Next Article in Journal
A Multi-Agent System-Based Approach for Optimal Operation of Building Microgrids with Rooftop Greenhouse
Previous Article in Journal
A Methodology Based on Cyclostationary Analysis for Fault Detection of Hydraulic Axial Piston Pumps
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Geometry Effect of Cathode/Anode Areas Ratio on Electrochemical Performance of Button Fuel Cell Using Mixed Conducting Materials

1
School of Energy and Power, Jiangsu University of Science and Technology, Zhenjiang 212003, China
2
School of Mechanical Engineering, Jiangsu University, Zhenjiang 212013, China
*
Author to whom correspondence should be addressed.
Energies 2018, 11(7), 1875; https://doi.org/10.3390/en11071875
Submission received: 10 June 2018 / Revised: 4 July 2018 / Accepted: 17 July 2018 / Published: 18 July 2018

Abstract

:
Intermediate temperature (IT) fuel cells using mixed conducting materials have been reported by many researchers by adopting different compositions, microstructures, manufacture processes and testing conditions. Most iop-Vop relationships of these button electrochemical devices are experimentally achieved based on anode or cathode surface area (i.e., Aan Aca). In this paper, a 3D multi-physics model for a typical IT solid oxide fuel cell (SOFC) that carefully considers detail electrochemical reaction, electric leakage, and e, ion and gas transporting coupling processes has been developed and verified to study the effect of Aca/Aan on button cell iop-Vop performance. The result shows that the over zone of the larger electrode can enhance charges and gas transport capacities within a limited scale of only 0.03 cm. The over electrode zone exceed this width would be inactive. Thus, the active zone of button fuel cell is restricted within the smaller electrode area min(Aan, Aca) due to the relative large disc radius and thin component layer. For a specified Vop, evaluating the responded iop by dividing output current Iop with min(Aan, Aca) for a larger value is reasonable to present real performance in the current device scale of cm. However, while the geometry of button cells or other electrochemical devices approach the scale less than 100 μ m , the effect of over electrode zone on electrochemical performance should not be ignored.

1. Introduction

In the past decade, low cost, clean and high efficiency energy conversation and storage devices, such as fuel cells [1], batteries [2] and super-capacitors [3], have been receiving more and more attentions. Solid oxide fuel cell (SOFC) has being recognized as a promising energy conservation device due to its efficiency [4] and capability to work with various fuels [5]. As high operating temperature might cause strict material compatibility constraints [6] and operational complexity [7], attentions have being devoted to the development of intermediate temperature (IT-) SOFC components (i.e., 350–650 °C) [8]. The key obstacles for reducing SOFC operation temperature are attributed to the insufficient activities of conventional cathode materials and low ionic conductivities of traditional electrolyte materials [9] (i.e., YSZ) in this temperature regime. Thus, mixed ion/e- conducting (MIEC) electrode materials [10] and alternative electrolyte materials [11] have received great attention for their potential applications in IT-SOFCs. Z. Shao et al. reported a mixed conducting Ba0.5Sr0.5Co0.8Fe0.2O3-δ as a potential cathode material, which can conduct both electron and O2− charges [10] Then, an in-situ photoelectron spectroscopy method was proposed to investigate the electrochemically active region within mixed conducting CeO2−x electrode [12]. S. Wang et al. compared the performances of various LSCF-based cathodes and found that LSCF-SDC exhibited a larger activation overpotential than did the single-phase LSCF cathode [13]. More interestingly, the LSM-coated LSCF composite electrode was reported to exhibit a lower activation overpotential compared with that in a pure LSCF cathode [13]. Furthermore, proton conducting oxides [14], such as BaZr0.7Pr0.1Y0.2O3_d [15] and BaZr0.1Ce0.7Y0.2O3_d [16] were also greatly invented to be used in IT-SOFCs because of their low activation energy and high ionic conductivity around IT-range.
Generally, the electrochemical reaction processed within an IT-SOFC using mixed conducting materials or proton conducting oxides are very different from those using conventional composite electrodes [17]. Taking the cathode of LSCF-SDC/SDC/NI-SDC IT-SOFC using mixed conducting materials as an example [13], the electrochemically active sites not only can be taken placed around the percolated three phase boundary sites (i.e., LSCF-SDC-pores and LSCF-dense electrolyte interfaces), but also can happen around the percolated double phase boundary sites (e.g., LSCF-pore) [18]. Up to now, many IT-SOFC button cells using the mixed conducting materials have been reported by many researchers by adopting different compositions (or materials), volume fractions, microstructure parameters, manufacture processes, operating conditions and different cell geometry sizes. It is interesting to note that different anode and cathode surface areas (i.e., different discs radii) were chosen during button cell fabricating and measuring. For a specified output voltage Vop, the corresponding output electric current density iop was always evaluated by dividing output current Iop with the relative smaller surface area between anode Aan and cathode Aca. This may lead to higher electrochemical performance results based on follow consideration. Generally, the performance of IT-SOFC button cell is a trade-off of electrochemical reacting, gas transporting, e and ionic conducing and their mutual coupling processes. The over zone of larger electrode can enhance charges and gases transport capacities within the button cells in a proper zone; and this may affect the Vop-iop performance measuring result. Thus, it is important to study the sensitivity of the SOFC button cell performance on different cathode and anode surface areas ratios.
3D multi-physics coupling numerical modeling is generally agreed to be an economic, valid and time saving approach for working detail investing [19], parameters-performance studying [20], geometric optimizing [21] and system operation optimizing [22]. In this paper, a 3D multi-physics model which carefully considers detail electrochemical reaction, electric leakage, and e, ion and gas transporting coupling processes within a typical IT-SOFC button cell is developed and verified. Then, the influences of different cathode/anode area ratios Aca/Aan on the button cell Vop-iop performances are carefully investigated, while different micro-structure parameters, electrode properties, component thicknesses and exchange current densities of reaction interfaces are varied within reasonable value ranges. The study results can help us achieve the valid affecting zone of the relative larger electrode; and assess the rationality that evaluating responded iop by dividing Iop with min(Aan, Aca) for a larger value, while Vop is specified. The achieved conclusions would provide good references for understanding the geometric effects of cathode/anode cross sections relationship on electrochemical performance of IT-SOFC button cell and similar electrochemical devices.

