Anda di halaman 1dari 24

Chapter 10 MODELING AND ANALYSIS

10.1 GENERAL A practical SOFC must have the electrical and electrochemical performance, along with the mechanical and structural integrity to meet operating requirements of specified power generation applications. In terms of electrical and electrochemical performance, the fuel cell must be designed to achieve high voltages (at the required current density) with low ohmic and polarization losses. In terms of mechanical and structural integrity, the fuel cell design must minimize thermal stresses during fabrication and operation to avoid exceeding the strength limits of the component materials. Mathematical modeling and analysis are common tools used in SOFC design and optimization to achieve the required electrical, electrochemical, mechanical, and structural performance. Mathematical modeling and analysis are used to predict cell behavior under a variety of conditions and to investigate the effect and the relative importance of various processing and operating parameters. Modeling and analysis can provide information on how various processes and parameters interrelate, allowing interpretation of cell behavior with a minimum of expensive testing. Modeling and analysis also provide a picture, for example, of stress, potential (or voltage), current density, and temperature as functions of position and time for various cell configurations and operating conditions. It may then be possible to use the information provided by the models and analysis to optimize cell designs and operating parameters. There are several levels of modeling and analysis: the molecular type model, electrode model, cell model, stack model, and system model [10.1]. The discussion in this chapter is limited mainly to thermal stress analysis, electrical analysis, and performance modeling (specifically current and temperature distribution) of SOFC cells and stacks. (i) Thermal stress analysis: Residual thermal stresses are generated due to differences in the coefficients of thermal expansion of various components.

308

Chapter 10

Such stresses have been analyzed in SOFC structures using mathematical models such as finite-element techniques. (ii) Electrical analysis" Analysis of ohmic and polarization resistances is performed to estimate cell voltage losses. Equivalent circuits have been developed to predict cell performance under a variety of operating conditions. (iii) Performance modeling: Mathematical modeling of SOFC performance is typically based on formulating and solving simultaneous mass, energy, and potential balance equations. Such a model provides voltage/current density characteristics, current distribution, and temperature distribution of the fuel cell. 10.2 STRESS ANALYSIS A SOFC is a layered ceramic structure composed of several different materials. During fabrication and operation, the structure is subject to nonuniform temperature distributions and various thermal cycles. Thus, matching thermal expansion in the cell components is critical. A small difference in the coefficients of thermal expansion of the layers can result in large thermal stresses in the fuel cell and cause cracking in the ceramic structure. In general, the problem of thermal expansion mismatch in present SOFCs mainly centers on the anode, since nickel has a higher coefficient of thermal expansion than Y203stabilized ZrO2 (YSZ). The anode typically must contain more than 30 vol% nickel to have sufficient conductivity. The coefficient of thermal expansion of this composition is higher than that of the YSZ electrolyte and other components, causing thermal expansion mismatch. Stress analysis has been performed to evaluate the effects of design, fabrication, and operation parameters on the fracture characteristics of SOFCs. Analysis of thermal stresses in the fuel cell structure typically uses simplified stress models or the finite-element method. For example, a simple model based on small-displacement linear elastic theory has been employed to analyze singlecell circular plates [10.2]. Simplified stress models generally involve establishing stress equations and deriving strain energy release rates and stress intensity factors. The finite-element method, a numerical technique of piecewise approximation, is conventionally used in stress analysis, especially for intricate structures. The finite-element method models a SOFC structure as an assemblage of elements with nodes indicating where elements are connected to one another. A finite element stress analysis typically involves the following steps"

Modeling and Analysis

309

9 Divide the structure into finite elements. 9 Formulate the properties of each element. (Determine nodal loads associated with all element deformation states that are allowed.) 9 Assemble elements to obtain the finite-element model of the structure. 9 Apply the known loads (nodal forces and/or moments). 9 Specify how the structure is supported. 9 Solve simultaneous equations to determine nodal displacements. 9 Calculate element strains from the displacements and the displacement field interpolation, and finally calculate stresses from strains. A simple double-layered model has been developed for the sealless tubular cell [10.3]. The cell is approximated as two layers -- an anode layer of thickness di~ and length 2e, and the rest of the cell (the support, cathode, and electrolyte) as a layer of thickness 62 and length 2f. The following equations are derived for the tensile stress, tri, in the anode and the shear stress, r, at its interface with the rest of the cell (~ is the coefficient of thermal expansion, T is the temperature, E is the elastic modulus, v is the Poisson's ratio, and x is the distance from the center of the double layer, x = 0 to ___e)"
o"1 =

