You are on page 1of 9

Electrochimica Acta 223 (2017) 100108

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

An Analysis of Contact Problems in Solid Oxide Fuel Cell Stacks Arising


from Differences in Thermal Expansion Coefcients
Ludger Blum*
Forschungszentrum Jlich GmbH, Institute of Energy and Climate Research, IEK-3, 52425 Jlich, Germany

A R T I C L E I N F O

Article history:
Received 22 August 2016
Received in revised form 4 December 2016
Accepted 5 December 2016
Available online 5 December 2016
Keywords:
Solid Oxide Fuel Cell
stack
contact problems
cathode contact
thermal expansion coefcient
performance

A B S T R A C T

The successful operation of solid oxide fuel cells (SOFC) imposes high demands on the similarity of the
thermal expansion coefcients of the materials used. This is not only due to thermo-mechanical factors
but also because of the risk of contact loss via the formation of a micro-gap between the electrodes and
adjacent contact layers caused by temperature changes. The origin of the formation of such a gap
between the different layers in an SOFC stack is investigated on the basis of various material
combinations. A comparison with successful and failed stack test results reveals that there is a high
probability of contact problems if the calculated gap on the cathode side exceeds 200 nm. This limit
relates to the maximum possible elastic deformation of the combination of the cathode and cathode
contact layer (CCL). Such contact problems can be avoided in the current design by utilizing an
appropriate combination of the cathode, CCL and glass-ceramic material that can be selected on the basis
of the methodology outlined in the current work. This optimization in materials selection will help to
establish a stack technology that allows high power density with a minimized risk of failure of single
layers.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
High temperature fuel cell stacks must meet the critical
requirement that all layers (all parts of a single plane or repeating
unit, like anode, electrolyte, cathode, contact layers, protective
coating, interconnect) have to retain good electrical contact,
although large temperature changes occur between sealing
temperatures (in this case, using a rigid glass sealing) or assembly
temperatures (e.g., in the case of using a mica sealing) and
operating temperatures (which can span a relatively wide range).
In an ideal case, all materials would have the same thermal
expansion coefcient (TEC) and no problem would arise. But in real
congurations, differences will emerge that can cause thermomechanical stress, which is investigated in numerous publications
[19]. A further problem can occur because of the differences in the
TECs of the different materials which result to a different change in
thickness of the various layers. While changing temperature from
the starting point (which can be the sealing or assembly
temperature, depending on the sealing concept used) to the
operating point, this different changes in thickness can lead to
contact loss at the mechanically weakest point which, in most

* Corresponding author. Tel.: +49 2461 61 6709, fax: +49 2461 61 6695.
E-mail address: l.blum@fz-juelich.de (L. Blum).
http://dx.doi.org/10.1016/j.electacta.2016.12.016
0013-4686/ 2016 Elsevier Ltd. All rights reserved.

cases, due to the limited reaction between these sintered layers,


will be between the electrodes and contact layer or interconnect
[10]. The following analysis is based on the standard stack design
used by the Forschungszentrum Jlich, the so called F-Design
[11,12], in which a rigid glass-ceramic sealing is used [13]. Here, at
certain operating conditions, unexpected voltage-current characteristics are observed (see Fig. 1): With increasing current density,
the cell voltage rst drops fast, then becomes stable and nally can
even increase. It is not exactly the same in all repeating units and
this could not be explained until now. But there is always an
inuence of temperature distribution in the stack and of
manufacturing tolerances between the single repeating units.
However, a clear tendency is visible. Because of the form these
curvatures take, we refer to this as a bathtub-like characteristic.
The characteristic can behave in such a way if the resistance (of the
cell or contacts) is reduced while current density increases. During
the measurement, the temperature inside the stack increases (see
the dotted lines in Fig. 1). This also has an effect on the resistances
of the various components, but this effect is small, as can be seen
by the behavior of the layers showing no bathtub. Another issue is
the different thermal expansion coefcients that can create a gap
between the different layers while cooling from sealing temperature (850-950  C) down to operating temperature (700-800  C).
This gap is reduced again during measurement, when the current

