University
stringclasses
19 values
Text
stringlengths
458
20.7k
Colorado School of Mines
5.1 12cm Diameter Quartz Reactor Model Setup Researchers have been developing a combustion model of the quartz flow reactor as shown in Figure 5.1 (Strebinger, et al., 2018). The model has a domain height of 12cm, which is the inner diameter of the quartz flow reactor, and length of 1.5m. In order to accurately model methane-air deflagration physics, researchers are using the following settings in ANSYS Fluent (v17.2): • 2D planar assumption for all 2D models • Pressure-Based Solver • Energy Equation • Viscous Standard k-ω Turbulence Model o Low Re Corrections o Shear Flow Corrections • Species Transport o Volumetric Reactions o Stiff Chemistry Solver o Finite Rate Chemistry ▪ Density solved using ideal gas theory ▪ Diffusion solved using kinetic theory ▪ Metghalchi and Keck laminar flame speed theory • Spark Ignition Model • PISO pressure-velocity coupling • CEI: 2 levels of mesh adaption on the gradient of temperature every 2-10 time steps • 2D Model: Residuals set to 10-6, dropping at least 3 orders or magnitude • 3D Model: Continuity and species residuals were set to 10-3, still dropping 3 orders of magnitude. All other residuals are set to 10-6 • Second order in time and space • Time step = 0.1ms for open-end ignition, 0.01ms for closed-end ignition • Boundary Conditions: o Closed End Wall – aluminum, no slip, adiabatic o Tube Walls – quartz, no slip, adiabatic o Outlets – 0 gauge pressure outlets 107
Colorado School of Mines
o Obstacles– no slip, adiabatic, assuming no heat transfer into the material Figure 5.1 CFD geometry of 12cm inner diameter quartz flow reactor and seeded lines at different vertical, y, positions used to find the flame front. Model height = 12cm, length = 1.5m. The compressible flow model is being used to capture the large density changes during combustion and the compression of the unburned and burned gas mixture upstream and downstream, respectively. The viscous-standard k-ω model is used because compared to other turbulence models, the k-ω model better predicts lower Reynolds number flows and flow separation, which often occurs when a flame passes over an obstacle; this assumption will be revisited in later sections of this manuscript. The transient time solver is a second order implicit time solver and the governing equations are solved using the PISO pressure-velocity coupling solver. The species transport model is used instead of the premixed combustion models because it allows for more control over mixture stoichiometry as well as the number of chemical reactions. The inlet and outlet boundary conditions are modeled as zero gauge pressure outlets to simulate ambient conditions in the laboratory and all solid boundaries are modeled as walls assuming no slip and adiabatic conditions. The obstacles are also modeled with a no slip boundary condition and adiabatic. Adiabatic conditions are assumed for the walls and obstacles because the flame is in contact with the walls of the tube for a very short time such that there is minimum heat transfer by conduction and convection. Radiation may play a role in flame enhancement and may need to be investigated in the future, however, Fig performed a preliminary investigation of including a radiation model, but found the error associated with the models is larger than not including radiation (Fig, 2019). Finally, the finite rate chemistry model is used in lieu of the turbulent chemistry interaction models, however, as will be shown in this 108
Colorado School of Mines
research, this will need to be revisited for highly turbulent flow interaction with obstacles as well as the inclusion of more detailed chemistry. Although previous research has shown that using a methane-air 2-step mechanism is not as accurate as solving the chemistry fully (Fig, 2019), it allows for faster simulation times which is important since these models typically run for a week or more. The first reaction of the 2-step mechanism is the reaction of methane and oxygen to form H 0 and intermediate species, CO: 2 (5.1) The second reaction is forwards and backwards and is the reaction of CO and oxygen to form CO : 2 (5.2) Table 5.1 Table showing the ANSYS Fluent 2-step methane-air chemical mechanism settings. Reaction 1 and 2 from (Dryer & Glassman, 1973). R stands for reactant and P stands for product. Pre- Activation Reaction Stoichiometric Temperature Molecule Path Exponential Energy Number Coefficient Exponent (K) Factor (J/kg-mol) 1 CH 1 R 5.012x1011 2x108 0 4 1 O 1.5 R 2 1 CO 1 P 1 H 0 2 P 2 2 CO 1 R 2.239x1012 1.7x108 0 2 O 0.5 R 2 2 CO 1 P 2 2 H 0 0 P 2 3 CO 1 R 5x108 1.7x108 0 2 3 CO 1 P 3 O 0.5 P 2 109
Colorado School of Mines
This research has explored two major ways of modeling the spark, referred to as the electrode model (EM) (Fig, Bogin, Brune, & Grubb, 2016; Fig, Bogin, Brune, & Grubb, 2017; Fig M. , 2019) and the spark model (SM) (Strebinger, Bogin, & Brune, 2019). The EM model uses an aluminum circle with a diameter of 1mm to represent the electrodes from the spark system. At time, t=0s, the aluminum wall has a heat flux boundary that produces 2.5x106 W/m2 of energy to the surroundings. After 25ms of simulation time, the heat flux boundary is set to 0 W/m2, assuming adiabatic. This produces approximately 5mJ of energy which is enough to ignite methane-air mixtures at the lean and rich limits. The SM model using the ANSYS Spark Model (v17.2) and the following settings shown in Table 5.2. Although, Table 5.2 shows the total ignition energy as 60mJ which is different than the 5mJ used in the EM, this difference will be explored in Section 5.1.2. Table 5.2 Table showing the ANSYS Fluent spark model (SM) settings. Start Time (s) 0 Duration (s) 0.001 Initial Spark Radius (m) 0.005 Ignition Energy (J) 0.06 Kernel Expansion Model Laminar 5.1.1 2D Mesh Independence Study Using these model settings, the first step of the modeling process is to determine mesh independence for an open-end ignition event. The spark was located at 11cm from the open end of the reactor. The ANSYS meshing client was used to develop the mesh for this model. The fluid body was meshed using quadrilaterals as shown in Figure 5.5. To compare the results of each body mesh size, researchers sampled five horizontal lines along the y-axis spanning the reactor as shown in Figure 5.1. Data was extracted from each line at different time steps and compared to one another to confirm the actual flame front; additionally, manually, researchers confirmed this was the actual flame front by comparing axial values to temperature and methane concentration contours. To determine the best method of finding the flame front, researchers compared defining the flame front as the maximum kinetic rate of reaction 1, the maximum total temperature at the gradient of the main flame brush, and different values of mass fraction of methane. All methods used were comparable and the difference error in determining the flame 110
Colorado School of Mines
front amongst methods was less than 1%. This was also compared to finding the maximum total temperature and the error between these methods is less than 1%. Thus, due to ease and comparability with Matt Fig’s work, researchers are determining the flame front by values of temperature along the gradient of the flame front (Fig, 2019). Results from the mesh independence study (Figure 5.2) show that a mesh size of 1mm is sufficient grid resolution to resolve the flame front. The error between the 1mm and 0.5mm mesh was 1% as shown in Figure 5.2. Results also show that a mesh size of 1mm takes 9.5days to run versus a 0.5mm mesh which takes over 20 days to complete as shown in Figure 5.3; note that the first 3-4 days is spent on flame development in the first 25cm from kernel initiation. Thus, all 2D models use a base mesh sizing of at least 1mm unless otherwise noted in this manuscript. Figure 5.2 Mesh independence study for the 2D, 12cm diameter quartz flow reactor. Mesh independence was achieved with a mesh size of 1mm and the average percent relative error of flame front location based on maximum temperature of 1%. Error bars represent the standard deviation of the relative error between mesh sizes. 111
Colorado School of Mines
Figure 5.3 Mesh size versus simulated time for the 2D, 12cm diameter quartz reactor model. Mesh independence is reached with a body mesh size of 1mm, taking approximately 9.5days to complete on an 8 core compute, 3.06GHz, 24GB RAM. After determining the mesh size, researchers investigated the effects of including wall inflation or wall edge sizing since resolving the boundary layer is important for modeling methane flame deflagrations in cylindrical reactors. Results from Table 5.3 show that adding edge sizing or wall inflation greatly affects the simulation. A uniform 1mm body mesh without any edge sizing or wall inflation does not accurately resolve the boundary layers as shown in Figure 5.4. Because of this the flame for the base case, 1mm body mesh, does not flip over until after 1s simulation time. However, in experiments it is observed that the flame turns over in the first 50ms before traveling halfway down the quartz reactor. Thus, from this study it was determined that resolving the boundary layer is of utmost importance and a 0.25mm edge sizing was used. Both edge sizing and inflation layer methods were tested and results are shown in Table 5.3 and Figure 5.4. As can be seen in these results, the edge sizing method and inflation layer methods help to better resolve the impact of the boundary layer in the turn-over of the flame. However, looking ahead to future modeling of obstacles inside the reactor, using an edge sizing can more easily be applied to meshing around irregular-shaped obstacles. Thus, the edge sizing was used over the inflation layers due to ease of meshing when obstacles are present in the reactor. 112
Colorado School of Mines
Table 5.3 Average percent error in flame front location between the 1mm body mesh with no wall sizing compared to adding edge sizing or wall inflation layers for the 2D, 12cm diameter quartz reactor model. 0.25mm Edge 40 layers, Single 80 layers, Single Sizing layer height = layer height = 0.5mm 0.5mm Average Error (%) 17 24 25 Figure 5.4 Temperature contours of 2D, 12cm diameter quartz reactor model comparing the impact of number of wall inflation layers on methane flame propagation. Simulation time=1.0s. Time step = 0.1ms. Ignition location is 11cm from the open end (left). Mesh cell size = 1mm, 0.5mm inflation cells, varying number of inflation layers. CH =9.5%. Temperature = 293K, 4 Pressure = 82kPa. One relief hole. An image of the 1mm body mesh with 0.25mm edge sizing is shown in Figure 5.5. As can be seen in this mesh, the mesh is unstructured and consists of both quadrilaterals and triangles. Mesh statistics are summarized in Table 5.4 and show a total cell count of approximately 250,000 cells. The orthogonality quality is close to 1 and skewness is close to 0 which means that flow quantities are transferred from one cell to the next well from one cell face to another. Also the aspect ratio is close to 1 which means the cells are not too stretched. 113
Colorado School of Mines
Altogether, these mesh statistics show a good quality mesh has been obtained. The same mesh was used for open-end ignition and closed-end ignition. Figure 5.5 Image of the 2D, 12cm diameter reactor mesh with a quadrilateral dominant 1mm body mesh and 0.25mm edge sizing on the tube walls. Table 5.4 2D, 12cm diameter reactor mesh statistics: number of elements, number of nodes, average orthogonality quality, average skewness, and average aspect ratio. Body mesh = 1mm. Edge sizing = 0.25mm. Average Number of Number of Average Average Aspect Orthogonality Elements Nodes Skewness Ratio Quality 254,148 258,367 0.99 ± 0.03 0.08 ± 0.1 1.2 ± 0.2 5.1.2 3D Mesh Independence Study Finally, a mesh independence study was undertaken for the 3D model of the 12cm diameter quartz reactor for a closed-end ignition using a base mesh of 8mm, 4mm, and 2mm with no edge sizing. An example of the mesh is shown in Figure 5.6 and mesh statistics for all meshes are summarized in Table 5.5. As can be seen, the mesh is a structured mesh and mesh statistics show low skewness and high aspect ratios. The high aspect ratios are mainly due to the cells at the tube walls, which is to be expected when using a structured mesh on a cylindrical body. Edge sizing may be used to help resolve the high aspect ratios, however, no edge sizing was used in order to improve the quality of the cells inside the 3D domain. Instead of specifying an edge sizing, mesh adaption was used. Previous work by M.K. Fig, 2019 investigated the best variable to adapt on to 114
Colorado School of Mines
better resolve the flame front (Fig M. , 2019). Results of Fig’s study showed that adapting on the temperature gradient helped to best resolve the flame front, using at least 2-3 levels of adaption. Therefore, this research also uses grid adaption on the gradient of temperature using 3 levels, every 2 time steps. This helped to resolve the flame front and flame propagation, but as will be discussed, there was some error at the tube wall boundaries. Figure 5.6 Image of a 3D, 12cm diameter reactor mesh with a quadrilateral dominant 4mm, cut cell body mesh. Table 5.5: 3D, 12cm diameter reactor mesh statistics for an 8mm, 4mm, and 2mm body mesh: number of elements, number of nodes, minimum orthogonality quality, maximum skewness, and maximum aspect ratio. No edge sizing. Minimum Number of Number of Maximum Maximum Orthogonality Elements Nodes Skewness Aspect Ratio Quality 8mm body mesh 26,240 28,497 0.4 0.4 7.1 4mm body mesh 194,560 201,909 0.5 0.2 5.1 2mm body mesh 1,520,640 1,554,425 0.4 0.4 6.4 Results of the mesh independence study are tabulated in Table 5.6 and shown in Figure 5.7 and Figure 5.8. Results show that a base mesh of 8mm is inadequate in capturing the flame front location compared to the 4mm or 2mm mesh. The average percent error between the 8mm and 4mm mesh was 25%, reducing to below 5% for the 4mm and 2mm mesh. To investigate this further, researchers made an estimation of the size of the boundary layer using turbulent fluid flow estimates. The mixture was approximated as air at standard temperature and pressure, 115
Colorado School of Mines
101kPa and 300K; the density was 1.2kg/m3, dynamic viscosity of 1.72x10-5 Pa-s, and a velocity range of 25-60m/s was used. The results of calculating the Reynold’s number based on the axial location indicate a turbulent regime (Re = 440,000 to 6,800,000). The boundary layer thickness x was calculated based on turbulent boundary layer theory and approximated using the following equation: (5.3) From this equation a range of boundary layer thicknesses, δ, dependent on the x location were calculated to be 7-24mm. Based on the results shown, it makes sense that the 8mm body mesh performed poorly at earlier times compared to the 4mm and 2mm since the 8mm cell can be larger than the boundary at certain locations. Therefore, based on flame front location and boundary layer thickness, the 4mm and 2mm body meshes are more accurate than the 8mm. Table 5.6 Table showing the percent error of the 3D, 12cm diameter quartz reactor model for different mesh sizes (8mm and 4mm, 4mm and 2mm). Time step = 0.01ms. Ignition location is 1.39m from the open end. CH =9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. 4 ign No relief hole. Time (s) Percent Error Percent Error 8mm : 4mm (%) 4mm : 2mm (%) 0.004 52 4.3 0.006 47 3.9 0.008 38 4.3 0.010 17 1.9 0.012 4.2 2.6 0.014 7.6 4.1 0.015 - 5.1 Average = 25±10% 4±0.4% 116
Colorado School of Mines
As discussed, evaluating the mesh based on the flame front location shows that a 4mm and 2mm body mesh sizing are adequate to resolve the boundary layers and the calculated error between the flame front locations are within 5%. However, the relative error of the flame front location was based on the axial movement of the flame and not all the flow quantities. Therefore, a study was performed investigating the percent difference in total temperature and flow velocity comparing the 4mm and 2mm body meshes for the 3D, 12cm diameter reactor model. To do this, five (5) XY planes were made 0.25m apart at z = 0.25, 0.5, 0.75, 1.0, 1.25m with a seeded line extending in the x direction as shown in Figure 5.9. Results of the flames propagating are shown in Figure 5.10, Figure 5.11, Figure 5.12, and Figure 5.13. As shown in these figures, there is little difference in the shape and propagation of the flame. However, extracting data from the seeded lines shows that there are differences in the total temperature and the fluid velocity magnitude as shown in Figure 5.14 through Figure 5.21. As can be seen, the difference in the total temperatures between the 4mm mesh and 2mm mesh are small. Although this is for a single time step, other time steps have been evaluated and the average percent difference across time steps is less than 5% and is mainly accumulated at the flow boundaries. This is not of surprise since the 2mm mesh has more cells in the boundary than the 4mm mesh, which allows the 2mm mesh to more accurately solve boundary effects. Evaluation of the flow velocities shows a much larger difference in predicted velocities between the 4mm mesh and 2mm mesh and the average percent difference across time steps is less than 13%. Note that in all these simulations mesh adaption on the gradient of temperature is employed every 2 time steps, 3 levels. Since the grid adaption is on the gradient of temperature, the percent difference of total temperature values between the 4mm and 2mm meshes are within 5%. However, because the adaption is on the gradient of temperature the flow velocities are not fully resolved using the 4mm mesh. Therefore researchers suggest using the 2mm mesh for more accurate predictions of flow velocities and turbulent quantities. If simulation time is of the utmost importance, a 4mm mesh can be used, but will provide less accurate predictions of the local flow and turbulence. 118
Colorado School of Mines
5.1.3 Modeling the Spark As previously discussed, two main methods of initiating combustion in the combustion models have been investigated: the EM uses a constant heat flux from a small disc and the SM uses the improved ANSYS Fluent Spark Model (v17.2). In order to compare these two methods of initiating combustion, a 2D model of the 12cm diameter reactor was setup and all parameters, meshes, were the exact same; researchers only changed the method of initiating combustion, EM versus SM. As shown in Figure 5.22 and Figure 5.23, both the EM and SM were able to capture the eventual flip over of the open-end ignition flame. However, both models predicted this flame turn over much later in the combustion process; this discrepancy is likely due to the fact that accurately modeling buoyancy is a difficult problem. This difference is also captured in the prediction of flame front propagation velocity as a function of distance. Figure 5.23 shows that the SM predicts a slightly faster flame than the EM, but both models predict the turnover too slowly resulting in a decreasing flame front propagation velocity instead of increasing as shown by the experimental results. Figure 5.22 Temperature contours of 2D, 12cm diameter quartz reactor model comparing the impact of methods modeling the spark electrodes. Simulation time=0.9s. Time step = 0.1ms. Ignition location is 11cm from the open end (left). Mesh cell size = 1mm, 0.25mm edge sizing. CH =9.5%. Temperature = 293K, Pressure = 82kPa. One relief hole H=1.2cm. 4 128
Colorado School of Mines
Figure 5.23 Comparison of flame front propagation velocity results from experiments to the 2D, 12cm diameter quartz reactor model using the spark model and electrode model. Time step = 0.1ms. Ignition location is 11cm from the open end (left). Mesh cell size = 1mm, 0.25mm edge sizing. CH =9.5%. Temperature = 293K, Pressure = 82kPa. One relief hole H=1.2cm. 4 Another important factor to consider in both of these methods is the amount of total simulation time as indicated on Figure 5.22. The SM takes 4 days to reach t = 0.9s versus the EM which takes 6 days to reach t = 0.9s. Further exploration into this difference has shown that the initial kernel expansion in the EM takes 3 days to simulate. This is important because for a confined ignition the time scales are much shorter, requiring a time step of 0.01ms to solve in a reasonable amount of time. For example, Figure 5.24 and Figure 5.25 show results of a confined, closed-end ignition (ignition 11cm from the closed end of the reactor). The geometry shown in these figures was cut in half to 0.75m instead of the full 1.5m. The base mesh was 1mm and the edge sizing on the walls was 0.25mm. Combustion was initiated using the EM and to simulate 36ms of flame propagation took over 21 days on an 8 core compute, 3.06GHz, 24GB RAM, compared to the SM which takes only 2-3 days to model the full 12cm diameter, 1.5m long reactor with the same mesh. Although the flame front propagation velocity results match the experiments within 5%, the large simulation times required of using the EM are infeasible for this modeling combustion in the large and full-scale models. In comparison, the SM matches the maximum flame front propagation velocity within 12% and matches the flame acceleration as will be discussed in Chapter 6. 129
