Next Article in Journal
Comparative Study of Seal Strength and Mechanical Behavior of Untreated and Corona-Treated Polymer Films
Previous Article in Journal
A Novel Maximum Power Point Tracking Method Based on Optimal Evaporation Pressure and Superheat Temperature for Organic Rankine Cycle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Physics Coupling Simulation of H2O–CO2 Co-Electrolysis Using Flat Tubular Solid Oxide Electrolysis Cells

by
Chaolong Cheng
1,2,†,
Wen Ding
3,†,
Junfeng Shen
1,
Penghui Liao
3,
Chengrong Yu
3,
Bin Miao
4,
Yexin Zhou
3,*,
Hui Li
5,
Hongying Zhang
5,* and
Zheng Zhong
3
1
Shenzhen Energy Group Co., Ltd., No. 2026 Jintian Rd., Shenzhen 518031, China
2
Shenzhen Energy Innovation Technology Co., Ltd., Fubao Street, Shenzhen 518048, China
3
School of Science, Harbin Institute of Technology, Shenzhen 518055, China
4
Energy Research Institute at NTU (ERI@N), Nanyang Technological University, 1 CleanTech Loop #06-04, Singapore 637141, Singapore
5
Qingpeng Hydrogen Technology (Shenzhen) Co., Ltd., No. 5 Gongye 5th Road, Shenzhen 518067, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Processes 2025, 13(10), 3192; https://doi.org/10.3390/pr13103192
Submission received: 25 August 2025 / Revised: 24 September 2025 / Accepted: 4 October 2025 / Published: 8 October 2025
(This article belongs to the Special Issue Recent Advances in Fuel Cell Technology and Its Application Process)

Abstract

Solid oxide electrolysis cells (SOECs) have emerged as a promising technology for efficient energy storage and CO2 utilization via H2O–CO2 co-electrolysis. While most previous studies focused on planar or tubular configurations, this work investigated a novel flat, tubular SOEC design using a comprehensive 3D multi-physics model developed in COMSOL Multiphysics 5.6. This model integrates charge transfer, gas flow, heat transfer, chemical/electrochemical reactions, and structural mechanics to analyze operational behavior and thermo-mechanical stress under different voltages and pressures. Simulation results indicate that increasing operating voltage leads to significant temperature and current density inhomogeneity. Furthermore, elevated pressure improves electrochemical performance, possibly due to increased reactant concentrations and reduced mass transfer limitations; however, it also increases temperature gradients and the maximum first principal stress. These findings underscore that the design and optimization of flat tubular SOECs in H2O–CO2 co-electrolysis should take the trade-off between performance and durability into consideration.

1. Introduction

In this era of rapid technological advancement, the energy crisis has emerged as one of the most critical challenges impeding the progress of human society. Traditional fossil fuels release substantial amounts of CO2 and other greenhouse gases during combustion, significantly contributing to global warming and severely disrupting the ecological balance of nature. Reducing atmospheric CO2 levels and mitigating the pace of global climate change have become imperative issues that humanity must address to achieve sustainable development. Therefore, a rising trend in the widespread adoption of renewable energy is observed. Nevertheless, the inherent intermittency and variability of renewable energies demand effective energy storage solutions to ensure grid stability and reliable power supply.
Solid oxide electrolysis cells (SOECs) are a high-temperature water electrolysis technology [1,2,3,4,5]. Table 1 compares the characteristics of three types of water electrolyzers, including alkaline electrolysis cells (AECs), proton exchange membrane electrolysis cells (PEMECs), and SOECs. Compared to low-temperature water-electrolysis technologies, such as AECs and PEMECs, SOECs exhibit higher electrolysis efficiency due to reduced activation loss. Therefore, a long-term, large-scale energy storage solution can be achieved by converting surplus renewable electricity into H2 using SOECs. H2 is an extensively utilized chemical product in traditional industrial applications, such as ammonia and methanol synthesis. With a wide availability of H2 enabled by renewable energy, the potential of H2 as an energy carrier can be further explored owing to its advantages, such as high gravimetric energy density (120 MJ/kg) and negligible environmental impact, during utilization [6,7]. Beyond water electrolysis for H2 production, SOECs can also co-electrolyze H2O and CO2 to produce CO and H2 (Figure 1) [8], or even CH4 with an in situ methanation reaction [9]. This process not only yields valuable synthetic gas fuels, but also enables CO2 utilization and contribution to carbon neutrality efforts. Figure 1 gives a typical I–V curve of SOEC and the three characteristic voltage losses, including activation loss, ohmic loss, and concentration loss.
In H2O–CO2 co-electrolysis via SOECs, multiple processes are coupled, such as the transport of gases and charged ions, electrochemical reactions, chemical reactions, and the generation/consumption and transfer of heat. In contrast to being capable of only characterizing the overall electrochemical performance by experiments, multi-physics coupling simulations are effective in probing the distribution of various physical fields. These simulations thereby provide useful information on the underlying physical mechanisms responsible for performance variations resulting from a change in materials and the structure of SOECs. Luo et al. [16] developed a two-dimensional (2D) model to analyze the performance and efficiency of H2O–CO2 co-electrolysis in tubular SOECs. Their calculations demonstrated that counter-flow configuration outperforms parallel-flow. Building upon this work, Luo et al. [17] found that tubular electrolysis cells, without additional hydrogen supply, cannot produce sufficient CH4 solely from H2 generated through H2O electrolysis. This limitation arises from the temperature difference between the optimal reaction rates of the water–gas shift reaction (WGSR) and methanation reaction (MR). They proposed that increasing hydrogen content in the inlet gas and elevating the electrolysis current could enhance CH4 production.
Chen et al. [18] reported that elevated pressure could improve CH4 conversion rates, with 3 bar identified as the optimal pressure. Beyond this threshold, no further conversion rate improvement was observed (achieving 2.5 times the conversion rate at 1 bar), suggesting that lower temperatures would be required for additional CH4 yield enhancement. Chi et al. [1] conducted a comparative study of various water electrolysis technologies, including alkaline water electrolysis, proton exchange membrane electrolysis, solid oxide electrolysis, and alkaline anion exchange membrane electrolysis. Their work examined ion transport mechanisms, operational characteristics, energy consumption, and process outputs, while also discussing the prospects of water electrolysis technologies.
Beyond electrochemical performance, researchers have also explored the mechanical properties of SOECs through multi-physical modeling. Cui et al. [19] calculated the thermal stress in an SOEC unit, evaluating the effects of voltage, flow direction, water mole fraction, and pressure. Their results indicate that the maximum principal stress occurs in the electrolyte layer. Fan et al. [20] observed that thinner electrolytes and thicker anodes lead to lower residual stresses in the PEN (Positive electrode–Electrolyte–Negative electrode) structure. They noted that tensile stresses in the electrolyte increase at higher SOFC operating temperatures, potentially causing anode cracking. Consequently, they recommended employing smaller operational temperature gradients, carefully matched coefficient of thermal expansion (CTE) materials, and thinner electrolyte layers.
The above studies focused on either planar or tubular SOECs. Recently, a novel flat tubular structure with a symmetrical double cathode was proposed, combining the advantages of both planar and tubular configurations [21]. In this work, based on the flat tubular SOEC, a three-dimensional multi-physics coupling model was built in the COMSOL Multiphysics simulation platform. The model took into account the complex chemical reactions, electrochemical reactions, gas diffusion, and internal thermal stress in the H2O–CO2 co-electrolysis process. The current density, gas distribution, reaction rate, temperature, and stress distribution of a single cell under different voltages were analyzed and are discussed herein.