,111 11 +
El61 E2~2
7 = #I(r

(r

(coshc coshcx) coshc

(Eq. 10.1)

(l-u)

r ) A T sinhcx

(Eq. 10.2)

/ilC

coshc

where c and/Xl are given as


r = I~1

1
E282

(Eq. 10.3)

81 E181

~1,1 =

(Eq. 10.4) 2(l+v)

310

Chapter 10

Using this simplified model, approximate equations for strain energy release rates for cracks in the anode have been derived. A finite-element stress analysis of the sealless tubular cell has also been performed [10.3]. The stress model consists of 166 four-node quadrilateral and three-node triangular isoparametric elements. Stress distribution in the individual components of the cell has been obtained from the model and has been compared with that from the simple double-layered model [10.3]. A stress and fracture analysis has been developed for SOFC single cells (anode/electrolyte/cathode trilayers) during cooldown after cofiring. This model considers a circular single-cell disk of radius r containing an annular crack of depth C c (Figure 10.1) [10.2]. The model assumes that the temperature, thermal expansion coefficient, and elastic modulus of the disk vary only in the thickness direction, if at all. The radial stress, r and tangential stress, frO(i), at section AiA i a r e given
as

Or(0 = Oo(0 - (l_v)l-~i-ot(zi)T(zi)

E(zi)[gx,

E(zi)zi My..
Di(l_v2) '

(Eq. 10.5)

where for each i = 1 to 3, E(z) is Young's modulus at zi, 1~ is Poisson's ratio, ~(z~ is the coefficient of thermal expansion at zi, T(z~ is the temperature range at zi, zi is the distance from centroidal axis.

I<
A,

6',
A3

~l

z~,
A,
r _

A3 A2

Figure 10,1. Geometry of circular single cell containing an annular crack [10.2]

Modeling and Analysis

311

The other parameters are defined by the following equations"

faEzi dz i = 0
xf. K~ - 1-v ,E dz~
Di -

(Eq. 10.6)

(Eq. 10.7)

1 1 - v 2 fa

Ez/2. d z i

(Eq. 10.8)

1 N~, _ l-v fA,Ea T dz~

(Eq. 10.9)

M~,

i
1-v

a Tz~ dz~

(Eq. 10.10)

Thus, the strain energy release rate per unit of extension of the crack area, G, is given as
G =

Gn +

G I =

~ +

6)2 ~3 + M2 + M3 r 2 r 3 D2(I+v) Da(l+v)

(Eq. 10.11)

where, for i = 2,3


GI =
4-

(Eq. 10.12)
D3(l+v)

/92(1+v)
2

6)2 ~3
GH =

(Eq. 10.13)

fAj %(0 dzi

(Eq. 10.14)

312
and

Chapter 10

M, =

-fa ~
l

dzi

(Eq. 10.15)

In this model, the opening mode component G~ is assumed to be more important than the shear component G,. The stress intensity factor, K~, is related to the mode I strain energy release rate by the following equation: [ E G t ] '/' K~ = 1-v 2 (Eq. 10.16)

The maximum K~ has been calculated as a function of the coefficient of thermal expansion of the anode for a symmetrical anode/electrolyte/anode trilayer (Figure 10.2). Correlation between calculated stress intensity factors and observed cracking behavior of trilayers indicates a fracture toughness of the YSZ electrolyte of about 1.25 MPa.n~. Figure 10.3 shows the boundaries (computed

,,

ANODE

~22
N x < N

1
O SURVIVED COOLDOWN 9 FRACTURED DURING COOLDOWN

0 11

12

13

14

THERMAL EXPANSION COEFFICIENT OF ANODE, 10 .6 cm/cm.K

Figure 10.2. Variation of maximum mode I stress intensity factor with coefficient of thermal expansion of anode [10.2]

Modeling and Analysis

313

ANODE

vol % Ni

o, 1 0 " e c m l c m K

E 100-

(A) (B)

40 30

13 12
CRACK AT ANODE(B) ~ELECTROLYTE ELECTROLYTE INTERFACE INTE_~F.a~E CRACK AT CATHODE-

u3
a o -1I"tL (A)

I
I

< (J O

z (...) "r"
I--

10
/ELECTROLYTE I ANODE/ M IDPLANE I ELTE;TFAR~)cL;TE ~, I ~'~

.....

