1. Introduction
The combustion of fuel gas mixtures containing hydrogen in small-scale devices is of great interest as a way to optimise the use of fossil fuels, make them more efficient, reduce greenhouse gas emissions, and promote sustainability and energy transition challenges [
1,
2]. The mixtures of fuel gases containing hydrogen have been carried out in large-scale systems; several companies worldwide are incorporating considerable amounts of hydrogen, usually around 20%, into their natural gas pipelines, which is a challenge for the devices that consume it, requiring, in many cases, redesigns in the dimensions of the burners and changes in the materials with which they are built and the geometries that optimise the combustion of these mixtures [
3,
4]. In many cases, traditional burners cannot maintain the combustion of the gaseous mixture owing to the requirement for high pressures, which increases the complexity of the operation [
5].
As more efficient devices, micro-combustors may function throughout a spectrum of system pressures [
2]. However, studies on the micro-combustion of natural gas–hydrogen mixtures are scarce and often do not contemplate environmental variables such as low pressures and oxygen deficiency conditions. Anstrom and Collier [
6] conducted a methane mixed combustion with different concentrations of hydrogen below 20% study, testing the effect on the variation in the caloric value of the flame and the emission of not-burned methane. They established that the system with fuel mixtures operates stably under normal operative conditions in traditional fossil fuel-powered vehicles. Fugitive methane emissions are lower than when operating on natural gas, probably because hydrogen promotes rapid and complete combustion of methane by providing an abundance of hydroxyl ions. However, the study does not contemplate systematic variations in the geometry of the device, nor does it determine the effect on the flue gases, especially regarding how CO production is affected.
Micro-combustion also has some aspects that must be considered seriously. The reduction of the channel dimension modifies conditions to achieve stable combustion. Studies carried out in shorter-scale systems via numerical simulation show that a low H
2 content of 20% flame dynamics is significantly modified [
7], evidencing that the amount of hydrogen affects the fluid stability of the gaseous system and the thermochemical properties that emerge from it. Additionally, the large surface-to-volume ratio can act as a heat sink into the reactor, although it can represent higher heat flux values from/to the combustion gases. Still, the energy liberated during fuel combustion can be efficiently dissipated through the combustor walls. The temperature gradients at the reactor wall can disturb the kinetic reaction, affecting its operation. A sudden temperature loss can lead to flame quenching and endanger combustion stability. Then, since the micro-combustor walls play the role of heat sinks, it is possible that the activation energy of the reaction will not be sufficient to continue with the chain reaction [
7,
8,
9].
Computational fluid dynamics is a powerful tool for understanding the micro-combustion process of methane/hydrogen mixtures, especially when experimental analysis methods do not bring enough specific information. Numerical simulations are essential to design and optimise micro-combustion devices. Still, when the fluid is immersed in a reaction, like combustion, the reaction mechanism limits obtaining good results in the simulation process. If the simulation accurately describes the reaction, the results are far from actual behaviour. Jiaqiang et al. [
10] determine an entire mechanism of methane combustion, which includes 320 steps with their respective activation energies and preexponential factors that reproduce the kinetic; however, when implemented in a reactive flux simulation, the computation time is raised. Therefore, many researchers take a semi-detailed mechanism reaction that, in many cases, produces results in good agreement with experimental measurements; usually, 20 to 60 reaction steps are enough to carry out simulations with acceptable results with reduced simulation times [
11]. For instance, Zhang et al. [
12] use a 29-step mechanism to simulate the behaviour of a laminar burning, comparing twelve different mechanisms; the results show that most mechanisms could reasonably predict the experimental laminar burning velocity (LBV) for stoichiometric and fuel-lean mixtures and mixtures with diluent ratios higher than 70%. Mendiara et al. [
13] use a set of mechanisms with 65, 22, and 18 chemical species and reproduce the experimental behaviour of a methane flame in a model of the combustor; they used these results to show that it is not necessary to use a hole mechanism to obtain good reproducibility concerning experimental data.