2. Numerical Model

2.1. Governing Equations

2.1.1. Electrochemical Reaction

In the co-electrolysis model, the gas in the cathode flow channel diffuses in the porous cathode and enters the three-phase interface through porous electrodes to enable electrochemical reactions, including the reduction of H2O and CO2 to H2 and CO, respectively, and the oxidation of O2− to O2 [22]:
H 2 O + 2 e H 2 + O 2
CO 2 + 2 e CO + O 2
  2 O 2 O 2 + 4 e
As the commonly used Ni catalyst in the cathode, the water–gas shift reaction (WGSR) and methane steam reforming reaction (MSR) also take place catalyzed by Ni [18,23].
WGSR : CO + H 2 O CO 2 + H 2
MSR : CH 4 + H 2 O CO + 3 H 2
Owing to various potential losses, including activation overpotential (ηact), ohmic overpotential (ηohm), and concentration overpotential (ηconc), the working voltage (V) of an SOEC is higher than the open circuit voltage (EOCV) [24]:
V = E OCV + η act , a + η act , c + η ohmic + η conc , a + η conc , c
Subscripts a and c denote anode and cathode, respectively.
During H2O–CO2 co-electrolysis, EOCV can be calculated based on the Nernst equation of thermodynamic reversible potential [25]:
E H 2 OCV = 1.253 0.00024516 T + R T 2 F ln P H 2 b P O 2 b 1 2 P H 2 O b
E CO OCV = 1.46713 0.0004527 T + R T 2 F ln P CO b P O 2 b 1 2 P CO 2 b
where R is the gas constant (8.3145 J·mol−1·K−1), T represents the reaction temperature (K), F is the Faraday constant (96,485 C·mol−1), and P H 2 , P O 2 , P H 2 O , P CO , and P CO 2 are the partial pressures of H2, O2, H2O, CO, and CO2, respectively. The superscript L indicates the electrolyte–electrode interface, while b indicates the electrode surface.
The relationship between ηact and current density can be obtained through the Butler–Volmer equation [24]:
i a , O 2 = A V a i 0 , O 2 exp 0.5 F η act , a R T exp 0.5 F η act , a R T
i c , H 2 = A V c i 0 , H 2 P H 2 P H 2 b exp 0.5 F η act , c R T P H 2 O P H 2 O b exp 0.5 F η act , c R T
i c , CO = A V c i 0 , CO P CO P CO b exp 0.5 F η act , c R T P CO 2 P CO 2 b exp 0.5 F η act , c R T
where AVa and AVc are the surface area of the anode and cathode electrochemical reaction interface and i0 is the exchange current density of the reaction.
The ηohm can be easily obtained via ohmic law.
The ηconc is caused by the slow mass transfer of gas species in the electrodes, leading to concentration differences between the electrode surface and the electrode–electrolyte interface. Thus, ηconc can be calculated with respect to these concentration differences with a Nernst-type expression [26]:
η conc , a = R T 2 F ln p H 2 O L p H 2 b p H 2 O b p H 2 L + R T 2 F ln p CO 2 L p CO b p CO 2 b p CO L
η conc , c = R T 2 F ln p O 2 b p O 2 L
The relevant parameters used in this section are shown in Table 2.