Colorado School of Mines
Figure 5.24 Temperature contours of 2D, 12cm diameter quartz reactor model for a closed-end ignition using the electrode model. Domain length = 0.75m. Simulation time=36ms. Time step = 0.01ms. Ignition location is 11cm from the closed end (left). Mesh cell size = 1mm, 0.25mm edge sizing. CH =9.5%. Temperature = 293K, Pressure = 82kPa. One relief hole. 4 Figure 5.25 Comparison of flame front propagation velocity results from experiments to the 2D, 12cm diameter quartz reactor model using the electrode model. Domain length = 0.75m. Time step = 0.01ms. Ignition location is 11cm from the closed end. Mesh cell size = 1mm, 0.25mm edge sizing. CH =9.5%. Temperature = 293K, Pressure = 82kPa. EM E = 5mJ. One relief hole. 4 ign Therefore, due to the large simulations times required by using the EM that would be required for full-scale longwall model explosion modeling it was decided to investigate using the t using the ANSYS Fluent Spark Model (SM). A sensitivity analysis was performed to determine which parameters of the ANSYS Fluent Spark Model is the flame most sensitive to: duration, initial kernel diameter, spark energy, or flame speed model. It was found that the model was sensitive to the initial kernel diameter and the amount of spark energy. 130
Colorado School of Mines
Ignition energies were tested between 5mJ and 1J and modeling results with the improved SM show no significant differences (<2%) between these energies as shown in Figure 5.26 , Figure 5.27, and Table 5.7. In 2014, Zipf, et al. (2013) performed experiments in a large 1.05m diameter, 73m long flame reactor investigating the detonability of natural gas-air mixtures and found transition to detonation and detonations for some experiments. In their experimental setup, the ignition source was an electric match that produced multiple sparks with a total energy of 2kJ. Therefore, researchers investigated the possibility of 1kJ of energy transfer to determine whether or not this large amount of energy impacts methane flame acceleration – results are shown in Figure 5.28 and Table 5.7. As can be seen, the 1kJ of energy shows differences in predicted methane flame speeds and flame shapes compared to 1J (6.8% difference); whether this amount of energy is realistic in a methane gas explosion accident is under investigation. This amount of energy could be the result of a possible lightning strike or an explosive, but in most methane gas explosion cases the ignition source is a result of machine friction, hot streaks, or rock friction, which all have lower energies than 1kJ. Figure 5.26 Temperature contours of 2D, 12cm diameter quartz reactor model comparing the impact of ignition energy of the SM on flame propagation. Simulation time=10ms. Time step = 0.01ms. Ignition location is 1.39cm from the open end. Mesh cell size = 1mm, 0.25mm edge sizing. CH =9.5%. Temperature = 293K, Pressure = 82kPa. E = 5mJ,120mJ,480mJ, and 4 ign 1000mJ. One relief hole. 131
Colorado School of Mines
Table 5.7 Table showing the average percent difference and standard deviation between different ignition energies in the 2D, 12cm diameter quartz reactor model for a closed-end ignition. Time step = 0.01ms. Ignition location is 1.39m from the open end. CH =9.5%. Temperature = 293K, 4 Pressure = 82kPa. SM E = 60mJ. One relief hole. ign Percent Difference (%) 60mJ : 120mJ : 5mJ : 60mJ 480mJ : 1J 1J : 1kJ 120mJ 480mJ Average 0.06% 0.06% 0.17% 0.78% 6.8% Standard 0.12% 0.12% 0.19% 0.45% 3.9% Deviation Maximum 0.48% 0.48% 0.64% 1.8% 11% In the experiments, the distance between the spark electrodes was measured to be between 5-10mm, which corresponds to the 5mm and 2.5mm spark radius settings. Initial spark radii of 5mm, 2.5mm, and 1.5mm were tested using the SM for a confined, closed-end ignition as shown in Figure 5.29 and Figure 5.30. Results show that the model is highly sensitive to the initial spark kernel radius and the difference in flame front results continue to grow as a function of time. Comparing the pressure results from Figure 5.30 to Figure 4.11 shows that the model captures a rise in the overpressure, but the model predicted overpressure is almost 3 times greater than the experiments (3.25kPa). Figure 5.29 Comparison of flame front propagation velocity results from experiments to the 2D, 12cm diameter quartz reactor model using the spark model for different initial kernel radii. R = ini 5, 2.5, 1.5mm. Time step = 0.01ms. Ignition location is 1.39m from the open end. Mesh cell size = 1mm, 0.25mm edge sizing. CH =9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 4 ign 60mJ. No relief hole. 133
Colorado School of Mines
initial spark kernel of 5mm, 2.5mm, and 1.5mm. Results are similar to the 2D model and the relative percent error between these cases ranges from 5-17%, the average being approximately 12%. Based on the 2D and 3D cases, this research determined using an initial spark kernel radius of 5mm because it better captures the continual acceleration of the flame shown in Figure 5.29 without significant differences in predicted overpressure and flame trends. To note, these were performed with no relief hole on the closed end; when the relief is included, the flame does slow down near the end, similar to the experiments which will be discussed in subsequent sections. Table 5.8 Table showing the percent error of the 3D, 12cm diameter quartz reactor model comparing the initial spark kernel radius for the 4mm body mesh (R = 5mm and 2.5mm). Time ini step = 0.01ms. Ignition location is 1.39m from the open end. CH =9.5%. Temperature = 293K, 4 Pressure = 82kPa. SM E = 60mJ. No relief hole. ign Percent Error (4mm body) Time (s) R =5mm : R =2.5mm (%) ini ini 0.004 5.1 0.006 7.1 0.008 11.1 0.010 13.3 0.012 13.3 0.014 16.4 0.015 16.9 Average = 12% 5.1.4 Turbulence Model and Parameter Study As discussed in Section 2.8, there are many different turbulence models which can be employed to model methane flame propagation. Previous research concluded for the 2D models of the small-scale 5cm diameter, 9.5cm diameter, and 71cm diameter reactors that the standard k-ε turbulence model best matched experiments (Fig, 2019). However, as discussed, that model uses a different method of initiating combustion which can significantly change the turbulence parameter settings of the model. This section aims at describing how this current research determined the appropriate turbulence parameters and turbulence model for initiating combustion via a spark model. This section also discusses the difference in assuming a 2D planar geometry versus axisymmetric. In ANSYS Fluent, using the Reynold’s Averaged Navier-Stokes equations with difference closure models requires model initialization of turbulence parameters such as turbulent kinetic energy (k), turbulent dissipation rate (ε), and specific dissipation rate (ω). To 135
Colorado School of Mines
calculate these quantities requires certain information about the flow including fluid properties and upstream flow properties. The first quantity which must be calculated is the Reynold’s number which depends on the mean flow velocity, kinematic viscosity, and hydraulic diameter shown in Equation (5.4). In this case, the hydraulic diameter of the cylindrical reactors is the reactor diameter, but for the experimental box, the hydraulic diameter is much more difficult to estimate and is estimated using a square duct equation. (5.4) Next, the turbulent length scale, l, as smaller than the hydraulic diameter of the reactor. A factor of 0.07 is recommended if there are obstacles in the flow, the length scale or hydraulic diameter may be more appropriately based on the obstacle size (ANSYS© Fluent, 2009). However, in order to directly compare the impact of obstacles on the flow, the current model uses the same initialization parameters for open reactors and those with obstacles. (5.5) After determining the flow regime and length scale, an estimation of the turbulent intensity can be made using Equation (5.6). The turbulent intensity (I) is the ratio of the RMS velocity fluctuations and the mean flow inside the reactor, but for a cylindrical reactor can be estimated from the following empirical correlation (ANSYS© Fluent, 2009): (5.6) Next, estimations for the turbulent kinetic energy (k), turbulent dissipation rate (ε), and specific dissipation rate (ω) are made using Equations (5.7), (5.8), and (5.9), where Cμ is an empirical constant usually 0.09 (ANSYS© Fluent, 2009). 136
Colorado School of Mines
(5.7) (5.8) (5.9) Unfortunately, the upstream conditions of the confined ignition are not well known, which requires a sensitivity analysis of initial turbulence parameters on flame front propagation. To do this, various mean flow velocities were estimated based on a range of velocities between 0 and the maximum flame front propagation velocities. From these, turbulent quantities were calculated for a closed-end ignition in the 12cm diameter reactor as shown in Table 5.9. In future work, the upstream turbulence conditions may be estimated using flow sensors and schlieren photography to gain a stronger understanding of local fluid fluctuations and the average size of the eddies in the flow. Table 5.9 Table showing calculated turbulent initialization parameters based off different flame front propagation velocities of a closed-end ignition in the 12cm diameter quartz reactor. u (m/s) 1 5 10 20 30 60 avg k (m2/s2) 0.004 0.07 0.24 0.8 1.6 5.5 ε (m2/s3) 0.005 0.37 2.3 14 40 250 ω (1/s) 14 58 106 195 278 509 As shown in the table, depending on how the mean flow velocity is defined, the turbulence parameters have quite a large range. Therefore, simulations were set up initializing with different turbulent parameters based on mean flow velocities of 1, 5, 10, 20, 30, and 60m/s. In the future, these flow velocities can be estimated based on the kernel expansion rate, but this will require additional imaging with a faster rate of frames per second than what was used in this research (above 240fps). Qualitative results of the simulations are shown in Figure 5.32, Figure 5.33, and Figure 5.34. Compared to experimental images in Figure 4.8, estimating the initial turbulence 137
Colorado School of Mines
parameters using 5m/s and 10m/s results in flame shapes which are different than those observed in experiments. Estimating the initial turbulence parameters based off mean flow velocities of 20m/s and 30m/s result in flame shapes which are more closely related to those observed in experiments. One interesting flame shape observed in experiments and the model is shown in Figure 5.35 at the open end of the reactor during a closed-end ignition. This shape was not observed in approximately 50% of the closed-end ignition experiments and in all of the 2D, 12cm diameter models of closed-end ignition. Investigation of this shape was performed and it was determined this is due to the boundary condition on the open end; the density differences between the ambient air and combustion products resulted in flame instabilities leading to a slight flame inversion. Figure 5.32 Temperature contours of the 2D, 12cm diameter quartz reactor model investigating the impact of different turbulence initialization parameters on flame propagation for a closed-end ignition. Simulation time=10ms. Time step = 0.01ms. Ignition location is 1.39cm from the open end. CH =9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole height = 4 ign 1.2cm. 138
Colorado School of Mines
Figure 5.35 Comparison of methane flame inversion at the open end of the reactor from a closed- end ignition in the 12cm diameter quartz reactor. Top – Temperature contour from Figure 5.35 initialization parameters, k = 1.5m2/s2 and ω = 250 1/s. Bottom – Experimental image showing flame inversion. Simulation time = 28ms. CH =9.5%. Temperature = 293K, Pressure = 82kPa. 4 SM E = 60mJ. Relief hole height = 1.2cm. ign After comparing the turbulent sensitivity results qualitatively, results were compared quantitatively as shown in Figure 5.36 and Figure 5.37. These figures compare the results of estimating turbulent properties based of 5, 10, 20, 30m/s as well as a case which rounds the turbulent values calculated based off 30m/s to k = 1.5m2/s2 and ω = 250 1/s. As can be seen in these figures, there is no one value which absolutely matches experiments perfectly. All of the modeling results predict flame arrival at the open end much faster than experiments. To investigate this, using the case of k = 1.5m2/s2 and ω = 250 1/s, the kernel diameter at time t = 2ms was calculated as dk=0.06m. Complimentary high-speed imaging at 480 fps, 720 pixels, found that the average kernel diameter of three experiments at time t = 2ms was an order of magnitude less, dk=0.008m. At time t = 12.5ms, the measured average flame kernel diameter of experiments was dk=0.06m. From these results, it was concluded that the ANSYS Fluent Spark Model (v17.2) overpredicts the initial kernel expansion of the flame in 2D. Despite this difference, when the results of Figure 5.36 are normalized as shown in Figure 5.37, the rate of increase of flame front propagation velocity in experiments match the 2D model when the turbulent initialization parameters are estimated based off a mean flow velocity of 30m/s, or simplified to k = 1.5m2/s2 and ω = 250 1/s (model – red circles compared to experiments – black diamonds). Additionally, when comparing the 2D model results to experiments, Figure 5.38, the model does a good job at predicting the flame propagation trends towards the open end of the 140
Colorado School of Mines
reactor. Although the model predicts more flame slow-down near the open end, this is difficult to conclude from the experiments because there were no additional ion sensors between x = 0-25cm from the open end. The flame slow down will be discussed in regards to modeling the relief hole in Section 5.1.6. Despite this difference, the flame shape of the model near the open end matches experiments. Therefore, it was concluded that initializing the model based off a mean flow velocity that is approximately 50% of the maximum measured flame front propagation velocity qualitatively and quantitatively matches the flame shape and flame acceleration down the reactor. Thus, 12cm diameter reactor models modeling closed-end ignition uses the following initial turbulent quantities: k = 1.5m2/s2 and ω = 250 1/s. This study was repeated for an open- end ignition in the 12cm diameter reactor and it determined the following quantities most accurately capture flame propagation: k = 0.004m2/s2 and ω = 0.1 1/s. This study was repeated for the other experimental reactors and will be discussed in subsequent sections. Figure 5.36 Flame front propagation velocity (FFPV) results of the 2D, 12cm diameter quartz reactor model investigating the impact of different turbulence initialization parameters on flame propagation for a closed-end ignition. Time step = 0.01ms. Ignition location is 1.39cm from the open end. CH =9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole 4 ign height = 1.2cm. 141
Colorado School of Mines
Figure 5.37 Normalized flame front propagation velocity (FFPV) results of the 2D, 12cm diameter quartz reactor model versus experiments for a closed-end ignition. Time step = 0.01ms. Ignition location is 1.39cm from the open end. CH =9.5%. Temperature = 293K, Pressure = 4 82kPa. SM E = 60mJ. Relief hole height = 1.2cm. ign Figure 5.38 2D, 12cm diameter quartz reactor model results compared to experiments for a closed-end ignition. Time step = 0.01ms. Ignition location is 1.39cm from the open end. CH =9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole height = 4 ign 1.2cm. After determining the most appropriate turbulent initialization parameters, a study was performed investigating the 2D planar assumption by modeling the reactor as axisymmetric. To do this requires modeling the 2D, planar 12cm diameter reactor without a relief hole since the axisymmetric case cannot assume this. Also explored was a comparison of using the k-ω turbulent model to the k-ε turbulence model using initial turbulence parameters based off the same mean flow velocity, 30m/s. Results of these studies are shown in Figure 5.39, Figure 5.40, 142
Colorado School of Mines
and Figure 5.41. Results show that not modeling the relief hole on the closed end of the reactor results in faster flame propagation than modeling the relief hole; this shall be discussed further in the subsequent section, however, it is introduced here in order to make a direct comparison between the 2D planar assumption and axisymmetric assumption. As shown in the figures, assuming axisymmetric using the same model setup results in much faster flame speeds than the 2D planar assumption. Additionally, using the k-ε model results in even faster flame speeds compared to the k-ω turbulence model, which agrees with previous findings (Fig M. , 2019). The difference between these turbulence models is due to the fact that the k-ε turbulence model uses approximations in the boundary layer and is more often used for flows with high turbulence in the bulk of the flow. Therefore, researchers recommend future work performing a sensitivity study of the k- ε model investigating different initial turbulence parameters. In contrast, the k-ω turbulence model is more well-suited for shear flows, which as it has been shown, has quite an impact on the flame shape observed in experiments. Therefore, this manuscript will continue using the k-ω turbulence model, but shall revisit this assumption as models continue to increase in scale. Figure 5.39 Temperature contours of the 2D, 12cm diameter quartz reactor model investigating the impact of mesh size, relief on the closed end, planar versus axisymmetric assumptions, and turbulence model on flame propagation for a closed-end ignition. Simulation time=8ms. Time step = 0.01ms. Ignition location is 1.39cm from the open end (left). CH =9.5%. Temperature = 4 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole height = 1.2cm. ign 143
Colorado School of Mines
5.1.5 Impact of Modeling Ion Sensors As described in Chapter 3, the 12cm diameter quartz flow reactor uses ion sensors to measure the flame front propagation velocity. The ion sensors are shown in Figure 3.5 on page 53, where the electrodes are enclosed in ceramic tubes that are 0.5cm in diameter. For open-end ignition experiments, the ion sensors were located flush with the top of the reactor, but as shown in Figure 3.8 on page 54, the flame resulting from a closed-end ignition requires the sensors to be located at least 2cm down into the reactor to more accurately measure the flame front propagation velocity. However, from the experiments, given the cross-sectional area of the ion probes relative to the cross-sectional area of the reactor it is expected that changing the depth of the ion sensors from 1-2cm will have very little impact on the flame characteristics and flame propagation velocities. To confirm this hypothesis, a study was performed using the 2D and 3D CFD models of the 12cm diameter reactor which included the ion sensors 0.5cm in diameter and 2cm long. Results of modeling the ion sensors in 2D are shown in Figure 5.42 and indicate that the ion sensors can impact methane flame propagation; increasing the overall speed of the flame as well as the turbulence induced downstream of the obstacles. However, modeling this scenario in 2D does not fully resolve turbulence in the third dimension. Results of modeling the ion sensors in 3D are shown in Figure 5.43 and show that the ion sensors do not impact methane flame propagation. Additionally, in 3D the turbulence induced by the ion sensors is much less than that induced by the ion sensors in 2D. This main difference is to the fact that in 2D, the turbulence is not fully being resolved, and thus is overestimated. Coupled with the fact that the ion sensors, in reality, do not take up the entire cross section of the reactor, the 2D model overpredicts the impact of the ion sensors. These results are important because 1) they show that the ion sensors do not have a significant impact on the experimental results and 2) they show that to model turbulence, especially turbulent flame propagation, requires solving the third dimension. Unfortunately, as the models continue to get larger and larger, they require more computational time and solving all scenarios in 3D require the use of parallelization over multiple compute nodes. Thus, it is important when interpreting 2D modeling results to have a strong knowledge of model sensitivity as shown in these studies. 145
Colorado School of Mines
Figure 5.43 Temperature and turbulent intensity contours of the 3D, 12cm diameter quartz reactor model investigating the impact the ion sensors on flame propagation for a closed-end ignition. Simulation time=12ms. Time step = 0.01ms. Ignition location is 1.39cm from the open end. CH =9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. No relief hole. 3D 4 ign Body mesh size = 2mm. 5.1.6 Impact of Modeling Relief Holes As discussed in the experimental setup, the relief holes size on the closed end of the reactor is 1±0.2cm. In most of the experiments, a single relief hole was opened in order to help stabilize the flame front which agrees with findings of other researchers (Andrews & Bradley, 1972; Rallis & Garforth, 1980). However, determining an adequate representation of the hole in the 2D model was done by modeling the relief hole height as a fraction of the total diameter of the reactor. This fraction was determined by the total area of the relief hole divided by the total area of the closed end also shown in the following equation: (5.10) 147
