Next Article in Journal
Multiband Multisine Excitation Signal for Online Impedance Spectroscopy of Battery Cells
Previous Article in Journal
Enhanced OCV Estimation in LiFePO4 Batteries: A Novel Statistical Approach Leveraging Real-Time Knee/Elbow Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Study of the Effects of Heat Loss and Solid Thermal Conductivity on Syngas Production for Fuel Cells

1
Ocean College, Binzhou Polytechnic, Binzhou 256603, China
2
School of Transportation and Vehicle Engineering, Shandong University of Technology, Zibo 255049, China
*
Author to whom correspondence should be addressed.
Batteries 2025, 11(5), 187; https://doi.org/10.3390/batteries11050187
Submission received: 12 March 2025 / Revised: 6 May 2025 / Accepted: 8 May 2025 / Published: 9 May 2025
(This article belongs to the Special Issue Challenges, Progress, and Outlook of High-Performance Fuel Cells)

Abstract

:
Syngas can be used as feedstock for efficient energy conversion in solid oxide fuel cells (SOFCs). In the current paper, the conversion efficiency of methane to synthesis gas (H2 and CO) within a two-layer porous media reactor is investigated by a one-dimensional two-temperature model. A detailed chemical reaction mechanism GRI-Mech 1.2 is used to describe the chemical processes. Attention is focused on CO2 content in the methane/air mixture, heat loss to the surroundings, and solid thermal conductivity on temperature distribution and conversion efficiency. Numerical results show that addition of CO2 to the methane/air mixture improves the conversion efficiency. For a molar ratio of CO2/CH4 = 1, the conversion efficiency reaches 44.8%. An increase in heat loss to the surroundings leads to a decrease in conversion efficiency. A greater solid thermal conductivity can improve the conversion efficiency.

Graphical Abstract

1. Introduction

Solid oxide fuel cells (SOFCs) offer advantages such as high efficiency and environmental friendliness. Compared to traditional power generation methods, SOFCs can achieve a power generation efficiency of up to 60% and do not emit pollutants like NO x , making them highly environmentally friendly.
The syngas for SOFCs is mostly obtained through the cracking of hydrocarbon fuel/air mixtures. H2 has the advantages of high calorific value per unit mass and almost no environmental pollution during combustion. Hence, it is considered as an alternative to hydrocarbon fuel in the future. Although the reserves of H2 are very rich, most of them exist in water and hydrocarbon fuels in the form of ions, which need to be extracted by certain means [1,2,3,4]. The primary methods for producing syngas include thermochemical water decomposition, biomass processing, electrolysis, coal and biogas utilization, steam reforming, partial oxidation, and autothermal reforming. Among these, the most prevalent method is the production of syngas through the reforming of hydrocarbon fuels [5,6].
Enhancing the conversion efficiency of hydrocarbon fuels is a hot topic in the field of SOFCs. Among various methods, the most advanced approach is to modify anodes by an active and stable coking catalytic layer as considered in the review by L. Fan et al. [7]. Another well-known approach is to transform fuel into syngas in the reactor with structured catalysts on heat-conducting substrates with nanocomposite active components possessing a high oxygen mobility and stability to coking [8]. The partial oxidation of rich fuels in porous media represents a promising technology. The porous media reactor has garnered significant attention due to its extensive internal surface area and intricate gas flow pathways, which enhance heat transfer between the gas and solid phases. Syngas production by partial oxidation of fuel rich in porous media has been extensively studied [9,10].
Toledo et al. [11] carried out an experimental study on syngas production in inert porous media for three different hydrocarbon fuels. A one-dimensional steady-state numerical model was established by using the two-temperature model and a GRI-Mech 3.0 detailed chemical reaction mechanism. The critical equivalence ratio of the three fuels for superadiabatic combustion was presented. Araya et al. [12] conducted experimental and theoretical studies on syngas production in an inert porous medium composed of Al2O3 spheres by adding a certain proportion of water vapor into natural gas. They found that the output of H2 increases with the increase in the proportion of steam in the mixture. Zeng et al. [13,14] carried out an experimental study on the fuel-rich combustion characteristics of a two-layer reactor by adding different proportions of CO2 into a CH4/air mixture, and they established a one-dimensional, steady-state, two-temperature model. The Peter mechanism of 17 species and 58 elementary reactions was adopted in their computations. The radiation heat loss of solids at the inlet and outlet to the environment was considered. Their results showed that CO2 injection can improve the conversion efficiency of syngas.
In the experiment, heat loss to the surroundings is inevitable and the solid thermal conductivity may have significant influence on combustion characteristics and syngas production. Therefore, in the process of numerical studies, the wall heat loss should be considered and the appropriate heat loss coefficient should be selected to reflect the actual combustion situation of the reactor. Although fuel-lean [15,16] and fuel-rich [17,18,19,20] combustion in porous media has been numerically investigated, these effects have not been fully considered in previous research. The performance and durability of a 100 W SOFC stack system was investigated by Alenazey et al. [21] under the conditions of steady-state and current load cycling. It was found that the total degradation rates were 54, 297, and 370 mV within 72 h for three different stack temperatures of 650, 700, and 750 °C, respectively. However, a steady drop as small as 0.5% was observed during the 24 current load cycles.
This paper aims to clarify the effects of heat loss and thermal conductivity on syngas production. The research results provide guidance for improving conversion efficiency through optimized combustion conditions. To this end, a one-dimensional two-temperature steady-state model of a two-layer porous medium reactor is established. The two-temperature model means that the gas and solid phases have their own temperatures. The GRI-Mech 1.2 detailed mechanism [22] is adopted to deal with the methane/air combustion in porous media. Different proportions of CO2 are added to the methane/air mixture to explore the effect of CO2 injection on conversion efficiency. A heat loss coefficient is introduced to account for the heat loss to the surroundings and its influence on syngas production. The impact of material properties on conversion efficiency is also examined.