2. Method and Theory

Taking a typical anode-supported LSCF-SDC/SDC/Ni-SDC IT-SOFC button cell in Figure 1a as an example, the relevant structure and geometry sizes of the distinct four different cell layers from the experiment report is illustrated. As shown in Figure 1b, the multi-physics working processes within these IT-SOFC button cells are complicated even in hydrogen fuel case. Oxygen within the air should be transported to the percolated LSCF-SDC-pores three phase boundaries (TPBs) or LSCF-pores double phase boundaries (DPBs) in the cathode side. ‘Percolated’ here is defined as a continuous connection through the entire electrode structure. At these places, O2 will react with the electrons transported by the electronic conducting paths, such as percolated LSCF network and the external current circuit or dense electrolyte that presences electronic conducting capability (e.g., SDC and CGO). The produced O2− will be conducted to the percolated Ni-SDC-pore TPBs in anode side through O2− conducting network which is constructed by both LSCF- and SDC-particles and dense electrolyte. These O2− will react with the fuels diffused through porous anode. Most of the produced electrons will be circuited back to cathode reaction sites through the external current circuit. But part of the produced electrons will be conducted from anode to cathode side through dense electrolyte directly due to the presence of electronic conducting property of electrolyte material. These electric currents are considered as idle work and cause complex relationships among microstructure parameters, effective electrode properties and multi-physics calculating processes.
As proposed in our previous paper [23], the characteristic properties of each SOFC component layer can be evaluated by the generalized percolation micro-model based on the thickness, composition, and microstructure parameters of each component layers. Taking the LSCF-SDC composite cathode as an example, the potential electrochemical active sites consists of the percolated LSCF-SDC-pores TPBs, percolated LSCF-dense electrolyte interfaces and percolated LSCF-pores DPB surface sites (illustrated in Figure 1b).
The percolated LSCF-SDC-pores TPBs per unit volume can be evaluated as
λ TPB ,   per V = γ LSCF , SDC n LSCF V Z LSCF ,   SDC P LSCF e P SDC O 2  
where subscript ‘per’ is used to represent ‘Percolated’. γ LSCF , SDC = π r c 2 ( r c = min ( r LSCF , r SDC ) sin θ ) is the electrochemical reaction site per contact between LSCF- and SDC-particles (explained in Figure 2a), n k V = ( 1 ϕ g ) ψ k / ( 4 π r k 3 / 3 ) is number of k-particles per unit volume within electrode. P LSCF e and P SDC O 2 are the probabilities of relevant particles belonging to the percolated electron and oxygen ion conducting paths, respectively. Both LSCF- and SDC-particles contribute to the O2− conducting path, thus P SDC O 2 = P LSCF O 2 = 1 . The probability of LSCF-particle belonging to percolated e conducting network can be estimated by Reference [16]
P LSCF e = 1 ( 4.236 Z LSCF , LSCF 2.472 ) 3.7  
Z k , is the number of contacts between k-particle and all of its neighboring -particles
Z k , = 0.5 ( 1 + r k 2 / r 2 ) Z ¯ ψ / r k = 1 M ψ k / r k  
where ψ k and r k are the corresponding solid volume fraction and radius of k -particles. ϕ g is the porosity of composite electrode.
The percolated LSCF-pores DPB surface sites per unit volume can be evaluated basing on the cathode microstructure parameters by [23]
S LSCF ,   per V = n LSCF V s es P LSCF e P LSCF O 2 , s es = 2 π r LSCF 2 [ 2 ( 1 cos θ LSCF ) Z LSCF ,   LSCF ( 1 cos θ LSCF ) Z LSCF ,   SDC ]  
as illustrated in Figure 2b, the exposed surface area of each LSCF-particle s es should be estimated by subtracting the overlap parts of neighboring particles from spherical surface area.
Similarly, the percolated LSCF-dense electrolyte interfaces per unit electrolyte surface area can be estimated by [23]
λ TPB , per S = γ LSCF , ele n LSCF S P LSCF e  
where n LSCF S = ( 1 ϕ g ) ψ LSCF / ( 2 π r LSCF 2 / 3 ) is LSCF-particles number per unit dense electrolyte surface. γ LSCF , ele = 2 π r LSCF sin θ is the electrochemical reaction site per connect between an LSCF-particle and a dense electrolyte.
More details about other effective electrode properties calculating, such as, the percolated Ni-SDC-pore TPBs, effective O2− and e electric conductivities, hydraulic radius of the porous electrodes and so on could also been found in our previous paper on percolation theory for details [23]. Combing with the electrode microstructure parameters of LSCF-SDC/SDC/Ni-SDC IT-button cell in experiment process [13], the corresponding effective characteristic properties of each layers of the above button cell are estimated and provided in Supplementary Materials. Based on these properties, the multi-scale predictive model that comprehensive considers the special characteristics of the typical mixed conducting SOFCs is developed to study the geometric effects of cathode/anode cross sections relationship on electrochemical performance.
According to anodic e-O2− charge transfer reaction, the electrochemical energy relationship at anode active sites, percolated Ni-SDC-pore TPBs (shown in Figure 1b), can be expressed as
H 2 ( g ) + O 2 ( SDC ) H 2 O ( g ) + 2 e ( Ni )  
μ H 2 + μ O 2 2 F Φ O 2 μ H 2 O 2 F Φ e  
where μ α = μ α st + R T ln p α is chemical potential of reactant α at local reaction sites. μ α st is chemical potential at standard condition p st = 1 atm. T and p α are the local temperature and partial pressure of species α , respectively. F is Faraday constant. Φ O 2 and Φ e are the local electrical potentials of O2− and e conducting phases, respectively.
‘=’ in Equation (4b) represents the energy equilibrium state at local place. In this case, the local electromotive force based on local working condition instead of the open circuit condition can be got as [18]
E an eq = Φ O 2 eq Φ e eq = ( μ H 2 + μ O 2 μ H 2 O ) / 2 F  