2.1.2. Gas Flow

Gas flow in the flow channels and in the porous electrodes is described by the Navier–Stokes equation modified with the Darcy term [31,32]:
  ρ v = S m a s s
ρ v v = p + μ v + v 2 3 μ v ε μ v k
anode :                                                                 S m a s s = M H 2 O M H 2 i 2 F
cathode :                                                                     S m a s s = M O 2 i 4 F
In the formula, ρ is the density of the mixed gas; ν is the velocity vector; Smass is the external quality source term; μ is the dynamic viscosity; ε and k are the permeability and porosity of the electrode structure of the electrolytic cell, respectively; and Mi is the relative molecular mass of i gas.
Gas diffusion in the porous electrodes is expressed using the Maxwell–Stefan model, modified with Knudsen diffusion:
j i = ρ D i m K ω i ρ ω i D i m K M M + ρ ω i k M i M D i m K x i
D i m k = ε τ 1 D i m + 1 D i k 1
M = i ω i M i 1
where ji and ωi are the flow rate and mass fraction of component i, respectively; τ is the tortuosity factor; and DmK is the total diffusion coefficient, which is determined by Fick diffusion coefficient Dm and Knudsen diffusion coefficient DK.

2.1.3. Heat Transfer

The heat transfer in the whole cell is based on the classical heat transfer and energy conservation equation:
λ e f f T + ρ C p v T = Q
where λeff represents effective thermal conductivity, and results are determined by the thermal conductivity of solid materials and fluid materials. The detailed material and structure parameters required for the simulation were extracted from the literature [33,34] and are shown in Table 3.
As the physical fields in the multi-physical field model are coupled, the thermal strain in the whole cell calculation formula can be transformed into:
ε t h = α T T r e f I
The thermal expansion coefficient of the material is affected by the type of material. Tref is the reference temperature without stress, which is 1023 K; I is a second-order unit tensor.
The equations of mechanics used are [35]:
ε = u + u / 2
σ + f = 0
σ = C : ε
where u represents displacement; σ is stress; f is physical strength; and C is the fourth-order elastic coefficient tensor. Each component in SOECs is assumed to be isotropic, and thus the material constitutive equation can be simplified into:
σ x x σ y y σ z z τ x y τ x z τ y z = E ( 1 + v ) ( 1 2 v ) 1 v v v 0 0 0 v 1 v v 0 0 0 v v 1 v 0 0 0 0 0 0 1 2 v 2 0 0 0 0 0 0 1 2 v 2 0 0 0 0 0 0 1 2 v 2 ε x x ε t h ε y y ε t h ε z z ε t h γ x y γ x z γ y z
where E is Young’s modulus and ν is Poisson’s ratio. Mechanical parameters are shown in Table 4.

2.1.4. Chemical Reaction

In the cathode of SOECs, due to the presence of Ni catalyst, chemical reactions such as the reverse WGSR and reverse MSR take place, as confirmed by Haberman and Young [37]. The reaction rates are calculated by:
R W G S R = K s f p H 2 O p C O p H 2 p C O 2 K p s mol m 3 s 1
K s f = 0.0171 × exp 103191 R T mol m 3 Pa 2 s 1
K p s = exp 0.2935 Z 3 + 0.6351 Z 2 + 4.1788 Z + 0.3169
R M S R = K r f p C H 4 p H 2 O p C O p H 2 3 K p r mol m 3 s 1
K r f = 2395 × exp 231266 R T mol m 3 Pa 2 s 1
K p s = 1.0267 × 10 10 × exp 0.2513 Z 4 + 0.36651 Z 3 + 0.5810 Z 2 27.134 Z + 3.277
Z = 1000 T K 1

2.2. Geometric Model and Boundary Conditions

Figure 2 shows the geometry of the flat tubular SOEC. As the SOEC is symmetric, a semi-structure of the cell is used in the model. The 3D multi-physical model is solved across the whole body of the geometry. Geometric parameters are shown in Table 5.
The boundary conditions of the model are as follows: the cathode inlet gas is 0.4 H2O, 0.4 CO2, and 0.2 H2; the flow rate is 0.3 m·s−1; the anode inlet gas is air (0.79 N2 and 0.21 O2); the gas flow rate is 1 m·s−1; and the outlet gas conditions of both are an atmospheric pressure of 1 atm. The external temperature is 1023 K. The voltage is 0 V on the surface of the supporting cathode. The model is solved with COMSOL Multiphysics 5.6 (COMSOL, Sweden).

3. Results and Discussion

3.1. Model Validation

Figure 3 compares I–V curves from simulation and from experimental results extracted from a previously published report [38]. These simulation results are in good agreement at operating temperatures of both 700 °C and 750 °C, validating the as-developed model.

3.2. Physical Field Distributions

3.2.1. Current Density Distribution

The current density distributions of the electrolyte at different voltages are shown in Figure 4. It can be clearly seen from the figure that as the operating voltage increases, the electrolyte current density gradually increases, and the distribution of the current density at the inlet and outlet gradually becomes uneven. This is because the electrolysis reaction rate at the entrance is accelerated, which consumes a large amount of H2O and CO2, resulting in insufficient reactant concentration in the middle and exit parts of the electrolytic cell.