2. Mathematical Model

2.1. Model Description

The research object of this paper is a two-layer porous medium reactor designed by Zeng et al. [14]. The inner diameter of the reactor is 30 mm, and its length is 80 mm, as illustrated in Figure 1. The porous medium material used in the reactor is Al2O3 balls. The balls used in the upstream section have a diameter of 2–3 mm, a filling height of 20 mm, and a porosity of 0.4, serving as the diffusion layer for the flame. In the downstream section, the balls have a diameter of 7.5 mm, a filling height of 60 mm, and a porosity of 0.46, functioning as the flame-supporting layer. For simplification, it is assumed that the diameter of the balls in the upstream section is 2.5 mm. The fuel used in the model is a methane/air mixture with varying proportions of CO 2 .

2.2. Governing Equation

In this paper, a one-dimensional model of the reactor is established. Considering the complexity of heat transfer between the gas and solid phases in the reactor, the model is simplified with the following assumptions:
  • Gas radiation is ignored;
  • Porous media are isotropic and have an inert optical thickness;
  • The gas in porous media is an ideal gas and the flow in porous media is laminar;
  • A heat loss coefficient β is used to account for the heat loss to the surroundings.
Based on the above assumptions, the following governing equations are established [15].
Continuity equation:
d ( ε ρ g u g ) d x = 0
where ε is the porosity, and ρ g and u g are the gas density and mixture velocity, respectively.
Momentum equation:
d ( ε ρ g u g u g ) d x = ε d p d x + d d x μ d u g d x Δ p Δ x
Gas energy equation:
d ( ε ρ g c g u g T g ) d x k ρ g ε Y k c gk D km d Y k d x d T g d x = d d x λ g ε d T g d x h v ( T g T s ) ε k W k h k ω ˙ k
where c g is the specific heat of the gas mixture, λ g is the thermal conductivity of the gas mixture, ω ˙ k and W k are the reaction rate and molecular weight of the i-th component in the mixture, and h v is the volumetric convective heat transfer coefficient, which is calculated by the following equations [15]:
h v = 6 N u v S λ g d 2
N u v = 2 + 1.1 P r 1 / 3 R e 0.6
Solid energy equation:
d d x ( λ s + λ rad ) d T s d x + h v ( T g T s ) β ( T s T 0 ) = 0
where β is the heat loss coefficient, and λ rad is the radiative thermal conductivity of Al2O3 spheres.
λ rad = 32 ε σ d 9 ( 1 ε ) T s 3
Species conservation equation:
d d x ε ρ g Y k u g = d d x ε ρ g D km d Y k d x + ε ω ˙ k W k
where Y k is the mass fraction of the k-th component in the gas mixture.
The equation of state for an ideal gas is as follows:
P = ρ g R T g
where R is the general constant of gas. The gas density is obtained from the ideal gas equation. The gas property parameters are computed using the Chemkin and Transport packages provided by Kee et al. [23].