L. Blum / Electrochimica Acta 223 (2017) 100108

Nomenclature

ASC
BSCF
CCL
CGO
LC
LCC
LCC10/12
LSCF
LSM
MMA
NCO
SOFC
TEC
YSZ

atec

dL0
d(T)
dc
Ddel

DdL
Ddr
E

e
emax
s
s max
T
Tmax

anode supported cell


Barium Strontium Cobalt Ferrite
cathode contact layer
Ceria Gadolinium Oxide
LaCoO3
La0.8Ca0.2CoO3
perovskite materials for contact layer
Lanthanum Strontium Cobalt Ferrite
Lanthanum Strontium Manganese
MagnesiaMagnesiumAluminate
Nickel Cobalt Oxide
solid oxide fuel cell
thermal expansion coefcient
yttria stabilized zirconia
technical thermal expansion coefcient
thickness of the layers at Tmax
thickness of the layer at temperature T
thickness of the CCL plus cathode
change in layer thickness caused by elastic
deformation
change in layer thickness caused by temperature
change
change in layer thickness until rupture
elastic modulus
elastic strain
deformation until rupture
stress
fracture strength
operation temperature
50 K below the sealing temperature

density is increased and the stack becomes hotter. From Fig. 1 it can
also be seen that the problem is quite complex, because not all
layers show the same behavior. This is affected by various
properties like variations in materials, the local temperature and
mechanical strength of the interfaces and the local dimensions
inuenced by manufacturing tolerances.
In the present work, as a rst approach the bathtub behavior is
analyzed concentrating on the differences in the thermal

101

expansion coefcient in the case of different materials for cathode,


cathode contact layers, anode contact layer and sealing material.
For the cathode side the strength of the connection of the layers is
analyzed to get a feeling for the impact of thermal expansion
differences on contact loss and thus on this bathtub phenomenon.
2. Approach
2.1. Test Items and Conditions
The rst time that this bathtub effect was observed was during
operation with natural gas and internal reforming. In particular,
during the current-voltage characteristic measurement this causes
a large temperature drop in the stack, especially at low current
densities, due to the pronounced effect of cooling from the
endothermal reforming while operating with a constant fuel ow.
Later, this phenomenon also occurred while operating with
hydrogen. The observation was that this tends to take place when
using LSCF (Lanthanum Strontium Cobalt Ferrite) cathodes instead
of LSM (Lanthanum Strontium Manganese). Therefore, the effect of
the differing thermal expansion coefcients of these materials on
the assembly was analyzed.
All tests analyzed utilized the standard F-design [11,12], which
can be F10 (with cells in 10  10 cm2 dimension) or F20 (with cells
in a 20  20 cm2 conguration). In Fig. 2 a simplied assembly
sketch of the F-design stacks is presented. The dimensions given
represent the status at Tmax which is 50 K below the sealing
temperature. The anode supported cell (ASC) is attached to the
interconnect on the cathode side with the help of a glass-ceramic
layer (200 mm) between the electrolyte at the outer rim of the cell
and the metallic interconnect. The thickness of this glass layer is
pre-determined by the nominal thickness of the cathode (50 mm)
and cathode contact layer (CCL) (150 mm), which may also include
the protective coating against chromium evaporation at the metal
interconnect. This means that a gap that might be created on the
cathode side depends on the differences in thermal expansion
between the glass-ceramic and the cathode with the CCL.
On the anode side, the inner assembly is composed of a glassceramic sealing of 200 mm (the same as on the cathode side),
anode substrate (warm pressed substrate of 1500 mm, including
the electrolyte) and nickel mesh (1400 mm) which, simultaneous
to the electrical contact, provides the gas channels. The outer
assembly is composed of two layers of glass-ceramic sealing
(275 mm each) and of a metal frame made of Crofer 22 APU

Fig. 1. Current-voltage characteristics of 5-plane stack F2005-05 and 10-plane stack F2010-06, with several planes showing bathtub behavior.

102

L. Blum / Electrochimica Acta 223 (2017) 100108

Fig. 2. Simplied assembly sketch of F-design stacks with a thickness of dL0 of the layers at Tmax.