‘>’ in Equation (4b) is essential to process forward reaction with electric current produced. Thus, the activation overpotential is called as the ionic-electronic voltage difference shifted from the local electromotive force η act an = E an eq ( Φ O 2 Φ e ) .
Similarly, the cathodic e-O2− charge transfer reaction and electrochemical energy relationship in local active sites (i.e., percolated LSCF-pore DPBs or LSCF-SDC-pore TPBs in Figure 1b) are
0.5 O 2 ( g )   +   2 e ( LSCF ) O 2 ( SDC   or LSCF )  
0.5 μ O 2 2 F Φ e μ O 2 2 F Φ O 2  
The corresponding electromotive force of equilibrium state at local cathode active sites and the activation overpotential shifted from this E ca eq are
η act ca = 1 4 F ( μ O 2 2 μ O 2 ) ( Φ e Φ O 2 ) = E ca eq ( Φ e Φ O 2 )  
Then, the relation between e-O2− charge transfer rate per unit TPBs and the activation overpotential can be evaluated by empirical Butler-Volmer equation
j TPB = j TPB , 0 [ exp ( 2 α f F R T η a c t ) exp ( 2 β r F R T η act ) ]  
α f (or β r ) is forward (or reverse) reaction symmetric factor. Local exchange current per unit TPB length at anode and cathode sides can be respectively estimated by Reference [24]
j TPB , 0 an = j TPB , 0 , ref an exp ( E H 2 R ( 1 T 1 T ref ) ) ( p H 2 p 0 )  
j TPB , 0 ca = j TPB , 0 , ref ca exp ( E O 2 R ( 1 T 1 T ref ) ) ( p O 2 p O 2 0 ) 0.25  
where E H 2 and E O 2 are activation energies for H2 oxidation and O2 reduction reactions, respectively. j TPB , 0 , ref is assigned empirically based on experiment at reference Tref. p α 0 is partial pressure of species α at open circuit state. Thus, the volumetric current sources for the transfer of e-O2− electric charges around the TPBs are i e O 2 , TPB V = j TPB λ TPB . per V in A m−3. The area metric current sources over electrode/electrolyte interfaces are i e O 2 , TPB S = j TPB λ TPB . per S in A m−2.
Similarly, the e--O2- charge transfer rate over per LSCF-pore DPB area can be evaluated as [18]
i LSCF = i LSCF , 0 [ exp ( 2 α LSCF F R T η act ca ) exp ( 2 β LSCF F R T η act ca ) ,   [ A   m 2 ] ]  
where i LSCF , 0 = i LSCF , 0 , ref ( p O 2 / p O 2 0 ) 0.25 exp ( E O 2 / ( 1 / T 1 / T ref ) / R ) . i LSCF , 0 , ref is assigned empirically at reference temperature Tref. And the volumetric current sources for the transfer of e-O2− electric charges around DPBs are i e O 2 , LSCF V , ca = i LSCF S LSCF ,   per V in A m−3.
The above items, expect μ O 2 , can be resolved by the local independent variables, such as, T, p α , Φ e and Φ O 2 . Generally, the constant potential shift does not alter e (or O2−) electric potential profiles within the electronic (or ionic) conducting phase. To exclude the influence of μ O 2 during calculating both η act and charge transfer rate, local electric potentials Φ e and Φ O 2 were always shifted by different reference amounts, as reported by D. Jean et al. [25] and S. Liu et al. [26]. However, it is necessary to mention that only limited assumptions reported in the above literatures can be used, while the electronic leaking property in dense SDC electrolyte is considered. Because both Φ e and Φ O 2 are continuously distributed throughout the whole cell structure (anode, electrolyte and cathode).
While keeping Φ e as it is, local Φ O 2 are shifted by a reference amount as Φ ^ O 2 = Φ O 2 + ( μ O 2 st 2 μ O 2 ) / ( 4 F ) . Then, the overpotential expresses should be adjusted accordingly
η act an = E st + ( Φ e Φ ^ O 2 ) R T 2 F ln p H 2 O p H 2  
η act cc = Φ ^ O 2 Φ e R T 4 F ln 1   atm p O 2  
where E st = ( μ H 2 st + 0.5 μ O 2 st μ H 2 O st ) / ( 2 F ) is the Nernst potential at standard state (1 atm).
Combing with the above e-O2 charge transfer rates within the composite electrodes and electrode/dense electrolyte interfaces, multi-physics model can be completed by further coupling momentum, mass, electronic and ionic electric current conservation equations. The relationship among electric current densities, electric potentials and e-O2 charge transfer rates can be solved by [18]
i O 2 = ( σ O 2 eff Φ ^ O 2 ) = { ( i e - O 2 ,   TPB V , ca + i e - O 2 , LSCF V , ca ) in   cathode 0 in   dense   electrolyte i e - O 2 V , an in   anode  
i e = ( σ e eff Φ e ) = { i e - O 2 ,   TPB V ,   ca + i e - O 2 , LSCF V ,   ca in   cathode 0 in   dense   electrolyte i e - O 2 V ,   an in   anode  
where i O 2 and i e are the O2− and e electric current densities within the button cell, respectively. σ O 2 eff and σ e eff are the effective O2− and e electric conductivities, respectively.
The dusty gas model is adopted to describe gas transport within porous anode and cathode layers [27]
N α = R α  
N α D α K eff + x α N β x β N α D α β eff = 1 R T ( p x α + x α p + x α B 0 p μ mix D α K eff p )  
N α and x α are molar flux and local molar fraction of species α , respectively. R α is reaction rate of each species. It can be evaluated through the e-O2 electric current transfer rates per unit electrode volume as
R O 2 = ( i e O 2 ,   LSCF V ,   ca + i e O 2 ,   TPB V ,   c ) / ( 4 F ) R N 2 = 0 } in cathode  
R H 2 = i e O 2 ,   TPB V ,   aa / ( 2 F ) R H 2 O = i e O 2 ,   TPB V ,   an / ( 2 F ) } in   anode  
The total gas pressure and permittivity within the porous structure can be evaluated by
p = c tot R T ,   B 0 = ϕ g 3 r g 2 8 τ 2
where r g is mean hydraulic pore radius of the specified porous electrode structure. τ is the corresponding tortuosity of the porous structure [28]. The effective dynamic viscosity of mixture gas μ mix can be predicted by ideal gas mixing law [29]
μ mix = α = 1 n x α μ α β = 1 n x β Φ α , β ,   μ α μ α 0 ( T T 0 ) 1.5 T 0 + S T + S
Φ α . β = 1 8 ( 1 + M α M β ) 1 / 2 [ 1 + ( μ α μ β ) 1 / 2 ( M α M β ) 1 / 4 ] 2  
where M α is the molar mass. μ α is the dynamic viscosity of species α , which can be evaluated based on Sutherland’s law based on the relevant parameters in Table 1.
The effective Knudsen diffusion coefficient [30] of species α and effective binary diffusion coefficient [31] can be estimated by
D α K eff = ϕ g τ 2 r g 3 8 R T π M α ,   D α β eff = ϕ g τ 3.24 × 10 8 T 1.75 p ( ν α 1 / 3 + ν β 1 / 3 ) ( 1 M α + 1 M β ) 0.5
where ν α is diffusion volume of species α , which is collected in Table 1.

