the following table.

84

Species
Carbon Monoxide (CO)
Nitrogen Oxides (NOx) as N O2
Organics
Paraffins (PAR)
Olefins (OLE)
Ethylene (C2 H4 )
Formaldehyde (HCHO)
Higher aldehydes (ALD2)
Toluene (TOL)
Xylene (XYL)
Isoprene (ISOP)
Total Non-Methane Organics
Methane (CH4 )
Sulfur Oxides (SOx) as SO2
Ammonia (N H3 )
Particulate Matter
Organic Matter (OM2.5 )
Black Carbon (BC2.5 )
Sulfate (SU LF2.5 )
Nitrate (N IT2.5 )
Other (OT H2.5 )
Total P M2.5
Organic Matter (OM10 )
Black Carbon (BC10 )
Sulfate (SU LF10 )
Nitrate (N IT10 )
Other (OT H10 )
Total P M10
Species
Carbon Dioxide (CO2 )
Water (H2 O)
CO2,equivalent (low)
CO2,equivalent (high)

HFCV and Natural Gas Gases


-55.29%
-33.18%
-27.10%
-32.43%
-24.86%
-19.68%
-51.98%
-16.71%
-27.69%
-49.54%
-26.24%
20.88%
2.05%
-5.25%
-1.78%
-15.08%
-0.32%
-0.82%
-0.15%
-1.26%
-1.18%
-10.87%
-0.35%
-0.38%
-0.03%
-0.41%
-15.03%
1.97%
-14.00%
-13.25%

Problem 11.11 This is an open-ended problem for which answers will


vary. Reasonable electrical efficiencies for fuel cell systems are between 40%
and 60%, and the efficiency of hydrogen generation can approach 1 (see
example 10.6). The efficiency of current US electricity generation is ap85

proximately 32%. Therefore, the efficiency could potentially almost double.


Emission reductions are estimated to be significant, based on emission estimates shown in the text in Table 11.1.
Problem 11.12 This is an open-ended problem for which answers will
vary. One fuel cell manufacturer produces fuel cell systems that have already
achieved electrical efficiencies of 50% in combination with heat recovery
efficiencies close to 40%. If 90% of the heating value of the fuel could be
usefully used, the efficiency of current US electricity and heat generation
combined could approximately triple. (The average boiler/furnace in the
US achieves an efficiency of about 80%. State-of-the-art condensing boilers
achieve efficiencies in the low to mid 90% range.)
Under this combined-heat-and-power scenario, emission reductions are
estimated to be even more significant than those shown in the text in Table
11.1.
Problem 11.13 This is an open-ended problem for which answers will
vary. Of all of the links in the upstream supply chains, one of the most
important is the energy required to put hydrogen into a form in which it
can be stored onboard a vehicle. For liquid H2 storage, approximately 30%
of the heating value of the fuel is consumed in the energy required to cool
H2 down to a liquid. For gaseous H2 storage, approximately 10% of the heat
value of the fuel is required for hydrogen compression. In this case, one can
assume the energy required to run the hydrogen compressors is provided by
electricity from the grid.
These upstream electricity generating plants can be assumed to have the
same mix of plant as shown in the text in Figure 11.2. Based on this scenario,
a reasonable set of results for the change in emissions is shown in the table
accompanying problem 11.10. The estimates in the table also assume 1%
CH4 leakage. The change in CO2 equivalent is also shown in the table.
All estimates with respect to total anthropogenic emissions. From these
emission estimates, one can calculate the change in health effects, external
costs due to air pollution, and external costs due to global warming.
First, as an example, we will calculate the health effects from this change.
Using the data in table 11.2 in the text for emissions from on-road vehicles,
and table 11.4 showing the monetized health effects, combined with the data
in the table of problem 11.10, we can find the change in health effects from
this switch. The low estimate is a reduction in costs of $36.8 million, and
the high estimate is a savings of $570 million.

86

Problem 11.14 This is an open ended problem for which answers will
vary.
Problem 11.15 In example 11.2, we find the quantity of hydrogen consumed by a hydrogen fuel cell vehicle fleet to be approximately 57.1 MT of
H2 /year. If 1% of H2 that was produced leaked, approximately 0.6 MT of
H2 /year would leak. According to table 1.2, conventional on-road vehicles
produce 0.16 MT of H2 /yr. Therefore, this hydrogen fleet might almost
quadruple H2 emissions from vehicles.

87

You might also like