(2500 mm). The differences in thermal expansion between the


inner and outer assembly are responsible for the gap formation on
the anode side. A smaller thickness of the anode substrate will be
compensated for by a smaller thickness of the outer metal frame.
As the two materials (of anode substrate and frame) do not differ a

great deal with respect to their thermal expansion coefcients, the


following investigation is also valid when thinner substrates are
used.
The essential physical property is the thermal expansion
coefcient of the materials utilized. Based on data in the literature

Fig. 3. Thermal expansion coefcient of components relevant for the cathode side and of Crofer 22 APU for reference [1424].

L. Blum / Electrochimica Acta 223 (2017) 100108

103

Fig. 4. Thermal expansion coefcient of components relevant for the anode side [14,15,24,25].

and our own measurements, the TEC as a function of temperature


for the different materials is given in Figs. 3 and 4. On the cathode
side (see Fig. 3), these materials are the LSM electrode [14,15], the
LSCF electrode [16], the contact layers comprised of perovskite
materials, called LCC10 and LCC12 [17,18], and the glass-ceramic
sealings B (sealing temperature 950  C for 10 hours), G (sealing
temperature 850  C for 10 hours), H with the addition of 20% YSZ
bers (sealing temperature 850  C for 10 hours (with a larger
amount of remaining amorphous phase) and 850  C for 100 hours
(higher crystallization status)) [1923].
On the anode side (see Fig. 4), these materials are, in addition to
the glass-ceramic sealing, the anode substrate [14,15], steel Crofer
22 APU [24] and nickel mesh [25].

The technical thermal expansion coefcient atec is the linear


thermal expansion coefcient of the material accounting for the
variation of it over temperature within the specied temperature
range (between 25C and T). This makes it relevant for the elastic
analysis conducted later on in the paper. atec is dened as:

atec

dT  d25 C
=T  25 C
d25 C

with T temperature in  C
d(T) thickness of the layer at temperature T
d(25  C) thickness of the layer at 25  C
The following assumptions have been made:

Table 1
Calculation sheet to obtain the value of the gap: here as example for the gap between the cathode and cathode contact layer (CCL).
[1]
Thickness of
layer b):

[2]
T

atec

mm

between
25  C and T
106 K1

200

900
860
800
750
700
650
600
500
400
300
200
25

12.33
12.22
12.11
12.03
11.96
11.90
11.85
11.76
11.73
11.60
10.99

[3]
Change of thickness of layer b)
between T and 25  C

[4]
Reduction of thickness of layer b)
compared to Tmax

[5] [6]
! Change of the thickness of a)
compared to Tmax

mm

mm

mm

[7]
a) b) >0 ! gap
cathode
- CCL
mm

2.157
2.041
1.877
1.745
1.615
1.487
1.361
1.118
0.880
0.638
0.385
0

0.000
0.116
0.280
0.412
0.542
0.670
0.796
1.039
1.277
1.519
1.772
2.157

0.000
0.124
0.296
0.432
0.566
0.700
0.837
1.116
1.399
1.669
1.909
2.203

0.000
0.008
0.016
0.019
0.023
0.030
0.041
0.077
0.122
0.150
0.137
0.046

104

L. Blum / Electrochimica Acta 223 (2017) 100108

 The sealing is assumed to be rigid at Tmax, which is 50 K below


the sealing temperature ! there, the gap is zero and the stress
amongst the stack components is also zero.
 The temperature range investigated more in detail is between
600 and 800  C (this is the operation window, where contact loss
is most critical for the performance of the stack). The effect of
thermal cycling between RT and operating temperature will be
investigated separately.
 The gap on the cathode side is not inuenced by a possible gap on
the anode side, because it is assumed that the anode substrate
does not bend due to the substrates stiffness.
 The gap is calculated under the assumption that there is no
bonding between the cathode and the contact layer respectively
between the anode and the nickel mesh. Both interfaces are
dened as the weak positions based on experimental observation [10] and thus gap formation is most likely located there.
 There is no elastic deformation or uneven contact due to
roughness or misalignment of the cells, and overall the layers are
assumed to be innitely stiff and do not deform due to
mechanical loading (the mechanical properties are investigated
in chapter 4).