Colorado School of Mines
As can be seen in Equation (5.10), the ratio of the area of the hole, ‘h’, to the area of the reactor, ‘r’, is 10%. For the model this was done by taking an edge split of 10% on the closed end of the reactor, which means the height of the relief hole modeled in 2D is 1.2cm. Results of modeling the relief hole with H=1.2cm versus not modeling the relief hole for an open-end ignition is shown in Figure 5.44 and Figure 5.45 and results are tabulated in Table 6.1. As can be seen in the following two figures, modeling the relief hole allows for improved flame stability and faster flame turn over. However, modeling the relief hole also allows for less pressure built up on the closed end of the reactor leading to slower overall flame speeds. When comparing the flame speeds, modeling the relief hole better estimates the maximum speed of the stoichiometric flame, but overestimates the rich flame speed and underestimates the lean flame speed. Not modeling the relief hole better estimates the lean flame speed and overestimates all other cases. Also, temperature magnitudes from the model agree with fundamental flame theory that a stoichiometric mixture (9.5% methane by volume) is the highest, then rich, and finally the lean case. Taking all of this into account, this research shows that modeling the relief hole in 2D must be taken with caution because the flame stability and propagation velocities are sensitive to the relief hole sized used. For the open-end ignition case without an obstacle, this shall be discussed further in subsequent sections. Figure 5.46 shows results of an open-end ignition with an obstacle wall located 37cm from the open end. This figure compares modeling flame propagation across the obstacle with and without a relief hole of H=1.2cm. Comparing Figure 5.46 to images of flames passing over the obstacle wall in Figure 6.18 shows that not modeling the relief hole results in more realistic flame shapes passing over the wall. Results of modeling the relief hole shows that the flame expands farther axially than radially, whereas images of the flame show that the flame tends to move both axially and radially. Based on these results, it was concluded that modeling an open- end ignition with an obstacle in 2D does not require modeling the current size of the relief hole because it may be artificially allowing faster flame propagation past the obstacle. In the future, a better representation of the relief hole may be obtained by measuring the rate of gas expansion out of the hole and setting the size of the hole for the model based on gas expansion. 148
Colorado School of Mines
Figure 5.44 Temperature contours of the 2D, 12cm diameter quartz reactor model without modeling the relief hole for an open-end ignition. Simulation time=1.4s. Time step = 0.1ms. Ignition location is 11cm from the open end (left). CH =9.5%. Temperature = 293K, Pressure = 4 82kPa. SM E = 60mJ. No relief hole. ign Finally, closed-end ignition results modeling the relief hole versus not modeling the relief hole are shown in Figure 5.47. As shown in this figure, not modeling the relief hole results in flame front propagation velocities over 115% greater than modeling the closed-end ignition with a relief hole. Closed-end ignition experiments predict a maximum flame front propagation velocity of 65m/s and the 2D model with a relief hole predicts a maximum flame speed of 72m/s, which is only 11% greater than experiments (Table 6.2, page 185). Additionally, the rate of flame front propagation increase (or acceleration) of the flame is greater without modeling the relief hole versus modeling the relief hole which was previously shown to match experiments well in Figure 5.37. Finally, as shown in Figure 5.34 and Figure 5.35, the 2D model predicts the flame stretching and flame shape well compared to experiments (Figure 4.8, page 67). Therefore, 149
Colorado School of Mines
Figure 5.47 Flame front propagation velocity location results of the 2D and 3D, 12cm diameter quartz reactor model investigating the impact of modeling the relief hole on the closed end of the reactor. Time step = 0.01ms. Ignition location is 1.39cm from the open end. CH =9.5%. 4 Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole H = 1.2cm. ign 5.1.7 Impact of Extending the Domain One of the major concerns with evaluating the flame shapes observed in Figure 5.35 was that the inversion was due to the zero gauge boundary condition on the open end of the reactor. To explore the impact of the open end boundary condition on a closed-end ignition, a study was performed extending the domain another 1m and initializing it to stagnant, air at 293K and 82kPa. In this scenario, the methane-air mixture was stoichiometric, 9.5% methane by volume, and the mixture was ignited from the closed end of the reactor. Results are presented in Figure 5.48-Figure 5.53 and show that extending the domain can change the exact location of the inversion, but results in similar flame propagation trends. Figure 5.48 compares the flame front propagation velocity versus time when adding the extension slows the flame front propagation velocity because as the volume expands, the flame slows in order to maintain mass flow. Although the flame front propagation velocities are slightly slow, the extended domain values are within 10% of the case not modeling an extension. In general, however, the simulation results are fairly similar and the time cost of running a longer model does not outweigh the results. Finally, another interesting finding was that, unlike the open-end ignition flame which consumes almost all the methane in the reactor domain, the closed-end flame generates enough pressure to push out a significant amount of methane. Unfortunately in the experimental setup, probes were not set up to measure the flow of methane out of the reactor, but other researchers have also noted that these types of flames could push out a significant amount of gas mixture (Bradley & 152
Colorado School of Mines
5.2 Box Reactor Model Setup To understand the impact of reactor shape on methane flame propagation and interaction with obstacles, CFD combustion models of the box experiments were setup as shown in 2D in Figure 5.54. The model settings of the experimental box reactor are the same as the 12cm diameter quartz reactor, including the spark model settings in Table 5.2 (ANSYS Fluent v17.2): • 2D planar assumption for all 2D models • Pressure-Based Solver • Energy Equation • Viscous Standard k-ω Turbulence Model o Low Re Corrections o Shear Flow Corrections • Species Transport o Volumetric Reactions o Stiff Chemistry Solver o Finite Rate Chemistry ▪ Density solved using ideal gas theory ▪ Diffusion solved using kinetic theory ▪ Metghalchi and Keck laminar flame speed theory • Spark Ignition Model • PISO pressure-velocity coupling • 2 levels of mesh adaption on the gradient of temperature every 2-10 time steps • 2D Model: Residuals set to 10-6, dropping at least 3 orders or magnitude • 3D Model: Continuity/velocity residuals set to 10-4, Energy/turbulent 10-6 • Second order in time and space • Time step = 0.1ms for open-end ignition, 0.01ms for closed-end & center ignition • Boundary Conditions: o Walls – no slip, adiabatic o Relief opening – 0 gauge pressure outlets o Obstacles– no slip, adiabatic, assuming no heat transfer into the material 156
Colorado School of Mines
Figure 5.54 Schematic of 2-D combustion model setup of the box experiments. In this model, ignition was located near the open end, closed end, and the center of the gob in order to understand how the flame might propagate if it were ignited from within the gob and might travel towards the longwall face. Although it is not well known if an ignition can travel from deep within the gob to the working face, a center ignition is most indicative if there is an ignition in the fallen roof gob near the tailgate entry, where explosive gas zones are known to exist. As discussed previously, determining the size of the relief opening in 2D is difficult and this research recognizes that this can be considered a tuning parameter in this model. For the experiment box, if one were to determine the size of the relief opening based on the ratio of areas then the opening would be 5.6cm. However, if the relief opening was determined based off the hydraulic diameter of the relief, then it would be 8.8cm. Averaging these two methods leads to a relief opening size of 7.2cm, which is approximately the height of the opening, 7cm. Since this average value is similar to the height of the opening, researchers are using a relief size of 7cm for the 2D model. Noted that for the 12cm diameter reactor the height of the relief was also used in the 2D model. However, to reiterate, the relief opening can be seen as a tuning parameter of the model. Additionally, estimating the turbulence parameters, as previous shown with the 2D 12cm diameter reactor model, is difficult and requires a sensitivity analysis to be performed. Unfortunately Equation (5.4) and Equation (5.5) are derived for pipe flow, thus, researchers determined the appropriate turbulence initialization parameters by comparing CFD flame 157
Colorado School of Mines
propagation results to the flame kernel growth and propagation recorded from experimental high- speed imaging. To begin, values from the 2D, 12cm diameter model were used: open-end ignition k = 0.004m2/s2 and ω = 0.1 1/s, closed-end ignition k = 1.5m2/s2 and ω = 250 1/s. The turbulent kinetic energies and specific dissipations were changed orders of magnitude away from these values. Based on flame shape and flame speed, it was determined the following turbulence values best represented what was observed from experiments: open-end ignition k = 0.001m2/s2 and ω = 0.1 1/s (corresponding to flow velocities less than 1m/s), closed-end ignition k = 1.5m2/s2 and ω = 25 1/s. Result comparisons are shown in subsequent sections of this manuscript. 5.2.1 2D Mesh Independence Study To begin modeling the experimental box as described in Section 3.1.2, a 2D model was created and a variety of meshes were compared: 2mm, 1mm, and 0.5mm. All of the meshes were quadrilateral dominant as shown in Figure 5.55 and mesh statistics are summarized in Table 5.10. As can be seen the mesh is primarily structured, but some of the cells are skewed and have aspect ratios larger than 1. However, despite these drawbacks, the meshes have fairly good quality. Additionally, for all models which include obstacles, the obstacles have a constant edge sizing of 0.5mm as shown in Figure 5.56. Edge sizing was used on the obstacles to help resolve the boundary layers as the flame interacts with the obstacles. Additionally, all models presented employ mesh adaption on the gradient of temperature, 2 levels every 2-10 time steps depending on whether the flame is highly turbulent or mainly wrinkled laminar. Figure 5.55 Image of the 2D, box reactor mesh with a quadrilateral dominant 0.5mm body mesh. 158
Colorado School of Mines
Table 5.10 2D, box reactor mesh statistics for a 2mm, 1mm, and 0.5mm body mesh: number of elements, number of nodes, minimum orthogonality quality, maximum skewness, and maximum aspect ratio. No edge sizing. Minimum Number of Number of Maximum Maximum Orthogonality Elements Nodes Skewness Aspect Ratio Quality 2mm body mesh 42,591 43,010 0.7 0.3 3.2 1mm body mesh 169,763 170,594 0.8 0.2 2.7 0.5mm body mesh 688,999 690,640 0.6 0.3 4.1 Figure 5.56 Image of the 2D, box reactor mesh with a quadrilateral dominant 1mm body mesh and 0.5mm edge sizing on obstacle boundaries. A mesh independence study for the 2D box model was performed alongside determining the appropriate turbulence initialization settings. Mesh independence results were determined by taking a seeded diagonal line as shown in Figure 5.57 and results are tabulated in Table 5.11. As shown, a mesh size of 1mm results in flame front locations less than 1% different than the 0.5mm case. Additionally, the standard deviation and maximum errors are less than 1% for the 1mm mesh. Therefore, all 2D models use a 1mm quadrilateral base mesh with a 0.5mm edge sizing on all obstacles to resolve the boundary layers. 159
Colorado School of Mines
Table 5.11 Average percent error, standard deviation, and maximum error in the flame front location between different body meshes for the 2D, box model. Percent Error (%) 2mm : 1mm 1mm : 0.5mm Average 1.9 0.0 Standard Deviation 2.5 0.6 Maximum Error 10 0.5 Figure 5.57 Schematic of the 2-D combustion model setup of the box experiments with a yellow line indicates seeded line used to compared flame front propagation towards the relief opening for different mesh sizes for a confined ignition. 5.2.2 3D Mesh Independence Study A 3D mesh was created for the experimental box using a cut cell, quadrilateral dominant, method as shown in Figure 5.58. Two different meshes were compared for the 3D, box model, a 5mm mesh and a 2.5mm mesh. Mesh statistics are presented in Table 5.12 and show that the orthogonality quality is very close to one with maximum skewness below 0.2. Also, the maximum aspect ratios are close to 2 which means the dimensions of the cells are mostly proportional. Altogether, these statistics show a very good quality mesh. As a note, the 2.5mm cut cell mesh resulted in over 1.5 million cells. A mesh of 1-1.25mm was attempted, but resulted 160
Colorado School of Mines
in a large amount of cells which was not feasible to run on an 8 core compute. In the future, running across nodes should be investigated so that finer meshes can be used. To help overcome some of the error of a 2.5mm mesh, aggressive mesh adaption was used: 3 levels every 2 time steps. As will be shown, this helped resolved the flame front propagation well, but resulted in some error associated with other flow quantities. Figure 5.58 Image of the 3D, box reactor mesh with a quadrilateral dominant 2.5mm body mesh. Table 5.12 3D, box reactor mesh statistics for a 5mm and 2.5mm body mesh: number of elements, number of nodes, minimum orthogonality quality, maximum skewness, and maximum aspect ratio. No edge sizing. Minimum Number of Number of Maximum Maximum Orthogonality Elements Nodes Skewness Aspect Ratio Quality 5mm body mesh 216,039 228,480 1 0 1.8 2.5mm body mesh 1,704,543 1,757,663 0.8 0.2 2.2 To begin, a mesh independence study was also undertaken for the 3D box model using the same initialization settings as the 2D model for a confined ignition (k = 1.5m2/s2 and ω = 25 1/s). A seeded line extending from the confined ignition towards the relief opening was used to determine flame front location and compare meshes. This research recognizes that in 3D the flame can propagate in all directions, but this analysis was performed to directly compare flame location to the average velocities observed in experiments., The 3D mesh independence study 161
Colorado School of Mines
shows that using mesh adaption on the temperature gradient, 3 levels every 2 time steps (0.01ms time step), with a coarse 5mm mesh is within 1±1% of the flame front location compared to a 2.5mm mesh, with a maximum relative error below 2%. Figure 5.59 Schematic of 3-D combustion model setup of the box experiments. XY plane is located in the middle of the reactor, at z = 0.075m. Yellow line indicates seeded line used to compared flame front propagation towards the relief opening for different mesh sizes for a confined ignition. Table 5.13 Average relative error and standard deviation in the flame front location between different body meshes of the 3D box model for a confined ignition. Adaption on the temperature gradient, 3 levels every 2 times steps. Time (s) Relative Error (%) 5mm : 2.5mm 0.02 0.4 0.045 0.0 0.06 -0.8 0.08 -1.6 0.105 -1.6 0.12 -0.5 Average -0.7 Standard Deviation 0.8 As discussed with the 3D model of the 12cm diameter reactor, evaluating mesh independence in 3D is quite complex and requires investigation of flow properties in 3D space. Thus far for the 3D, box reactor model, the flame front location as a function of time has been 162
Colorado School of Mines
evaluated and the relative error between a 5mm and 2.5mm mesh is less than 2%. The next step of the mesh independence process is to plot and compare flow quantities. To do this, three (3) XZ planes were compared with two (2) seeded lines at different locations as shown in Figure 5.60. Temperature contours are shown as a function of time in Figure 5.61, Figure 5.62, and Figure 5.63; as shown in these figures, the temperature contours are similar, but the 5mm mesh shows a much coarser resolution. Results of probing the seeded lines (999 points were taken and averaged) are tabulated in Table 5.14 and Table 5.15. Although these results are only tabulated for time t=11ms, they have been reviewed over several time steps. In general, results show that evaluating the 5mm and 2.5mm mesh based on the total temperature results in average percent differences less than 5%, though the standard deviations range from 7-11%. Note that grid adaption is used in these models and resolves the flame front by employing 3 levels of mesh adaption on the gradient of temperature every 2 time steps. Since the grid adaption resolves the temperature with higher accuracy compared to the flow velocities, the large average percent differences calculated based off the fluid velocity magnitude is of no surprise, 15-40% differences. This was also observed with the 3D model of the 12cm diameter reactor. Based on these results, this research has concluded that for the 3D box model, a base mesh of 5mm with aggressive mesh adaption is adequate to resolve the flame front, but a base mesh of 2.5mm is highly recommended to better resolve local flow velocities and temperature gradients. Figure 5.60 Schematic of 3-D combustion model setup of the box experiments showing three XZ planes with two seeded lines per plane used to extract data for further mesh independence investigation. 163
Colorado School of Mines
Figure 5.63 Temperature contours of the 3D, box reactor model investigating the impact of mesh size on flame propagation. Temperature and velocity data was extracted from the seeded lines on XZ planes. Simulation time = 11ms. Time step = 0.01ms. Confined ignition located opposite the relief. CH = 9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Body mesh size = 4 ign 5mm and 2.5mm. Table 5.14 Average percent difference and standard deviation in the total temperature between different body meshes of the 3D box model for a confined ignition at time, t = 11ms. Body mesh sizes compared are 5mm and 2.5mm. Lines are seeded with 999 points. XZ Plane at XZ Plane at XZ Plane at y = 0.125m y = 0.25m y = 0.375m Top Bottom Top Bottom Top Bottom Line Line Line Line Line Line Average Difference (%) 0 1 3 3 4 4 Standard Deviation (%) 8 7 9 9 11 11 Table 5.15 Average percent difference and standard deviation in the fluid velocity magnitude between different body meshes of the 3D box model for a confined ignition at time, t = 11ms. Body mesh sizes compared are 5mm and 2.5mm. Lines are seeded with 999 points. XZ Plane at XZ Plane at XZ Plane at y = 0.125m y = 0.25m y = 0.375m Top Bottom Top Bottom Top Bottom Line Line Line Line Line Line Average Difference (%) 40 32 23 15 23 22 Standard Deviation (%) 43 32 18 13 16 16 165
Colorado School of Mines
5.2.3 Mesh Adaption Study As shown in the previous section, for the 3D model of the experimental box a coarse mesh size of 5mm is adequate to resolve the flame front within 2% of a 2.5mm mesh using mesh adaption on the temperature gradient. A further study was performed to determine the best settings to use for the mesh adaption by comparing the frequency at which the model is adapting and the levels of adaption. Results of this study are shown in Table 5.16 and Figure 5.70. As can be seen, using 2 levels of adaption results in shorter simulations times than using 3 levels of adaption and the less frequent the mesh adaption the less total simulation time. Based on the total temperature on the diagonal line (shown in Figure 5.59 on page 162), for this coarse grid using 3 levels of adaption every 2 time steps results in flames that are almost 5% faster than using 2 levels of adaption. However, the simulation time of modeling the box with 3 levels every 2 times steps is twice the amount of simulation time when using 2 levels of adaption. Therefore, it is recommended that for flame front accuracy, 3 levels every 2 times steps is used, but 2 levels every 2 time steps may be appropriate when simulation time is of concern. This is important as models continue to increase in size and simulation time becomes more important to balance. Figure 5.70 Total temperature versus location results of the 3D, box reactor model investigating the impact of mesh adaption settings. Simulation time = 5ms. Time step = 0.01ms. Confined ignition. CH = 9.5%. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. 4 ign 169
Colorado School of Mines