2.3. Boundary Conditions

The boundary conditions used in the model are given as follows.
Reactor inlet:
u g = u g , in , Y k = Y k , in , T g = T g , in , λ s + λ rad d T s d x = ξ σ T s 4 T 0 4
Reactor outlet:
d u g d x = d Y k d x = d T g d x = d T s d x = 0 , λ s + λ rad d T s d x = ξ σ T s 4 T 0 4
Here, ξ is the surface emissivity of porous media.

2.4. Parameter Definition

The conversion efficiency of syngas is specified as
η = Y H 2 , out × LHV H 2 + Y CO , out × LHV CO Y CH 4 , in × LHV CH 4
where LHV is the low calorific value of components.

2.5. Chemical Mechanism

The chemical reaction mechanism has a significant effect on the temperature distribution and composition in porous media. The simplified mechanism is not sufficient to fully account for the partial combustion of methane in the reactor. Therefore, the detailed chemical reaction mechanism of GRI-Mech 1.2 is adopted in this paper, which consists of 32 species and 175 elementary reactions. ANSYS 15.0 is used in the numerical calculation.

2.6. Meshing and Solution Strategy

Based on the volume-averaged method, the burner is simplified as a regular geometric shape. A uniform square grid of 0.25 mm is applied throughout the computational domain, and a grid independence test is performed. The commercial software Fluent 15.0 is employed to solve the conservation equations with prescribed boundary conditions. The pressure–velocity coupling utilizes the SIMPLE algorithm. Second-order difference schemes are adopted for the gas and solid energy equations and the momentum equation. To simulate methane ignition, a 20 mm long high-temperature zone (1800 K) is initialized at the interface between sections. The convergence criteria are set to 1 × 10 6 for energy equations and 1 × 10 3 for other equations.