2.2. Process of calculation


In a spreadsheet calculation (see Table 1), the differences of the
various material combinations are calculated according to the
following schematic (the single steps are also indicated in Table 1
as steps [1] to [7]):
[1] The thickness dL0 of the single layer at Tmax is dened (for
values, see Fig. 2).
[2] Denition of the thermal expansion coefcient as a function
of temperature (according to Figs. 3 and 4).
[3] Calculation of the change in layer thickness DdL between T
and 25  C:

DdL T d0L  atec T  T  25 C

[4] Calculation of the change in layer thickness as a function of


temperature compared to Tmax:

DdL T max  T DdL T max  DdL T

[5] Calculation of the change of thickness for the other relevant


layers (according to Fig. 2) as a function of temperature, also with
steps [14].

Fig. 5. Formation of gap on the cathode side in the case of different glass-ceramic types and sealing temperatures using an LSM cathode and LCC10 as the contact layer: a) total
temperature range; b) in a 600-900  C temperature range.

L. Blum / Electrochimica Acta 223 (2017) 100108

[6] Calculation of the resulting change in the thickness of all


layers on top of each other (in Table 1 this is shown for the cathode
side):
Cathode side: a) cathode + CCL
b) glass layer
Anode side: c) glass layer + anode substrate + Ni-mesh
d) glass layer + metal frame + glass layer
[7] Calculation of the difference in thickness between a) and b)
(respectively, between c) and d)) which results to a gap between
cathode and CCL (respectively, between anode and Ni-mesh):
if a) b) (respectively, c) d)) is positive, a gap is formed
between the cathode and CCL (respectively, between the anode
and Ni-mesh) during cooling down from the reference temperature Tmax to temperature T.
3. Results
At rst, the standard material combination, using LSM as the
cathode material and LCC10 as the cathode contact layer, was
investigated.
As can be seen from Fig. 5, there is a big difference in the
resulting gap size depending on the sealing material. At the
beginning, glass-ceramic G was used. This glass-ceramic type was
relatively well adapted to the thermal expansion of the interconnect material and the LSM cathode. Therefore, the gap on the
cathode side is less than 0.2 mm at 600  C (see Fig. 5b). However,

105

because the mechanical strength was not sufcient, another glassceramic type was developed: glass-ceramic H with the addition of
20%YSZ bers. If this glass-ceramic is sealed for 10 hours at 850  C,
a relatively high amount of amorphous glass phase remains. This
can be seen in the curvature of the thermal expansion coefcient of
between 600 and 700  C. In the operating temperature range, the
resulting gap is relatively small but increases down to room
temperature and becomes similar to glass-ceramic G (see Fig. 5a).
However, because of the high amorphous content, the mechanical
properties are still not good enough and, in addition the glassceramic changes its behavior during operation by increasing the
content of its crystal phase which changes the thermal expansion
behavior again. To improve both, the sealing time at 850  C was
increased to 100 hours, resulting in higher mechanical strength
[26]. However, the thermal expansion coefcient goes down,
resulting in a somewhat bigger gap. Another very promising glassceramic material development was glass-ceramic B, which has a
sealing temperature of 950  C and a much better adapted thermal
expansion coefcient, resulting in the smallest gap size. However,
the experiments revealed that this glass-ceramic has too high a
viscosity during the sealing process, so it cannot be used to seal the
cell. Therefore, it was combined with glass-ceramic H, which
resulted in good thermomechanical stability and gastight stacks
[13]. However, the processing of glass-ceramic H at 950  C results
in a much larger gap, especially in the operating temperature range
(see Fig. 5b).

Fig. 6. Formation of the gap on the cathode side in the case of different glass-ceramic types and sealing temperatures using an LSCF cathode in a 600-900  C temperature
range: a) with LCC10 as the contact layer, b) with LCC12 as the contact layer.

106

L. Blum / Electrochimica Acta 223 (2017) 100108

Fig. 7. Formation of a gap on the anode side in the case of different glass-ceramic types and sealing temperatures.