Table 5.16 Comparison of mesh adaption settings and simulation times of the 3D box model for a confined ignition. Adaption on the temperature gradient with varying levels and frequency. Models run on an 8 core compute, 3.06GHz, 24GB RAM. Levels 2 2 2 3 3 Frequency 1 2 5 1 2 Simulation Time 14hr, 42min 12hr, 33min 11hr, 27min 48hr 30hr 5.3 2D 71cm Diameter Reactor Model Setup A 2D combustion model of the 71cm diameter, 6.1m long reactor was developed as shown in Figure 5.71. Figure 5.71 Schematic of 2D combustion model geometry setup of the 71cm diameter steel reactor located at Edgar Experimental Mine in Idaho Springs, CO. The model settings of the 71cm diameter reactor are the same as the 12cm diameter quartz reactor and box model, including the spark model settings in Table 5.2 (ANSYS Fluent v17.2): • 2D planar assumption for all 2D models • Pressure-Based Solver • Energy Equation • Viscous Standard k-ω Turbulence Model o Low Re Corrections o Shear Flow Corrections • Species Transport 170
Colorado School of Mines
o Volumetric Reactions o Stiff Chemistry Solver o Finite Rate Chemistry ▪ Density solved using ideal gas theory ▪ Diffusion solved using kinetic theory ▪ Metaghalchi and Keck laminar flame speed theory • Spark Ignition Model • PISO pressure-velocity coupling • Continuity/energy residuals set to 10-5, Velocity/species residuals set to 10-3 • Second order in time and space • Time step = 0.01ms • Boundary Conditions: o Walls – steel, 5mm roughness height, adiabatic o Obstacles– no slip, adiabatic, assuming no heat transfer into the material 5.3.1 2D Turbulence Model Settings & Mesh Independence Study To begin modeling the 71cm diameter reactor presented in Section 3.2, a 2D quadrilateral dominant mesh was created as demonstrated in Figure 5.72. As can be seen in this image, the mesh is primarily structured, but based on the mesh statistics shown in Table 5.17 some of the cells are slightly skewed and can have large aspect ratios. Two main base meshes were compared, a 4mm and 2mm mesh, however a third mesh was created with slightly larger cells: 5mm base mesh with 1mm edge sizing on the walls as shown in Figure 5.73. The 5mm base mesh with 1mm edge sizing results in a similar number of cells as the 4mm mesh, but the maximum skewness increases due to the unstructured mesh. This 5mm mesh with 1mm edge sizing was primarily used when considering flame propagation across obstacles; unfortunately, Fluent’s meshing client had difficulties meshing at 4mm and 2mm when considering obstacles in the flow, thus a slightly coarser mesh was used with edge sizing on the obstacles to help resolve boundary layers. Additionally, mesh adaption was employed to help resolve the flame front in the bulk flow. 171
Colorado School of Mines
Table 5.17 2D, 71cm diameter reactor mesh statistics for a 2mm body mesh, a 4mm body mesh, and a 5mm body mesh with 1mm edge sizing: number of elements, number of nodes, minimum orthogonality quality, maximum skewness, and maximum aspect ratio. Minimum Number of Number of Maximum Maximum Orthogonality Elements Nodes Skewness Aspect Ratio Quality 2mm body mesh 979,191 982,147 0.4 0.4 6.1 4mm body mesh 250,358 251,866 0.4 0.4 5.3 5mm body mesh, 254,386 258,813 0.5 0.7 3.2 1mm edge sizing Determining the turbulence parameters, as previous shown with the 12cm diameter reactor model and box model, can be difficult, but researchers used their knowledge of flame front propagation velocities and flame shape to help determine the most appropriate parameters. For example, for the 12cm diameter reactor the most appropriate turbulent quantities were based off an average flow velocity of 30m/s, which was slightly less than half the maximum flame front propagation velocity. Thus, to begin this for the 2D, 71cm diameter reactor model, researchers began with estimating the flow quantities based off an average flow velocity of 40m/s, corresponding to k = 1m2/s2 and ω = 45 1/s (maximum flame front propagation velocity was approximately 125m/s). After this, a parametric study was performed changing the kinetic energy and the dissipation rate orders of magnitude. Results of this study are shown in Figure 5.74 and show that estimating the turbulence parameters as k = 0.1m2/s2 and ω = 45 1/s gives the closest agreement between the 2D CFD model and experiments. After determining the best turbulence parameters, a 2D mesh independence study was performed and results are shown in Table 5.18. Results show that the flame front calculated using a 4mm body mesh is within 15% the flame front for a 2mm body mesh. Although an error of 15% seems large, Figure 5.75 and Figure 5.76 show there is a very close agreement between the 4mm and 2mm body meshes. Unfortunately comparing these meshes to a 1mm body mesh was infeasible with the current computational power. Additionally, the total simulation time for the 4mm body mesh was 5 days and the total simulation time for the 2mm body mesh was 15 days. Extrapolating this to a 1mm body mesh would mean an estimating simulation time of 45 days. Taking all of this into account, researchers recommend a 4mm body mesh due to ease of meshing, reduced simulation times, and reasonable flame front propagation velocity estimates. 173
Colorado School of Mines
Figure 5.74 Flame front propagation velocity versus distance for the 2D 71cm diameter reactor model for a closed-end ignition with varying turbulence initialization parameters. Ignition location is 28.5cm from the closed end. Time step = 0.01ms. CH = 9.5%. Body mesh size = 4- 4 5mm. Temperature = 295K, Pressure = 76kPa. SM E = 60mJ. ign Table 5.18 Average percent error, standard deviation, and maximum error in the flame front location between different body meshes for the 2D, 71cm diameter reactor model. k = 0.1m2/s2 and ω = 45 1/s. 4mm mesh uses grid adaption on the temperature gradient, 2 levels every 2 time steps. 4mm : 2mm Average Error (%) 15 Standard Deviation (%) 7 Maximum Error (%) 25 Figure 5.75 Flame front propagation velocity versus time for the 2D 71cm diameter reactor model for a closed-end ignition with varying body mesh sizes. k = 0.1m2/s2 and ω = 45 1/s. Ignition location is 28.5cm from the closed end. Time step = 0.01ms. CH = 9.5%. Body mesh 4 size = 4mm and 2mm. Temperature = 295K, Pressure = 76kPa. SM E = 60mJ. ign 174
Colorado School of Mines
Figure 5.78 Temperature contours of the axisymmetric 2D, 71cm diameter quartz reactor model using k = 1.5m2/s2 and ω = 45 1/s. Simulation time=14ms. Time step = 0.01ms. Ignition location is 28.5cm from the closed end. Time step = 0.01ms. CH = 9.5%. Body mesh size = 4-5mm. 4 Temperature = 295K, Pressure = 79kPa. SM E = 60mJ. ign Researchers also investigating possibly using an asymmetric model instead of a 2D planar model. Results of the investigating are presented in Figure 5.77 and Figure 5.78. As can be seen in these figures, the axisymmetric case predicts faster flame front propagation velocities than the 2D planar case agreeing with previous observations modeling the 12cm diameter quartz reactor. Although the axisymmetric case predicts faster flames than assuming a 2D planar geometry, the predict flame shape shown in Figure 5.78 shows flame inversion at the line of symmetry. However, if the shape of the flame is assumed to look more like a closed-end ignition flame (Figure 2.7) then the axisymmetric case does not predict this well. Unfortunately there is no photographic evidence of the shape of this flame to help determine whether the axisymmetric assumption is valid. Also, one of the reasons the axisymmetric case is predicting faster flames than the 2D planar case is because the same turbulence parameters were used. This is important because in the planar case the turbulence parameters compensate for some of the turbulence in the radial direction, whereas the axisymmetric case already takes this swirl into account. 176
Colorado School of Mines
CHAPTER 6 CFD MODELING OF PREMIXED METHANE-AIR DEFLAGRATIONS IN EXPERIMENTAL REACTORS The previous chapter, Chapter 5, discusses the model setups for the 12cm diameter reactor, experimental box, and 71cm diameter reactor. Additionally, Chapter 5 discusses the combustion model settings such as the turbulence model, turbulence initialization parameters, and spark initiation methods and how these settings impact model predicted methane gas deflagrations. This chapter summarizes the major modeling results of the 12cm diameter reactor, experimental box, and 71cm diameter reactor in 2D and 3D and compares these results to flame front propagation velocities, flame trends, and flame shapes observed from experiments presented in Chapter 4. 6.1 2D 12cm Diameter Quartz Reactor Modeling Results Experiments in the 12cm diameter reactor are presented in Section 4.1-4.5 and the CFD model settings are presented in Section 5.1. This section shall present complimentary 2D and 3D modeling results and discuss the model performance compared to experiments and other researchers. 6.1.1 Empty Reactor: Impact of Mixture Stoichiometry It is well known that EGZs exist underground in longwall coal mines and although there has been significant research modeling these EGZs (Juganda, Brune, Bogin, Grubb, & Lolon, 2017; Ren & Edwards, 2000; Tanguturi, Balusu, & Bongani, 2017), still not a lot is known about the location, movement, and actual composition since a mine environment is inherently transient. Therefore, it is important to capture the impact of stoichiometry on methane flame propagation in the CFD models in order to build a comprehensive model capable of modeling methane gas explosions in underground coal mines. Previous open-end ignition experiments investigating the impact of methane-air mixture stoichiometry on flame propagation in the 12cm diameter quartz flow reactor found that the stoichiometric flame (9.5% methane by volume) was fastest, followed by the lean (7.5%) and rich flame (11.5%) (Figure 4.3 and Figure 4.4 in Section 4.1). However, according to laminar flame theory, a rich 11.5% laminar flame at standard temperature and pressure is faster than a 177
Colorado School of Mines
stoichiometric or 7.5% lean flame as shown in Figure 2.4. The reason for this difference is that in the experiments, a diffusion flame was observed on the open end of the reactor during the rich flame propagation. This diffusion flame acted similar to a counter balance, slowing down the main rich flame front, reducing the overall speed of the flame. Figure 6.1: 2D 12cm diameter reactor results of investigating the impact of mixture stoichiometry on methane flame front location versus time for an open-end ignition. Ignition location is 11cm from the open end. Time step = 0.1ms. CH = 7.5, 9.5, 11.5%. Body mesh size 4 = 1mm, 0.25mm edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. No relief ign hole. Table 6.1 Table comparing experimental data to the 2D 12cm reactor model predictions of maximum flame front propagation velocity (FFPV) and standard deviation of the mean for an open-end ignition with different mixture stoichiometries. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. ign 2D Model with Relief 2D Model without Relief Experiments H=1.2cm Mixture 7.5% 9.5% 11.5% 7.5% 9.5% 11.5% 7.5% 9.5% 11.5% Stoichiometry Max FFPV 1.05 1.65 1.65 0.85 1.39 1.46 1.05 1.32 0.95 (m/s) Standard Deviation - - - - - - 0.04 0.04 0.14 (m/s) These experiments were modeled using the 2D 12cm diameter reactor model (without a relief hole) and results are presented in Figure 6.1 and Table 6.1. Results show that the rich and 178
Colorado School of Mines
stoichiometric flames propagate at similar speeds, but near the end of the reactor the rich flame propagates faster than the stoichiometric flame. The lean flame propagates slowest, as expected, and the maximum flame front propagation velocity calculated from the lean flame matches the experiments well. The 2D model without a relief hole predicts a 25% increase in the maximum flame front propagation velocity of the stoichiometric flame measured in experiments. This is mainly due to the simplified 2-step methane-air mechanism summarized in Table 5.1: the first reaction of methane of oxygen was experimentally determined at high temperatures (>1000C) and an equivalence ratio range of 0.05-0.5 and the second reaction of CO and oxygen is validated over an equivalence ratio range of 0.04-0.5 (Dryer & Glassman, 1973). Since these mechanisms have been validated under the lean conditions, it makes sense that the model results would more accurately predict the lean flame propagation. In summary, the increased speeds predicted by the 2D model for the stoichiometric and rich case can be attributed to the reduced 2-step methane-air mechanism employed coupled with the simplicity of the 2D model. Since the chemistry is reduced to a 2-step mechanism, the heat release by the reaction of methane and oxygen is overpredicted leading to faster flame speeds for the stoichiometric and rich flame. As discussed, one of the main advantages of the 12cm diameter quartz reactor is the optical access to the flame, allowing validation of the models based on flame speeds as well as flame shape and trends. Thus, in addition to capturing the overall flame trends and approximately flame front propagation velocities, the 2D 12cm diameter reactor model (without a relief) also captures the general flame shape as shown in Figure 6.2, Figure 6.3, and Figure 6.4. As can be seen in these transient images, the flame propagates and eventually turns over due to hot, buoyant product gases rising to the top of the reactor and pushing over the flame front agreeing with observations by other researchers (Guenoche & Jouy, 1953). Compared to experimental images in Figure 4.5, Section 4.1, the model captures this turn over well, but later in the flame development than observed in experiments. This is due to the fact that buoyancy is a slow process and it is difficult to accurately capture this phenomenon, which is a 3D process. Additionally, the 2D model without a relief shows that the stoichiometric flame does not fully turn over into the classic angled flame. This is because the model does not include the relief hole on the closed end of the reactor which increases the pressure build up such that it is more difficult for the flame to turn over. These results also help demonstrate the need to accurately 179
Colorado School of Mines
model the experiment and the sensitivity of the model to the pressure relief hole, and if ignored, will provide results which overpredict the maximum flame front propagation velocities. To investigate this flame turnover discrepancy, additional 2D models were run with the relief hole, H=1.2cm; flame shapes are presented in Figure 6.5 and maximum flame front propagation velocities are shown in Table 6.1. As shown, modeling the relief allows for some additional flame stability and slightly less pressure build up which allows the buoyant gases time to rise and push over the top of the flame, decreasing the overall speed of the flame. Thus, the resulting maximum flame front propagation velocity of the modeled stoichiometric flame is within 5% of the measured flame speed. These results are important because they show that the model accurately captures the flame propagation shape and trends and predicts the maximum flame front propagation velocity well in 2D. These model results also capture the flame temperature trends well: the highest being stoichiometric, then rich, followed by the lean case, which agrees with fundamental flame theory (Turns, 2012). Additionally, the results show how small changes to the 2D model can result in large changes to the predicted flame shapes and speeds. Most of these differences can be attributed to the model being 2D and assuming a 2-step methane-air chemistry mechanism, but these differences should be noted and taken into account as the model is transformed to 3D. Experiments showed that a confined ignition at the closed end of the reactor resulted in flame front propagation velocities approximately 50 times greater than an unconfined, open-end ignition. Additionally, the closed-end flame had a very different shape than the open-end flame as shown in Figure 4.5 and Figure 4.8 in Section 4.1. Since the confined, closed-end flame was significantly faster than the open-end flame, buoyancy was essentially negligible, and the flame looks more parabolic which agrees with observations made by other researchers (Clanet & Searby, 1996; Ellis & Wheeler, 1928; Guenoche & Jouy, 1953). As discussed in Chapter 5, modeling a closed-end ignition in the 12cm diameter reactor in 2D is most accurate by modeling the relief hole on the closed end of the reactor. Results of modeling a closed-end ignition with varying stoichiometry is shown in Figure 6.6-Figure 6.10 and Table 6.2 on pages 184-188. Comparing modeling results in Figure 6.6 to experimental results shown in Figure 4.6 and Figure 4.7 of Section 4.1, the trends of flame acceleration match experiments, but the rich flame travels faster than observed in experiments. This is due to the 2- step methane air mechanism predicting faster conversion of methane to CO and CO instead of 2 180
Colorado School of Mines
intermediate species such as OH and other radicals, leading to faster rich flames. Despite this simplification, the maximum flame front propagation velocities predicted by the 2D model are within 12% of the experiments. Additionally, the flame shapes predicted by the 2D model match well with experimental images; the flame is the traditional “finger shape” and due to flame stretching, the flame tears apart. Overall, the 2D model predicts the lean and stoichiometric flame well, but overpredicts the speed of the rich flame. The model does a good job of predicting flame acceleration and flame shape, but shows a significant flame slow-down near the open end of the reactor. This slow down occurs in the last 25-35cm of the reactor, where there is only one sensor so it is difficult to determine whether the slow down occurs experimentally. Also, using high-speed imaging at 240fps only shows 1 or 2 frames near the open end which is not sufficient to adequately resolve the change in propagation velocity during the last 25cm of the reactor. Figure 6.2 Temperature contours of methane flame propagation for an open-end ignition with varying stoichiometry at time, t=0.2s. Ignition location is 11cm from the open end. Simulation time=0.2s. Time step = 0.1ms.CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. No relief hole. ign 181
Colorado School of Mines
Figure 6.10 Temperature contours of methane flame propagation for a closed-end ignition with varying stoichiometry at time, t = 25ms. Ignition location is 1.39cm from the open end. Simulation time = 25ms. Time step = 0.01ms.CH = 9.5%. Body mesh size = 1mm, 0.25mm 4 edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole, H=1.2cm. ign 6.1.2 Modeling an Obstacle Wall As described in Section 2.6 and Section 2.3, obstacles can greatly accelerate flames and in some cases, can transition a flame from a deflagration to a detonation. Therefore, it is important to capture the effects of obstacles on methane gas deflagrations since methane gas explosions in longwall coal mines can occur in working areas where equipment, miners, etc. are located. Experiments have shown flame acceleration across an obstacle wall made of 6.35mm diameter spheres as shown in Chapter 4, Section 4.3. In the experiments, 6.35mm diameter solid glass spheres were arranged in a wall geometry such that the flame could only pass through the void space above the obstacle wall. The height of the walls is 7.62cm and the axial width of the walls are 6.35mm. The obstacle walls were modeled in two difference ways as shown in Figure 6.11: a wall of spheres (which is best representative of experiments) and a smooth, solid, rectangular wall. The sphere wall was made of 6.35mm diameter spheres and solid connections 188
Colorado School of Mines
were made between the spheres to not allow for any flame propagation between the spheres and also reduces meshing complexities. If the spheres were left with one point of contact this would create a pinch point, which is difficult to mesh and would require many small triangle cells. To compare to the sphere wall, a solid wall of H=7.62cm and W=6.25mm was created and both walls were meshed with an edge sizing of 0.25mm to help resolve boundary layers forming on the obstacles. Results of comparing the sphere wall to the solid wall are shown in Figure 6.11, Figure 6.12, and Figure 6.13. As can be seen, as the flame approaches the wall there is a very slight difference in flame shape and the turbulent intensity contours are different as well. In general, the sharp edges of the solid wall tend to increase the local turbulence near the top of the wall due to vortex shedding. However, because these geometries are small and thin, the flame front location is unaffected by the slight changes in local turbulence across the obstacle (<1% difference). Due to ease of meshing, researchers will continue using the solid wall geometry to investigate the impact of the obstacle walls on methane flame propagation. Figure 6.11 Temperature contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle wall modeled as solid, connected spheres versus a straight, solid rectangular wall. Obstacle: 6.35mm spheres arranged in a wall geometry and a solid, smooth wall, H = 7.62cm, L = 6.25mm. Ignition location is 11cm from the open end (left). Simulation Time = 0.1s. Time step = 0.1ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. No relief hole. ign 189
Colorado School of Mines