3. Results and Analysis

3.1. Effect of CO2 Injection

In this paper, the volume flow rate and equivalence ratio of air are fixed at 5 L/min and 1.5, respectively. Figure 2 depicts the gas and solid temperatures for four different molar ratios of CO2/CH4. It is evident that the gas temperature increases sharply within the reaction zone. The heat released during combustion is rapidly transferred to the solid porous media through convection heat transfer, resulting in a swift increase in solid temperature as well. However, a temperature difference exists between the gas and solid phases, which is referred to as the thermal nonequilibrium phenomenon. Consequently, it is necessary to model the energy equations for gas and solid separately in the numerical study. In the upstream of the combustion zone, the solid temperature is higher than the gas temperature, and the fresh mixture is preheated by the solid porous media. In the combustion area, the gas temperature rises above that of the solid temperature. In the downstream of the combustion area, the temperature difference between the gas and the solid gradually decreases, indicating that heat is being transferred from the gas to the solid [24].
As the molar ratio of CO2/CH4 increases from 0 to 1, the flame front gradually moves downstream. Specifically, when the molar ratio of CO2/CH4 reaches 1, the flame is located approximately 17 mm downstream of the interface between the two sections, and the peak gas temperature reaches 1950 K. The preheating zone is lengthened, and more heat is stored in the solid matrix, which enhances the preheating effect on the fresh mixture, resulting in an increase in the peak gas temperature. As the molar ratio of CO2/CH4 increases, the temperature difference between the gas and solid phases also increases. This phenomenon can be attributed to two main reasons. Firstly, increasing the proportion of CO2 injection leads to an increase in the inlet air velocity, which shortens the residence time of the gas in the porous media and weakens the heat transfer effect between the gas and solid. Secondly, as the flame propagates downstream, the diameter of the small ball increases to 7.5 mm. With the increase in the diameter of the ball, the convective heat transfer coefficient decreases under the same conditions.
Figure 3 illustrates the mole fraction of the main components near the reaction zone for molar ratios of CO2/CH4 = 0 and 1. The chemical reaction occurs in a very narrow area, approximately 3 mm thick. Within the reaction zone, the mole fractions of CH4 and O2 decrease rapidly, indicating that the combustion reaction is taking place. The main components of syngas quickly reach their peak values and remain stable. For molar ratios of CO2/CH4 = 0 and 1, the distribution of each component is similar, except for CO2 in the reaction zone. For a molar ratio of CO2/CH4 = 1, the mole fraction of CO2 initially decreases through the main elementary reaction reactions 2H + CO2 ↔ H2 + CO2 and CH + CO2 ↔ HCO + CO, and then increases via elementary reaction O2 + CO ↔ O + CO2, suggesting that CO2 first participates in the reaction as a reactant and is subsequently generated as a product.
Figure 4 presents the molar fraction of the main syngas components in the products and the conversion efficiency for different molar ratios of CO2/CH4, with experimental results included for comparison.
It can be observed from the figure that the numerical results in this study show good agreement with the experimental data. The experimental results indicate that conversion efficiency increases with higher molar ratios of CO2/CH4. However, predictions show a significant drop in reforming efficiency at a CO2/CH4 ratio of 0.5. This discrepancy may be attributed to the model used in this study. We employed the volume-averaged method for modeling, which could lead to deviations between the predictions and experimental results. Additionally, the predictions utilize GRI 1.2 rather than the most recent chemical kinetics, which may also contribute to the observed deviations.
As the molar ratio of CO2/CH4 increases, the molar fractions of CO2 and CO rise, while the molar fraction of H2 slightly decreases. Although the molar fraction of CO2 at the outlet increases, the net production of CO2 decreases because CO2 participates in the reverse reaction of water to gas, namely CO2 + H2 ↔ H2O + CO. In this reaction, H2 also acts as a reactant, leading to the formation of CO and a consequent reduction in H2 concentration. As the CO2/CH4 ratio increases from 0 to 1, the conversion efficiency of methane/air improves from 37.1% to 44.7%.

3.2. Effect of Heat Loss

Figure 5 shows the temperature curves of gas and solid phases under different heat loss coefficients. The heat loss coefficient, denoted as β , was experimentally evaluated to be 330 W / ( m 3 · K ) for a well-insulated porous burner [25]. The experimental values of Zeng et al. [14] are also shown for comparison. For the molar ratio of CO2/CH4 = 0, the numerical results for β = 1000 W / ( m 3 · K ) and β = 2000 W / ( m 3 · K ) are in good agreement with the experimental values. With the increase in β , the heat loss through the wall to the surroundings increases, and the temperatures of gas and solid phases decrease, which are obviously lower than Zeng’s experimental values. When the molar ratio of CO2/CH4 is 0 and β exceeds 1000 W / ( m 3 · K ) , the flame propagates downstream and finally moves out to the reactor outlet. This phenomenon is not shown in Figure 5 for the sake of clarity. This β value is defined as a critical value for stable combustion in the reactor. For the molar ratio of CO2/CH4 = 1, the critical value of β is 2500 W / ( m 3 · K ) . This is consistent with the results of Shi et al. [17] that the wall heat loss should not exceed 70% of the total heat of the reactor. The flame propagates downstream by predictions for CO2/CH4 = 1 with different heat loss coefficients compared to the experiments. This discrepancy may stem from the inappropriate heat loss coefficient applied in the computations.
Figure 6 shows the molar fractions of the main syngas components in the products and conversion efficiency as a function of β for molar ratios of CO2/CH4 = 0 and 1.
It can be observed from the figure that the increase in β leads to a decrease in the temperatures of gas and solid phases. At the same time, a decrease in the molar fraction of H2 and a slight increase in the molar fraction of CO are observed as β is increased. However, the conversion efficiency decreases with β, which indicates that a greater conversion can be obtained for the lower β.