3.2.2. Gas Composition Distribution

Figure 5 and Figure 6 show the distributions of H2 and CO concentrations at different voltages. Obviously, H2 and CO concentrations increase with an increase in applied voltage, due to the progress of electrochemical reactions. When the applied voltage is 1.2 V, the average H2 concentration reaches 0.27, corresponding to a conversion ratio of H2O of 2.5%. As the voltage continues to increase, the conversion ratio increases to 82.5% at an applied voltage of 1.6 V. Simultaneously, the conversion of CO2 also enhances. However, contributions to the CO2 reduction from the direct CO2 electrochemical reduction and reverse WGS reaction need to be further investigated. Furthermore, at the investigated temperature of 750 °C, the amount of CH4 production is minimal [39]. Thus, the impact of MSR on the overall performance can be considered negligible.

3.2.3. Temperature Distribution

Figure 7 shows the temperature distribution in the flat tubular SOEC under different voltages. It can be seen from the figure that with an increase in operating voltage, the average temperature increases and the distribution becomes more uneven. Inside the SOEC, a few electrochemical and chemical reactions take place simultaneously, including the water–gas shift reaction, methanation reaction, and the electrochemical reduction of H2O and CO2. Among these, the electrochemical reduction reactions are both endothermic processes, while the two chemical reactions are both exothermic reactions. Thus, the temperature distribution in the SOEC is a combined result of the interplay of these competing thermal effects. At low voltages of 1.0 V and 1.2 V, concentrations of H2 and CO are relatively low, as shown in Figure 5 and Figure 6. Thus, WGS and MR reactions are hindered due to the lack of reactants and the existence of a high concentration of products. As a result, the temperature of SOEC is low, close to the boundary temperature of 1023 K. In contrast, with applied voltages increasing to 1.4 V and 1.6 V, chemical reactions are dominant and more heat is released, increasing the SOEC temperature. Furthermore, it can be noted that the highest temperature is observed at the middle part of the SOEC.

3.2.4. Stress Distribution

The distribution of the first principal stress inside the SOEC at different voltages is shown in Figure 8. The first principal stress can be used to predict the crack and fail of brittle materials such as ceramic along planes perpendicular to the direction of maximum tensile stress. Thus, in this work, the larger the first principal stress, the higher the failure probability of an SOEC. The main origin of the stress is thermal stress due to the temperature gradient, shown in Figure 7, generated inside the SOEC during operation. Therefore, with an increase in the applied voltage, the maximum first principal stress also increases, from 0.19 MPa at an applied voltage of 1 V to 301 MPa at an applied voltage of 1.6 V. The compressive strength of YSZ is around 1.6 GPa [40]. Considering that the anode is highly porous and the electrolyte layer is very thin (~10–20 μm), a higher stress increases the failure probability. In addition, at 1.6 V, the first principal stress is higher near both inlet and outlet regions, while the central section presents a lower stress level. This distribution corresponds to the fact that the highest temperature occurs in the middle part, leading to the highest temperature gradients near the inlet and outlet areas.

3.2.5. Effect of Operating Pressure

In this section, the effect of operating pressure on SOEC performance and physical field distributions are presented. Figure 9 gives the variation in I–V curves with the operating temperature. Clearly, a high operating pressure is beneficial to electrochemical performance; the higher the operating pressure, the higher the current density at the same applied voltage. With an operating pressure of 1 atm, the average current density is 7116 A/m2 at an applied voltage of 1.6 V; the value increases to 9817 A/m2 under an operating pressure of 2 atm, and further reaches 12,586 A/m2 under 10 atm. This trend is in accordance with the previous modeling study on a planar SOEC stack [41,42] and the experimental study [43]. The high operating pressure could improve cell performance through the following two mechanisms. First, a high operating pressure improves the reactant volumetric concentration. Thus, especially in regions near the outlet, the increased concentration of reactants could improve electrochemical reaction rates. Second, the higher operating pressure may also promote the transport of gas molecules, and thus lower the concentration overpotential. This is demonstrated by the more substantial performance improvement observed at higher current densities.
Figure 10 gives temperature distributions at different operating pressures. Evidently, with an increase in operating pressure, the temperature distribution becomes more uneven. Figure 11 summarizes the highest and lowest temperatures at each operating pressure. At 1 atm, the temperature difference between the highest and the lowest is 380 K, while at 10 atm, the difference increases to 490 K. Furthermore, a shift in the peak temperature region from near the inlet to near the outlet can also be observed. As mentioned above, the electrochemical reactions are endothermic and the chemical reactions are exothermic. The shift in temperature distribution implies a shift in the distributions of sites where the various reactions take place.
Correspondingly, increasing uneven distribution leads to higher maximum first principal stress, as shown by Figure 12. At 1.6 V, the maximum first principal stress increases from 301 MPa at an operating pressure of 1 atm to 465 MPa at an operating pressure of 10 atm. Thus, when aiming to achieve higher performance by operating the SOEC at elevated pressures, the potential cell degradation caused by the higher first principal stresses should also be taken into consideration.