' i

'

'

' 9 - ' ' ]

'

10 THICKNESS

100 OF A N O D E , ,urn

Figure 10.3. Effect of coefficient of thermal expansion of anode on cracking of asymmetrical single cell with electrolyte 38 I~m thick [10.2]

from the model) between the domains of cathode and anode thickness combinations that are predicted to survive a cooldown after cofiring (lower left) and those that are predicted to fail (upper right) for an electrolyte thickness of 38/zm and two different values of coefficient of thermal expansion of the anode. The finiteelement method has also been applied to analyzing the residual stresses developed in a single cell bonded to anode and cathode corrugations [10.4]. An eight-node layered-shell element is used to model the anode/electrolyte/cathode trilayer (single cell). A quadrilateral-shell element is used to model the anode and cathode corrugations. Based on the analysis, failure-zone and safe-zone maps have been developed to provide guidelines in selecting the proper thicknesses of electrodes and electrolyte to prevent failure of the fuel cell during fabrication. An example of such a map is shown in Figure 10.4.

314

Chapter 10
CORRUGATIOh THICKNESS = 2 5 0 p m I

175

/
E 150 ...........
LU

..

............................................................

r 125 z v
-r- lOO

........................................

............................

............................ m ~

!natle s .....................

~ o
~
_J

/
75

l.............................

5O

.......................................

. . . . . . . . . . . . . . . . . . . . . . . . .

25 0

25

50

75

100

125

150

175

ELECTROLYTE THICKNESS, pm

Figure 10.4. Failure- and safe-zone map for single cell bonded to anode and cathode corrugations [10.4]

10.3 ELECTRICAL ANALYSIS One of the objectives in stack design is to maximize electrochemical performance (minimize ohmic and polarization losses) in the fuel cell. Electrical analysis is often used to estimate cell internal resistances and to predict cell performance potentials under specified conditions. Electrical analysis of a particular SOFC design considers electron/ion paths in the stack to model ohmic resistances of cell components and contact resistances at interfaces. Several examples of electrical analysis are discussed in this section to illustrate the analysis principles and techniques. In banded SOFCs, one approach in electrical analysis is to approximate cell resistance [10.5]. The resistance of an unit cell of the banded configuration (Figure 10.5) is considered to be the sum of the resistances of the active cell, the interconnect, the cathode across the anode gap, and the anode under the cathode gap. Figure 10.6 shows plots of power, unit cell resistance, and current density as functions of electrolyte and interconnect lengths calculated by the approximation method. In this case, at maximum power per unit of volume, the value for electrolyte and interconnect length is 0.21 cm. Under those conditions, the resistance and the current density of a unit cell are 0.12 fl and 0.66 A/cm 2, respectively.

Modeling and Analysis

315

ANODE GAP ACTIVE CEkk~iNTERCONNECTION| ELECTROLYTE ..~ CATHODE GAP

-I ~.~s :.~-... : ~"..-~.2,.".q ;..5 :~;-"


'--,
. .

~~:-:,:"~::
...........

".,,'.-~.-~ INTERCONNECT

UNIT C E L L POROUS SUPPORT

~(('~t, r,-,,,,'z ~I~(( I


Figure 10.5. Unit cell of banded configuration [10.5]

~o:~ ,,0F,
,,=
.~

~ :~176
50 oi
0.4 - ~
|

"'~

i~
i ;

-~

l ,
I

,,,~
z r
~
~

o3-\
02-

:
i t

i t
I

0.1 0 0.8 >: t-ILl

0.6 ,,, -~ 0.4 oE


0.2 0

LU

rr re ::) U

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 ACTIVE CELL OR INTERCONNECT LENGTH, cm

Figure 10.6. Power, unit cell resistance, and current density as functions of electrolyte and interconnect lengths for banded SOFCs [10.5]

316

Chapter 10

A more exact approach to assessing cell resistance in banded SOFCs is to use a model of distributed constants [10.5-10. 7]. In this approach, the model considers an element, Ax, of the cell and its equivalent circuit (Figure 10.7). The following equations can be derived from the equivalent circuit: (Eq 10.17)

(Eq. 10.18) where subscripts o, f, and e denote cathode, anode, and electrolyte. Each current (]) value in these equations can be replaced by voltage and resistance. As Ax --,0, the following equations can be obtained (d is the diameter)"
dj~,~ _ -rid v'c~) dx p~8~

(Eq. 10.19)

dJ~

- ~d

v~x)
p~8~

(Eq. 10.20)

dx

where v is the voltage, p the resistivity, and c5the thickness. On the other hand, the relationship between the voltages is given as
p~
ve(x) +
8 o= d j~ +Ax) - v ~x ax)

PlAXjg~+,~) : 0 (Eq. 10.21)