3.3. Effect of Thermal Conductivity

In the packing bed, the gas–solid convective heat transfer coefficient and thermal conductivity are very important for heat transfer in a porous media reactor. The effects of thermal conductivity on the temperature distribution of gas and solid phases and the conversion efficiency of syngas are discussed. In this study, the heat loss coefficient β = 1000 W / ( m 3 · K ) and the solid surface emissivity ξ = 0.8 are kept unchanged, while λ s is varied.
Figure 7 shows the temperature distributions with CO2/CH4 = 0.5 for different solid thermal conductivities. The thermal conductivity of the benchmark is 0.32 W / ( m · K ) . As shown in Figure 7, the thermal conductivity of the solid has little effect on the gas temperature, but it has an obvious effect on the solid’s temperature. The effective thermal conductivity of the packing bed consists of two parts, which are the thermal conductivity of solid material λs and radiative thermal conductivity λrad. With the increase in λs, the heat conduction in the packing bed is enhanced, the solid temperature gradient becomes smaller, the preheating effect of the fresh mixture in the upstream is reduced, and the flame moves downstream. For λs = 0.16 W / ( m · K ) , the flame is located at the interface of two porous media.
Table 1 presents the main components of syngas in the products and the conversion efficiency. It can be observed that the lowest molar fractions of H2 and CO are obtained for the benchmark thermal conductivity value of 0.32 W / ( m · K ) . When λs is increased from 0.32 W / ( m · K ) to 0.64 W / ( m · K ) , the conversion efficiency increases from 39.5% to 42.9%. However, further increase in λs from 0.64 W / ( m · K ) to 1.28 W / ( m · K ) leads to a negligible decrease in conversion efficiency. This indicates that increases in λs lead to a slight increase in conversion efficiency.

4. Conclusions

The partial oxidation of rich fuels in porous media is a promising technology for supplying the syngas used in solid oxide fuel cells. In this paper, a one-dimensional steady-state model of a two-layer porous media reactor is established and the detailed chemical reaction mechanism of GRI-Mech 1.2 is used to study the partial combustion of methane/air with injection of CO2 in the mixture. The effects of molar ratio of CO2/CH4, heat loss coefficient and solid thermal conductivity on gas and solid temperature distribution, syngas production, and conversion efficiency are analyzed. The main conclusions are as follows:
  • Injection of CO2 into methane/air mixture can improve the conversion efficiency. As the CO2/CH4 ratio increases from 0 to 1, the conversion efficiency of the methane/air mixture improves from 37.1% to 44.7%;
  • The conversion efficiency decreases from 0.15 to 0.12 when the heat loss coefficient is increased from 1000 W / ( m 3 · K ) to 2000 W / ( m 3 · K ) ;
  • The thermal conductivity of porous media primarily affects the temperature of the solid and the conversion efficiency. When the thermal conductivity is increased from 0.32 W / ( m · K ) to 0.64 W / ( m · K ) for a molar ratio of CO2/CH4 = 0.5, the conversion efficiency increases from 39.5% to 42.9%.

Author Contributions

Conceptualization: X.W.; methodology: X.W.; formal analysis and investigation: X.W., M.Y., Z.L. and Z.W.; writing—original draft preparation: X.W.; writing—review and editing: X.Z., J.S., X.K. and J.L.; funding acquisition: X.Z. and J.S.; supervision: X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (NSFC Grant No. 51876107).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors are grateful to many colleagues with whom they had the privilege to interact and collaborate over the years and whose work is partially referenced in this article.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