4. Conclusions

This study established a multi-physics model to evaluate the performance and mechanical behavior of a flat tubular SOEC during H2O–CO2 co-electrolysis. Results reveal that higher operating voltages cause more uneven distributions of current density and temperature. Furthermore, the highest temperature shifts toward the middle region, leading to increased temperature gradients near the inlet and outlet, and thus the first principal stress. Elevated operating pressure is beneficial to enhancing electrochemical performance, while simultaneously increasing the maximum first principal stress from 300 MPa to 465 MPa as pressure rises from 1 atm to 10 atm, potentially compromising cell stability. In the future, a more detailed optimization should be performed considering the trade-offs between performance and stability.

Author Contributions

Conceptualization, C.C., Y.Z., H.Z. and Z.Z.; methodology, C.C., P.L., C.Y. and Z.Z.; validation, W.D., J.S., C.Y., B.M. and Y.Z.; formal analysis, C.C., W.D. and J.S.; investigation, C.C., W.D., P.L., C.Y. and H.L.; resources, Y.Z., H.Z. and Z.Z.; data curation, C.C., P.L. and H.Z.; writing—original draft preparation, C.C., W.D., J.S., P.L. and C.Y.; writing—review and editing, B.M., Y.Z., H.Z. and Z.Z.; visualization, P.L. and C.Y.; supervision, Y.Z. and Z.Z.; project administration, H.Z.; funding acquisition, Y.Z. and Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding from Shenzhen Energy Group Co., Ltd.

Data Availability Statement

Original contributions presented in this study are included in this article. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors acknowledge the management and research facilities of the Harbin Institute of Technology, Shenzhen.

Conflicts of Interest

Chaolong Cheng was employed by the Shenzhen Energy Group Co., Ltd. and Shenzhen Energy Innovation Technology Co., Ltd. Junfeng Shen was employed by the Shenzhen Energy Group Co., Ltd. Hui Li was employed by the Qingpeng Hydrogen Technology (Shenzhen) Co., Ltd. Hongying Zhang was employed by the Qingpeng Hydrogen Technology (Shenzhen) Co., Ltd. The remaining authors declare that this research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The authors declare that this study received funding from the Shenzhen Energy Group Co., Ltd. The funder was not involved in this study’s design, the collection, analysis, or interpretation of data, the writing of this article, or the decision to submit it for publication.