Figure 6.12 Turbulent intensity contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle wall modeled as solid, connected spheres versus a straight, solid rectangular wall. Obstacle: 6.35mm spheres arranged in a wall geometry and a solid, smooth wall, H = 7.62cm, L = 6.25mm. Ignition location is 11cm from the open end (left). Simulation Time = 0.1s. Time step = 0.1ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. No relief hole. ign Figure 6.13 2D 12cm diameter reactor results of investigating the impact of modeling the obstacle wall on methane flame front location versus time for an open-end ignition. Obstacle: 6.35mm spheres arranged in a wall geometry and a solid, smooth wall, H = 7.62cm, L = 6.35mm. Ignition location is 11cm from the open end. Time step = 0.1ms. CH = 9.5% Body 4 mesh size = 1mm, 0.25mm edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = ign 60mJ. No relief hole. Experimental results in Figure 4.24 and Figure 4.25 in Section 4.2.3 show that methane flames accelerate across obstacle walls, but as the obstacle is moved further from ignition, the 190
Colorado School of Mines
relative acceleration decreases. This trend was investigated using the 2D 12cm diameter combustion model by modeling an obstacle wall, H = 7.62cm L = 6.35mm, at different locations in the reactor, 37cm, 62cm, and 87cm from the open end. Results of modeling the obstacle wall at difference locations is shown in Figure 6.14, Figure 6.15, and Figure 6.16. Results show that an obstacle wall at 37cm results in the greatest acceleration across the obstacle, but the flame arrives at the closed end slower compared to an obstacle wall at 62cm or 87cm. As can be seen, as the obstacle wall is moved from 37cm from the open end to 62cm and 87cm, the relative acceleration across the obstacle decreases. Additionally, Figure 6.16 shows how the modeled flame passes over the obstacle; the flame tends to move in all directions, meaning both axially and radially as it passes over the obstacle wall. These results show that the modeled flame shape and trends across the obstacle match experiments, which is important to capture because the mechanism for flame acceleration across an obstacle for an unconfined (open-end ignition) is much different than a confined (closed-end ignition) which will be presented subsequently. Figure 6.14 2D 12cm diameter reactor results of investigating the impact of obstacle wall location on methane flame front location versus time for an open-end ignition. Obstacle: solid, smooth wall, H = 7.62cm, L = 6.35mm. Ignition location is 11cm from the open end. Time step = 0.1ms. CH = 9.5% Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 293K, 4 Pressure = 82kPa. SM E = 60mJ. No relief hole. ign 191
Colorado School of Mines
Figure 6.15 2D 12cm diameter reactor results of investigating the impact of obstacle wall location on methane flame front propagation velocity versus distance for an open-end ignition. Obstacle: solid, smooth wall, H = 7.62cm, L = 6.35mm. Ignition location is 11cm from the open end. Time step = 0.1ms. CH = 9.5% Body mesh size = 1mm, 0.25mm edge sizing. Temperature 4 = 293K, Pressure = 82kPa. SM E = 60mJ. No relief hole. ign Figure 6.16 Images of stoichiometric methane-air flame passing over a glass sphere wall, H=7.62cm, L=12.35mm compared to 2D modeling results. Flame travels from left to right. CH 4 = 9.5%. Operating conditions 293K, 82kPa. E =60mJ. No relief hole modeled. ign Experiments of a closed-end ignition flame propagation over an obstacle wall was presented in Figure 4.16 and Figure 4.17 in Section 4.2.1. Results showed that even a short obstacle wall greatly accelerated the closed-end ignition flame across the obstacle wall. Additionally, the method of flame acceleration past an obstacle wall for an open-end versus closed-end ignition flame were different. As discussed, the open-end flame moves both radially and axially across the obstacle wall. The closed-end flame passed over the wall, accelerating 192
Colorado School of Mines
mainly in the axial direction, creating flow separation past the obstacle wall. This flow separation is important because it forms eddies on the downstream side of the wall, trapping unburned gases and increasing fluid motion, agreeing with observations made by other researchers (Chapman & Wheeler, 1926) (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). The experiments presented in Figure 4.16 were modeled and results are shown in Figure 6.17, Figure 6.18, and Figure 6.19. 2D modeling results show that as the flame approaches the obstacle wall it is slowed down and then accelerated up to 88m/s across the wall. In the experiments, the recorded flame front propagation velocity across the wall was 81m/s, which is a 9% difference compared to the 2D model. Also the modeling results show that the shape of the flame propagating across the wall matches the flow separation observed from experiments. Turbulent intensity contours at 24ms show that there is more fluid motion downstream of the wall as the flame passes over. Overall the 2D closed-end ignition model accurately captures the flame acceleration across the wall both quantitatively and qualitatively. Figure 6.17 2D 12cm diameter reactor results of investigating the impact of an obstacle wall on methane flame front propagation velocity versus time for a closed-end ignition. Obstacle: wall H = 3.81cm, L = 6.35mm located at 0.37m from the open end (1.13m from the closed end). Ignition location is 1.39cm from the open end. Time step = 0.01ms. CH = 9.5% Body mesh size 4 = 1mm, 0.25mm edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. Relief ign hole H = 1.2cm. 193
Colorado School of Mines
Figure 6.18 Images of stoichiometric methane-air flame passing over a glass sphere wall, H = 3.81cm, L = 6.35mm compared to 2D modeling results. Flame travels from right to left. CH = 4 9.5%. Operating conditions 293K, 82kPa. E = 60mJ. Relief hole H = 1.2cm. ign Figure 6.19 Turbulent intensity contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle wall for a closed-end ignition. Obstacle: wall H = 3.81cm, L = 6.35mm located at 37cm from the open end. Ignition location is 1.39cm from the open end. Time step = 0.01ms. CH = 9.5% Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 4 293K, Pressure = 82kPa. SM E = 60mJ. Relief hole H = 1.2cm. ign 6.1.3 Modeling a Checkerboard Obstacle After modeling the impact of an obstacle wall on methane flame enhancement, researchers also modeled a checkerboard obstacle. The main purpose of the checkerboard obstacle was to isolate and understand the impact of porosity on flame enhancement. This is important because in a real mine explosion the flame can interact with piles of rock rubble that have varying porosities. Similar to the obstacle wall, the checkerboard obstacles, as described in Chapter 4, also resulted in flame acceleration across the obstacle, creating local fluid motion near 194
Colorado School of Mines
the checkerboard obstacle. To model this in 2D, researchers modeled the obstacle as nine (9) solid spheres with a diameter of 6.35mm in the vertical direction (this equates to the maximum amount of spheres in the y-direction as used in the experiments experiments). The checkerboard obstacle was modeled at three different locations, 37cm, 62cm, and 87cm from the open end. Results are shown in Figure 6.20 through Figure 6.28 and the general trends observed in the obstacle wall experiments were captured with the model: the flame passing over the checkerboard obstacle located 37cm from the open end reached the closed end of the reactor before those located at 62cm and 87cm. These flame propagation trends are similar to those observed with moving the non-reacting metal cage and obstacle wall at different locations in the reactor as shown in Figure 4.24, Section 4.2.3. As expected from the cage and checkerboard obstacles, the obstacles induce movement in the nearby gases resulting in slight upstream turbulence as shown in Figure 6.22. This is important to capture because it shows that discretely modeling the objects can improve local turbulence prediction and thus, more accurate flame propagation trends. This shall be explored further in the following section as well as modeling the porous medium in the experimental box. Figure 6.20 2D 12cm diameter reactor results of investigating the impact of checkerboard obstacle location on methane flame front location versus time for an open-end ignition. Obstacle: 6.35mm spheres in a checkerboard pattern (9 total in the y direction) located at 37, 62, and 87cm from the open end. Ignition location is 11cm from the open end. Time step = 0.1ms. CH = 9.5% 4 Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = ign 60mJ. No relief hole. 195
Colorado School of Mines
Figure 6.28 Temperature contours of the 2D 12cm diameter reactor model showing flame propagation across a checkerboard obstacle (77% porosity) for an open-end ignition. Obstacle: 6.35mm spheres in a checkerboard pattern (9 total in the y direction) located at 37, 62, and 87cm from the open end. Ignition location is 11cm from the open end. Simulation Time = 1.5s. Time step = 0.1ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 294K, 4 Pressure = 82kPa. E = 60mJ. No relief hole. ign 6.1.4 Modeling a Simulated Gob Bed Previous experiments by Fig (2019) have shown flame acceleration across a pile of rock rubble as described and shown in Figure 4.1 and Figure 4.2 on page 62 and page 63. However, the main purpose of this research is to understand what property of the rock pile has the largest impact on methane flame acceleration. As previously shown, the obstacle location, void spacing, and porosity can all impact methane flame acceleration. Also, closed-end ignition experiments with a simulated gob bed have been presented in Chapter 4 and results in Figure 4.28, Figure 4.29, and Table 4.1 (pages 82-84) show that the flame accelerates across the simulated gob bed made of 1cm diameter spheres, creating a turbulent boundary layer above and slow burning down into the gob. An example of the simulated gob bed used in experiments is reproduced in 203
Colorado School of Mines
Figure 6.29. To replicate these experiments in the 2D model, a comparison was made modeling the simulated gob bed as solid spheres versus a solid rectangle, H = 2cm, L = 30cm located 44cm from the open end as shown in Figure 6.30. In the experiments, the spheres were aligned next to each other, but in the model, the spheres were spaced 0.2cm apart to allow for at least 2 cells between each sphere to help model the flame propagation down into the simulated gob bed. Figure 6.29 Images of glass spheres and example of simulated gob bed. 2D modeling results are shown in Figure 6.30 through Figure 6.34. Modeling results accurately capture the flame acceleration across the obstacle, but overestimates the maximum flame front propagation velocity by 25%. Some of this overprediction is due to the fact that in the experiments, the flame burns above the simulated gob bed as well as down into the gob bed. Because the model is in 2D, the void spacing between the spheres (circles) is not truly indicative of the void spaces in the experimental simulated gob and thus, the flame does not burn down into the gob as quickly as the experiments. Modeling the simulated gob bed as a rectangle gob gives a good approximation of the acceleration, but produces much higher turbulence in the turbulent boundary layer above the obstacle than the modeled sphere gob (Figure 6.34). Overall, 2D modeling of flame propagation past a simulated gob bed matches experimental trends, captures the local fluid motion within and above the gob, and burning down into the modeled gob. It is important that the model captures the movement of nearby gases and burning down into the gob because the additional burning within the gob increases local temperatures and pressures, which feeds back to the main flame brush, accelerating combustion rates. This flame acceleration phenomena has been observed by other researchers who found that the properties of the “bead layer” can impact the main flame brush (Babkin, Korzhavin, & Bunev, 1991) (Howell, Hall, & Ellzey, 1996). These results also indicate that modeling a simulated gob will require discretely modeling of the obstacles, which shall be revisited in later sections.. 204
Colorado School of Mines
Figure 6.30 2D 12cm diameter reactor results of investigating the impact of an obstacle on methane flame front location versus time for a closed-end ignition. Obstacle: a solid rectangle H=2cm L=30cm located 44cm from the open end, and a 1cm diameter sphere gob H = 2 spheres L=30cm located 44cm from the open end. Ignition location is 1.39cm from the open end. Time step = 0.01ms. CH = 9.5% Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 293K, 4 Pressure = 82kPa. SM E = 60mJ. Relief hold H=1.2cm. ign Figure 6.31 Temperature contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle for a closed-end ignition. Obstacle: a solid rectangle H=2cm L=30cm located 44cm from the open end, and a 1cm diameter sphere gob H = 2 spheres L=30cm located 44cm from the open end. Ignition location is 1.39cm from the open end. Simulation Time = 10ms. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. No relief hole. ign 205
Colorado School of Mines
Figure 6.32 Temperature contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle for a closed-end ignition. Obstacle: a solid rectangle H=2cm L=30cm located 44cm from the open end, and a 1cm diameter sphere gob H = 2 spheres L=30cm located 44cm from the open end. Ignition location is 1.39cm from the open end. Simulation Time = 16ms. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. No relief hole. ign Figure 6.33 Temperature contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle for a closed-end ignition. Obstacle: a solid rectangle H=2cm L=30cm located 44cm from the open end, and a 1cm diameter sphere gob H = 2 spheres L=30cm located 44cm from the open end. Ignition location is 1.39cm from the open end. Simulation Time = 24ms. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. No relief hole. ign 206
Colorado School of Mines
Figure 6.34 Turbulent intensity contours of the 2D 12cm diameter reactor model showing flame propagation across an obstacle for a closed-end ignition. Obstacle: a solid rectangle H=2cm L=30cm located 44cm from the open end, and a 1cm diameter sphere gob H = 2 spheres L=30cm located 44cm from the open end. Ignition location is 1.39cm from the open end. Simulation Time = 16ms. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. No relief hole. ign 6.1.5 In-gob Modeling The main purpose of the in-gob ignition experiments presented in Section 4.3 was to understand how a flame propagates between obstacles. This is important for longwall coal mining because EGZs can exist within the gob and have the potential to be ignited by rock-on- rock friction (Brune, 2014). In order to model the in-gob ignition experiments shown in Figure 4.39, the checkerboard geometries were modeled as spheres, granite rock, and hexagons centered on a vertical line 7.62cm on either side of the spark location 25cm from the open end. The spheres and hexagons were modeled with a hydraulic diameter of 6.35mm and a no slip, adiabatic boundary condition on the surface. The granite pebbles were modeled as irregular shapes as shown in Figure 6.35 with no additional surface roughness and a no slip boundary condition. The diameter of the circle enclosed inside the rock is 4.6mm and the average length, L, of the irregularities is 1.7 ± 0.3mm, bringing the size of the irregular granite rock to within the size of the spheres and hexagons. Also to note, these simulations were run using the EM model and were initialized assuming zero-flow laminar flame expansion, k = 0.001 m2/s2 and ω = 0.001 1/s. These same simulations were run using the SM, but because the SM overpredicts the kernel expansion, the resulting flame shapes were unrealistic. Future work must be done to rerun these simulations with the SM and appropriate turbulence initialization parameters. 207
Colorado School of Mines
Figure 6.35 Image of irregularly shaped granite rock (left) and glass spheres (right) used for comparison of in-gob ignition models. Sphere diameter = 6.35mm. Thermal properties of glass were used for direct comparison between all cases. Note: not to scale. Figure 6.36 Temperature contours of the 2D 12cm diameter reactor model showing flame propagation during an in-gob ignition. Obstacle: spheres with diameter 6.35mm, rock with an average diameter of 4.6mm, and hexagons with diameter 6.35mm. Obstacles located 15cm apart on center with ignition. Ignition location is 25cm from the open end. Simulation Time = 0.11s. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 4 294K, Pressure = 82kPa. EM E = 5mJ. No relief hole. ign Results of modeling the in-gob ignition experiments in 2D are shown in Figure 6.36, Figure 6.37, and Figure 6.38. As can be seen in these figures, at the same time, the rock and hexagons produced significantly more turbulence in the local gases than the spheres. Although this only shows a single time step, this was confirmed at several time steps as the flame was passing through the obstacles. Also interestingly, the rock and hexagons resulted in significantly more CO production than the spheres which indicates there was incomplete combustion. It is also likely the increased CO production is due to the 2-step chemistry mechanism. To investigate this further would require a more complex chemistry mechanism capable of predicting the intermediate species and radicals produced by the initial reaction of methane and air. This is not within the scope of this research, but even these preliminary results show that modeling the gob 208
Colorado School of Mines
as different shapes can impact methane flame propagation which will be important when modeling a large-scale ignition from within the gob. Figure 6.37 Turbulent intensity contours of the 2D 12cm diameter reactor model showing flame propagation during an in-gob ignition. Obstacle: spheres with diameter 6.35mm, rock with an average diameter of 4.6mm, and hexagons with diameter 6.35mm. Obstacles located 15cm apart on center with ignition. Ignition location is 25cm from the open end. Simulation Time = 0.11s. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 4 294K, Pressure = 82kPa. EM E = 5mJ. No relief hole. ign Figure 6.38 CO mass fraction contours of the 2D 12cm diameter reactor model showing flame propagation during an in-gob ignition. Obstacle: spheres with diameter 6.35mm, rock with an average diameter of 4.6mm, and hexagons with diameter 6.35mm. Obstacles located 15cm apart on center with ignition. Ignition location is 25cm from the open end. Simulation Time = 0.11s. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.25mm edge sizing. Temperature = 4 294K, Pressure = 82kPa. EM E = 5mJ. No relief hole. ign 209
Colorado School of Mines
6.1.6 Ignition Location Modeling It is often thought that a faster flame means greater pressure generation, but the ignition location experiments presented in Section 4.4 helped show that this is not always the case. For example, an ignition in the center of the reactor resulted in greater pressure generation than a closed-end ignition, but slower flame front propagation velocities. This is important to capture in the CFD model because methane gas explosions can occur in a variety of places in a mine with varying degrees of confinement. Therefore, 2D models were run exploring the impact of ignition location on methane flame front propagation. To directly compare the experimental results shown in Figure 4.43, Section 4.4 to CFD results, the maximum flame front propagation velocity towards the closed end of the reactor of each experiment (i.e. ignition in Port 1, ignition in Port 2, ignition in Port 3) was normalized to the maximum flame front propagation velocity of ignition in Port 1. For example, the maximum flame front propagation velocity towards the closed end in Port 1 was 148cm/s, thus 148cm/s divided by 148cm/s is 1. For ignition in Port 2, the maximum flame front propagation velocity towards the closed end was 416cm/s; dividing 416cm/s by 148cm/s is 2.8. And so-on for Port 3. The normalized experimental data was then compared to the normalized CFD data, summarized in Table 6.3. The normalized data further demonstrates that as ignition is moved further from the open end, the flame front velocity increases. The differences between the experimental and numerical data from ignition in Port 2 and Port 3 can be attributed to the fact that acoustic effects have not yet been considered in the model since the flame acceleration in Port 2 and 3 was mainly due to acoustic-flame interactions. Since the acoustic model has not been implemented, there are qualitative differences in flame shape and oscillation as shown in Figure 6.39. As can be seen in this figure, in the experiments the flame tends to move forward and then be pushed backwards (t=0.07s to t=0.08s) versus the model predicting the flame continues towards the closed end. Overall, the model predicts faster flames propagations towards the open end and slower towards the closed end, but will require an acoustic model to capture the impact of pressure oscillations on flame front stretching observed in experiments. 210
Colorado School of Mines