Nomenclature

SymbolDefinition
c g specific heat capacity, J·kg−1·K−1
dthe diameter of the ball, mm
h k the molar enthalpy of component i, J·mol−1
h v the volume convective heat transfer coefficient
Mmole fraction
N u Nusselt number
Ppressure, Pa
P r Prandtl number
R e Reynolds number
Ttemperature, K
ugas velocity, m·s−1
W k molecular weight of the k component
Y k mass fraction of component k
ω ˙ k the reaction rate of the k component
β heat loss coefficient, W·m−3·K−1
ε porosity
η conversion efficiency
λ eff effective thermal conductivity, W·m−1·K−1
λ rad radiative thermal conductivity, W·m−1·K−1
λ s thermal conductivity for solid, W·m−1·K−1
ρ density, kg·m−3
σ Stefan–Boltzmann constant, W·m−2·K−4
ξ surface emissivity of porous media

References

  1. Al-Hamamre, Z.; Al-Zoubi, A. The use of inert porous media based reactors for hydrogen production. Int. J. Hydrogen Energy 2010, 35, 1971–1986. [Google Scholar] [CrossRef]
  2. Toledo, M.G.; Utria, K.S.; González, F.A.; Zuñiga, J.P.; Saveliev, A.V. Hybrid filtration combustion of natural gas and coal. Int. J. Hydrogen Energy 2012, 37, 6942–6948. [Google Scholar] [CrossRef]
  3. Dhamrat, R.S.; Ellzey, J.L. Numerical and experimental study of the conversion of methane to hydrogen in a porous media reactor. Combust. Flame 2006, 144, 698–709. [Google Scholar] [CrossRef]
  4. Wang, Y.; Zeng, H.; Shi, Y.; Cao, T.; Cai, N.; Ye, X.; Wang, S. Power and heat co-generation by micro-tubular flame fuel cell on a porous media burner. Energy 2016, 109, 117–123. [Google Scholar] [CrossRef]
  5. Smith, C.H.; Leahey, D.M.; Miller, L.E.; Ellzey, J.L. Conversion of wet ethanol to syngas via filtration combustion: An experimental and computational investigation. Proc. Combust. Inst. 2011, 33, 3317–3324. [Google Scholar] [CrossRef]
  6. Dorofeenko, S.O.; Polianczyk, E.V. Conversion of hydrocarbon gases to synthesis gas in a reversed-flow filtration combustion reactor. Chem. Eng. J. 2016, 292, 183–189. [Google Scholar] [CrossRef]
  7. Fan, L.; Li, C.; Aravind, P.V.; Cai, W.; Han, M.; Brandon, N. Methane reforming in solid oxide fuel cells: Challenges and strategies. J. Power Sources 2022, 538, 231573. [Google Scholar] [CrossRef]
  8. Fukuhara, C.; Yamamoto, K.; Makiyama, Y.; Kawasaki, W.; Watanabe, R. A metal-honeycomb-type structured catalyst for steam reforming of methane: Effect of preparation condition change on reforming performance. Appl. Catal. A 2015, 492, 190–200. [Google Scholar] [CrossRef]
  9. Bingue, J.P.; Saveliev, A.V.; Fridman, A.A.; Kennedy, L.A. Hydrogen production in ultra-rich filtration combustion of methane and hydrogen sulfide. Int. J. Hydrogen Energy 2002, 27, 643–649. [Google Scholar] [CrossRef]
  10. Toledo, M.; Gracia, F.; Caro, S.; Gómez, J.; Jovicic, V. Hydrocarbons conversion to syngas in inert porous media combustion. Int. J. Hydrogen Energy 2016, 41, 5857–5864. [Google Scholar] [CrossRef]
  11. Toledo, M.; Bubnovich, V.; Saveliev, A.; Kennedy, L. Hydrogen production in ultrarich combustion of hydrocarbon fuels in porous media. Int. J. Hydrogen Energy 2009, 34, 1818–1827. [Google Scholar] [CrossRef]
  12. Araya, R.; Araus, K.; Utria, K.; Toledo, M. Optimization of hydrogen production by filtration combustion of natural gas by water addition. Int. J. Hydrogen Energy 2014, 39, 7338–7345. [Google Scholar] [CrossRef]
  13. Zeng, H.Y.; Wang, Y.Q.; Shi, Y.X.; Cai, N.S. Experimental study for partial oxidation reforming of CH4/CO2 with rich combustion in a porous media burner. J. Eng. Thermophys. 2017, 38, 2269–2275. [Google Scholar]
  14. Zeng, H.; Wang, Y.; Shi, Y.; Ni, M.; Cai, N. Syngas production from CO2/CH4 rich combustion in a porous media burner: Experimental characterization and elementary reaction model. Fuel 2017, 199, 413–419. [Google Scholar] [CrossRef]
  15. Jiang, L.; Liu, H.; Suo, S.; Xie, M.; Bai, M. Simulation of propane-air premixed combustion process in randomly packed beds. Appl. Therm. Eng. 2018, 141, 153–163. [Google Scholar] [CrossRef]
  16. Jiang, L.; Liu, H.; Wu, D.; Wang, J.; Xie, M.Z.; Bai, M. Pore-scale simulation of hydrogen–air premixed combustion process in randomly packed beds. Energy Fuels 2017, 31, 12791–12803. [Google Scholar] [CrossRef]
  17. Shi, J.; Mao, M.; Li, H.; Liu, Y.; Liu, Y.; Deng, Y. Influence of chemical kinetics on predictions of performance of syngas production from fuel-rich combustion of CO2/CH4 mixture in a two-layer burner. Front. Chem. 2020, 7, 902. [Google Scholar] [CrossRef]