~~ d

p,,Ax I

%o,)

f po~,

v ~,.~)

Figure 10.7. Equivalent circuit of banded SOFC cell [10.5]

Modeling and Analysis


As Ax --, 0, Eq. 10.21 becomes
dYe(x) _ 1 r P O dx ; d t-~o J~ + Ax)
9

317

Pf.
-5--/J1~xAx)] +

(Eq. 10.22)

Combining and solving Eqs. 10.19, 10.20, and 10.22 give V~x) = I [Amsinh(Ox)+Bmcosh(Ox)] (Eq. 10.23)

Je-

I [Amsinh(Ox) + BmC~ p~8~

(Eq. 10.24)

where je is the current density in the electrolyte, I is the current, and 0, Am, and Bm are given as (le is the electrolyte length)
02 = ~ 1 (Po+ P_~_f) PeSe 8 o 6f

(Eq. 10.25)

An, =

P/

(Eq. 10.26)

0~d 81

Bm

1
-

P o + P/cosh(0l,) 8 0 8f

(Eq. 10.27)

sinh(0/)

The following equation has also been derived for the resistance R c (in fl):

R~ =

6/

o/

. (Eq. 10 281

60

8/

This equation illustrates how the various parameters affect unit cell resistance, thus cell power output per unit of volume. For banded SOFCs, the most desirable number of single cells on a support tube (radius r) can be obtained at the condition of minimum resistance and can be calculated from the thicknesses of the electrolyte and electrodes [10.8].

318 The number of cells is given by


--.

Chapter 10

ll ~ 1

.[. ~

(Eq. 10.29)

The minimum resistance is thus given as


-

R~.

1 + 1 ) ~-r o, o15/ Oo6 o

2 ~ 5__( e

(Eq. 10.30)

The maximum power of the banded SOFC tube, Pwr the following equation (E~ is the reversible voltage): P ~max)

can be estimated from

4R~.

8~ ~ ( ~ . 4 - ~ 1 1 ) Oe O71 008 0

| 6e

(Eq. 10.31)

An analytical model has been developed to evaluate performance potentials of various SOFC configurations [10.9,10.10]. In this model, the current flow through a single cell is governed by a second order differential equation of the form A2V_

~2

(Eq. 10.32)

where V is the voltage between the two interconnect contact points, and = ~ P e6e (Eq. 10.33)

P/5/+ P08 o
Eq. 10.32 can be solved for one-dimensional single-cell strips and disks. The general solution for a single cell strip is as follows"

V(x) = A vsinh(x/f~ ) + Bvcosh(x/f~ )

(Eq. 10.34)

where x is the strip coordinate in the direction of current flow, and Av and By are integration constants determined by the boundary conditions. Two contact

Modeling and Analysis

319

geometries are possible for the single cell resistance (Figure 10.8). For directly opposed line terminals, the effective specific ohmic resistance is given as
R,,, :

(Eq. 10.35)

with

Jv- X

ff

(Eq. 10.36)

For diagonally opposed line terminals, the effective specific ohmic resistance is R,~ = (p~;+p~6 ~+po8 o)Jv(coth(Jv) + Bv[Jv-2tanh(Jvl2)]} where (Eq. 10.37)

Po 8/ By =
8oPf

(Eq. 10 38)

PoSf2 (1+ ~)
8oPf
From these equations and information on comact and polarization resistances, potential cell voltage, power density, and cell efficiency can be estimated for various cell configurations. Some examples of predicted voltage/current curves are given in Figures 10.9 and 10.10.

ELECTR CA1

7 / F-=-~~--ax
0 x

Figure 10. 8. Current collection contact configurations [10.9]

320

Chapter 10

--=--

700oc 800oc

--,--

900oC 1000oC

--,--

1100o0

I
> 0.80

+ I
(,.9 0.60 < I.-.,,_,1 0 >

0.40 020 o.00 J 0 \ 2 4 6 8 CURRENT DENSITY, k A / m 2

-SISTANCE
lOSS-PLANE .ANE ONNECT 1100

-I,
~

\
10

Figure 10.9. Projected contributions to ohmic resistance, and voltage~current characteristics of sealless tubular SOFCs [10.9]

i
!