This study replicates and analyses this type of combustion at micro-scale levels, considering different methane–hydrogen–air mixtures and the effect of the wall reactor conditions (e.g., heat losses), inlet velocity, and pressure outlet conditions on reaction performance. The microchannel reactive flow of methane–hydrogen–air mixtures under laminar flow conditions was studied using numerical simulations. A homogenous mechanism was used to analyse the chemical reaction inside the channel by solving the species transport model, including volumetric reactions. The results showed micro-combustion reactor conditions that produced a flame stabiliser with high flame stability while lowering nitrogen oxides, carbon monoxide, and partially oxidised hydrocarbons. This study introduces several innovative aspects compared to the existing research on methane–hydrogen combustion. Unlike most studies on large-scale or industrial combustion systems, this work investigates millimetre-scale combustion, providing new insights into flame stabilisation mechanisms in confined spaces. This is particularly significant for developing micro-combustors for portable power generation and high-efficiency energy systems. Additionally, the research conducts a detailed numerical assessment of critical parameters, such as equivalence ratio, hydrogen volume fraction, inlet velocity, and gas pressure, to optimise the combustion conditions in small reactors. While previous studies primarily explore methane–hydrogen combustion at macro scales, this work refines the design of micro-combustion reactors through precise numerical analysis, contributing to advancements in micro-energy systems.
The following sections present the fundamental transport equations for mass, momentum, energy, species transport, and reaction processes. The simulation methodology, geometry, and procedure parameters will be stated, followed by presentation of the results and their respective analyses being performed.
2. Materials and Methods
The phenomena present in reacting flows are studied with equations that combine all the types of transports (i.e., mass, momentum, energy) involved and allow us to account for the reactions inside the flow by solving them with numerical methods [
14]. Its application can describe how the fluids act under different operational conditions and react under certain environmental modifications. This kind of analysis is performed through computational fluid dynamics (CFD) and, more specifically, using the finite volume method through the CFD commercial software Ansys Fluent 2023 R1. This approach is one of CFD’s most versatile discretisation techniques. It is based on the definition of the control volume for the analytical formulation of fluid dynamics [
15].
Equations (1)–(4) present the steady-state principles of mass, linear momentum, energy, and species transport expressed in the tensorial form.
where
indicates the velocity components,
is the density,
the enthalpy,
the temperature,
is the pressure,
is the stress tensor which describes the advective momentum transport,
is the Cartesian coordinates,
is an enthalpy flux source,
the
species net production rate,
is the flux diffusion of species, and
is the species mass fraction. On the previously stated expressions, the sub-index,
, and
take the values of 1, 2, and 3 [
15,
16]. Equations (1) and (2) are numerically solved to describe the flow development, attaining the conservation of mass and momentum. Meanwhile, Equations (3) and (4) are used to describe the conservation of energy and the transport of species within this flow.
Table 1 presents the conditions for the numerical simulation of the reaction flow inside the microchannel implemented in Ansys Fluent 2023 R1.
The numerical simulations were conducted using a double-precision solver with 10 parallel processes on a local machine. The laminar viscous model was selected given that the Reynolds numbers are low for the flow conditions, and the energy equation was activated to account for heat transfer effects. The SIMPLE algorithm was used for pressure-velocity coupling, with second-order upwind discretisation for improved accuracy. Initial conditions were set using standard initialisation from the inlet, with a patched fluid zone temperature of 2000 K. The reacting flow was modelled using the species transport approach with volumetric reactions, employing a kinetic mechanism consisting of 16 species and 41 reactions. The gas mixture properties were defined using the ideal gas law for density, mixing law for specific heat, and mass-weighted formulations for thermal conductivity and viscosity, while species-specific heat was determined through piecewise polynomials.
The presence of reactions inside the fluid flow will also introduce several factors that will show variations in the flow transport properties: density, viscosity, and thermal conductivity, as well as the specific heat and other properties of the species inside the reactive flow and the mixture. Some parameters will be established as constant for the species (i.e., thermal conductivity and viscosity); others will be defined as temperature-dependent polynomials (i.e., the specific heat). Meanwhile, the thermal conductivity and viscosity of the mix are accounted for by the amount of each species in the mixture using a mass-weighted approach and the specific heat by the mixing law. In contrast, the mix density is evaluated using the ideal gas model as a function of the pressure inside the channel.
Table 2 summarises the property models used for the reacting flows studied with Ansys Fluent 2023 R1 in this work.