  18. Shi, J.; Mao, M.; Li, H.; Liu, Y.; Sun, Y. Pore-level study of syngas production from fuel-rich partial oxidation in a simplified two-Layer burner. Front. Chem. 2019, 7, 793. [Google Scholar] [CrossRef]
  19. Shi, J.; Mao, M.; Li, H.; Liu, Y.; Lv, J. A pore level study of syngas production in two-layer burner formed by staggered arrangement of particles. Int. J. Hydrogen Energy 2020, 45, 2331–2340. [Google Scholar] [CrossRef]
  20. Zheng, C.; Cheng, L.; Cen, K.; Bingue, J.P.; Saveliev, A. Partial oxidation of methane in a reciprocal flow porous burner with an external heat source. Int. J. Hydrogen Energy 2012, 37, 4119–4126. [Google Scholar] [CrossRef]
  21. Alenazey, F.; Alyousef, Y.; AlOtaibi, B.; Almutairi, G.; Minakshi, M.; Cheng, C.K.; Vo, D.-V.N. Degradation behaviors of solid oxide fuel cell stacks in steady-state and cycling conditions. Energy Fuels 2020, 34, 14864–14873. [Google Scholar] [CrossRef]
  22. Bowman, C.T.; Hanson, R.K.; Davidson, D.F.; Gardiner, W.C.; Lissianski, J.V.; Smith, G.P.; Golden, D.M.; Frenklach, M.; Goldenberg, M. GRI-Mech 3.0 Chemical Kinetic Mechanism. Combustion Research Group at Berkeley. 2009. Available online: http://combustion.berkeley.edu/gri-mech/new21/version12/text12.html (accessed on 6 September 2023).
  23. Kee, R.J.; Dixon, G.L.; Warnatz, J.; Coltrin, M.E.; Miller, J.A. A Fortran Computer Code Package for the Evaluation of Gas-Phase Multicomponent Transport Properties; Report SAND86-8246; Sandia National Laboratories: Albuquerque, NM, USA, 1986. [Google Scholar]
  24. Barra, A.J.; Diepvens, G.; Ellzey, J.L.; Henneke, M.R. Numerical study of the effects of material properties on flame stabilization in a porous burner. Combust. Flame 2003, 134, 369–379. [Google Scholar] [CrossRef]
  25. Contarin, F.; Saveliev, A.; Fridman, A.; Kennedy, L.A. Reciprocal flow filtration combustor with embedded heat exchangers: Numerical study. Int. J. Heat Mass Transf. 2003, 46, 949–961. [Google Scholar] [CrossRef]
Figure 1. Schematic of the two-layer reactor.
Figure 1. Schematic of the two-layer reactor.
Batteries 11 00187 g001
Figure 2. Temperature distributions for different molar ratios of CO2/CH4 (solid line presents gas temperature Tg; dashed line presents solid temperature Ts).
Figure 2. Temperature distributions for different molar ratios of CO2/CH4 (solid line presents gas temperature Tg; dashed line presents solid temperature Ts).
Batteries 11 00187 g002
Figure 3. The distribution of major species near the reaction zone. (a) CO2/CH4 = 0. (b) CO2/CH4 = 1.
Figure 3. The distribution of major species near the reaction zone. (a) CO2/CH4 = 0. (b) CO2/CH4 = 1.
Batteries 11 00187 g003
Figure 4. The molar fraction of major species and reforming efficiency for different CO2 dilution concentrations [14].
Figure 4. The molar fraction of major species and reforming efficiency for different CO2 dilution concentrations [14].
Batteries 11 00187 g004
Figure 5. Temperature distributions for different β (solid line is gas temperature Tg; dashed line is solid temperature Ts). (a) CO2/CH4 = 0. (b) CO2/CH4 = 1.
Figure 5. Temperature distributions for different β (solid line is gas temperature Tg; dashed line is solid temperature Ts). (a) CO2/CH4 = 0. (b) CO2/CH4 = 1.
Batteries 11 00187 g005
Figure 6. Major syngas and reforming efficiency as a function of β . (a) CO2/CH4 = 0. (b) CO2/CH4 = 1.
Figure 6. Major syngas and reforming efficiency as a function of β . (a) CO2/CH4 = 0. (b) CO2/CH4 = 1.
Batteries 11 00187 g006
Figure 7. Temperature distributions for different λ s ( CO 2 / CH 4   =   0.5 ,   β   =   1000   W / ( m 3 · K ) ,   ξ   =   0.8 ).
Figure 7. Temperature distributions for different λ s ( CO 2 / CH 4   =   0.5 ,   β   =   1000   W / ( m 3 · K ) ,   ξ   =   0.8 ).
Batteries 11 00187 g007
Table 1. Molar fraction of major species and reforming efficiency for different λ s ( CO 2 / CH 4 = 0.5, β = 1000 W / ( m 3 · K ) , ξ = 0.8).
Table 1. Molar fraction of major species and reforming efficiency for different λ s ( CO 2 / CH 4 = 0.5, β = 1000 W / ( m 3 · K ) , ξ = 0.8).
λ s  (W/m · K) X CO 2 X H 2 X CO X CH 4 η  (%)
0.160.08670.05840.08940.003141.7
0.320.08380.05150.08850.002439.5
0.640.08940.06340.08700.001942.9
1.280.08570.06120.08970.003142.5
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