References

  1. Chi, J.; Yu, H. Water electrolysis based on renewable energy for hydrogen production. Chin. J. Catal. 2018, 39, 390–394. [Google Scholar]
  2. Kang, S.; Pan, Z.; Guo, J.; Zhou, Y.; Wang, J.; Fan, L.; Zheng, C.; Cha, S.W.; Zhong, Z. Scientometric analysis of research trends on solid oxide electrolysis cells for green hydrogen and syngas production. Front. Energy 2024, 18, 583–611. [Google Scholar] [CrossRef]
  3. Yuan, B.; Zhang, X.; Tang, C.; Wang, N.; Ye, S. How AI guided the development of green hydrogen production: In the case of solid oxide electrolysis cell? J. Mater. Inform. 2025, 5, 25. [Google Scholar] [CrossRef]
  4. Ferrete, F.; Molina, A.; Cabello González, G.M.; Moreno-Racero, Á.; Olmedo, H.; Iranzo, A. Solid Oxide Electrolyzers Process Integration: A Comprehensive Review. Processes 2025, 13, 2656. [Google Scholar] [CrossRef]
  5. Pan, Z.; Liu, Q.; Ni, M.; Lyu, R.; Li, P.; Chan, S.H. Activation and failure mechanism of La0. 6Sr0. 4Co0. 2Fe0. 8O3− δ air electrode in solid oxide electrolyzer cells under high-current electrolysis. Int. J. Hydrogen Energy 2018, 43, 5437–5450. [Google Scholar] [CrossRef]
  6. Cunjin, Z.; Shuaibo, Q.; Hui, G.; Peng, J. High hydrogen evolution activities of dual-metal atoms incorporated N-doped graphenes achieved by coordination regulation. J. Mater. Inform. 2024, 4, 1. [Google Scholar]
  7. Wang, S.; Chen, H.; Hu, S.; Zhao, W.; Chen, M.; Chen, D.; Xu, X. Characterization of Hydrogen-in-Oxygen Changes in Alkaline Electrolysis Hydrogen Production System and Analysis of Influencing Factors. Processes 2025, 13, 2517. [Google Scholar] [CrossRef]
  8. Dogu, D.; Gunduz, S.; Meyer, K.E.; Deka, D.J.; Co, A.C.; Ozkan, U.S. CO2 and H2O electrolysis using solid oxide electrolyzer cell (SOEC) with La and Cl-doped strontium titanate cathode. Catal. Lett. 2019, 149, 1743–1752. [Google Scholar] [CrossRef]
  9. Pan, Z.; Duan, C.; Pritchard, T.; Thatte, A.; White, E.; Braun, R.; O’hAyre, R.; Sullivan, N.P. High-yield electrochemical upgrading of CO2 into CH4 using large-area protonic ceramic electrolysis cells. Appl. Catal. B 2022, 307, 121196. [Google Scholar] [CrossRef]
  10. Cifrain, M.; Kordesch, K.V. Advances, aging mechanism and lifetime in AFCs with circulating electrolytes. J. Power Sources 2004, 127, 234–242. [Google Scholar] [CrossRef]
  11. Gülzow, E. Alkaline fuel cells: A critical view. J. Power Sources 1996, 61, 99–104. [Google Scholar] [CrossRef]
  12. International Energy Agency. 2014 Annual Report; International Energy Agency: Paris, France, 2014. [Google Scholar]
  13. Litster, S.; McLean, G. PEM fuel cell electrodes. J. Power Sources 2004, 130, 61–76. [Google Scholar] [CrossRef]
  14. Sishtla, C.; Koncar, G.; Platon, R.; Gamburzev, S.; Appleby, A.J.; Velev, O.A. Performance and endurance of a PEMFC operated with synthetic reformate fuel feed. J. Power Sources 1998, 71, 249–255. [Google Scholar] [CrossRef]
  15. Gardner, F.J.; Day, M.J.; Brandon, N.P.; Pashley, M.N.; Cassidy, M. SOFC technology development at Rolls-Royce. J. Power Sources 2000, 86, 122–129. [Google Scholar] [CrossRef]
  16. Luo, Y.; Shi, Y.; Li, W.; Cai, N. Comprehensive modeling of tubular solid oxide electrolysis cell for co-electrolysis of steam and carbon dioxide. Energy 2014, 70, 420–434. [Google Scholar] [CrossRef]
  17. Luo, Y.; Li, W.; Shi, Y.; Cao, T.; Ye, X.; Wang, S.; Cai, N. Experimental characterization and theoretical modeling of methane production by H2O/CO2 co-electrolysis in a tubular solid oxide electrolysis cell. J. Electrochem. Soc. 2015, 162, F1129. [Google Scholar] [CrossRef]
  18. Chen, B.; Xu, H.; Ni, M. Modelling of SOEC-FT reactor: Pressure effects on methanation process. Appl. Energy 2017, 185, 814–824. [Google Scholar] [CrossRef]
  19. Cui, T.; Xiao, G.; Yan, H.; Zhang, Y.; Wang, J.-Q. Numerical simulation and analysis of the thermal stresses of a planar solid oxide electrolysis cell. Int. J. Green Energy 2023, 20, 432–444. [Google Scholar] [CrossRef]
  20. Fan, P.; Li, G.; Zeng, Y.; Zhang, X. Numerical study on thermal stresses of a planar solid oxide fuel cell. Int. J. Therm. Sci. 2014, 77, 1–10. [Google Scholar] [CrossRef]
  21. Guan, W.; Du, Z.; Wang, J.; Jiang, L.; Yang, J.; Zhou, X.-D. Mechanisms of performance degradation induced by thermal cycling in solid oxide fuel cell stacks with flat-tube anode-supported cells based on double-sided cathodes. Int. J. Hydrogen Energy 2020, 45, 19840–19846. [Google Scholar] [CrossRef]
  22. Zheng, Y.; Chen, Z.; Zhang, J. Solid oxide electrolysis of H2O and CO2 to produce hydrogen and low-carbon fuels. Electrochem. Energy Rev. 2021, 4, 508–517. [Google Scholar] [CrossRef]
  23. Sun, X.; Chen, M.; Jensen, S.H.; Ebbesen, S.D.; Graves, C.; Mogensen, M. Thermodynamic analysis of synthetic hydrocarbon fuel production in pressurized solid oxide electrolysis cells. Int. J. Hydrogen Energy 2012, 37, 17101–17110. [Google Scholar] [CrossRef]
  24. Yu, C.; Pan, Z.; Zhang, H.; Chen, B.; Guan, W.; Miao, B.; Chan, S.H.; Zhong, Z.; Zhou, Y. Numerical multi-physical optimization of operating condition and current collecting setup for large-area solid oxide fuel cells. Front. Energy 2024, 18, 356–368. [Google Scholar] [CrossRef]
  25. Xu, H.; Chen, B.; Irvine, J.; Ni, M. Modeling of CH4-assisted SOEC for H2O/CO2 co-electrolysis. Int. J. Hydrogen Energy 2016, 41, 21839–21849. [Google Scholar] [CrossRef]
  26. Xiong, X.; Liang, K.; Ma, G.; Ba, L. Three-dimensional multi-physics modelling and structural optimization of SOFC large-scale stack and stack tower. Int. J. Hydrogen Energy 2023, 48, 2742–2761. [Google Scholar] [CrossRef]
  27. Kim, Y.J.; Lee, M.C. Numerical investigation of flow/heat transfer and structural stress in a planar solid oxide fuel cell. Int. J. Hydrogen Energy 2017, 42, 18504–18513. [Google Scholar] [CrossRef]
  28. Pianko-Oprych, P.; Zinko, T.; Jaworski, Z. A Numerical Investigation of the Thermal Stresses of a Planar Solid Oxide Fuel Cell. Materials 2016, 9, 814. [Google Scholar] [CrossRef] [PubMed]
  29. Stevenson, J.W.; Armstrong, T.R.; Carneim, R.D.; Pederson, L.R.; Weber, W.J. Electrochemical properties of mixed conducting perovskites La(1-x)M(x)Co(1-y)Fe(y)O(3-delta) (M = Sr, Ba, Ca). J. Electrochem. Soc. 1996, 143, 2722–2729. [Google Scholar] [CrossRef]
  30. Xu, H.R.; Chen, B.; Liu, J.; Ni, M. Modeling of direct carbon solid oxide fuel cell for CO and electricity cogeneration. Appl. Energy 2016, 178, 353–362. [Google Scholar] [CrossRef]
  31. Ali, S.; Sørensen, K.; Nielsen, M.P. Modeling a novel combined solid oxide electrolysis cell (SOEC)-Biomass gasification renewable methanol production system. Renew. Energy 2020, 154, 1025–1034. [Google Scholar] [CrossRef]
  32. Baxter, S.J.; Rine, M.; Min, B.; Liu, Y.; Yao, J. Near 100% CO2 conversion and CH4 selectivity in a solid oxide electrolysis cell with integrated catalyst operating at 450 °C. J. CO2 Util. 2022, 59, 101954. [Google Scholar] [CrossRef]
  33. Zhang, Z.; Yue, D.; Yang, G.; Chen, J.; Zheng, Y.; Miao, H.; Wang, W.; Yuan, J.; Huang, N. Three-dimensional CFD modeling of transport phenomena in multi-channel anode-supported planar SOFCs. Int. J. Heat Mass Transf. 2015, 84, 942–954. [Google Scholar] [CrossRef]