Figure 6.39 Comparison of numerical to experimental methane flame front propagation of ignition in port 3 (75 cm from the open end). CH = 9.5%. Operating conditions 294K, 82kPa. 4 EM E = 5mJ. One (1) relief hole. ign Table 6.3 Comparison of normalized experimental flame front propagation velocities at the closed end of the quartz reactor compared to maximum flame front propagation velocities predicted by the 2D model. CH = 9.5%. Body mesh size = 0.001m, 0.25mm edge sizing. 4 Temperature = 293K, Pressure = 82kPa. Normalized Flame Front Propagation Velocity Port 2D Model Experiments 1 1 1.0 2 2.1 2.8 3 2.8 3.6 6.2 3D 12cm Diameter Quartz Reactor Modeling Results Although the 2D 12cm diameter reactor model successfully captured methane flame propagation trends, maximum speeds, and interaction with obstacles, turbulence is inherently a 3D process. Therefore, this research also developed a 3D model of the 12cm diameter. As discussed in Chapter 5, a 2mm body mesh for the 3D 12cm diameter reactor is sufficient in predicting the flame front location and capturing the local flow velocities as well. The 3D, 12cm diameter reactor model uses similar settings to the 2D model, summarized in Section 5.1, except for the following changes: • 2 levels of mesh adaption on the gradient of temperature every 2 time steps • Residuals set to 10-4, dropping at least 3 orders or magnitude • Closed-end ignition: k = 1.5m2/s2 and ω = 250 1/s 211
Colorado School of Mines
The main difference between the 2D and 3D model is that the 3D model does not include modeling the relief hole on the closed end of the reactor. A comparison of the 2D model with and without a relief hole versus the 3D model without a relief hole is shown in Figure 5.47, Section 5.1. The relief hole was not modeled in 3D because modeling the relief hole resulted in reverse flow in this region due to the pressure oscillations creating low pressure zones. To begin comparing experimental trends with 3D modeling results, a model was developed to capture the observed phenomena of a flame passing over an obstacle wall. Experimental results show that an obstacle wall in the path of the flame will tend to increase the velocity of the flame. Upstream of the obstacle the flame may be slightly retarded by the pressure resistance felt by the obstacle. Downstream of the obstacle, the flame is accelerated by a combination of the reduced void spacing and a large eddy of unburned gas trapped downstream of the obstacle that feeds the flame. Figure 6.40 3D 12cm diameter reactor results of investigating the impact of an obstacle on methane flame front location versus time for a closed-end ignition. Obstacle: a wall H=6cm L=2cm located 1.13m from the open end and a wall H=6cm L=2cm located 37cm from the open end. Ignition location is 11cm from the closed end. Time step = 0.01ms. CH = 9.5% Body mesh 4 size = 2mm. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. No relief hole. ign To model this in 3D, a solid obstacle wall with height H = 6cm and L = 2cm was located 37cm from the open end and then moved to 1.13m from the open end. Flame front propagation results as a function of time are presented in Figure 6.40 and results indicate that an obstacle closer to ignition source (0.37m) accelerates the flame over the entire length of the reactor. 212
Colorado School of Mines
Results also show that when the obstacle wall is further from the ignition source (1.13m), the pressure resistance from the obstacle slightly retards the flame, but once the flame passes over the obstacle it is accelerated. 2D model results are presented in Figure 6.41, Figure 6.42, Figure 6.43, Figure 6.44, and Figure 6.45 and predicts the flame shape and combustion acceleration mechanisms well, referring to flow over a step and the fluid motion that is formed behind the obstacle. This is important for model development as researchers continue to validate larger models in the transition to a mine-scale model which will be discussed in Chapter 7. Additionally these figures show that as time continues, the flame accelerates past the obstacle similar to flow over a step as shown in Figure 6.46. This is important because it forms eddies on the downstream side of the wall, which increases temperatures and fluid motion promoting flame acceleration. Figure 6.41 Temperature contours of the 2D 12cm diameter reactor model investigating the impact of an obstacle on methane flame front location versus time for a closed-end ignition. Obstacle: a wall H=6cm L=2cm located 1.13m from the open end and a wall H=6cm L=2cm located 37cm from the open end. Ignition location is 11cm from the closed end. Simulation Time = 6ms. Time step = 0.01ms. CH = 9.5%. Body mesh size = 2mm. Temperature = 293K, 4 Pressure = 82kPa. E = 60mJ. No relief hole. ign 213
Colorado School of Mines
6.3 Experimental Box Modeling Results As discussed in Chapter 3, Section 3.1.2, the main goal of the experimental box was to understand the impact of reactor shape on methane flame propagation and interaction with a porous medium. Experimental results showed that, similar to the cylindrical reactors, ignitions from a more confined space results in much faster flame velocities than an unconfined space. Additionally, images from the flame propagation in the box reactor showed residual burning of methane-air mixture in the corners of the box (Section 4.6). Finally, experimental results with a porous medium showed flame acceleration across the porous medium and the flame tended to propagate faster through the porous medium than in the open spaces inside the reactor. These results are important to capture in the 2D and 3D box models because in a real mine explosion, the propagation direction will be unknown and so developing a validated, robust model will be important for understanding flame propagation at the mine-scale. 6.3.1 2D Modeling Results As discussed in Chapter 5, in 2D, the experimental box is modeled with a 7cm relief and a mesh size of 1mm with 0.25mm edge sizing on walls and obstacles and uses the following turbulence parameter settings: open-end ignition k = 0.001m2/s2 and ω = 0.1 1/s, closed-end ignition k = 1.5m2/s2 and ω = 25 1/s. In order to accurately model flame propagation across a porous medium or gob, it is first important to understand how previous researchers have modeled the gob. Many of the researchers using ANSYS Fluent to perform ventilation studies on the air and methane distribution in a longwall coal mine typically model the gob as a Darcy flow porous medium with varying permeabilities and resistances to account for the different levels of gob compaction (Gilmore, et al., 2016; Ren & Edwards, 2000; Tanguturi, Balusu, & Bongani, 2017; Yuan, Smith, & Brune, 2000). However, modeling the gob as a porous media with laminar flow assumes Darcy flow where there is a linear relationship between flow through the porous media and pressure drop (Whitaker, 1986). Evidence has shown that EGZs are likely to exist near the gob fringes where the porosity can be almost 40% (Gilmore, et al., 2016; Marts, et al., 2014) and assuming a laminar Darcy flow may not account for the differences in void space and rock size/distribution in these areas. As this research has previously shown, the void spacings, porosity, and location of voids can have a significant impact on the resulting methane gas deflagration. This research has also shown that modeling a simulated gob bed requires discrete 218
Colorado School of Mines
modeling of the spheres instead of a solid rectangle (Section 6.1.4). Therefore, a study was performed to investigate whether the Darcy flow assumption is valid and if so, under what conditions. To do this, the 2D experimental box model was first validated for ignition without a porous medium and then the model was used to investigate how to best represent the porous medium. Results of an open-end ignition without a simulated gob shows that the flame propagates freely in all directions in the experimental box (Figure 6.47). 2-D combustion model results agree with this trend, but at later times slightly overpredicts the methane flame speed. This is to be expected since the model is in 2-D and assumes a 2-step methane-air mechanism. Also, as previously mentioned these slight differences may also be attributed to the spark kernel diameter and/or the turbulence initialization parameters. However, in general, the model is able to accurately predict flame trends and even though it is still in 2D, it is capable of predicting flame front wrinkling observed in experiments. Figure 6.47 Temperature contours of the experimental box setup compared to the 2-D combustion model for an open ignition with no gob. CH = 9.5%. Body mesh size = 1mm. 4 Temperature = 294K, Pressure = 82kPa. SM E = 60mJ. ign 219
Colorado School of Mines
Figure 6.48 Temperature contours of the box setup comparing modeling the gob as a Darcy flow porous media with 66% porosity (left), porous media with opening 66% porosity (center), versus modeling the gob as discrete spheres with 66% porosity (right). Center ignition. Simulation time = 15ms. Time step = 0.01ms. CH = 9.5%. Body mesh size = 1mm, 0.5mm edge sizing. 4 Temperature = 294K, Pressure = 82kPa. E = 60mJ. ign After demonstrating model performance without a gob, Figure 6.48 shows the comparison of modeling the gob as a Darcy flow porous media with 66% porosity (left), with a small opening and 66% porosity (center), to modeling the gob as discrete spheres (D=0.025m) with a porosity of 66% (right). As can be seen, ignition within the Darcy porous media with a porosity of 66% did not allow the methane flame to propagate even when a small opening is made to help allow for flame expansion (center image). Additionally, porosities between 25-90% have been tested, permeabilities between 1.5x10-3m2 and 4.7x10-11m2 have been tested, and spark energies between 60mJ and 1kJ have been tested and also show no flame propagation. However, when the porous media has a 100% porosity, meaning all fluid space, the flame is able to propagate. This condition, however, is unrealistic for a longwall coal mine environment where the porosities typically range between 14-40%. Also to note, the flame propagation through the 100% porous media does not take into account any void spacing and so the flame front is very smooth compared to modeling the gob discretely as shown on the right of Figure 6.48. This is important because flame front stretching can increase combustion rates thereby increasing flame speed and overpressure. Results of a closed-end ignition show that the flame travels much faster through the simulated gob than in the open spaces (Figure 6.49). This is because the pressure waves from the confined ignition are enough to disturb the upstream unburned gases in the gob area. This 220
Colorado School of Mines
increased turbulence enhances mixing and transport of unburned mixture to the flame front, as well as increases the flame surface area; thereby increasing combustion rates. The combustion model captures this well, but overpredicts the speed of the flame as seen in the sequence of images in Figure 6.49. This overprediction is mainly due to the fact that the model is still in 2D and the initialization of turbulence results in an overprediction of flame speed since the model is confined to two dimensions. Despite these differences, the 2D model performs fairly well at capturing the general flame dynamics. Figure 6.49 Temperature contours of the experimental box setup compared to the 2-D combustion model for a closed ignition with a gob. Sphere diameter = 0.025m, porosity = 65.9%. Time step = 0.01ms.CH = 9.5%. Body mesh size = 1mm, 0.5mm edge sizing. Temperature = 4 294K, Pressure = 82kPa. SM E = 60mJ. ign After determining that to accurately model methane flame propagation through a gob requires modeling the rock discretely, a study was performed to determine the impact of discrete 221
Colorado School of Mines
object shape on methane flame propagation through a simulated gob. The shapes investigated are detailed in Table 6.4; circles, hexagons, and squares were compared and sizes were made to match the same porous media bed porosity of 66%. Interestingly though, matching the porosity resulted in similar sizes to the hydraulic diameter, which is typically used in fluid mechanics to scale objects geometrically. Table 6.4 Table summarizing the discrete objects compared in this study including the size/number of the shapes, the porosity, and the index of sphericity. *Calculated based on Ionescu-Tirgoviste et al. 2015 as the ratio of the perimeter to hydraulic diameter, divided by π (Ionescu-Tirgoviste, et al., 2015). Object Shape Circle Hexagon Square Diameter/Side Length (m) 0.025 0.014 0.022 Hydraulic Diameter (m) 0.025 0.024 0.022 Number of Objects 42 42 42 Bed Porosity (%) 66 66 66 Index of Sphericity* 1 1.1 1.2 Figure 6.50, Figure 6.51, Figure 6.52, and Table 6.5 show results comparing how the shape of the discrete object can affect methane flame propagation through the gob. In all cases, the porosity of the discrete shape porous medium was 66%. As can be seen in Table 6.5, in all cases, the average flame front propagation velocity with a discretely modeled gob is faster than a closed-end ignition with a gob, matching experimental observations. Additionally, the shape of the discrete object can significantly affect methane flame propagation; the squares produced the fastest flame and most tortuous flame path, followed by the hexagons and the circles (increased flame speeds with increasing index of sphericity). The squares also produced the highest turbulent intensity ahead of the flame and the most eddies. This is important because higher turbulent intensities mean there are larger fluctuations in the average flow, which can increase unburned gas velocities and enhance methane flame propagation. The eddies produced by the shapes is important because one major concern in longwall coal mining, are dead zones where methane and air mix and accumulate (i.e. an EGZ). As can be seen, the square gob produced much more eddies, which is due to the sharp corners of the shape disrupting the boundary layer resulting in vortex shedding. This is important because typical rock found down in coal mines may have sharp edges and sphericity not equal to 1. Thus far, results indicate that modeling the 222
Colorado School of Mines
Table 6.5 Table summarizing the average methane flame front propagation velocity from the box experiments compared to 2D modeling results with and without a simulated gob for a confined, closed-end ignition (ignition in the top-right corner of the experimental box). CH = 9.5%. 4 Temperature = 293K, Pressure = 82kPa. 2D Model Results Experiments Average Flame Simulated Gob Average Flame Simulated Gob Front Propagation (Porous Medium) Front Propagation Shape Velocity (m/s) Conditions Velocity (m/s) None 26 No Gob 7.7 ± 1 Sphere Gob 44 Gob 12 ± 4 Hexagon Gob 54 Square Gob 68 The main advantage of CFD modeling is the ability to gain a better understanding of methane flame interaction with the gob. Thus, in addition to investigating the average velocity of the methane flame front propagation through the discretely modeled gob, this research also investigated methane flame propagation in the open areas around the discretely modeled gob. To do this, horizontal and vertical seeded lines were taken as shown in Figure 6.53 and results of tracking the flame front location are presented in Figure 6.64 and Figure 6.65. As shown, in all cases, the discretely modeled gob enhanced flame propagation in both the vertical and horizontal directions around the gob. Complimentary to previous results, the squares had the fastest flame propagation in the open spaces, followed by the hexagons, and finally the circles. These results are important because they show that the gob can impact flame propagation in the open spaces around the gob which is important for longwall coal mining because it shows that an ignition near or around the gob may be impacted by the gob or nearby obstacles. This is also important for future modeling of the gob because it helps to demonstrate how the discrete object used to represent the gob area can impact flame propagation through and around the gob. 225
Colorado School of Mines
Figure 6.55 2D box model results for a closed-end ignition investigating the impact of gob shape on methane flame propagation in the ‘y’-vertical direction away from ignition. Simulated gob with 66% porosity modeled as discrete circles (dashed green line), discrete hexagons (dotted blue line), and discrete squares (dash-dot red line). Time step = 0.01ms. Ignition in the top-right. CH 4 = 9.5%. Body mesh size = 1mm, 0.5mm edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. ign In a typical laminar methane flame, the quenching distance of the flame can be as small as 1-2mm depending on the ambient conditions, however, a turbulent flame has a much smaller quenching distance. A concern in longwall coal mining is whether or not an ignition deep within the gob can propagate towards the longwall face. Thus, researchers used the 2D CFD, combustion model to investigate the smallest possible distance between objects that allowed for flame propagation. Researchers investigated separation distances of 10, 5, and 1mm as shown in Figure 6.56. It is important to note that the mesh in-between the obstacles had at least 2 cells and in many cases 3-4 cells. Results show that a more turbulent flame propagation from a confined space can propagate through much smaller void space than a typical laminar flame. This is important because it is typically assumed that in a highly compacted area the flame cannot travel from deep within the gob towards the face. However, results indicate that a highly turbulent flame can propagate through small cracks less than a millimeter in size. Although these results agree with turbulent flame theory, future work must be done to confirm these results are physically accurate which will require future modeling of more complex chemistry mechanisms as well as translation of the model to 3D. Also, the boundary conditions on the obstacles were adiabatic, assuming no heat loss to the obstacle. Additional future work includes investigating 227
Colorado School of Mines
the assumption of an adiabatic boundary condition because flame quenching can occur due to heat loss to a solid. Figure 6.56 Temperature contours of methane flame propagation for closed end ignition across a simulated gob consisting of squares with spacing of 10, 5, and 1mm. Simulation Time = 8.5ms. Time step = 0.01ms. Ignition in the top-right. CH = 9.5%. Body mesh size = 1mm, 0.5mm edge 4 sizing. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. ign In Figure 6.50, the methane flame had some room to propagate around the gob as well as through the gob. Although the information from these types of simulations are important for understanding flame propagation in and around the gob, it does not fully explain the entire impact of shape on methane flame propagation. Therefore, a study was performed where the entire box model was filled with discrete objects: circles, hexagons, and squares with the same dimensions as shown in Table 6.4. Results of this study are shown in Figure 6.57, Figure 6.58, Figure 6.59, and Figure 6.60. As expected, the square gob resulted in the fastest flame speeds and turbulent intensities compared to the circles and hexagons. Additionally, the square gob produces CO at a much faster rate than the hexagon and circle gob, but after the flame has exited the reactor the amount of CO decays to almost the same value for all cases. This is important because the model predicts a larger release of CO at a faster rate for the square gob, however this is most likely an artifact of the methane-air 2-step chemistry mechanism as shown in Table 5.1, Equation (5.1), and Equation (5.2) on page 109. Note that in these equations, the methane and oxygen react to form CO and then the second reaction is the reaction of CO with O to form CO . However, in a real situation, after methane 2 2 228
Colorado School of Mines
reacts with oxygen there would be many intermediate species such as OH and other radicals. The reaction of the OH radical with CO to form CO is a much faster process than the reaction of CO 2 and O to form CO . Although these results show there may be a trend of gob shape to CO 2 2 production, this must be reevaluated with more complex chemistry mechanisms that include more intermediate species such as OH. Future recommended work includes investigating different methane-air 2-step mechanisms as well as 13-step and full GRI methane air mechanisms and repeating these experiments/modeling. These results are important because they demonstrate that shape of the object can have a large impact on methane gas combustion. This is extremely important for building the coupled, mine-scale CFD combustion model because it shows that gob shape will be important in helping predict the severity of the explosion. These results may also help investigative teams better understand the heat damage or CO concentration in some of these large-scale explosions, leading to better explosion prediction. Figure 6.57 Temperature contours of methane flame propagation for closed end ignition across a simulated gob modeled as discrete circles (left), discrete hexagons (center), and discrete squares (right). Ignition in the top-right. Simulation Time = 4.5ms. Time step = 0.01ms.CH = 9.5%. 4 Body mesh size = 1mm, 0.5mm edge sizing. Temperature = 293K, Pressure = 82kPa. SM E = ign 60mJ. 229