A wide variety of methane combustion mechanisms have been reported in the literature, including hydrogen as one of the intermediate molecular and radical species. Of the different mechanisms, some can be found with around 300 intermediate reactions, and others are simplified with four reactions. The more stages there are, the more details are possible to evidence the combustion process. Still, when implementing them in simulations such as the present one, it becomes unwieldy because the computation times are very long; therefore, we try to implement a mechanism that has enough detail but allows obtaining accurate results in reasonable times. One of the most used is the one used in the present work, which has been reported in previous works. Niklas et al. [
17] used different mechanisms from the literature to conduct methane combustion studies, determining that 41 reactions allow obtaining results that adequately reproduce the experimental observables. This type of mechanism evaluation is not new, as shown in a study from 1977 by Olson et al. [
18], who evaluated different mechanisms and found that the one that contains more species and reactions is not necessarily the one that best fits a computer simulation.
In the present work, the software CHEMKIN 2023 R1 [
14] was used to import 16 species and 41 reactions’ skeletal oxidation mechanisms into Ansys Fluent [
19]. The reaction mechanism starts with methane reacting with oxygen; these reactions produce a set of intermediate radical species like CH
3*, O
2*, O*, H*, OH*, and other species that are very reactive and are produced and consumed very fast. Still, they are the clue point to model the flame behaviour of intermediate chemical species, including radicals, which are presented in
Appendix A. The main intermediate reactions involving hydrogen are highlighted in it, and it can be observed that the preexponential factor
in them is usually more significant than in the other reactions. Additionally, it is observed that many of them occur without an associated activation energy, which indicates that they are processes that occur with great speed. Therefore, they are not the limiting stages, but it suggests that if the initial conditions for them to happen are not present, the reaction is delayed, extinguished, or occurs through paths that lead to unburned fuel, soot, or the production of carbon monoxide. This mechanism is crucial because it allows both fuels, methane and hydrogen, to be considered, which is necessary to model the flame generated when the ignition process occurs appropriately. The mechanism used has been used in the studies by Gautham et al. [
20], Qing et al. [
21], and Yin et al. [
22], among others, which have presented results that reasonably fit the experimental results in various experimental and simulation conditions, therefore we consider that it is an appropriate mechanism to carry out the present study since it has been widely tested, involves hydrogen as a fundamental species in the development of products, and allows obtaining results in short times.
The boundary conditions defined in this work are presented in
Figure 1. At the input, uniform velocity and temperature are employed. At the same time, the molar fractions of CH
4, O
2, and H
2 are specified based on the air-to-fuel equivalence ratio and volume fraction (
R) (i.e., percentage of hydrogen in the mixture). Axis symmetry boundary conditions are imposed at the centre line, and no diffusion or convective flux exists across the symmetry plane. The no-slip boundary condition and no-species flux conditions are used at the wall. A set of pressures is provided at the combustor outlet (see
Table 3). Lastly, the heat losses from the outer wall of the combustor to the ambient are defined by a heat loss coefficient, as observed in
Figure 1, where the convective heat transfer heat coefficient varies, as presented in
Table 3. Values for all these parameters are reported in
Table 3 and used to analyse their relevance to the overall reaction. The combustor wall is made of stainless steel (
= 16.6 W/m·K,
= 515 J/kg·K and
= 7900 kg/m
3). The same table also reports the channel height and pressure outlet variation. A homogeneous constant inlet temperature of 300 K is defined for the gas mixture. Finally, the combustion process was ignited by imposing an initial temperature of 2000 K on the entire computational area.
Turkeli-Ramadan et al.’s [
23] mesh sensitivity analysis was used as a reference in this work to perform the mesh independence study to ensure the grid resolution did not influence the numerical results. Six different meshes were tested (i.e., Mesh 1: 15,000, Mesh 2: 17,500, Mesh 3: 57,000, Mesh 4: 72,000, Mesh 5: 120,000, and Mesh 6: 204,000 elements), evaluating the outlet average temperature of gases and the molar fraction of CO
2 as parameters of reference.