Wang, X.; Yu, M.; Li, Z.; Wang, Z.; Zhang, X.; Shi, J.; Kong, X.; Lv, J. Numerical Study of the Effects of Heat Loss and Solid Thermal Conductivity on Syngas Production for Fuel Cells. Batteries 2025, 11, 187. https://doi.org/10.3390/batteries11050187

AMA Style

Wang X, Yu M, Li Z, Wang Z, Zhang X, Shi J, Kong X, Lv J. Numerical Study of the Effects of Heat Loss and Solid Thermal Conductivity on Syngas Production for Fuel Cells. Batteries. 2025; 11(5):187. https://doi.org/10.3390/batteries11050187

Chicago/Turabian Style

Wang, Xiaolong, Mengmeng Yu, Zunmin Li, Zhen Wang, Xiuxia Zhang, Junrui Shi, Xiangjin Kong, and Jinsheng Lv. 2025. "Numerical Study of the Effects of Heat Loss and Solid Thermal Conductivity on Syngas Production for Fuel Cells" Batteries 11, no. 5: 187. https://doi.org/10.3390/batteries11050187

APA Style

Wang, X., Yu, M., Li, Z., Wang, Z., Zhang, X., Shi, J., Kong, X., & Lv, J. (2025). Numerical Study of the Effects of Heat Loss and Solid Thermal Conductivity on Syngas Production for Fuel Cells. Batteries, 11(5), 187. https://doi.org/10.3390/batteries11050187

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