--u--

700oc 800oc

"--

900~

----o---- 1000~ 1100~

,oo 9 ~ _ _ / > o6o


< o 60

~
ooo
Ix: < 700

\
\
, .~ \ I
m

> o,o

+oo

(..) 0.20 0.oo 0 2 4

"'-:~..<.._/2/~--__.~"-..:..~//~ ~ IN-PLANE 800 ,.,^~ ~...~~------~~ INTERCONN=.,.,T ~uu ~<-~ 7/ L.~ ~ TEMPERATURE, o C 1000 ~ CONTACT 1100

10

CURRENT DENSITY, k A / m 2

Figure 10.10. Projected contributions to ohmic resistance, and voltage~current characteristics of bell-and-spigot SOFCs [10.9]

Modeling and Analysis

321

10.4 M O D E L I N G OF CURRENT AND T E M P E R A T U R E DISTRIBUTION At each point in a SOFC, a potential balance equation relates the cell voltage E to the reversible potential Er and to the voltage terms that depend on the local current I. Thus, cell voltage (assuming E is constant over the cell face) is given by
E = ErIRoj= - (ha+nO

(Eq. 10.39)

where Rohm is the local ohmic impedance of the cell, and r/a and r/~ are the anodic and cathodic overpotentials, respectively. To model the current distribution of a SOFC, it is necessary to solve Eq. 10.39. Knowledge of Er, Rohm, ~/a, and r/c is required to solve Eq. 10.39. 9 Reversible voltage: The reversible voltage Er is given by the reversible potentials at the cathode and anode. These potentials can be calculated using the Nernst equation, along with the local gas composition. The local gas composition can be calculated from mass balances on all reacting species, usually in terms of a local degree-of-conversion variable for each electrode. 9 Ohmic impedance: The ohmic impedance can be obtained from measured cell resistance or can be estimated from the effective conductivity of the cell components and the effective distance between the components. 9 Overpotentials: In general, the overall polarization behavior of a SOFC system is approximately linear over the current density range of practical interest. The quasilinear local polarization characteristics may be determined from cell measurements or electrode modeling. In a fuel cell stack, current and temperature distributions are strongly coupled. Practically every chemical, material, and transport property involved in current generation is significantly temperature dependent. Thus, the local temperature plays an important role in determining local current density, which, in turn, determines the local heat flux. To model temperature distribution, knowledge of a heat balance is required. Energy balance equations can be formulated to describe the heat transfer processes occurring in the local heat balance. The predominant heat transfer processes in a SOFC include: 9 Heat release or absorption arising from the electrochemical reactions, electrical resistances, and anode fuel chemistry (for example, internal reforming)

322

Chapter 10

9 Convective heat transfer between the cell components and the anode and cathode gas streams In-plane heat conduction through the cell components 9 Heat exchange between the cell and either gas stream due to the combined effect of conductive heat transfer and a mass transfer contribution associated with exchange of reacting species at the cell temperature. Many factors affect the relative importance of these heat transfer processes, and hence, the thermal characteristics of the fuel cell. Some of the factors include gas flow channel shape and size, gas flow configuration (e.g., coflow, crossflow, counterflow), gas inlet temperature, gas flow rate, and cell area, voltage, and pressure. Current and temperature distributions have been calculated for the various cell designs [10.11-10.36]. For example, a mathematical model has been developed to simulate a SOFC of crossflow configuration (Figure 10.11). The model considers a unit cell placed deep within the stack (such that end effects may be neglected). In this model, the horizontal plane of the cell is divided into small rectangular areas or elements, each defined as a node. The elements are sufficiently small such that the nodal values of temperature, pressure, concentration, etc., can be used over the element for the calculation of local current densities and heat fluxes.
9

ELECTROLYTE NTERCONNECT

Figure 10.11. SOFC of crossflow configuration [10.20]

Modeling and Analysis


The Nernst potential at any node (i,j) is given by the relationship

323

4. eI~ o ,av

(Eq. 10.40)

For a specific cell operating voltage, E, the current generated at the node is Itj = Er,q - E (Eq. 10.41)

Ri/

where the total resistance at any node R Ucan be calculated from the resistivities of the cell components and the electron and ion path length. Heat is generated in the node from the resistance to current flow and from the entropy change of the electrochemical reaction. The total heat released is assumed in the electrolyte layer and is given by Q~j = -~-ff(-A , (Eq. 10.42)

The heat is removed as sensible heat with the flowing fuel and oxidant gases. The heat transfer between the SOFC component layers and the flowing gases is assumed to occur only by convection. For heat removal from the electrolyte, the fuel- and oxidant-side heat transfer coefficients (h/and ho) are calculated based on a limiting Nusselt number of 3.0. For heat transfer between fuel and oxidant across the interconnect, an overall heat transfer coefficient, U, is used. The following energy balances are solved to determine the unknown electrolyte temperature, T e, at each node, fuel temperature at the outlet of each node, T/,u, and oxidant temperature at the outlet of each node, To,u: Cathode (oxidant) energy balance

ro, l_ijG,o(L,l_ij- Td) - ro,ijG, o(To, ij- Td) + Saho(L, ij- ro,a) + S a e (
Anode (fuel) energy balance
-

ro,a) - o

(Eq. 10.43)

+ SahtT,.,j- G )

* SaVG.av- G )

- 0

(Eq. 10.44)

324

Chapter 10
Electrolyte energy balance

Qij - Saho(T,,ij - To,a) - Sahy(Te,ij- T/a) = 0

(Eq. 10.45)

where Sa is the heat transfer area, Td the datum temperature, r the gas flow rate, Cp the specific heat capacity, and

_ To,i,,za- To,~j
2

(Eq. 10.46)

Tl, inta- Tf, iJ


2

(Eq. 10.47)

The model solution starts at the node where the fuel and oxidant inlet edges meet, because both inlet streams are completely defined. The average partial pressures and exit temperatures of the node are assumed, and a first estimate of the Nernst potential is obtained. The resistance and current are calculated; then the energy balance equations are solved to determine the average electrolyte temperature in the node and the temperatures of the gas streams leaving the node. The amount of fuel and oxidant consumed is calculated from the current to determine the new compositions of the gases. The compositions are used to calculate the partial pressures of the gas components at the node exit, which are, in turn, used to update the estimates of the average partial pressures. The iteration scheme is repeated until convergence of the partial pressures has been reached. The model has been used to simulate a 9-cm square cell in a SOFC stack. An example of input geometric and process data is given in Table 10.1. The results of the model calculations based on the input data in Table 10.1 are summarized in Table 10.2. Figures 10.12 and 10.13 show the current density distribution and electrolyte temperatures across the cell, respectively. In Figure 10.12, the current density is highest near the intersection of the fuel inlet and oxidant outlet edges because (i) the hydrogen concentration is high in the fuel stream, leading to higher Nernst potential, and (ii) the temperatures in this region are higher, resulting in better conductivity in the electrolyte, thus lower resistances. The temperature is also highest in this region because (i) the oxidant entering these nodes has picked up sensible heat from previous nodes, and (ii) the high current densities generate large amounts of heat, as seen in Figure 10.13.

Modeling and Analysis

325

TABLE 10.1 Model Input Data [10.20]

Geometric data Electrolyte thickness, mm Electrode thickness on electrolyte, mm Interconnect thickness, mm Electrode thickness on interconnect, mm Channel wall thickness, mm Channel height, mm Channel width,mm Fuel edge length, mm Oxidant edge length,mm Process data Hydrogen content in fuel inlet, volume fraction Water content in fuel inlet, volume fraction Inlet temperature, ~ Fuel inlet flow, standard L/h Oxygen content in oxidant inlet, volume fraction Nitrogen content in oxidant inlet, volume fraction Oxidant inlet flow, standard L/h Cell operating voltage, V

0.180 0.025 0.120 0.025 0.15 1.00 1.00 90 90 0.97 0.03 900 11.5 0.21 0.79 434 0.75

TABLE 10.2 Model Output Data [10.20]

Fuel utilization, % Oxygen utilization, % Fuel efficiency, % Gross power, W Net power, W Specific power, kW/kg Average Nernst potential, V Average current density, mA/cm 2 Average cell temperature, ~ Maximum cell temperature, ~ Maximum temperature gradient, ~

85 5.2 49 17.0 16.3 0.60 0.88 280 934 1050 187

326

Chapter 10

TO0"

~E u < E o~ z w o z oll oc oc

8o

8
"'~'~tVT EDG'E~'*t.E

0.00

u.

Figure 10.12. Current distribution across cell layer (Geometric and process data given in Table 10.1.) [10.20]
.. : : . . ::.: : .:::,. :.~::~-~. . .. . . . .

'

/
CT"lo,,w ,,,

Figure 10.13. Temperature distribution across cell layer (Geometric and process data given in Table 10.1.) [10.20]

As mentioned earlier, gas flow configuration plays a major role in determining current and temperature distributions in a fuel cell stack. For example, a model has been developed to simulate current and temperature distributions in SOFCs having crossflow, coflow, and counterflow configurations (Figure 10.14) [10.35]. This model considers a stack of square geometry with metallic interconnect. The stack has 20 channels, each 4 mm wide, 1 mm high, and 1 dm long, for fuel and oxidant gases. The anode, electrolyte, and cathode are 120, 200, and 50/zm thick, respectively. The gas composition is 54% H2,

Modeling and Analysis

327

29% H20, 12% CO, and 5 % CO2; the air stoichiometric ratio is 2; and the inlet temperature is 800~ As can be seen from Figure 10.14, the counterflow configuration shows the highest temperature; the coflow has a more uniform current density profile; and the crossflow presents a larger thermal gradient. Internal reforming of hydrocarbons significantly affects current and temperature distributions in SOFC stacks [10.21,10.33,10.35]. Figure 10.15 shows, as an example, the differences in temperature distribution in a SOFC operated with hydrogen (no internal reforming) (90 % H2, 10 % H20) and with methane (internal reforming) (25 % CH 4, 75 % H20 ) [10.21]. In this case, the inlet temperature is 900~ and the air stoichiometric ratio is 12.
TEMPERATURE DISTRIBUTION
CROSSFLOW COFLOW COUNTERFLOW

I/lll/lllIl[[/!1'0
2 I
4

TEMPERATURE, K

1~0

"

1~0

. 1370 v

1370

,1.1o 12~o

~"
=~
\
i

1310

1250 1190

~
,
r /

OXIDANT

: 1130

0
0 2 4 6 6 10 0 DISTANCE FROM OXIDANT INLET, c.m

!, _

soup
1

i lO7O
"
cm

1070

'

"

o"

i"

" ;"

DISTANCE FROM FUEL INLET,

DISTANCE FROM FUEL INLET, c m

CURRENT DISTRIBUTION
CURRENT DENSITY, A/cm"
,o

1.0
"N,,

"E
_

9 \\, "~,, "\,

COFLOW - ....... COUNTERFLOW

~,

. 6
4

o 0.8

~
~:

>~ 0.6 t~ Q ~" 0.4 II:

0
0 2 4 6 8 10
cm

10
cm

DISTANCE FROM OXIDANT INLET,

DISTANCE FROM OXIDANT INLET,

Figure 10.14. Simulated current and temperature distributions in SOFCs having crossflow, coflow, and counterflow configurations [10.35]

328

Chapter lO

WITHOUT INTERNAL REFORMING (90% H2, 10% H20) 91o~

WITH INTERNAL REFORMING (25% CH4, 75% H20) ~Jo. 9671


LU

i
t1"1._:3 94~] '~il IX."rt" 977] '

,,,:~ 907]

I,--

887J

.................

Figure 10.15. Temperature distributions in SOFC operated with hydrogen (no reforming) and with methane (internal reforming) [10.211

SOFC systems can be designed to influence current and temperature distributions within the stack, thus maximum current output. For example, a sealless tubular SOFC system with conventional air cooling would produce an average current density of 370 mA/cm 2 at 0.606 V with 80 % fuel utilization and 25 % air utilization [10.36]. Fuel recycling modifies the current and temperature distributions of the fuel cell. Thus, a fuel recycling system would produce 0.926 mA/cm z at 0.811 V with 10 % fuel utilization per cycle and 100 % air utilization. As a result, power generation density is improved.

References
10.1 J.R. Selman, in Proceedings of the International Symposium on Solid Oxide Fuel Cells, November 13 - 14, 1989, Nagoya, Japan, O. Yamamoto, M. Dokiya, and H. Tagawa (eds.), Science House, Tokyo, Japan, 1989, p. 212. S. Majumdar, T. Claar, and B. Flandermeyer, J. Am. Ceram. Soc., 69 (1986) 628. Westinghouse Electric Corporation, High-Temperature Solid Oxide Electrolyte Fuel Cell Power Generation System, Quarterly Summary Report, January 1, 1984March 31, 1984, Report No. DOE/ET/17089--2217, U.S. Department of Energy, Washington, DC, 1984. A. Saigal and S. Majumdar, presented at 1992 ANSYS Technology Conference, May 4-7, 1992, Pittsburgh, PA, CONF-9205129--1, U.S. Department of Energy, Washington, DC, 1992. E.F. Sverdrup. C.J. Warde, and R.L. Eback, Energy Convers., 13 (1973) 129.