  34. Rui, W.; Lucas, R.P.; Yu, Z. Sulfur poisoning mechanism of LSCF cathode material in the presence of SO2: A computational and experimental study. J. Mater. Inform. 2023, 3, 3. [Google Scholar]
  35. Yu, C.; Guo, Z.; Pan, Z.; Zhou, Y.; Zhang, H.; Chen, B.; Tan, P.; Guan, W.; Zhong, Z. Performance and stress analysis of flat-tubular solid oxide fuel cells fueled with methane and hydrogen. Acta Mech. Solida Sin. 2024, 38, 402–414. [Google Scholar] [CrossRef]
  36. Jiang, C.; Gu, Y.; Guan, W.; Ni, M.; Sang, J.; Zhong, Z.; Singhal, S.C. Thermal Stress Analysis of Solid Oxide Fuel Cell with Z-type and Serpentine-Type Channels Considering Pressure Drop. J. Electrochem. Soc. 2020, 167, 044517. [Google Scholar] [CrossRef]
  37. Haberman, B.; Young, J. Three-dimensional simulation of chemically reacting gas flows in the porous support structure of an integrated-planar solid oxide fuel cell. Int. J. Heat Mass Transf. 2004, 47, 3617–3629. [Google Scholar] [CrossRef]
  38. Xi, C.; Sang, J.; Wu, A.; Yang, J.; Qi, X.; Guan, W.; Wang, J.; Singhal, S.C. Electrochemical performance and durability of flat-tube solid oxide electrolysis cells for H2O/CO2 co-electrolysis. Int. J. Hydrogen Energy 2022, 47, 10166–10174. [Google Scholar] [CrossRef]
  39. Xie, K.; Zhang, Y.; Meng, G.; Irvine, J.T.S. Direct synthesis of methane from CO2/H2O in an oxygen-ion conducting solid oxide electrolyser. Energy Environ. Sci. 2011, 4, 2218. [Google Scholar] [CrossRef]
  40. Shimonosono, T.; Ueno, T.; Hirata, Y. Mechanical and thermal properties of porous yttria-stabilized zirconia. J. Asian Ceram. Soc. 2019, 7, 20–30. [Google Scholar] [CrossRef]
  41. Du, Y.; Qin, Y.; Zhang, G.; Yin, Y.; Jiao, K.; Du, Q. Modelling of effect of pressure on co-electrolysis of water and carbon dioxide in solid oxide electrolysis cell. Int. J. Hydrogen Energy 2019, 44, 3456–3469. [Google Scholar] [CrossRef]
  42. Henke, M.; Willich, C.; Kallo, J.; Friedrich, K.A. Theoretical study on pressurized operation of solid oxide electrolysis cells. Int. J. Hydrogen Energy 2014, 39, 12434–12439. [Google Scholar] [CrossRef]
  43. Bernadet, L.; Gousseau, G.; Chatroux, A.; Laurencin, J.; Mauvy, F.; Reytier, M. Influence of pressure on solid oxide electrolysis cells investigated by experimental and modeling approach. Int. J. Hydrogen Energy 2015, 40, 12918–12928. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of co-electrolysis by SOEC; (b) schematic illustration of a typical I–V curve of SOEC and corresponding losses; (c) schematic image and photograph of flat tubular SOECs; (d) photograph of planar SOECs.
Figure 1. (a) Schematic diagram of co-electrolysis by SOEC; (b) schematic illustration of a typical I–V curve of SOEC and corresponding losses; (c) schematic image and photograph of flat tubular SOECs; (d) photograph of planar SOECs.
Processes 13 03192 g001
Figure 2. SOEC structure diagram.
Figure 2. SOEC structure diagram.
Processes 13 03192 g002
Figure 3. I–V curves of H2O–CO2 co-electrolysis using flat tubular SOEC.
Figure 3. I–V curves of H2O–CO2 co-electrolysis using flat tubular SOEC.
Processes 13 03192 g003
Figure 4. Current density distribution of electrolyte at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Figure 4. Current density distribution of electrolyte at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Processes 13 03192 g004
Figure 5. H2 distribution at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Figure 5. H2 distribution at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Processes 13 03192 g005
Figure 6. CO distribution at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Figure 6. CO distribution at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Processes 13 03192 g006
Figure 7. Temperature distribution under different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Figure 7. Temperature distribution under different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Processes 13 03192 g007
Figure 8. First principal stress distribution at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Figure 8. First principal stress distribution at different voltages of (a) 1 V; (b) 1.2 V; (c) 1.4 V; and (d) 1.6 V.
Processes 13 03192 g008
Figure 9. Variation in I–V curves with operating pressure.
Figure 9. Variation in I–V curves with operating pressure.
Processes 13 03192 g009
Figure 10. Temperature distributions under different operating pressures of (a) 1 atm; (b) 2 atm; (c) 5 atm; and (d) 10 atm.
Figure 10. Temperature distributions under different operating pressures of (a) 1 atm; (b) 2 atm; (c) 5 atm; and (d) 10 atm.
Processes 13 03192 g010
Figure 11. Temperature variation with a change in operating pressure.
Figure 11. Temperature variation with a change in operating pressure.
Processes 13 03192 g011
Figure 12. Variation in maximum first principal stress along operating pressure.
Figure 12. Variation in maximum first principal stress along operating pressure.
Processes 13 03192 g012
Table 1. Comparison of characteristics between different electrolyzers.
Table 1. Comparison of characteristics between different electrolyzers.
TechnologyAECsPEMECsSOECs
ElectrolytePotassium hydroxideProton exchange membraneSolid oxide
Operating temperature70–90 °C80–150 °C600–1000 °C
Charge carrierOHH+O2−
Electrolyte stateImmobilized liquidHydrated solidSolid
ElectrodesTransition metalsCarbonCeramic/Metal cermet
CatalystPlatinumPlatinumNi-Cermet
ReactantPure H2OPure H2OH2O(g), CO2
Efficiency
(% LHV of H2)
56–6956–83>80
References[10,11,12][12,13,14][12,15]
Table 2. Model parameter table.
Table 2. Model parameter table.
ParametersValueUnit
Ionic conductivity of cathode [27] 9.5 × 10 7 T exp 1150 T S·m−1
Ionic conductivity of anode [27] 4.2 × 10 7 T exp 1200 T S·m−1
Ionic conductivity of electrolyte [27] 33.4 × 10 3 exp 10300 T S·m−1
Electronic conductivity of anode [28]30,300S·m−1
Electronic conductivity of cathode [29]17,000S·m−1
Electronic conductivity of interconnect [28]769,000S·m−1
AVa [30]2.14 × 105m−2
AVc [30]2.14 × 105m−2
Table 3. Material properties used in the model [33].
Table 3. Material properties used in the model [33].
ComponentPorosityPermeability
(m2)
Thermal Conductivity
(W·m−1·K−1)
Thermal Capacity
(J·Kg−1·K−1)
Anode functional layer0.231 × 10−1211450
Anode support layer0.461 × 10−1011450
Electrolyte--2.7550
Cathode layer0.31 × 10−126430
Interconnect--20550
Table 4. Mechanical property parameters of solid mechanics in SOEC [36].
Table 4. Mechanical property parameters of solid mechanics in SOEC [36].
ComponentE (GPa)ν (-)CET (10−6·K−1)
Anode functional layer2130.311.4
Anode2200.312.5
Electrolyte2050.310.3
Cathode1600.311.4
Table 5. Geometric parameters of solid oxide electrolytic cell.
Table 5. Geometric parameters of solid oxide electrolytic cell.
SOFC_S
Ni-3YSZ supporting layer (mm3)98.6 × 46 × 4.7
NiO + 8YSZ anode functional layer (mm3)85.5 × 41 × 0.02
YSZ electrolyte layer (mm3)85.5 × 41 × 0.01
LSCF perovskite cathode layer (mm3)85.5 × 41 × 0.02
Number of steam channels13
Number of metallic alloy interconnect ribs26
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cheng, C.; Ding, W.; Shen, J.; Liao, P.; Yu, C.; Miao, B.; Zhou, Y.; Li, H.; Zhang, H.; Zhong, Z. Multi-Physics Coupling Simulation of H2O–CO2 Co-Electrolysis Using Flat Tubular Solid Oxide Electrolysis Cells. Processes 2025, 13, 3192. https://doi.org/10.3390/pr13103192