3. Result and Discussion

Figure 3 shows five calculated iop-Vop curves at different operation T for a LSCF-SDC/SDC/Ni-SDC button cell with the ratio of anode and cathode discs radii around rca/ran = 0.8 cm/1 cm. In other words, the surface areas ratio of anode and cathode is Aca/Aan = 0.64. The corresponding parameters are illustrated in Supplementary Materials. It Is necessary to mention that the deviation of iop-Vop curves between the calculating and experiment results in high current density zone at 700 °C was considered to be an error caused by some unknown factors during the testing process based on follow considerations. (i) The sharp drop of iop-Vop curve at high current density zone is considered to be caused by concentration overpotential. However, the limited current density at 700 °C smaller than that at 600 °C is unreasonable. (ii) These deviations were happened around the up boundary operation zone (up operation temperature and current density zones). The deviation could be caused by abnormal factors. (iii) Good agreements between calculated and experiment results [13] at several other T can well illustrate that the modeling parameters can well represent the electrochemical properties of the button cell; and the cell-level multi-physics model can well describe the working details within it. It should be note that the feature of electronic leakage of dense electrolyte would lead to a sharp decrease of the open circuit voltage (shown in Figure 3).
For a button cell, the support component layer is always fabricated with a relative lager surface area compared with the measured electrode [32] (i.e., the anode surface area in current button cell is 3.14 cm2 and the corresponding surface area of measured cathode is only 2 cm2). Obviously, for a specified output voltage Vop, the responded operating current density iop can be evaluated by two ways, divided the output current Iop by Aan for a larger value or Aca for a lower value. To improve the performance quality, most of the reported iop-Vop curves of button cells were always obtained based on the relative smaller area between Aan and Aca. Thus, it is important to evaluate the influence of the over zone from support layer on the experiment measuring and numerical calculating iop-Vop performance results for IT-SOFC button cell using mixed conducting material.
Effect of different Aca/Aanratio at 700 and 600°C:Figure 4 compares the Iop-Vop performances of LSCF-SDC/SDC/Ni-SDC IT-SOFC button cells with respectively cathode area 2 and 0.5 cm2 (labeled as cells 1 and 2), while kept the surface area of support anode as 3.14 cm2. Generally, the experimentally measured iop-Vop performances may be obtained by the following two steps. Firstly, the responded output currents Iop should be measured while the output voltages Vop are specified. Then, the corresponding output current densities iop can be obtained through divided Iop by the electrode surface. Table 2 compares the iop-Vop relations of cells 1 and 2 based on both anode and cathode cross section surfaces, respectively. Obviously, using the relative larger electrode surface (i.e., Aca in current anode supported case) means larger iop value. Using the relative lower electrode surface (i.e., Aan) means lower iop value. It is interesting to find that there are very similar iop-Vop relationships between cells 1 and 2, while estimated iop based on the relevant smaller electrode surface area (i.e., Aca in current case). There is no obvious difference between the performance results even at the maximum power density case. Taking Vop = 0.5 V and T = 700 °C as an example, the maximum power densities are 1.412 W cm−2 for cathode area 0.5 cm2 and 1.397 W cm−2 for Aca = 2 cm2 cases, respectively.
Thus, it can be concluded that although the over zone of larger electrode is generally considered can enhance charges and gases transport capacities in a proper zone, the active zone of the button cell will be restricted at the electrode zone with relevant small area (i.e., cathode layer area in current anode support case), instead of the support layer surface areas (i.e., Aan). The influence of the over zone of anode surface area on iop-Vop performance would be negligible. Taking the small area between anode and cathode surfaces to calculate iop-Vop performance is more reasonable to indicate the electrochemical properties of the tested IT-SOFC button cell.
Effects of Electrolyte thickness on cell performance: Generally, too thin electrolyte layer is considered as a key factor to weaken the influence of the over zone of larger electrode on iop-Vop performance. Figure 5 further compares four different cell performances at 600 °C among the button cells with different combination of cathode surface areas (0.5 and 2 cm2) and electrolyte thicknesses (i.e., 15 and 50 μ m ), while keeps the anode surface areas and other operating parameters values. Obviously, we can find that increase the dense electrolyte thickness in a limited value from 13 to 50 μ m would not affect the above conclusions. Taking the small areas between anode and cathode surfaces to calculate iop-Vop performance is more reasonable to indicate the electrochemical properties of the tested IT-SOFC button cell.
Effects of exchange current density: The exchange current density j TPB ,   0 ,   ref an based on TPBs is an important factor to character the electrochemical property of the button cells. A higher j TPB ,   0 ,   ref an means that smaller activation overpotential is needed to convert same amount of charge between e and O2− electric currents. As shown in Figure 6a, for a specified iop, Vop increases with the increasing j TPB ,   0 ,   ref an . However, it is interesting to get that for all the three cases with j TPB ,   0 ,   ref an = 8.0 × 10−2, 8.0 × 10−3 and 8.0 × 10−4 A m−1, the maximum power density differences between Aca = 0.5 and 2 cm2 cases are less than 100 W m−2, at 0.4 V and 600 °C. Therefore, for those button cells with exchange current densities within reasonable range, the effect of the over zone from larger electrode surface on button cell performance is insignificant.
Effects of different component support cases:Figure 6b shows the sensitivities of button cell performances on different anode/cathode surface area ratios at 600 °C, while different component support cases are adopted. The geometry parameters of the anode, cathode, electrolyte and component-self-support button cells are respectively listed in Table 3. Although different component support cases would lead to very different button cell performances, it should be noted that the sensitivities of iop-Vop performances on different Aca/Aan are still insignificant while evaluating iop based on the smaller electrode surface area between anode and cathode layers.
The real affection zone of the over electrode surface area: It is theoretically agreed that the over anode (or cathode) surface area due to A ca / A an 1 will decrease the potential losses of the gas, electron and ion transports in the corresponding electrode. The above study results, however, show that the geometric effect of anode and cathode surface areas ratio on the IT-button cell iop-Vop performance is quite limited, while the working parameters vary in a reasonable range. The active zone of the button cell is restricted within the smaller electrode area zone between anode and cathode (i.e., min (Aca, Aan)). This should be caused by the geometric characteristics of button cell with relative thin component layer. Taking dense electrolyte layer as an example, the length-width ratio between thickness and radius of disc surface is 13 μm/1 cm = 0.0013. Thus, finding out the real influence width of this over electrode zone on the button cell performance would be very helpful to understand the working properties of IT-SOFCs.
Taking T = 600 °C and Vop = 0.4 V as an example, the concentration distributions of H2 and H2O within porous composite anode, c H 2 and c H 2 O , are shown in Figure 7. c H 2 on anode side is 13.563 mol m−3 initially and drops sharply along z-axis to 6.81 mol m3 due to the oxidation reaction of H2. In contrast, the concentration of product vapor c H 2 O on anode side increases along z axis from 0.42 to 12.4 mol m3. Obviously, c H 2 and c H 2 O have opposite distribution characteristics. Theoretically, the over zone of the current anode surface (i.e., Aan = 3.14 cm2, Aca = 2 cm2, and Aca/Aan = 0.64 in current button cells) would enhance the hydrogen and vapor transports within porous anode in a proper width because of the enlarged cross section.
Figure 8a further shows a c O 2 distribution within porous cathode. The O2 concentration is consumed from 2.93 to 1.41 mol m3. As shown in Figure 8c, the over zone of the anode surface area can also enhance the conducting capacity of electronic current density in y direction due to the enlarged cross section. Combining Figure 7 and Figure 8, we can find that the real effective width of the over electrode zone in an IT-SOFC button cell is only in a scale of 0.03 cm. The over electrode zone exceed this width would be inactive. This can well explain that why the IT-SOFC button cells iop-Vop performance is insensitive to the Aca/Aan ratio; and a smaller Aca/Aan may not greatly increase the measured iop-Vop electrochemical quality. Because the button cell is fabricated in a radius scale of rdisc = 1 cm. These real effective width of the over electrode zone reference to the button cell disc overall radius is less than 5%.
To further confirm this conclusion, a button cell in a smaller scale (i.e., rdisc = 0.1 cm, Aan = 3.14 × 10−2 cm2) is developed. Figure 9 compares the iop-Vop performances between the buttons cells with different cathode/anode surface areas ratios (i.e., Aca = 2 × 10−2 and 0.5 × 10−2 cm2, Aca/Aan = 0.64 and 0.16). The corresponding maximum power densities are 0.9749 and 0.8814 W cm−2 for Aca/Aan = 0.64 and 0.16 cases, respectively. Obviously, while the geometry of the button cells or other electrochemical devices approach the scale less than 100 μm, the effect of the over electrode zone on the electrochemical performance should not be ignored. Taking the smaller electrode surface area to evaluate iop for a specified Vop would cause an improper over evaluation of the electrochemical performance.