Colorado School of Mines
Figure 6.60 Total moles of CO over time of methane flame propagation for a closed end ignition 2 across a simulated gob modeled as discrete circles (green dashed line), discrete hexagons (blue dotted line), and discrete squares (red dash-dot line). Ignition in the top-right of the box. Time step = 0.01ms.CH = 9.5%. Body mesh size = 1mm, 0.5mm edge sizing. Temperature = 293K, 4 Pressure = 82kPa. SM E = 60mJ. ign 6.3.2 3D Modeling Results The 3D, 12cm diameter reactor model uses the same ANSYS Fluent (v 17.2) as described in Section 5.1 except for the following settings: • Continuity/velocity residuals set to 10-4, Energy/turbulent 10-6 • 3 levels of mesh adaption on the gradient of temperature every 2 time steps The 3D model the continuity and velocity residuals are set to 10-4 for model stability, but the mesh adaption on the temperature gradient has been increased to 3 levels every 2 time steps. These changes have been noted to help convergence and model predicted flame front location on a coarse grid (5mm or 2.5mm) as discussed and shown in Chapter 5, Section 5.2.2 and Section 5.2.3. The first experiment that was modeled was an open-end ignition without a simulated gob as shown in Figure 4.53 on page 101. In the experiments, it was observed that the flame moved in all directions and preferentially away from the relief opening. This was modeled in 2D and compared to experiments in Figure 6.47 (page 219), concluding that the 2D model accurately captured the flame propagation at early stages, but overestimated the flame progression later in 231
Colorado School of Mines
development. Figure 6.61, Figure 6.63, Figure 6.64, Figure 6.65, and Figure 6.66 (pages 233- 236) show the 3D modeling results compared to the 2D model and experiments for an open-end ignition. 3D results show similar flame propagation trends compared to the 2D model; the 3D model predicts the flame propagation in all directions well, but compared to the 2D model the 3D flame propagating towards the closed end moves faster than towards the walls. In general, the 3D model captures the early stages of flame development well, from t = 0s to t = 0.15s, but at t = 0.20s the 3D flame tends to move faster than the 2D model and experiments. Also, after t = 0.20s, the 3D model does not show any significant flame instabilities along the flame front, which has been observed in experiments and captured in the 2D model. This is likely due to the fact that the 3D model uses gradient adaption on temperature to resolve the flame front location on a coarse, 2.5mm mesh; this method predicts the flame front location well, but as shown in Chapter 5, Table 5.14 and Table 5.15 on page 165, temperature and the flame front are resolved well, but the fluid velocities are not resolved as well. This is important because in Chapter 2 it was shown that the fluid velocities change through a wrinkled flame front (Figure 2.3 on page 11) and that hydrodynamic and thermo-diffusive instabilities are what can lead to wrinkled flame fronts. It was concluded that the lack of flame front wrinkling in the 3D model was due to the fact that the 3D model does not fully resolve the fluid velocities coupled with the simplified chemistry mechanism not perfectly resolving the diffusive effects of species. Overall, the 3D model does a good job of capturing the flame trends and flame shape despite the coarse mesh and simplified chemistry mechanism. Experiments of a closed-end ignition without a gob was shown in Figure 4.55 (page 102) and results indicated a significantly faster flame than the open-end ignition. Figure 6.67, Figure 6.68, Figure 6.69, Figure 6.70, and Figure 6.71 (pages 236-238) compare experimental images to 2D and 3D modeling results of a closed-end ignition without a gob. Results show that at early times, t = 2-6.3ms, the models predict the flame kernel expansion and propagation well. Due to the overestimation of initial flame kernel expansion by the spark model (discussed in Chapter 5, Section 5.1.2) the 2D and 3D model times are an order of magnitude different than the experiments, but the acceleration of the flame at early times is estimated very well by the models. Good estimation of the acceleration of a closed-end flame has also been observed in the 2D and 3D 12cm diameter reactor models, which also showed flame slow down near the reactor outlet. In the 3D box model, after 6.3ms, the flame front location is underpredicted compared to 232
Colorado School of Mines
experiments (Figure 6.70, page 238), which agrees with observations made in the 2D and 3D 12cm diameter reactor models. However, the shape of the flame in Figure 6.71 (page 238) at time t = 12ms matches that observed in experiments and because the flame is resolved in the 3rd dimension, researchers can see how the flame tends towards the relief which is near the bottom of the box. Overall, the 2D and 3D models of a closed-end ignition inside the experimental box capture general flame propagation shapes and trends as well as acceleration rates at early stages in the flame development. These results are important because as complexity, such as obstacles, is added to the models it is important that general flame trends and propagation velocities are captured in these simple cases. Also these results help show that as model volume increases, the ability to resolve all flow quantities reduces and certain flame shapes or trends may be lost when using more coarse meshes to reduce computational time. Figure 6.61 Temperature contours of the experimental box setup compared to the 2D and 3D combustion model for an open-end ignition with no gob. 3D isocontour at T=2200K. Simulation time = 0.05s. Time step = 0.1ms. CH = 9.5%. 2D Body mesh size = 1mm, 0.25mm edge sizing. 4 3D Body mesh size = 5mm. Temperature = 294K, Pressure = 82kPa. SM E = 60mJ. ign 233
Colorado School of Mines
Experiments of a confined methane gas explosion across a simulated gob was presented in Figure 4.56 (page 102) and showed significant flame speed enhancement. It was also shown that the flame moved through the gob significantly faster than around the gob in the open passages. This is important because an ignition in or around the gob area might tend to interact with the gob itself, other structures such as pillars, machinery, etc. enhancing combustion, leading to more violent explosions. Complimentary 2D modeling was presented and it was found that 1) the rock rubble must be modeled discretely to capture flame stretching and propagation through the porous gob and 2) results in Figure 6.50, Figure 6.51, and Figure 6.52 (pages 223- 224) showed that the shape used to represent the rock rubble does affect local turbulence and flame speed. 3D modeling of a confined ignition with rock rubble is presented in Figure 6.72, Figure 6.73, Figure 6.74, and Figure 6.75 on pages 239-240. Due to ease of meshing, the squares in the 3D model are 2x2x2cm instead of 2.2cm as shown in Table 6.4 on page 222. Despite this small difference, the 3D model results follow the trends of the 2D model and experiments well. The 3D model, similar to the 2D model, overestimates the actual time values of the flame expansion, but the acceleration of the flame at early times, t = 2.5-4.2ms, matches the acceleration of the flame. Compared to the 2D model, the 3D model better captures the impact of the third dimension on flame propagation through the gob. For example, in Figure 6.73, the experiments show the flame starting to propagate through the porous medium. The 2D model shows that some of the flame passing through the obstacles, but most of the flame is still in the open regions. Compared to the 2D model, the 3D model flame is able to expand and freely propagate in the vertical direction, showing significant flame propagation through the gob, matching experimental observations better. This can also be observed in Figure 6.75; the 2D model shows significant flame stretching versus the 3D model better predicting the bulk flame brush exiting the reactor. Overall, the 3D model of the experimental box shows good agreement with experiments. The 3D model better predicts the flame shape and flame front propagation trends for confined ignitions compared to the 2D model, but due to the coarser mesh of the 3D model, does not predict the local fluid velocities or instabilities observed in an unconfined ignition without a gob. These results are important because they demonstrate the need to always balance model accuracy and simulation time. Although a coarser 2.5mm cut cell body mesh was used with 3 levels of mesh adaption on the gradient of temperature every 2 times steps to resolve the flame front, other 241
Colorado School of Mines
flow quantities were not as well predicted resulting in certain flame inaccuracies. As the models become larger and continue to scale, it will become more important to keep in mind the loss of model accuracy to improve simulation times. However, this demonstrates why it is important to continue to validate models across a variety of scales and conditions to ensure a robust model is produced. 6.4 2D 71cm Steel Reactor Modeling Results The main purpose of the 71cm diameter reactor is to develop an understanding of methane flame dynamics at the large-scale for improved model development towards the mine- scale. Experimental results in the 71cm diameter steel reactor showed that a rock pile at the back of the reactor (closed end) resulted in faster flame front propagation velocities than a rock pile located at the front (open end) of the reactor and faster velocities than no rock pile at all (Section 4.7, Figure 4.58 and Figure 4.59 on pages 104-105). In the experimental setup, the length of the rock pile was L=1.8m and the height of the rock pile was 24cm. Previous results of modeling a rock pile in the 2D, 12cm diameter quartz reactor showed that the obstacle must be modeled as discrete objects (Section 6.1.4). Thus to model a rock pile in the 71cm diameter reactor, the rock pile was modeled as idealized circles with a diameter of 10cm, which was based on the average size of the rocks in the experiments. The discrete circles were spaced 2cm apart in the vertical direction so that the total height of the simulated gob was 24cm. Results of modeling a rock pile at the front and back of the 71cm are presented in Figure 6.76 and Figure 6.77 and flame shape and propagation trends are presented in Figure 6.78 through Figure 6.82. As can be seen in these figures, a rock pile at the back of the reactor results in faster flame front propagation velocities than a rock pile at the front or no rock pile at all. Compared to a rock pile at the front, a rock pile at the back of the reactor accelerates the flame along the entire length of the reactor. When the rock pile is at the front, it only tends to accelerate the flame across the obstacle. Also to note, the reason that the 2D model does not exactly match the measured velocities of flame acceleration across a rock pile is because 1) the model uses a coarse mesh with errors up to 15% as discussed in Section 5.3, 2) the discrete circles used to represent the rock rubble does not fully capture the size/shape of the rock, 3) the spacing used between the discrete circles is not representative of the actual void spaces in a pile of rock rubble, and 4) the model is in 2D which means that turbulence is not fully resolved. Despite 242
Colorado School of Mines
these differences, the 2D model does a good job at capturing the overall trends of methane flame acceleration across a rock pile. Figure 6.76 2D, 71cm diameter reactor model results for a closed-end ignition with a sphere gob investigating the impact of gob location on methane flame front propagation velocity versus distance. Obstacle: 10cm diameter sphere gob, H = 24cm, L = 1.8m. Time step = 0.01ms. CH = 4 9.5%. 2D Body mesh size = 5mm, 1mm edge sizing. Temperature = 295K, Pressure = 76kPa. SM E = 60mJ. ign Figure 6.77 2D, 71cm diameter reactor model results for a closed-end ignition with a sphere gob investigating the impact of gob location on methane flame front location versus time. Obstacle: 10cm diameter sphere gob, H = 24cm, L = 2m. Time step = 0.01ms. CH = 9.5%. 2D Body mesh 4 size = 5mm, 1mm edge sizing. Temperature = 295K, Pressure = 76kPa. SM E = 60mJ. ign Finally, although these researchers did not experiment with a rock pile in the middle of the reactor (Fig, Strebinger, Bogin, & Brune, 2018; Fig M. , 2019), this condition is of 243
Colorado School of Mines
Figure 7.1 Diagram showing the multiple pathways the ventilation model and combustion model have been combined to simulate a 3D Sub-Section, a 3D Full-Scale, and a 2D methane gas explosion. As shown in Figure 7.1, three different full-scale models have been developed: 1) a 3D full-scale sub-section model, 2) a 3D, full-scale mine model, and 3) a 2D mine model. The 3D, sub-section models were developed to help balance computational time versus accuracy. As the models have increased in size, it has become more and more difficult to determine mesh independence, which has forced researchers to use mesh adaption on coarse grids as described in Section 5.2.3 and Section 5.3. The sub-section models were built to model different methane gas ignition scenarios as shown in Figure 7.2; the scenarios shall be described in subsequent sections. The main idea is that once the flame propagates in this small domain, the information can be translated over to the next domain, such that, researchers can track flame and pressure propagation throughout the entire mine. In general, the sub-section models help prove the viability of using detailed modeling to simulate a methane-gas explosion. Additionally, the sub- 250
Colorado School of Mines
section models are run on a conservative number of cores, 8-12, on a single compute, demonstrating that these models can run on a typical desktop computer. Figure 7.2 Schematic taken from the 3D full-scale ventilation model showing dimensions, flow directions and velocities, and sub-sections used for methane gas explosion modeling. In addition to developing 3D, sub-section models, a 3D, full-scale simulation was also performed. The main purpose of the full-scale simulation is to model a full-scale methane-gas explosion in an underground coal mine, which has never been done to the knowledge of these researchers. However, one of the drawbacks of this model is that it requires 96 cores over 4 compute on a supercomputer. Details of this model shall be given and results will be discussed in Section 7.4. Finally, complimentary to these 3D models, a 2D model has been developed using inlet, outlet, dimensions, and gob conditions from the full-scale, 3D ventilation model. The main purpose of the 2D model is to show the capabilities of a reduced order model to predict methane flame and pressure propagation in a longwall coal mine. Additionally, this reduced order model can be run on a typical desktop and takes less than 1 week to solve, making it a good predictive tool for initial flame and pressure wave propagation. 251
Colorado School of Mines
7.1 3D Full-Scale Sub-Section 1 After validating the 2D and 3D reactor CFD combustion models under a variety of conditions, researchers have begun modeling a small methane gas ignition under a single longwall coal mine shield as shown in Figure 7.3. This model is taken as a subsection of the full- scale longwall ventilation model; the height of the coal face is 3m and the length of the shield is 6m. In the previously developed 3D models presented in this research, mesh independence was typically found at a mesh size of 2.5-2mm for the smaller, laboratory-scale reactors. Unfortunately meshing this large volume would require millions of cells and researchers have found modeling the 3D 12cm diameter reactor with 1 million cells takes a little less than 2 weeks on a supercomputing node with 8-12 cores, 2.7-3.02GHz, 24-64GB RAM. However, it was also found that using aggressing mesh adaption can help in using larger cell sizes while still predicting physically accurate methane flame behavior (Section 5.2.3). Therefore, the single shield model was meshed with approximately 500,000 tetrahedral elements, size 7-10cm. To accurately model the deflagration physics, researchers are using the following settings in ANSYS Fluent (v17.2): • Pressure-Based Solver • Energy Equation • Viscous Standard k-ω Turbulence Model o Low Re Corrections o Shear Flow Corrections • Species Transport o Volumetric Reactions o Stiff Chemistry Solver o Finite Rate Chemistry ▪ Density solved using ideal gas theory ▪ Diffusion solved using kinetic theory ▪ Metghalchi and Keck laminar flame speed theory • Spark Ignition Model o Initial kernel radius = 5mm o Duration = 1ms o Energy = 60mJ 252
Colorado School of Mines
Results of ignition underneath the single shield are shown in Figure 7.4, Figure 7.5, Figure 7.6, and Figure 7.7 and took approximately 15 days to model flame expansion to the model boundaries (note that parallelization across computes can help speed-up this simulation time). As can be seen, the flame extends towards the path of least resistance, along the longwall face in the x directions. Calculations of the average flame speed predicts a flame expanding at 30m/s. According to the Upper Big Branch explosion in 2010, investigators estimated the flame speeds directly near the explosion to be close to 90m/s (Page, et al., 2011). What is positive about these results is that the flame expansion is captured well and the approximate speed of expansion is the same order of magnitude as the UBB explosion. As previously discussed, two main reasons for the discrepancy in speeds is the large cells required to run this simulation and the fact that the chemistry model assumes a 2-step methane-air mechanism. In general, 2-step mechanism underestimates the flame speed as compared to more complex chemistry mechanisms (13-step, full-GRI) as shown by Fig (2019). Additionally, the velocities and pressures from the UBB were estimated based off investigative evidence, which means the error bars on those values are not well known. Therefore, what this sub-section model shows is that these full-scale methane gas explosions can be modeled in a reasonable amount of simulation time on a typical desktop computer. Using ANSYS Fluent also allows data from this model to be interpolated onto another model, or simply transferred to the next shield along the longwall face (Figure 7.3). 7.2 3D Full-Scale Sub-Section 2 After running a preliminary simulation of ignition under a single shield, a second model was developed which includes a discrete gob behind the shields as shown in Figure 7.9. 3D hexagons were created to represent the gob and are 30cm in diameter. Some of the benefits of modeling the gob as discrete hexagons is that 1) the hexagons capture the turbulence induced by rock observed in experiments, 2) because they have flat faces they are easy to mesh, and 3) they are easily arranged in different packing orientations to change gob porosity or resistance. The mesh for this study is the same as the single shield model, 7-10cm tetrahedral cells, approximately 1 million cells total. Model settings are the same as the sub-section 1 model except for the following: • Residuals set to 10-3, dropping at least 3 orders or magnitude • Turbulence parameters: k = 1.5m2/s2 and ω = 25 1/s 256
Colorado School of Mines
Figure 7.12 Isocontour of temperature at 2200K showing methane flame propagation behind the longwall face within a discrete hexagonal gob. Ignition in the center of the gob. Simulation Time = 8ms. Time step = 0.01ms.CH = 9.5%. Body mesh size = 7-10cm with 2 levels of mesh 4 adaption every 2 time steps. Temperature = 293K, Pressure = 82kPa. SM E = 60mJ. ign Results are shown in Figure 7.9 through Figure 7.12 and show the flame expanding at an average velocity of 22m/s. Although these simulations are on-going, they demonstrate that modeling these large-scale explosions is possible and requires a complete understanding of model settings in addition to multiple points of model validation during development. In the future, researchers are urged to investigate ways to obtain mesh independence, while also maintaining reasonable simulations times, most likely by parallelization which is used in the full- scale simulation presented in Section 7.4. Future work also includes testing these models using an LES turbulence model in lieu of the k-ω turbulence model since the two-equation moment models require an estimation of the initial turbulence. Since the initial turbulence is unknown and the scales of these models are significantly larger than previously modeled, the LES turbulence model may be more appropriate for higher fidelity simulations. 7.3 3D Full-Scale Sub-Section 3 The previous sub-section models described in Section 7.1 and Section 7.2 modeled a methane gas explosion under stagnant conditions under a single shield along the longwall face and in the gob behind the shields. These models helped demonstrated that the sub-section models can be used to simulate methane-gas ignitions and solved on a typical-sized desktop computer. 259
Colorado School of Mines