Figure 2 presents the mesh independence study results, assessing the two parameters selected. The results show that as the mesh is refined, the variation in both parameters becomes negligible, suggesting that further mesh refinement would not significantly alter the results. The average variation between the parameters evaluated for each mesh was 3.29% and 1.61% for the temperature and the CO
2 molar fraction, respectively. In this regard, the mesh selected to ensure accuracy while minimising computational cost and following the work of Turkeli-Ramadan et al. [
23] was Mesh 3, which was composed by 75 grids × 600 grids (Fluid zone) and 20 grids × 600 grids (Solid zone) for a total of 57,000 elements.
Validation of the Model
The validation of two-dimensional modelling is verified by comparing the adiabatic flame temperature and species concentrations with the results reported by Betchel [
24]. Although the present two-dimensional model does not match the experimental setup reported in [
24], their results are for a one-dimensional flame geometry with a stoichiometric methane–air mixture of a stoichiometric, premixed laminar methane–air flame at atmospheric pressure.
Figure 3 shows the distribution of species and temperature profiles at the channel centre line for the case of an inlet temperature and velocity of 300 K and 0.3 m/s, respectively. In this case, an air/fuel equivalence ratio of 1.0 is assumed with no hydrogen addition to the mixture (i.e., the hydrogen to methane ratio
R is 0); an adiabatic wall is also assumed (i.e., no heat loss to the ambient). The micro-combustor studied has a 3 mm diameter and 12 mm length, and results are shown over the first 4 mm from the inlet since the combustion occurred over a very short distance from the inlet.
The numerical approach correlates well with the experimental data, allowing us to forecast the rapid rise observed at the combustor’s inlet. However, Turkeli-Ramadan et al. [
19] observed that the experimental temperatures are slightly lower than those predicted with numerical modelling at distances more than 3 mm from the start of the temperature rise. This behaviour is due to lateral heat losses in the experiment that were not considered in the computer model (i.e., an adiabatic boundary condition was used for the side walls). Regarding the prediction for the species’ behaviour inside the micro-combustor, a reasonable agreement between the numerical model and experiments is observed for CH
4 and O
2, it even being possible to depict the maximum mole fraction and the location for H
2 and CO. However, some discrepancies in the mole fractions for H
2O are observed in a region just beyond where the temperature profile changes slope (i.e., between approximately 0.8 mm and 2.5 mm from the inlet).
Figure 3 shows that the kinetic model adequately represents the behaviour of methane combustion. It is worth mentioning that the species that involve H do not appear in significant concentrations at the reactor outlet and, although their production can be followed, they are immediately consumed since the activation energy is zero in the reactions that consume H the rate of consumption is so high that they cannot remain in time. Still, they are essential to start the reaction since they reproduce species, temperatures, and concentrations of the products in agreement with previous reports.
Figure 4 presents the temperature, velocity, CH
4 mass fraction, and heat of reaction contour plots for a stoichiometric methane–air flame at an incoming flow velocity of 0.3 m/s and initial reactant temperature of 300 K, obtained employing the reaction mechanisms presented in
Appendix A. The predicted behaviour of flame temperature, velocity, methane, and heat of reaction profiles follows the results from peers [
19]. Additionally,
Table 4 presents the error metrics variation for CFD simulation variables compared with data from Betchel [
24], showing acceptable error magnitudes for these numerical approaches and the simplified mechanism used. The trim error levels for temperature and species mole fraction alongside the centre line of the reactor show that the reaction mechanisms can be applied to predict laminar methane–air flame characteristics at atmospheric pressure. The rapid temperature rise obtained from numerical modelling matches the experimental data well. However,
Figure 3 also shows that beyond 3 mm from the onset of the temperature increase, the experimentally measured temperatures are slightly lower than those predicted by numerical modelling. This discrepancy is likely due to the lateral heat losses inherent in the experiment, which are absent in the numerical model because of the adiabatic boundary condition applied to the side walls. In addition to the accuracy of the solution, the computational cost also needs consideration in any numerical-based work. As reported by Turkeli-Ramadan et al. [
23], the reaction mechanism that predicts a sufficiently accurate solution in the shortest amount of time is selected in this work (see
Appendix A).