Because of the much improved electrochemical behavior of the


LSCF cathode, stacks were built using the various glass-ceramic
materials with LCC10 as a contact layer. Because of the higher
thermal expansion coefcient of LSCF compared to LSM, the
resulting gap increased by about 0.1 mm at 600  C (see Fig. 6a). In
the meantime, another CCL material was developed with better
electric conductivity, called LCC12. However, this material also has
a higher thermal expansion coefcient, which results in a further
increase in the gap, of about 0.05 mm at 600  C (see Fig. 6b).
The same considerations also apply on the anode side. In that
case, three different glass-ceramics are considered (see Fig. 7). The
gap between anode substrate and Ni-mesh, which is created
during cooling, is much bigger than on the cathode side. This gap
does not depend so much on the glass-ceramic material for the
thick Ni-mesh is dominating due to its high thermal expansion
coefcient (see Fig. 3). While changing from nickel to a ferritic
stainless steel such as Crofer 22 APU, the gap size is much reduced.
On the contrary, over a wide temperature range (depending on the
glass-ceramic type) a negative gap is formed, meaning that the
anode contact is under compression.
A comparison of numerous stack tests (54 in total) performed at
the Forschungszentrum Jlich utilizing the F10 and F20-type
stacks led to the result presented in Fig. 8. On this basis, the
following statement can be derived: When the calculated gap on
the cathode side is smaller than 100 nm, there is a marginal risk for
delamination, between 100 and 200 nm in some cases bathtub-

behavior can be observed, and in calculated gaps wider than


200 nm, problems very often arise.
4. Discussion
As can be seen in Fig. 7, the risk of delamination because of gap
formation seems to be far more critical on the anode side. However,
because of the interdiffusion between the Ni-mesh and Ni-anode
and the meshs high degree of exibility, the delamination effect is
insignicant. As also shown in Fig. 7, the difference in thermal
expansion can be very much reduced by replacing the Ni-mesh
with a ferritic chromium steel mesh. In that case though, the
interaction of the mesh and anode must be investigated with
respect to whether a good bond is created and whether e.g., the
electronic conductivity of the anode substrate is inuenced by
nickel diffusion into the steel.
Whether the difference in thermal expansion can really cause a
gap between cathode and CCL depends on the strength of the
connection between cathode and CCL. To investigate this, rstly
the elastic deformation is checked and secondly the quality of the
connection between cathode and CCL in case of further deformation.
On the cathode side, sintering occurs between CCL and the
cathode during sealing and subsequently during operation. Both
layers are highly porous and mechanically not very robust. Lipinska
at al. [27], Pecanac et al. [28] and Atkinson et al. [29] measured the
mechanical properties of porous ceramic layers based on BSCF
(Barium Strontium Cobalt Ferrite), LSCF, CGO (Ceria Gadolinium
Oxide) and MMA (MagnesiaMagnesiumAluminate) mainly at
RT, in a few cases at 800  C (see Table 2). For comparison, the values
of dense CCL material (LCC = La0.8Ca0.2CoO3, LC = LaCoO3), measured by Orlovskaya et al. [30], are also given. Based on these data,
the elastic deformation Ddel of the given porous ceramic layers
under compression (during the sealing process, when the glass is
soft and so all the load is on the contact layer) in F10 and F20
designs can be calculated by utilizing:

Ddel dc  e dc 

Fig. 8. Bathtub behavior as a function of gap height on the cathode side.

s
E

where dc is the thickness of the CCL plus cathode, which in our case
is 200 mm (see Fig. 2).
The compression stress s can be calculated based on the
compression force during the sealing process, while the glassceramic sealing is still soft and the contact area on the cathode side,
which is based on the channel structure, makes up about 40% of the

L. Blum / Electrochimica Acta 223 (2017) 100108

107

Table 2
Mechanical properties of porous ceramic layers and resulting deformation during sealing, taking the E-modulus at RT and at 800  C.
Material

BSCF

Porosity

Sintering

E-Modulusa
ERT E800

Fracture Strengthb

GPa

MPa
800  C

RT

800  C

24

31
34

50

1200 C/3h
1000  C/10h

35
33

LSCF

46%

1200 C/5h

12

MMA

24,5%

1500  C

70

LCC
LC

dense
dense

CGO

43%

a
b

18
54

30

Lit.