However, these models do not include the high velocity airflow typical along the longwall face (5-7m/s) which can change the local initial turbulence and thus, solution. Therefore, a 3rd sub- section model was created to more realistically model a methane gas explosion. For example, in the UBB explosion, the shearer was at the tailgate and hot streaks left from worn shearer bits ignited a pocket of methane at the tailgate (Page, et al., 2011). Thus, the sub-section modeled is at the tailgate corner of the longwall face as shown in Figure 7.2 and Figure 7.13. The settings for this model were the same as the sub-section 2 model (Section 7.2) except for the model initialization. The model was no longer initialized to stagnant conditions; instead, the 3D full-scale ventilation model was run to steady state and the flow and pressure data was extracted. This data was then interpolated onto the sub-section model and the pressure profiles from the steady state model were used as boundary conditions for the pressure inlet and outlets. In the ventilation model used for this condition, there was not any methane at the tailgate corner, so to simulate an EGZ a small box (0.5m cube) was filled with stoichiometric methane and was ignited inside the box (volume of methane = 0.125m3). An example of the pressure profiles and flow conditions are shown in Figure 7.13 and Figure 7.14. Figure 7.13 Schematic of the 3D, Sub-Section 3 model showing the geometry used to simulate a methane gas explosion at the tailgate corner. Pressure profiles were extracted from the 3D, ventilation model and used as pressure boundary conditions as shown. Ignition is initiated in a small methane box (0.5m cube) was initialized at 9.5% methane by volume. 260
Colorado School of Mines
Figure 7.20 3D sub-section 3 results of the first pressure wave expansion location as a function of time. Time step = 0.01ms.CH = 9.5%. Body mesh size = 7-10cm with 2 levels of mesh 4 adaption every 2 time steps. SM E = 60mJ. ign Figure 7.21 3D sub-section 3 results of the maximum overpressure of the first pressure wave as a function of time. Time step = 0.01ms.CH = 9.5%. Body mesh size = 7-10cm with 2 levels of 4 mesh adaption every 2 time steps. SM E = 60mJ. ign Results from this study are presented in Figure 7.15 through Figure 7.21. Results show that the pressure wave expands towards the boundaries of the domain much faster than the flame itself. Calculations of the expansions reveal that the flame front expands at approximately 22m/s and the pressure wave expands at 350m/s. The overpressure of the first expanding pressure wave is plotted as a function of time in Figure 7.21 and results show an initial pressure wave of approximately 18kPa at 2ms. However, it is important to note that there may have been higher 264
Colorado School of Mines
pressures generated between spark initiation and 2ms. Therefore, this research recommends future work saving time steps more frequently during the initial kernel expansion in order show the initial pressure rise likely missed here. Additionally, results show that the expanding pressure wave disturbs the flow in nearby areas, such that by 30ms there is no flow from the longwall face to the tailgate entries. This can also be seen in Figure 7.19 which shows that the model is able to predict the pressure wave reflection off the mine walls and pressure wave interaction. From the UBB explosion it was estimated that near the ignition the flame front propagation velocity was near 90m/s and the overpressure was 27kPa (Page, et al., 2011). Although this initial simulation predicts overpressure slightly less than 27kPa, they are still the same order of magnitude and again, the actual peak overpressure may have been between 0-2ms. More importantly these results show that even for these small overpressures, they are enough to change the airflow patterns in this area. These results are important because they demonstrate the viability of modeling methane gas explosions using data interpolated from a steady state ventilation model. These results also show how the flow in the tailgate can be disturbed from the explosion overpressure, which can be important for understanding movement of EGZs or even entrainment of coal dust (which would require a multiphase model). Additionally, models such as this can be used to help estimate the amount of impact force on nearby mine structure, which may help in the design of pillars or seals. In general, all of the sub-section models have demonstrated the potential for modeling these explosions on smaller domains. They have shown reasonable results under totally stagnant conditions and have proven that steady state results can be used to initialize and model a methane gas explosion. They have also demonstrated the need to continue improving the models by investigating the use of an LES model, more complex chemistry mechanisms, sensitivity analysis on turbulence parameters, parallelization, and coupling with the ANSYS Mechanical to understand the stresses on nearby mine equipment/structure. Altogether, these results have enormous potential for accurately predicting methane gas explosions in a full-scale, sub-section model. Data extracted from these models could even be used to help further track the flame and pressure waves in the 2D mine model, which will be discussed in Section 7.5. 7.4 3D Full-Scale Simulations Researchers have also simulated an ignition and subsequent flame propagation in the full- scale ventilation model. The ventilation model contains the full longwall face (300m long), 6m 265
Colorado School of Mines
of gob behind the shields (152 shields), and part of tailgate and tailgate bleeder entries. The height of the coal seam is 3m and the distance from the coal face to the back of the shields is 6.5m. Pressure profiles obtained from the steady state ventilation model are used as boundary conditions for the inlets and outlets, to maintain model accuracy. For the ventilation scenario, it is assumed that the flow at the tailgate corner is directed toward the open crosscut outby the face due to the tailgate entry inby the face blocked by the roof fall. The shearer is cutting the tailgate corner, with half of the tailgate drum exposed at the tailgate entry. Both shearer drums are 1.8m in diameter and are rotating at 30 rpm (shearer cowls are also modeled). For ventilation, 85,000 cfm of fresh air enters the longwall face at the headgate. The majority of the air leaks into the gob, resulting in 30,000 cfm remains inside the face by the time it reaches the tailgate side. Figure 7.22 shows the ventilation condition used for this test, while Figure 7.23 shows volume rendering of methane mass fraction around shearer drums. The rendering is limited to methane around shearer drums for visual purposes. Figure 7.22 Volume rendering of velocity inside longwall face from plan view (top) and velocity contour plot showing close-up view of flow around shearer drums (bottom). Blue arrows indicate flow direction. 266
Colorado School of Mines
Figure 7.23 Volume rendering of methane mass fraction near the shearer drums. This scenario represents the case when there is insufficient fresh air to dilute the methane inflow from the coal face, resulting in the formation of explosive gas zones of methane and air around the shearer drums. From the figure, methane accumulation can be observed in the small gap between the shearer body and uncut coal face. There is also notable methane accumulation in areas between the headgate drum and coal face, and between the tailgate drum and cowl. For this test, it is assumed that the ignition occurred when the headgate drum is cutting the coal face, as shown in Figure 7.23. This ignition location is chosen to test the viability of initiating combustion in a region of high turbulence. Before ignition, the shearer drums are rotating until the flow is fully developed. After the onset of ignition, the drums are switched to stationary and considering the time scale of the explosion, the continuous rotation of the drums should not have any significant impact on the flame expansion. The combustion model settings for this simulation are the same as those used for the sub-section 3 model (Section 7.3) except for the following: • First order implicit time stepping (for simulation stability) • 3 levels of mesh adaption on the gradient of temperature every time step (for increased accuracy in predicting the flame front) 267
Colorado School of Mines
• Initial spark kernel radius is 2cm which matches the mesh size near ignition • Spark duration is 2ms Figure 7.24 through Figure 7.28 shows volume rendering of the total pressure and temperature after ignition near the headgate drum, looking from inside the longwall face towards the uncut coal face. As can be seen in these images, the overpressure from the explosion develops very quickly at 350m/s and expands to a much larger radius than the main flame front which is moving at approximately 30-35m/s as shown by the expansion of temperature in Figure 7.28. This is important because the quickly expanding pressure wave is increasing the pressures and temperatures inside the volume such that as the main flame front expands, the unburned upstream gases are slightly preheated. Subsequently, increased preheating in the unburned gases can increase combustion rates and flame acceleration, which is well known from fundamental flame theory (Andrews & Bradley, 1972). If these processes continue, this can lead to significant flame acceleration and possible transition to a detonation as described in Section 2.3. Figure 7.24 Volume rendering of total pressure showing ignition and explosion overpressure. 268
Colorado School of Mines
Figure 7.25 Volume rendering of total pressure showing ignition at the headgate drum and explosion overpressure overlaid with black streamlines for flow visualization. Also shown in Figure 7.25 and Figure 7.26 are black streamlines used to visualize the flow in the longwall face. As can be seen in these figures, the overpressure from the explosion diverts the main airflow in the longwall face such that there is little airflow near the coal face and tailgate drum. This is important because this pushes more flow into the gob area, which can potentially mix with pockets of methane creating more areas of EGZs. The diverted flow around the shield can also entrain more methane from the face, creating an environment which could lead to secondary or tertiary explosions. Figure 7.27 shows the decrease in pressure of the first pressure wave as it expands away from ignition. At 1ms the model predicts an overpressure of approximately 13kPa which decreases proportionally to the inverse of distance from ignition squared, as expected from a pressure wave. Similar to the full-scale sub-section 3 model presented in Section 7.3, the maximum overpressure is the same order of magnitude as those estimated from the UBB explosion (27kPa) (Page, et al., 2011). Also, although 13kPa was recorded at 1ms, there may 269
Colorado School of Mines
have been higher pressures generated between 0-1ms. Thus, researchers again recommend future work recording more frequent time steps in the initial flame kernel expansion in order to more accurately capture overpressure of the explosion as a function of time. This test successfully demonstrates the viability of integrating the combustion model into the full-scale bleeder ventilation model. One major issue that needs to be addressed is balancing computational time versus model accuracy. This current simulation has taken approximately 4 days to simulate 2.25ms of methane gas combustion using 4 x 24 cores nodes of computational power. It is important to note that this model has ~22.5 million base cells before using mesh adaptation. Mesh adaption is employed on the gradient of temperature to better resolve the flame front on such a coarse grid, but this increases the total simulation time. In the future, mesh coarsening and dynamic meshing can be used so that, as cells are added to the model to refine flow/temperature gradients, other mesh areas are coarsened thereby reducing the total number of cells in the model while still maintaining model accuracy and reasonable simulation times. Figure 7.26 Volume rendering of total pressure showing ignition at the headgate drum and black streamlines showing how the flow is diverted away from the coal face and tailgate drum. 270
Colorado School of Mines
7.4.1 Discussion of Potential Future Work To the knowledge of these researchers, the simulations presented in this Chapter are the first 3D, full-scale simulations of methane gas explosions in an underground longwall coal mine. More specifically, the simulations presented in Section 7.4 are monumental because they show that modeling a methane gas explosion in a longwall coal mine is feasible using a commercial CFD software, which has enormous potential for future research areas including: • Explosion prevention and mitigation • Design of mine layout • Shearer and drum design • Design of mine structures and seals • Design of water spray systems For explosion prevention and mitigation, for example, if there is an EGZ in a hanging roof behind the longwall shields, researchers could model different explosion scenarios to estimate flame speeds and overpressures. This type of information could be used to better distribute inert rock dust in these areas or perhaps include water sprays, increasing the humidity thereby decreasing the flame speed and pressure. Inclusion of a multiphase model (which is an option in ANSYS Fluent) could help in modeling the transition of a gaseous explosion to a coal dust explosion, like what happened in the UBB explosion (Page, et al., 2011). Running different mine explosion scenarios can also help research understand the potential transition of a deflagration to a detonation for improved mine design and layout. As described in Section 2.3, a deflagration can transition to a detonation by different flame acceleration mechanisms. For example, flame stretching can increase combustion rates, accelerating the flame. If there was an explosion in the gob, the turbulence induced by the nearby rock rubble and mine equipment could potentially aid in DDT. Also, deflagrations can transition to detonations if there is enough run-up distance; for methane-air mixtures this is typically around a L/D ratio of 50, but can be shortened by roughened walls (Ciccarelli & Dorofeev, 2008; Lee J. , 1984). In a mine, the entryways are typically 6m wide and can be 3m high, which means the hydraulic diameter of the entryways are 4m. Active longwall panels continue to get longer, and can be upwards of 1000m, corresponding to an L/D ratio of 250. In general this demonstrates how a mine environmental can be inherently dangerous, but the models developed in this work can aid in better designing certain areas of the mine with potential for methane gas explosions. 272
Colorado School of Mines
The models developed can also aid in the design of the shearer, drums, mine structures, and seals. ANSYS also owns ANSYS Mechanical which is a program that can estimate mechanical stresses and strains on solid objects. This can be coupled with ANSYS Fluent quite easily, such that researchers can estimate the forces and impacts of an explosion on nearby mine equipment and structure. This kind of information can also aid in designing better seals or perhaps the layout of seals. Finally, another potential research area could be the design of water spray systems either on the shearer, along the longwall face, or in the gob. Water sprays are important because they can capture entrained coal dust reducing the risk of a coal dust explosion. Water sprays can also provide enough pressure to mitigate the accumulation of EGZs and they can also make the nearby air humid, thereby decreasing the potential explosion hazard. In ANSYS Fluent, this could be done by incorporating the multiphase model in Fluent and running different spray scenarios. 7.5 2D Mine Model Simulation Section 7.1 through Section 7.4 presented high-fidelity, 3D models of methane gas explosions in underground coal mines. The models and simulations presented thus far have shown great potential in modeling these large-scale explosions in 3D, capturing the flame and pressure wave propagation trends as well as predicting reasonable flame speeds. However, they take a significant amount of simulation time and computational resources. Therefore, researchers have also created a 2D model of the ventilation conditions in a underground longwall coal mine and have combined it with the 2D combustion model developed in this research. The main purpose of this model is to simulate the flame and pressure wave propagation throughout the entire mine, but in a more user-friendly, reasonable amount of time. To begin, the 2D mine model was run as a steady state to obtain the velocity profiles in the mine; a diagram of the setup is shown in Figure 7.29. Airflow was initialized as a fluid with no methane while the gob was a porous media with 9.5% methane by volume (stoichiometric). With the velocity profile from the steady state case imported, the ANSYS Spark Model was turned on with an initial radius of 10cm at the location marked with a yellow star in Figure 7.29. As shown in Section 6.3.1, a flame cannot propagate in a Darcy flow porous media unless the porosity is set to 100%. Therefore, to obtain realistic flame propagation in this area, the box in the gob around the spark location was set as a fluid zone to allow the flame kernel to expand. 273
Colorado School of Mines
Figure 7.31 Absolute pressure contours of the 2D mine model with an ignition behind the shields, in an area of the gob which is not fully caved. Black lines are the forward and backwards streamlines of airflow along the longwall face. Time step = 0.01ms. CH = 9.5%. 4 Body mesh size = 10cm. Temperature = 293K, Pressure = 82kPa. E = 60mJ. ign Initial results at 10ms show that the flame propagates through the fluid zone of the gob and has resistance as it reaches the porous zone of the gob, shown in Figure 7.30. The black regions in Figure 7.30 are the areas where methane reacts with oxygen, representing the flame front. The pressure contour in Figure 7.31 shows that at a given time, a large pressure wave will propagate ahead of the flame front by several meters. As time progresses, the difference between the flame front and pressure front increases. Additionally, the black lines in Figure 7.31 represent velocity streamlines, showing that the pressure wave diverts the flow along the longwall face. These results were also observed modeling an explosion using the sub-section 3 model (Section 7.3) and the full-scale methane explosion model (Section 7.4). This is important because even in 2D, the model predicts general flame and pressure wave propagation trends observed in high- fidelity simulations, and in a fraction of the time, approximately 5 days. The 2D model is on-going, but future modeling will include a mesh independence study and a study investigating model parallelization for reduced computational times. Further studies including discrete modeling of obstacles in the gob rather than using a porous media as well as adding obstacles representing the hydraulic roof jacks along the longwall face. However, what is positive about this initial 2D model is that researchers will be able to run a variety of mine explosion scenarios and evaluate general flame and pressure wave propagation trends in under a 275
Colorado School of Mines
CHAPTER 8 SUMMARY OF IMPACTFUL RESULTS 8.1 Summary of Impactful Experimental Results Open-End Ignition versus Closed-End Ignition – Section 4.1 ➢ Purpose: These experiments demonstrate the difference between an un-confined (OEI) and a fully confined (CEI) methane gas explosion. ➢ Outcomes: • Ignition from a confined space increases flame speeds 5000% and peak overpressures 1200%. This increase is due to increased temperatures and pressures during flame kernel expansion, which increases fluid motion ahead of the flame thereby increasing turbulence and combustion rates. • CEI results in large overpressures and pressure oscillations, over 6 times greater than an OEI. ➢ Impact: In longwall coal mining, EGZs typically exist near the working face, behind the shields, and the corridors. Therefore, ignitions can occur in variety of locations and it is extremely important to capture these effects in the combustion model. Additionally, large overpressures from a methane gas explosion can cause serious damage to nearby workers and large pressure oscillations can damage ventilation controls and reverse airflow in a mine which will be important to capture in the combustion model. ➢ Novelty: Many researchers have compared open- versus closed-end ignition, but not across the wide range of scales under investigation in this project. Due to the variety of experimental setups and methods of measuring flame speed and overpressure, it would be extremely difficult to develop a comprehensive combustion model using other researcher’s experimental data. Even for those researchers who use different sized reactors, many of them have a different ratio of length to diameter, which makes scaling difficult and many of them did not have optical access to the flame. These results are novel in that they help us further complete our understanding of the effect of scale on methane flame dynamics as we continue validating larger combustion models. ➢ Presented at the SME Annual Conference, Minneapolis, Minnesota, 2018: 277