4. Conclusions
This study advances the understanding of methane–hydrogen–air combustion in micro-combustors through two-dimensional numerical simulations, evaluating key parameters such as hydrogen enrichment, the air/fuel ratio, heat losses, inlet velocity, outlet pressure, and reactor geometry. The findings show that increasing the hydrogen content enhances flame temperatures, promotes complete oxidation, and reduces CO emissions, with peak heat release at 40% hydrogen. A leaner air/fuel ratio improves combustion efficiency, while excessive wall heat losses, smaller reactor diameters, and lower temperatures increase incomplete combustion. Additionally, higher outlet pressures elevate flame temperatures and CO2 production, while hydrogen enrichment proves beneficial under low-oxidant conditions, making it particularly relevant for high-altitude cities.
Various factors significantly influence the performance of methane–hydrogen combustion. Increasing the hydrogen concentration from 0 to 90% enhances flame temperatures by about 300 K and promotes complete oxidation, leading to higher CO2 formation and reduced CO emissions. The results show that maximum heat combustion reaches 40% of hydrogen. A leaner air/fuel ratio improves the combustion efficiency and minimises emissions, while excessive wall heat losses lower temperatures and reduce fuel conversion efficiency. Higher inlet velocities can shorten residence time, potentially limiting reaction completeness, though velocity distribution remains relatively unaffected by other parameters. Elevated outlet pressures increase the flame temperatures (by almost 15%) and CO2 production due to enhanced reaction rates and equilibrium shifts. Finally, larger microreactor diameters reduce heat losses, improving temperature profiles and combustion efficiency by about 20%, as based on CO2 and CO production. In contrast, smaller diameters lead to incomplete combustion and higher CO emissions due to more thermal dissipation. These effects underscore the importance of carefully analysing hydrogen enrichment, the air/fuel ratio, heat transfer, operating pressures, flow dynamics, and reactor geometry for efficient, low-emission combustion in microreactor systems. It was observed that the combustion of methane–hydrogen–air mixtures remains stable under the conditions studied, despite the significant changes in the exterior heat losses and pressure outlet. Additionally, CO and CO2 emissions were reduced by almost 20% as the amount of H2 in the fuel–air mixture was increased to 50%. The results concerning the combustion of air/fuel and pressure suggest that the addition of H2 is significant under conditions with low oxidant levels, which indicates that it is beneficial for cities located at a considerable altitude above sea level. This study shows that CFD is an essential tool for understanding combustion and heat loss mechanisms at micro-millimetre scales, where it may be challenging to physically measure the flow parameters inside the microreactor itself. However, the kinetic mechanisms used can represent the combustion of methane–hydrogen–air mixtures; more studies should be developed to ensure the accuracy of these results.
The results obtained in this study allow us to understand the operating conditions under which a microreactor must operate so that the flame is not extinguished. It allows the establishment of the maximum temperatures and those reached by the reaction chamber; the inlet flows and H2-CH4-air composition allow a microreactor to operate with the least possible variability. Therefore, this study allows obtaining essential elements necessary to decide on the construction characteristics of a microreactor, such as the type of material, its thickness, and the internal diameter of the reaction chamber, so that all the fuel is burned. An adequate temperature is maintained with a controlled heat transfer to the outside without extinguishing the flame or increasing the amount of unburnt fuel or incomplete combustion. Another relevant aspect is that it provides elements to decide how to inject the fuel into the chamber, the flow at which this should be performed, and the optimal operating pressure.
In particular, it is expected to design a reactor that transmits around 20 W/m2K, with a slight excess of air at the inlet and operating above ambient pressure. This is relevant, primarily if the reactor is intended to be used in cities above sea level, as it is evident that its operation at low pressures is inefficient. In other words, it provides elements to know that a microreactor should be designed based on the city’s height in which it will operate.
Finally, the models used to assess combustion quality, pollutant reduction, and heat transfer rely on physicochemical approximations. These assume, for example, that the reaction mechanism remains valid under all conditions, which may limit accuracy, particularly in extreme cases. As numerical approximations, the species and heat transport models have inherent limitations tied to the solution method, leading to results that are close to, but not identical to, experimental values. Additionally, the simulations assume uniform heat loss properties along all reactor walls, whereas, in reality, material wear can cause variations. However, since the same types of modelling and numerical errors are consistently present, they do not affect the observed trends, which are often more valuable than the absolute numerical results.