4. Conclusions

The comprehensive multi-physics model of IT-SOFC button cells considers special features, such as using mixed conducting materials, electric leakage and complex multi-physics mutual coupling processes have been developed and verified. The geometry effect of different anode and cathode surface area ratios on iop-Vop performance of IT-SOFC button cells are investigated and many conclusions are reached,
(i).
The over zone of the larger electrode can only enhance charges and gas transport capacities within a limited scale of only 0.03 cm, an over electrode zone exceeding this width would be inactive.
(ii).
The active zone of button cell is restricted within the smaller electrode area min(Aan, Aca) due to the relatively large disc radius in scale of cm and the thin component layer.
(iii).
For a specified Vop, evaluating the responded iop by dividing output current Iop with min(Aan, Aca) for a larger value is reasonable for presenting the real performance in a current device scale.
(iv).
While the geometry of button cell or other electrochemical device approaches a scale of less than 100 μ m , taking the smaller electrode surface area to evaluate iop for a specified Vop would cause an improper over evaluating of the electrochemical performance.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1073/11/7/1875/s1.

Author Contributions

Data curation, B.H.; Investigation, K.D.; Writing-Original Draft Preparation, C.Y.; Writing-Review and Editing, D.C. and L.L. All authors read and approved the final manuscript.

Funding

This research was funded by the National Science Foundation of China (51776092 and 21406095) and the Natural Science Foundation of Jiangsu Province BK20151325.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