Deformation during sealing


Ddel
F10
nm

RT
34%
38%

1050  C/4h

s max,RT s max,800

26

F20
nm

F10
nm

F20
nm

800  C

RT
[28]
[27]

1.8
1.9

2.6
2.7

[28]

5.1

7.5

[29]

0.9

1.3

141
76

140
70

[30]
[30]

0.4
0.8

0,6
1.2

57

55

[28]

1.1

1.6

2.6

3.8

1.1

1.7

Mean values.
Loading rate 100 N/min (T-change within 30 to 60 min).

cathode area. The resulting values for the F10 and F20 designs are
outlined in Table 3.
Based on these values, the possible elastic deformation during
sealing is calculated. The results are given in Table 2. The maximum
deformation result was 7.5 nm for porous LSCF with mechanical
properties for 25  C. Assuming the same temperature dependency
as reported for BSCF [28], the resulting deformation at 800  C
would be 10.4 nm. This value is at least one order of magnitude
below the deformation of the ceramic layers needed to avoid gap
formation due to the differences in thermal expansion while
relaxing compression during cooling to operating temperature.
This means the connection between CCL and the cathode
(respectively the layers themselves) will be subject to tensile
stress during this cooling process. The deformation until rupture
Ddr can be calculated by

Ddr dc  emax dc 

s max

This fracture strength could not be measured till now though.


Therefore, the calculated values for the deformation until rupture
represent only an upper limit. The following considerations can
therefore only provide a rough estimation for determining an
upper threshold because the weak point will be the interface
between the different porous ceramic layers, which would
probably have lower fracture strength than the sintered ceramic
layers themselves.
Considering the materials, with their fracture strength and
elastic modulus outlined in Table 2, deformation under tensile
conditions is calculated, taking the mechanical properties for RT
and for 800  C (as far as available). The results are shown in Table 4.
There are only a few values at increased temperature available. At
800  C, they vary from 96 to 421 nm. The lower values result from
the MMA material, which was sintered at much higher temperatures than the CCL used in the F-design stacks. The fracture
strength of layers comparable to CCL is difcult to measure, as is
also reported by Lu et al. [31]. For a NiCo-oxide (NCO) powder
sintered between LSM, they measured a fracture strength of

6.8 MPa. Till now though, no E-Modulus of such material could be


determined. However, the resulting deformation until rupture of
these various porous ceramic layers t quite well to the
observation that above a gap formation of 200 nm under operation
conditions, there is a high risk of delamination of CCL from the
cathode or steel surface. And already at about 100 nm the risk of
delamination is visible. In real stacks, there will be some scattering
because the operating parameters are not really identical in all
cases because of deviations in furnace temperature, temperature
gradients within the stack and tolerances in the thicknesses of the
various layers resulting from the manufacturing processes.
5. Conclusions
This paper investigated the tendency of a gap formed between
the various layers of an SOFC stack with glass-ceramic sealing
while cooling from sealing temperature to operating temperature
as a result of the differing thermal expansion coefcients. This can
cause delamination which results in a loss in power output.
Various material combinations were analyzed in an effort to
illuminate the nature of this issue. Based on the Jlich F-Design
stacks a comparison of stack test results revealed that there is a
high probability of contact problems in the case of a theoretical gap
on the cathode side above 200 nm. This value can be explained by
the maximal possible deformation before rupture of the combination of the cathode and CCL in that range. Mechanical properties of
porous ceramic layers indicate that there is a high probability for
delamination in that range of deformation. These contact problems
Table 4
Resulting elastic deformation (relative and absolute) for different materials under
tensile conditions until rupture in case of mechanical properties at RT and at 800  C.
Material

Deformation until rupture

emax

Dd_r
(related to
200 mm)

%
Table 3
Compression stress s during the sealing process in case of design F10 and F20.
Loading