AMA Style

Cheng C, Ding W, Shen J, Liao P, Yu C, Miao B, Zhou Y, Li H, Zhang H, Zhong Z. Multi-Physics Coupling Simulation of H2O–CO2 Co-Electrolysis Using Flat Tubular Solid Oxide Electrolysis Cells. Processes. 2025; 13(10):3192. https://doi.org/10.3390/pr13103192

Chicago/Turabian Style

Cheng, Chaolong, Wen Ding, Junfeng Shen, Penghui Liao, Chengrong Yu, Bin Miao, Yexin Zhou, Hui Li, Hongying Zhang, and Zheng Zhong. 2025. "Multi-Physics Coupling Simulation of H2O–CO2 Co-Electrolysis Using Flat Tubular Solid Oxide Electrolysis Cells" Processes 13, no. 10: 3192. https://doi.org/10.3390/pr13103192

APA Style

Cheng, C., Ding, W., Shen, J., Liao, P., Yu, C., Miao, B., Zhou, Y., Li, H., Zhang, H., & Zhong, Z. (2025). Multi-Physics Coupling Simulation of H2O–CO2 Co-Electrolysis Using Flat Tubular Solid Oxide Electrolysis Cells. Processes, 13(10), 3192. https://doi.org/10.3390/pr13103192

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