Acathe cross section area of the cathode layer, m2
Aanthe cross section area of the anode layer, m2
B0the flow permeability, m2
c α 0 the concentration of species α at the channel inlet, mol m−3
DPBsthe LSCF-pore double phase boundaries
D α β eff the effective binary diffusivity, m2 s−1
D α K eff the effective Knudsen diffusivity of species α , m2 s−1
Eeqthe local equilibrium electric potentials difference at working state, V
Estthe Nernst potential at the standard state, V
E H 2 the activation energy for H2 oxidation reaction, J
E O 2 the activation energy for O2 reduction reaction, J
Fthe Faraday constant, C mol−1
GDCthe Gd0.1Ce0.9O1.95
j TPB the local e-O2− charge transfer rate per unit TPB length, A m−1
j TPB , 0 the local exchange transfer current per unit TPB lengths, A m−1
j TPB , 0 , ref the value assigned empirically based on experiment at reference temperature
iethe local e electric current density, A m−2
i O 2 the local O2− electric current densities, A m−2
iopthe output current density, A m−2
Iopthe output current, A
i LSCF the e-O2− charge transfer rate per unit percolated DPB area, A m−2
i LSCF , 0 the local exchange transfer current per unit percolated DPB area, A m−2
i e O 2 , TPB V the e-O2− charge transfer rate per unit volume based on percolated TPBs, A m−3
i e O 2 , LSCF V the e-O2− charge transfer rate per unit volume based on percolated LSCF-pore DPBs, A m−3
i e O 2 , TPB S the e-O2− charge transfer rate per unit dense electrolyte surface, A m−2
LSCFthe La0.6Sr0.4Co0.2Fe0.8O3−δ
LSMthe La1-xSrxMnO3
M α the mole mass of species α , kg mol−3
N α the molar flux of species α , mol m−2 s−1
n k V the number of k-particles per unit volume
n k S the number of k-particles per unit dense electrolyte surface area
p α the partial pressure of gas species α at the local reaction sites, atm
p α 0 the partial pressure of gas species α in the channel inlet, atm
P SDC O 2 the probabilities of SDC-particles belonging to percolated O2− conducting path
P LSCF O 2 the probabilities of LSCF-particles belonging to percolated O2− conducting path
P LSCF e the probabilities of LSCF-particles belonging to percolated e conducting path
rgthe mean hydraulic pore radius of porous electrode structure, m
r k the radius of k-particle, m
rcthe neck radius between two connected particles, m
Rthe universal gas constant, J mol−1 K−1
R α the sources/leak of species α , mol m−3 s−1
SDCthe Sm0.2Ce0.8O2−δ
S LSCF ,   per V the percolated LSCF-pore DPBs per unit volume, m−1
s es the exposed surface area of each LSCF-particle, m2
Tthe operating temperature, K
TPBsthe three phase boundary sites
Vopthe output voltage at working state, V
ν α the diffusion volume for species α , m3 mol−1
x α the molar fraction of species α
YSZthe yttrium-stabilized zirconia
Z k , l the average number of contacts between k- and all of its neighboring l-particles
Z ¯ the average coordination number of all particles
Greek letters
α f , β r the forward and reverse reaction symmetric factors
γ LSCF , SDC the 1D circular length per contact between LSCF- and SDC-particles, m
γ LSCF , ele the 1D circular length per contact between LSCF-particle and the dense electrolyte, m
λ TPB , per V the percolated TPB length per unit volume, m−2
λ TPB , per S the percolated TPB length per dense electrolyte surface area, m−1
ϕ g the porosity of porous structure
Φ e the local e electric potential, V
Φ O 2 the local O2− electric potential, V
Φ ^ O 2 the shift of Φ O 2 by a reference amount, V
ψ k the solid volume fraction of k-particles
η act the local activation overpotential, V
θ the smaller contact angle between two particles
τ the tortuosity of gas transport path within the porous electrode
σ e eff the effective electronic conductivity, S m−1
σ O 2 eff the effective O2− ionic conductivity, S m−1
μ mix the viscosity of gas mixture, kg m−1 s−1
μ α the mole chemical potential of reactant α , J mol−1
Superscripts and subscripts
ananode
actactivation
cacathode
eqequilibrium
ststandard condition (1 atm)
refreference value