Stress s

Area

F10
N

F20
N

F10
mm2

F20
mm2

F10
MPa

F20
MPa

1000

6500

3240

14440

0.309

0.450

RT
BSCF
LSCF
MMA
LCC
LC
CGO

0.089
0.103
0.150
0.043
0.099
0.092
0.096

nm
800  C
0.210
0.048

RT
177
206
300
86
199
184
193

800  C
421
96

108

L. Blum / Electrochimica Acta 223 (2017) 100108

can be minimized by using the right combination of cathode,


cathode contact layer, and glass-ceramic material which can be
selected using the method presented in this paper. This optimization in materials selection will help to establish a stack technology
that allows high power density with a minimized risk of failure of
single layers.
Acknowledgment
The author would like to express his gratitude to Dr. Sonja GroBarsnick and Mr. Andreas Koppitz for providing the thermal
expansion coefcients for the various glasses, Dr. Frank Tietz for
information about the properties of the contact layers and
cathodes and Dr. Jrgen Malzbender for providing data on the
mechanical properties of the various materials and for discussing
the mechanical aspects. The author also would like to thank all
colleagues at Jlich who were responsible for materials development, design, manufacturing and stack assembling and testing.
This research did not receive any specic grant from funding
agencies in the public, commercial, or not-for-prot sectors.
References
[1] P. Singh, L.R. Pederson, S.P. Simner, J.W. Stevenson, V.V. Viswanathan, Solid
oxide fuel cell power generation systems, GA; United States, Proc. of the 36th
Intersociety Energy Conversion Engineering Conference Savannah, vol. 22001,
pp. 953958.
[2] K.S. Weil, J.S. Hardy, Development of a compliant seal for use in planar solid
oxide fuel cells, Ceramic Engineering and Science Proceedings 25 (3) (2004)
321326.
[3] J. Malzbender, T. Wakui, R.W. Steinbrech, Curvature of Planar Solid Oxide Fuel
Cells during Sealing and Cooling of Stacks, Fuel Cells 06 (02) (2006) 123129.
[4] J. Malzbender, J. Mnch, R.W. Steinbrech, T. Koppitz, S.M. Gross, J. Remmel,
Symmetric shear test of glass-ceramic sealants at SOFC operation
temperature, J. Mater. Sci. 42 (2007) 62976301.
[5] W.D. Callister, D.G. Rethwisch, Fundamentals of Materials Science and
Engineering: An Integrated Approach, 3rd ed. (2008), Sections 7.3, 17.3, 17.5.
[6] J. Malzbender, Curvature and stresses for bi-layer functional ceramic materials,
J. European Ceramic Society 30 (2010) 34073413.
[7] J. Malzbender, R.W. Steinbrech, L. Singheiser, A review of advanced techniques
for characterising SOFC behavior, Fuel Cells 09 (06) (2009) 785793.
[8] K. Fischer, J.R. Seume, Thermo-mechanical stress in tubular solid oxide fuel
cells: Part II Operating strategy for reduced probability of fracture failure, IET
Renewable Power Generation 6 (3) (2012) 194205.
[9] A. Al-Masri, M. Peksen, L. Blum, D. Stolten, Optimization of heating-up process
in SOFC stacks to reduce thermomechanically induced stresses by means of
simulation tools, EFC 2013 Proceedings of the 5th European Fuel Cell Piero
Lunghi Conference, Rome Italy, 2013, pp. 419420.
[10] N.H. Menzler, P. Batfalsky, S.M. Gro, V. Shemet, F. Tietz, Post-Test
Characterization of an SOFC Short Stack after 17,000 hours of Steady
Operation, ECS Transaction 35 (2011) 195206.

[11] R. Steinberger-Wilckens, I.C. Vinke, L. Blum, J. Remmel, F. Tietz, W.J.


Quadakkers, Progress in SOFC stack development at Forschungszentrum
Jlich, Proc. of the 6th European SOFC Forum, Luzern/CH, 2004, pp. 1119.
[12] L. Blum, L.G.J. de Haart, J. Malzbender, N.H. Menzler, J. Remmel, R. SteinbergerWilckens, Recent results in Jlich solid oxide fuel cell technology
development, J. Power Sources 241 (2013) 477485.
[13] L. Blum, S.M. Gro, J. Malzbender, U. Pabst, M. Peksen, R. Peters, I.C. Vinke,
Investigation of solid oxide fuel cell sealing behavior under stack relevant
conditions at Forschungszentrum Jlich, J. Power Sources 196 (2011) 7175
7181.
[14] F. Tietz, Thermal Expansion of SOFC Materials, Solid State Ionics 5 (1999) 129
139.
[15] F. Tietz, Peculiarities in the thermal expansion behavior of ceramic fuel cell
materials, P. Vincenzini (Ed.), Proc. of the 9th Cimtec-World Forum on New
Materials, Symposium VII (1999) 6170.
[16] A. Petric, P. Huang, F. Tietz, Evaluation of La-Sr-Co-Fe-O perovskites for solid
oxide fuel cells and gas separation membranes, Solid State Ionics 135 (2000)
719725.
[17] F. Tietz, A. Schmidt, M. Zahid, Investigation of the quasi-ternary system
LaMnO3-LaCoO3-LaCuO3 I: the series La(Mn0.5Co0,5)1-xCuxO3-d, J. Solid State
Chemistry 177 (2004) 745751.
[18] F. Tietz, I.A. Raj, Q.X. Fu, M. Zahid, Investigation of the quasi-ternary system
LaMnO3-LaCoO3-LaCuO3 II: the series LaMn0.25-xCo0.75-xCu2xO3-d and
LaMn0.75-xCo0.25-xCu2xO3-d, J. Mater. Sci. 44 (2009) 48834891.
[19] S.M. Gross, T. Koppitz, J. Remmel, J.-B. Bouche, U. Reisgen, Joining properties of
a composite glass-ceramic sealant, Fuel Cells Bulletin (2006) 1215.
[20] J. Malzbender, T. Koppitz, S.M. Gross, R.W. Steinbrech, Studies on the thermal
expansion of glass-ceramic sealants, Proc. of the 7th European SOFC Forum
(2006) P0418.
[21] S.M. Gross, U. Reisgen, Glaslten -eine anspruchsvolle Fgetechnik nicht nur
fr die Hochtemperatur-Brennstoffzelle, Proc. Schweien und Schneiden 59
(2007) 7077.
[22] S.M. Gross, D. Federmann, J. Remmel, M. Pap, Reinforced composite sealants
for solid oxide fuel cell applications, J. Power Sources 196 (2011) 73387342.
[23] B. Cela, S. Sillapawatana, S.M. Gross, T. Koppitz, R. Conradt, Inuence of ller
additives on the effective viscosity of glass-ceramic composite sealants, J.
University of Chemical Technology and Metallurgy 47 (4) (2012) 449458.
[24] VDM Metals GmbH, Crofer 22 APU, Material Data Sheet No. 4046 May 2010
Edition.
[25] VDM Metals GmbH, Nickel 99.2alloy 200, Material Data Sheet No. 1001
October 2002 Edition.
[26] J. Malzbender, Y. Zhao, T. Beck, Fracture and creep of glass-ceramic solid oxide
fuel cell sealant materials, J. Power Sources 246 (2014) 574580.
[27] M. Lipinska-Chwalek, J. Malzbender, A. Chanda, S. Baumann, R.W. Steinbrech,
Mechanical characterization of porous Ba0.5Sr0.5Co0.8Fe0.2O3-d, J. European
Ceramic Society 31 (2011) 29973002.
[28] G. Pecanac, S. Foghmoes, M. Lipinska-Chwaek, S. Baumann, T. Beck, J.
Malzbender, Strength degradation and failure limits of dense and porous
ceramic membrane materials, J. European Ceramic Society 33 (2013) 2689
2698.
[29] A. Atkinson, P. Bastid, Q. Liu, Mechanical Properties of MagnesiaSpinel
Composites, J. Am. Ceram. Soc. 90 (8) (2007) 24892496.
[30] N. Orlovskaya, M. Lugovy, S. Pathak, D. Steinmetz, J. Lloyd, L. Fegely, M. Radovic,
E.A. Payzant, E. Lara-Curzio, L.F. Allard, J. Kuebler, Thermal and mechanical
properties of LaCoO3 and La0.8Ca0.2CoO3 perovskites, J. Power Sources 182
(2008) 230239.
[31] Z. Lu, G. Xia, J.D. Templeton, X. Li, Z. Nie, Z. Yang, J.W. Stevenson, Development
of Ni1-xCoxO as the cathode/interconnect contact for solid oxide fuel cells,
Electrochemistry Communications 13 (2011) 642645.

You might also like