10.2 10.3

10.4

10.5

Modeling and Analysis

329

10.6 10.7

10.8 10.9

10.10

10.11 10.12 10.13 10.14 10.15

10.16 10.17 10.18 10.19

10.20 10.21

10.22 10.23 10.24 10.25 10.26

S. Nagata, K. Shirai, and H. Sato, Bull. Electrotech. Lab., 40 (1976) 974. S. Nagata, Y. Ohno, Y. Kasuga, Y. Kaga, and H. Sato, in Proceedings of the 19th IECEC, August 19-24, 1984, San Francisco, CA, Vol. 2, American Nuclear Society, La Grange Park, IL, 1984, p. 827. Y. Ohno, Y. Kaga, and S. Nagata, Electr. Eng. Jpn., 107 (1987) 59 (translated from Denki Gakkai Ronbunshi, 106B (1986) 693). U.G. Bossel, Performance Potentials of Solid Oxide Fuel Cell Configurations, Report No. EPRI TR-101109, Electric Power Research Institute, Palo Alto, CA, 1992. U.G. Bossel, in Proceedings of the Third International Symposium on Solid Oxide Fuel Cells, May 16-21, 1993, Honolulu, HI, S.C. Singhal and H. Iwahara (eds.), Electrochemical Society, Pennington, NJ, 1993, p. 833. T. Kudo and H. Obayashi, Energy Convers., 15 (1976) 121. P.G. Debenedetti and C.G. Vayenas, Chem. Eng. Sci., 38 (1983) 1817. P.G. Debenedetti, C.G. Vayenas, I. Yentekakis, and L.L. Hegedus, in American Chemical Society Symposium Series, 237 (1984) 171. W.J. Wepfer and M.H. Woolsey, Energy Convers., 24 (1985) 477. W.R. Dunbar and R.A. Gaggioli, in Proceedings of the 23rd IECEC, July 31-August 5, 1988, Denver, CO, American Society of Mechanical Engineers, New York, 1988, p.257. W.J. Marner, J.W. Suitor, and C.R. Glazer, see Ref. 10.15, p. 265. C.Y. Lu and T.M. Maloney, in 1988 Fuel Cell Seminar Abstracts, October 23-26, 1988, Long Beach, CA, Courtesy Associates, Washington, DC, 1988, p. 78. T.M. Maloney and G.A. Coulman, in 1990 Fuel Cell Seminar Abstracts, November 25-28, 1990, Phoenix, AZ, Courtesy Associates, Washington, DC, 1990, p. 239. S. Ahmed, R. Kumar, H. Walden, and K.P. Barr, in Proceedings of the 26th IECEC, August 4-9, 1991, Boston, MA, Vol. 3, American Nuclear Society, La Grange Park, IL, 1991, p. 594. S. Ahmed, C.C. McPheeters, and R. Kumar, J. Electrochem. Soc., 138 (1991) 2712. E. Erdle, J. GroB, H.G. Miiller, W.J.C. Miiller, H.-J. Reusch, and R. Sonnenschein, in Proceedings of the Second International Symposium on Solid Oxide Fuel Cells, July 2-5, 1991, Athens, Greece, F. Grosz, P. Zegers, S.C. Singhal, and O. Yamamoto (eds.), Commission of The European Communities, Luxembourg, 1991, p. 265. E. Arato and P. Costa, see Ref. 10.21, p. 273. I.V. Yentekakis, S. Neophytides, S. Seimanides, and C.G. Vayenas, see Ref. 10.21, p. 281. D.G. Walhood and J.R. Ferguson, see Ref. 10.21, p. 289. A. Solheim, R. Tunold, and R. OdegArd, see Ref. 10.21, p. 297. J.R. Ferguson, see Ref. 10.21, p. 305.

330

Chapter 10

10.27 10.28

10.29 10.30 10.31 10.32 10.33 10.34 10.35 10.36

Z. Takehara, K. Kanamura, A. Hirano, and M. Ippommatsu, see Ref. 10.21, p. 313. N.F. Bessette II and W.J. Wepfer, in 1992 Fuel Cell Seminar Abstracts, November 29-December 2, 1992, Tucson, AZ, Courtesy Associates, Washington, DC, 1992, p. 519. J. Hartvigsen, S. Elangovan, and A. Khandkar, see Ref. 10.28, p. 532. A. Solheim, see Ref. 10.10, p. 841. T. Sira and M. Ostenstad, see Ref. 10.10, p. 851. C. Bleise, J. Divisek, B. Steffen, U. K6nig, and J.W. Schultze, see Ref. 10.10, p. 861. H. Karoliussen, K. Nisancioglu, A. Solheim, and R. Odeghrd, see Ref. 10.10, p. 868. J. Hartvigsen, S. Elangovan, and A. Khandkar, see Ref. 10.10, p. 878. A. Malandrino and M. Chindemi, see Ref. 10.10, p. 885. A. Hirano, M. Suzuki, and M. Ippommatsu, J. Electrochem. Soc., 139 (1992)2744.

Anda mungkin juga menyukai