References

  1. Kong, W.; Gao, X.; Liu, S.; Su, S.; Chen, A.D. Optimization of the Interconnect Ribs for a Cathode-Supported Solid Oxide Fuel Cell. Energies 2014, 7, 295–313. [Google Scholar] [CrossRef] [Green Version]
  2. Fotouhi, A.; Auger, D.J.; O’Neill, L.; Cleaver, T.; Walus, S. Lithium-Sulfur Battery Technology Readiness and Applications—A Review. Energies 2017, 10, 1937. [Google Scholar] [CrossRef]
  3. Repp, S.; Harputlu, E.; Gurgen, S.; Castellano, M.; Kremer, N.; Pompe, N.; Worner, J.; Hoffmann, A.; Thomann, R.; Emen, F.M.; et al. Synergetic effects of Fe(3+) doped spinel Li4Ti5O12 nanoparticles on reduced graphene oxide for high surface electrode hybrid supercapacitors. Nanoscale 2018, 10, 1877–1884. [Google Scholar] [CrossRef] [PubMed]
  4. Kupecki, J.; Motylinski, K.; Milewski, J. Dynamic analysis of direct internal reforming in a SOFC stack with electrolyte-supported cells using a quasi-1D model. Appl. Energy 2017. [Google Scholar] [CrossRef]
  5. Papurello, D.; Iafrate, C.; Lanzini, A.; Santarelli, M. Trace compounds impact on SOFC performance: Experimental and modelling approach. Appl. Energy 2017, 208, 637–654. [Google Scholar] [CrossRef]
  6. Fang, X.; Zhu, J.; Lin, Z. Effects of Electrode Composition and Thickness on the Mechanical Performance of a Solid Oxide Fuel Cell. Energies 2018, 11, 1735. [Google Scholar] [CrossRef]
  7. Chen, D.; Xu, Y.; Tade, M.O.; Shao, Z. General Regulation of Air Flow Distribution Characteristics within Planar Solid Oxide Fuel Cell Stacks. ACS Energy Lett. 2017, 2, 319–326. [Google Scholar] [CrossRef]
  8. Park, J.; Kim, D.; Baek, J.; Yoon, Y.-J.; Su, P.-C.; Lee, S. Numerical Study on Electrochemical Performance of Low-Temperature Micro-Solid Oxide Fuel Cells with Submicron Platinum Electrodes. Energies 2018, 11, 1204. [Google Scholar] [CrossRef]
  9. Tarancón, A. Strategies for Lowering Solid Oxide Fuel Cells Operating Temperature. Energies 2009, 2, 1130–1150. [Google Scholar] [CrossRef] [Green Version]
  10. Shao, Z.P.; Haile, S.M. A high-performance cathode for the next generation of solid-oxide fuel cells. Nature 2004, 431, 170–173. [Google Scholar] [CrossRef] [PubMed]
  11. Ni, M.; Leung, M.K.H.; Leung, D.Y.C. Theoretical analysis of reversible solid oxide fuel cell based on proton-conducting electrolyte. J. Power Sources 2008, 177, 369–375. [Google Scholar] [CrossRef]
  12. Zhang, C.; Grass, M.E.; McDaniel, A.H.; DeCaluwe, S.C.; Gabaly, F.E.; Liu, Z.; McCarty, K.F.; Farrow, R.L.; Linne, M.A.; Hussain, Z.; et al. Measuring fundamental properties in operating solid oxide electrochemical cells by using in situ X-ray photoelectron spectroscopy. Nat. Mater. 2010, 9, 944–949. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, S.-F.; Wang, Y.-R.; Yeh, C.-T.; Hsu, Y.-F.; Chyou, S.-D.; Lee, W.-T. Effects of bi-layer La0.6Sr0.4Co0.2Fe0.8O3−δ-based cathodes on characteristics of intermediate temperature solid oxide fuel cells. J. Power Sources 2011, 196, 977–987. [Google Scholar] [CrossRef]
  14. Liu, H.; Akhtar, Z.; Li, P.; Wang, K. Mathematical Modeling Analysis and Optimization of Key Design Parameters of Proton-Conductive Solid Oxide Fuel Cells. Energies 2014, 7, 173–190. [Google Scholar] [CrossRef] [Green Version]
  15. Kim, J.; Sengodan, S.; Kwon, G.; Ding, D.; Shin, J.; Liu, M.; Kim, G. Triple-Conducting Layered Perovskites as Cathode Materials for Proton-Conducting Solid Oxide Fuel Cells. ChemSusChem 2014, 7, 2811–2815. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, D.; Zhang, Q.; Lu, L.; Periasamy, V.; Tade, M.O.; Shao, Z. Multi scale and physics models for intermediate and low temperatures H+-solid oxide fuel cells with H+/e/O2− mixed conducting properties: Part A, generalized percolation theory for LSCF-SDC-BZCY 3-component cathodes. J. Power Sources 2016, 303, 305–316. [Google Scholar] [CrossRef]
  17. Papurello, D.; Lanzini, A. SOFC single cells fed by biogas: Experimental tests with trace contaminants. Waste Manag. 2018, 72, 306–312. [Google Scholar] [CrossRef] [PubMed]
  18. Chen, D.; Wang, H.; Zhang, S.; Tade, M.O.; Shao, Z.; Chen, H. Multiscale model for solid oxide fuel cell with electrode containing mixed conducting material. Aiche J. 2015. [Google Scholar] [CrossRef]
  19. Chen, D.; Xu, Y.; Hu, B.; Yan, C.; Lu, L. Investigation of proper external air flow path for tubular fuel cell stacks with an anode support feature. Energy Convers. Manag. 2018, 171, 807–814. [Google Scholar] [CrossRef]
  20. Liu, K.; Liu, B.; Villavicencio, R.; Wang, Z.; Guedes Soares, C. Assessment of material strain rate effects on square steel plates under lateral dynamic impact loads. Ships Offshore Struct. 2018, 13, 217–225. [Google Scholar] [CrossRef]
  21. Su, S.; He, H.; Chen, D.; Zhu, W.; Wu, Y.; Kong, W.; Wang, B.; Lu, L. Flow distribution analyzing for the solid oxide fuel cell short stacks with rectangular and discrete cylindrical rib configurations. Int. J. Hydrog. Energy 2015, 40, 577–592. [Google Scholar] [CrossRef]
  22. Pianko-Oprych, P.; Hosseini, S. Dynamic Analysis of Load Operations of Two-Stage SOFC Stacks Power Generation System. Energies 2017, 10, 2103. [Google Scholar] [CrossRef]
  23. Chen, D.; He, H.; Zhang, D.; Wang, H.; Ni, M. Percolation theory in solid oxide fuel cell composite electrodes with a mixed electronic and ionic conductor. Energies 2013, 6, 1632–1656. [Google Scholar] [CrossRef]
  24. Zhu, H.Y.; Kee, R.J. Modeling distributed charge-transfer processes in SOFC membrane electrode assemblies. J. Electrochem. Soc. 2008, 155, B715–B729. [Google Scholar] [CrossRef]
  25. Jeon, D.H.; Nam, J.H.; Kim, C.J. Microstructural optimization of anode-supported solid oxide fuel cells by a comprehensive microscale model. J. Electrochem. Soc. 2006, 153, A406–A417. [Google Scholar] [CrossRef]
  26. Liu, S.X.; Song, C.; Lin, Z.J. The effects of the interconnect rib contact resistance on the performance of planar solid oxide fuel cell stack and the rib design optimization. J. Power Sources 2008, 183, 214–225. [Google Scholar] [CrossRef]
  27. Tseronis, K.; Kookos, I.K.; Theodoropoulos, C. Modelling mass transport in solid oxide fuel cell anodes: A case for a multidimensional dusty gas-based model. Chem. Eng. Sci. 2008, 63, 5626–5638. [Google Scholar] [CrossRef]
  28. Kong, W.; Zhang, Q.; Xu, X.; Chen, D. A Simple Expression for the Tortuosity of Gas Transport Paths in Solid Oxide Fuel Cells’ Porous Electrodes. Energies 2015, 8, 13953–13959. [Google Scholar] [CrossRef] [Green Version]
  29. Ni, M.; Shao, Z.; Chan, K. Modeling of Proton-Conducting Solid Oxide Fuel Cells Fueled with Syngas. Energies 2014, 7, 4381–4396. [Google Scholar] [CrossRef] [Green Version]
  30. Veldsink, J.W.; Vandamme, R.M.J.; Versteeg, G.F.; Vanswaaij, W.P.M. The Use of the Dusty-Gas Model for the Description of Mass-Transport with Chemical-Reaction in Porous-Media. Chem. Eng. J. Biochem. Eng. J. 1995, 57, 115–125. [Google Scholar] [CrossRef]
  31. Todd, B.; Young, J.B. Thermodynamic and transport properties of gases for use in solid oxide fuel cell modelling. J. Power Sources 2002, 110, 186–200. [Google Scholar] [CrossRef]
  32. Nguyen, X.-V.; Chang, C.-T.; Jung, G.-B.; Chan, S.-H.; Huang, W.; Hsiao, K.-J.; Lee, W.-T.; Chang, S.-W.; Kao, I.-C. Effect of Sintering Temperature and Applied Load on Anode-Supported Electrodes for SOFC Application. Energies 2016, 9, 701. [Google Scholar] [CrossRef]
Figure 1. (a) The sketch figure of a typical LSCF-SDC/SDC/Ni-SDC IT-SOFC, (b) the corresponding multi-physics working processes within it.
Figure 1. (a) The sketch figure of a typical LSCF-SDC/SDC/Ni-SDC IT-SOFC, (b) the corresponding multi-physics working processes within it.
Energies 11 01875 g001
Figure 2. (a) Sketch of electrochemical reaction site per contact between LSCF- and SDC-particles γ LSCF , SDC ; (b) illustrated of the exposed surface area of per LSCF-particle s es .
Figure 2. (a) Sketch of electrochemical reaction site per contact between LSCF- and SDC-particles γ LSCF , SDC ; (b) illustrated of the exposed surface area of per LSCF-particle s es .
Energies 11 01875 g002
Figure 3. Comparison between the numerical modeling and experiment [13] results at various operating temperatures.
Figure 3. Comparison between the numerical modeling and experiment [13] results at various operating temperatures.
Energies 11 01875 g003
Figure 4. Effects of different cathode areas on cell, (a) Iop-Vop performances at 700 °C; (b) Iop-Vop performances at 600 °C.
Figure 4. Effects of different cathode areas on cell, (a) Iop-Vop performances at 700 °C; (b) Iop-Vop performances at 600 °C.
Energies 11 01875 g004
Figure 5. The effects of different Aca/Aan on the button cell performances at 600 °C while increases the thickness of dense electrolyte from 15 to 50 μ m .
Figure 5. The effects of different Aca/Aan on the button cell performances at 600 °C while increases the thickness of dense electrolyte from 15 to 50 μ m .
Energies 11 01875 g005
Figure 6. The sensitivity of button cell performances on different Aca/Aan: (a) Under various exchange current densities, (b) while different component support cases are considered.
Figure 6. The sensitivity of button cell performances on different Aca/Aan: (a) Under various exchange current densities, (b) while different component support cases are considered.
Energies 11 01875 g006
Figure 7. The concentration distributions of H2 and H2O within anode at T = 600 °C and Vop = 0.4 V as an example.
Figure 7. The concentration distributions of H2 and H2O within anode at T = 600 °C and Vop = 0.4 V as an example.
Energies 11 01875 g007
Figure 8. (a,b) c O 2 distribution within porous cathode at T = 600 °C and Vop = 0.4 V as an example; (c) i e distribution throughout the whole button cell.
Figure 8. (a,b) c O 2 distribution within porous cathode at T = 600 °C and Vop = 0.4 V as an example; (c) i e distribution throughout the whole button cell.
Energies 11 01875 g008
Figure 9. iop-Vop performances between the buttons cells with different cathode/anode surface areas ratios (i.e., Aan = 3.14 × 10−2 cm2, Aca/Aan = 0.64 and 0.16).
Figure 9. iop-Vop performances between the buttons cells with different cathode/anode surface areas ratios (i.e., Aan = 3.14 × 10−2 cm2, Aca/Aan = 0.64 and 0.16).
Energies 11 01875 g009
Table 1. Gas composition and parameters for viscosity calculations by Sutherland’s law.
Table 1. Gas composition and parameters for viscosity calculations by Sutherland’s law.
Gas ν α ( × 10 6 m 3 mol 1 )   μ α 0   ( × 10 6   kg   m 1   s 1 ) T0 (K)S (K)
H26.128.41127397
vapor13.111.23501064
O216.319.19273139
N218.516.63273107
Table 2. For a specified Vop, comparing the responded iop that are evaluated by divided Iop with Aan for larger value and Aca for lower value, respectively.
Table 2. For a specified Vop, comparing the responded iop that are evaluated by divided Iop with Aan for larger value and Aca for lower value, respectively.
VopI (A)iop Based Aan (A cm−2)iop Based on Aca (A cm−2)
Cell 1Cell 2Cell 1Cell 2Cell 1Cell 2
0.28.1392.0472.5920.6524.0694.095
0.37.5851.9062.4160.6073.7933.811
0.46.8381.7172.1780.5473.4193.434
0.55.5871.4121.7790.4502.7942.824
0.63.8600.9711.2290.3091.9301.942
0.71.9050.4810.6070.1530.9530.962
0.780.3020.0780.0960.0250.1510.155
Aan = 3.14 cm2. Aca = 2 and 0.5 cm2 for cell 1 and 2, respectively.
Table 3. Geometry parameters of the anode, cathode, electrolyte and component-self-support button cells in the unit of mm.
Table 3. Geometry parameters of the anode, cathode, electrolyte and component-self-support button cells in the unit of mm.
ItemAnode SupportFunctional LayerDense ElectrolyteCathode
Anode support46061340
Cathode support20613480
Electrolyte support20646040
Self support10010050200

Share and Cite

MDPI and ACS Style

Chen, D.; Hu, B.; Ding, K.; Yan, C.; Lu, L. The Geometry Effect of Cathode/Anode Areas Ratio on Electrochemical Performance of Button Fuel Cell Using Mixed Conducting Materials. Energies 2018, 11, 1875. https://doi.org/10.3390/en11071875

AMA Style

Chen D, Hu B, Ding K, Yan C, Lu L. The Geometry Effect of Cathode/Anode Areas Ratio on Electrochemical Performance of Button Fuel Cell Using Mixed Conducting Materials. Energies. 2018; 11(7):1875. https://doi.org/10.3390/en11071875

Chicago/Turabian Style

Chen, Daifen, Biao Hu, Kai Ding, Cheng Yan, and Liu Lu. 2018. "The Geometry Effect of Cathode/Anode Areas Ratio on Electrochemical Performance of Button Fuel Cell Using Mixed Conducting Materials" Energies 11, no. 7: 1875. https://doi.org/10.3390/en11071875

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop