University
stringclasses
19 values
Text
stringlengths
458
20.7k
Colorado School of Mines
Table 6.1 Slopes calculated for each test group with respective R2 values Sample Group Slope: [(mg/m3) / (kW-hr/m3)] / mm R2 values Concrete 1 4.40 0.952 Concrete 2 2.59 0.996 Concrete 3 12.07 0.914 Limestone 1 0.557 0.796 Limestone 2 6.75 0.946 Limestone 3 2.56 0.885 Sandstone 1 0.265 0.790 Sandstone 2 2.69 0.989 Sandstone 3 3.41 0.375 Average 2.8 0.908 (excluding outliers) As shown in Figure 5.2 and Table 6.1, there is a positive linear relationship between dust concentration and specific energy input into the cutting as the pick becomes more worn with a greater tip (cid:3288)(cid:3282) (cid:4674) (cid:4675) (cid:4678) (cid:3288)(cid:3119) (cid:4679) (cid:3286)(cid:3272)(cid:3127)(cid:3283)(cid:3293) (cid:3417) (cid:4674) (cid:4675) radius. With the average slope at 2.8 (cid:3288)(cid:3119) 𝑚𝑚 it can be expected that the concentration of dust per specific energy unit will increase at a faster rate as the wear of the pick increases concerning the symmetrical millimeter radius. This value should be considered just a data point, not a concrete number used to quantify the dust generated in the rock-cutting processes as the picks wear out. This is due to rock texture and strength changes, pick geometry, cutter head design, and operational settings. Much more data and perhaps field studies are needed to offer a more suitable model for calculating the dust concentration generated as a function of SE. In practice, a system for monitoring power (perhaps an installed power transducer on the input cable of the machine or at the transformer hook up for the excavation unit) can be used to obtain the amount of energy used for excavating bulk volume of rock such that cutting specific energy can be calculated. Then, this value can be compared with the ambient airborne dust across different parts of the mine, from the heading to major access points into the mine workings. This combination of data could then unveil the relationship between the dust at the monitoring points with the amount of energy used for production. With further investigation, this information can subsequently be used to link the power draw for excavation of the rock of a given volume (which can be calculated from the location of the machine) to the bit wear conditions and anticipated dust in the mine workings. Ultimately, these preliminary 84
Colorado School of Mines
CHAPTER 7: CONCLUSIONS AND RECOMMENDATIONS 7.1 Contributions This study made many contributions, and the findings have addressed the knowledge gaps discussed in previous sections. The contributions are highlighted in this section and include the study of various methods along with the design and fabrication of the dust collection system, coordination of the full-scale cutting experiments, collection of the airborne and fines dust, analysis of the collected samples, measurement of the quantity of dust, analysis of particle size distributions, analysis of particle shapes, confirmation of findings in previous studies, examining dust characteristic trends, and comparing the dust characteristics with the specific energy of rock cutting. The new and novel methodology for collecting dust in the Linear Cutting Machine can provide rapid and representative dust samples. This automated dust collection system removed a researcher from turning knobs and switches while cutting, and therefore, this new system:  Eliminated human error with the exact timing set to control multiple pumps and vacuums;  Allowed for mass collection of dust samples with four ports for four different instruments;  Provided consistent dust collection times for all four instruments with the optimized timed system;  Decreased the collection time with the consistently optimized timed system;  Ensured a clean environment to collect representative samples before collecting the following samples with the real-time dust concentration monitor and HEPA filter system; The automated dust collection system can be utilized at the EMI lab and in similar studies for future experiments to increase dust collection output, decrease dust collection time, and consistently provide representative dust samples. With an increase in rapid dust sample collection, this new methodology can increase the rate of dust studies and understanding of airborne rock dust. Additionally, with limited previous dust experiments cutting rock in a full-scale environment, these experiments provided a foundation for studying dust characteristics when cutting samples at the full-scale level. The results from these experiments provided new dust characteristic data never seen before because cutting samples at full-scale on concrete, limestone, and sandstone hasn’t been analyzed before at this level. These experiments also contributed to the body of knowledge in dust studies by confirming findings in previous studies. For example, previous studies found that coal dust concentration increases 86
Colorado School of Mines
with increased symmetrical pick wear for conical picks. The thesis research agrees with this statement, and the analysis in this study has extended the available information from coal to other rock types, including concrete, limestone, and sandstone. The findings in this study offer an incremental improvement in understanding dust characteristics and their quantitative connection to pick wear. The observations confirmed that dust concentration increases with the pick tip radius increase. Results also suggest a quantitative relationship and future testing can expand comparisons and understanding. Additionally, the dust concentration compared to the cutting forces and specific energy and pick wear reveals future areas for investigation. This quantitative analysis was one of the novelties of the current studies and can lead to the development of methods for bit wear management strategies to control respirable airborne dust. The current study and measurements complimented the limited previous studies and analyses of the particles generated in the cutting process. These experiments included studying and presenting the particle size distributions and particle shape analysis as a function of symmetrical pick wear. The quantitative particle shape analyses performed on the collected dust samples from concrete, limestone, and sandstone can be used in future studies on the impact of dust on workers’ health in the construction and mining industries. There is also an understanding that particle shape characteristics are connected to rock type instead of pick wear. In the end, the research achieved all the objectives outlined in Section 1.4 Research Objectives which are as follows:  Verified that the respirable dust concentrations increase as the pick wear increases;  Determined the quantitative measure to track the rise in dust concentration as the pick wear increases;  Examined the presence of silica within the respirable airborne dust generated from the rock samples during the cutting process;  Investigated the shifts, or lack thereof, in airborne respirable dust particle size distributions generated from the three pick wears;  Investigated the particle size distributions of the fines material generated from the three pick wears;  Investigated the change, or lack thereof, in respirable airborne particle shapes generated from the three pick wears. 87
Colorado School of Mines
7.2 Conclusions For this thesis, full-scale cutting tests generated airborne dust and fines with a new, moderately worn, and fully worn conical pick. The LCM cut three medium-strength samples, which included a concrete, limestone, and sandstone sample. Various equipment and instruments collected the dust and determined the dust concentrations, silica contents, particle size distributions, and particle shapes. A comparative analysis examined the impacts of pick wear on various airborne and fines dust characteristics. Following is a summary of the main conclusions of this study concerning the effect of the symmetrical pick tip wear on characteristics of dust from mechanical excavation systems:  Airborne respirable dust concentrations increased as the pick wear increased for all rock types.  The concentration of dust samples collected in full-scale testing increased by about 50 mg/m3 for every millimeter increase in pick tip radius in concrete, limestone, and sandstone samples.  The silica content of the dust was a function of the rock type and was not impacted by pick wear levels. The suspended respirable dust containing quartz or cristobalite was a function of the original content of these minerals in the rock.  Statistical evidence suggests that the airborne particle size distributions are different from one another due to pick wear. However, there is no clear trend or connection to pick wear. The pick wear did not influence significant shifts in airborne particle size distributions within the respirable size range.  Strong statistical and visual evidence shows that the size distribution of the fines increased as the pick wear increased. Additionally, the pick wear did not significantly shift the fines material particle size distributions within the respirable size range.  There is strong statistical evidence that the pick wear did not influence any change in the particle shapes. Instead, the rock type influenced the particle shapes in the dust samples. All the picks consistently generated airborne respirable particles in all the rock types with particle shapes that were slightly oval with mostly smooth edges.  Comparison of the measured dust concentrations with the specific energy of cutting to the pick wear radii showed a positive linear increase of around 3.0 [(mg/m3) / (kW-hr/m3)] / mm during symmetrical wear. 88
Colorado School of Mines
7.3 Recommendations With this project's scope set to cut and analyze dust generated from three rock samples with three pick wears, it would be beneficial to continue work in this realm. Following is a list of recommended follow-up studies:  Cut additional rock types, such as coal and shale, to confirm and extrapolate trends found. This is because there is a significant focus in research and industry on coal dust. Extending the methods and analyses in this thesis to coal would be of interest.  Create and use additional pick wears while cutting future samples for dust collection and analyses. Adding a fourth or fifth pick wear level would increase the resolution and add more data to verify trends. For example, adding more pick tip radii will improve the dust concentration analyses by extending the outer bound of wear and increasing the concentration accuracy versus radius slope rate values.  Cut the samples at different attack angles to the sample surface to see if there is a correlation between the angle of pick and the dust characteristics.  Perform cutting tests in the field with a water spray system to analyze how the addition of dust control changes the research results.  Perform cutting tests on highly variable, rough, and angular surfaces to observe any changes in the mechanisms of dust generation and if there are any changes in the results.  Additional sample collection instruments would increase the number of duplicate samples in the testing matrix and add informative, real-time results. Adding more cyclones and Tsai Diffusion Samplers would confirm findings and add duplicates to the studies to improve the accuracy. Additionally, incorporating a Thermo Scientific Personal Dust Monitor 3700 into the dust collection setup would provide accurate real-time measurements of the dust concentrations. Adding in this one instrument can provide a time-dependent study with peak dust concentration measurements during cutting.  Compare the results of this study with field observations, such as using various dust measurement and monitoring systems to establish the impact of pick wear on dust generation in operations in connection to specific energy input to the machines.  Analyze the particle shapes of the fines material. This research only analyzed the airborne particle shapes, as using the FE-SEM to analyze particles down to 0.25 µm was feasible with the PC filters. It was more challenging to capture the fines particles on a surface possible for FE-SEM analysis. With this, the research in this study did not use the Microtrac SYNC laser diffraction instrument optical analysis feature for samples because the optical range goes as low as 4µm. It 89
Colorado School of Mines
APPENDIX A PUBLICATIONS AND PRESENTATIONS Journal Publications Slouka, S., J. Brune, and J. Rostami. "Characterization of Respirable Dust Generated from Full-Scale Laboratory Igneous Rock Cutting Tests with Conical Picks at Two Stages of Wear." Mining, Metallurgy & Exploration 39.4 (2022): 1801-1809. Slouka, Syd, et al. "Characterization of Respirable Dust Generated from Full Scale Cutting Tests in Limestone with Conical Picks at Three Stages of Wear." Minerals 12.8 (2022): 930. Slouka, Syd, et al. "Preliminary Characterization of Respirable Rock Dust Generated from Cutting Potash in Laboratory Full-Scale Tests with Radial Picks at Different Stages of Wear." Journal of Environmental Science and Engineering A11 (2022): 213-219. Peer-Reviewed Publications Slouka, S., J. Rostami, and J. Brune. "Characterization of respirable dust samples generated from picks at differing stages of wear." Mine Ventilation. CRC Press, 2021. 198-207. Slouka, S., Sidrow, E., C. Tsai, and J. Brune. “Comparing respirable dust characteristics from full scale cutting tests of three rock samples with conical picks at three stages of wear.” Underground Ventilation. CRC Press, 2023. 264-273. Conferences Presentations “Characterization of respirable dust samples generated from picks at differing stages of wear”. North American Mine Ventilation Symposium. June 2021. 97
Colorado School of Mines
ABSTRACT Fires and explosions in confined spaces are extremely dangerous, destroying homes and buildings, damaging infrastructure, and posing a fatal risk to civilians and fire responders. In 2016 alone, The National Fire and Protection Association estimated over a million reported fires, killing 14,650 civilians, 81% of which were home structural fires (Association, 2017). However, fires and explosions have been a problem across many industries including oil and gas, textiles, sugar refineries, and retail. A common denominator of the majority of incidents is that the fires and explosions occurred in confined spaces with complex geometries (i.e. apartment buildings, homes, industrial facilities, pipelines). This is important because explosions in confined spaces can quickly accelerate and result in catastrophic events; and obstacles in the path of the flame could generate a significant amount of turbulence accelerating a high-speed deflagration resulting ultimately in a detonation. This is especially important for the coal mining industry where methane gas explosions are a serious risk in underground mines and can be devastating such as the Upper Big Branch (UBB) explosion in West Virginia in 2010 which killed 29 miners (Page, et al., 2011), the Willow Creek explosions in Utah in 2000 which killed 2 miners and injured 8 more (McKinney, et al., 2001), and the Buchanan Mine in Virginia in 2005 which produced overpressures large enough to knock down miners (Carico, 2005). Although significant work has been done over the years to help mitigate these explosions, they still pose a fatal risk to workers. To gain a full understanding of these gas explosions requires detailed knowledge of mine ventilation schemes, the movement of methane gas in the mine, and high-speed methane gas deflagrations in the presence of various obstacles and run-up lengths. Therefore, the main objective of this research is to build a full-scale, 3D CFD model of a methane explosion in a longwall coal mine and to help assess risk and potential mitigation methods. The knowledge and experience gained in this research can easily be applied to other large-scale fires and explosions such as the 2017 Qishayan tunnel explosion in the Guizhou Province of China which killed 12 workers and injured more (PTI, 2017). The knowledge gained can also help assess the potential risk and provide guidance for stronger prevention strategies against such disasters; as well as guide future designs that minimize the potential for such disasters from occurring. iii
Colorado School of Mines
Methane deflagration experiments are performed to build a stronger understanding of how methane flames propagate and interact with obstacles under conditions typically found in a longwall coal mine. Subsequently the data obtained from these experiments is used to validate the combustion model across various scales, providing for a robust and accurate model. Researchers ignited methane-air mixtures in 12cm and 71cm diameter horizontal, cylindrical flame reactors and a rectangular, experimental box with and without obstacles used to simulate various gob characteristics. Measurements of methane flame front propagation velocities, explosion overpressure, and high-speed imaging was performed to develop a comprehensive understanding of flame behavior and provide multiple points of validation for the continual improvement of the CFD combustion model. Experimental results show that methane gas deflagrations in confined spaces are sensitive to the simulated gob characteristics investigated, which includes but are not limited to ignition location, obstacle location, void spacing, and obstacle surface topology. Key experiments were reproduced using a two-dimensional (2D) and three-dimensional (3D) CFD models and results demonstrate the ability of the model to capture methane flame propagation trends seen in experiments, matching maximum flame front propagation velocities within 7.5% in some cases. The models predict flame acceleration across obstacles, capturing the recirculation zone downstream of solid wall-type obstacles and flame propagation through porous gobs. Additionally, the models capture the effects of surface topology on local mixing and demonstrates the importance of modeling the gob area discretely instead of approximating it as Darcy flow porous media when modeling methane flame propagation in this area. After validating the different reactor models in 2D and 3D, the CFD combustion model was combined with a full-scale, 3D ventilation model of an underground longwall coal mine. Using this combined model, researchers have successfully modeled a methane gas explosion in a full-scale mine which is the first time this has ever been modeled. The scenario that was modeled was an ignition at the longwall face, near the headgate drum of the shearer. Results show a large pressure wave traveling at 350m/s leading an expanding flame front with a velocity of 30-35m/s. Results show the pressure wave compressing the air ahead of the flame, increasing the temperature of the unburned gases. This is an important result because increased temperatures can increase combustion rates and accelerate the flame which could potentially transition the flame from a deflagration to a detonation. Results from this study also show that the pressure iv
Colorado School of Mines
ACKNOWLEDGEMENTS Firstly, I would like to thank my thesis committee members: Dr. Gregory E. Bogin, Jr., Dr. Jürgen F. Brune, Dr. Hugh Miller, Dr. Jason Porter, and Dr. Neal Sullivan for their help and guidance in bringing this project to fruition. I thank my fellow researchers on this project including Aditya Juganda for his expertise and helpful discussions. My sincere gratitude to my parents, Richard and Jean Anne, and my siblings, Martha, Mary, and Joseph for all their support. Additional thanks to my partner, Patrick Nuessly, and all my friends for listening to my struggles and making sure I practiced self-love. Thank you to all the dogs I have sat over the years, for their cuddles, kisses, and warmth. Finally, this research is made possible with support from the National Institute for Occupational Safety and Health (NIOSH). Contract #211-2014-600050, and I would like to thank NIOSH and Dr. Gerrit Goodman for all their support and hard questions. Also, thank you to HPC at Mines for their use of supercomputing nodes on Mio. xxx
Colorado School of Mines
CHAPTER 1 INTRODUCTION Gas explosions are a hazard in the United States and often result in serious injury, death, and damage to structures. For example, in 2014 a natural gas explosion in East Harlem, NY killed 8 people, injured dozens, and leveled two buildings which were over 4 stories tall (Dunlap, 2015; Santora, 2014). In 2010, the San Bruno pipeline exploded in a residential area, resulting in flames that spread to nearby houses. The Pipeline and Hazardous Materials Safety Administration (PHMSA), which oversees 2.7 million miles of pipeline, estimates 131 public fatalities from 2005 to 2018 due to significant incidences which include fatalities or injuries requiring hospitalization, over $50k in costs, volatile liquid released from 5+ barrels or more, and liquid released that results in an unintentional fire or explosion (PHMSA, 2018). These statistics do not include pipeline incidents where a fire/explosion was the source of the incident, which may increase these numbers. Many of these explosions occur in confined spaces and are often exacerbated by nearby obstacles. This is especially true for the underground coal mining industry where methane gas explosions can occur deep within a mine, often in working areas. It is well known that explosive gas zones (EGZs) of methane and air are present in underground longwall coal mines and can be a potential hazard to mine equipment, structures, and workers (Brune, 2014; Karacan, Ruiz, Cote, & Phipps, 2011). In extreme cases these EGZs can ignite and result in fatal methane gas explosions, as evidenced by the recent mine explosions detailed subsequently: • In 2000, the Willow Creek mine in Utah, USA, experienced 4 explosions which reversed airflow in the mine and resulted in a fire burning behind the shields near the longwall gob (McKinney, et al., 2001). The mine fire and explosion at Willow Creek was likely due to a roof fall igniting an EGZ in the longwall gob, coupled with a lack of ventilation air, resulting in an explosion which killed 2 miners and severely injured 8 others. • In 2005, the Buchanan mine in Virginia, USA had a mine fire resulting from a roof fall of thick sandstone releasing methane near the shearer (Carico, 2005). Investigators hypothesized that the EGZ was ignited from sparks from the shearer or heat from cutting 1
Colorado School of Mines
bits. Although the Buchanan mine fire resulted in mine damage, no miners were killed in this accident. • In 2006, an explosion occurred the Sago mine in West Virginia, USA which killed 12 miners (one from CO poisoning) and injured 1 (Gates, et al., 2006). The investigative report concluded that an EGZ had formed in an area that had been previously mined and sealed. The seal was not built properly and investigators believe it may have been able to only withstand 20psi of pressure forces whereas pressure forces of the explosion were estimated to be greater than 93psi based on damage in the mine area. The ignition source was investigated by Sandia National Laboratories who found that likely a lightening strike caused an arc in a nearby pump cable which ignited the EGZ. • In 2010, there was an explosion at the Upper Big Branch Mine in West Virginia, USA, which entrained coal dust leading to a deadly explosion killing 29 miners (Page, et al., 2011). Reports found that a recent roof fall near the tailgate restricted airflow, allowing for an EGZ to form near the shearer. The shearer was cutting sandstone and worn shearer bits left hot smears, which ignited the EGZ resulting in an explosion. The pressure waves entrained coal dust, due to a lack of rock dust in this area, resulting in a massive coal dust explosion. Investigators also found that a water barrier did help to stop some of the flame propagation, but recommended looking into different methods for future sealing. • In 2014, the Soma Mine explosion in Turkey killed 301 people, trapping hundreds of miners underground (Tuysuz, Watson, & Smith-Spark, 2014). Also, in 2016 there were two coal mine explosions in China which killed a total of 65 miners (Luu, 2016; Wang & Dong, 2016). Unfortunately, there is not enough evidence of these explosions to determine the EGZ location and cause for ignition. These examples help demonstrate that coal mine explosions are not just a problem in the United States, but are an international concern. These examples also show the diversity of explosions and causes, as well as the need for improved understanding of these explosions. Current methods used to mitigate the likelihood of an explosion include ventilation schemes, nitrogen inertization, water sprays, and gob ventilation boreholes among some. However, these methods do not totally prevent methane accumulation in the mine since coal beds contain methane at high pressures which naturally migrates and outgases during the mining process (Karacan, Ruiz, Cote, & Phipps, 2011). Thus, one of the goals of this project is to leverage the 2
Colorado School of Mines
models during the design and developmental stage of the underground longwall mine to try to reduce the potential for future explosion disasters. The amount of methane gas emission depends on a variety of factors both geological and operational including the depth of the coal seam, longwall panel size, amount of coal production, mining height, and degasification (Karacan, Ruiz, Cote, & Phipps, 2011). These complexities added to the fact that mining is a transient process makes it difficult to understand exactly where these EGZs are located and migrate. Detailed investigative accident reports and research has proven that EGZs exist in the under- and over-lying strata as well as the collapsed strata/gob and can migrate towards the working longwall face (Brune, 2014). The active panel of a longwall coal mine consists of an entry, the longwall face, the headgate and tailgate, the conveyor/belt, and the longwall face as shown in Figure 1.1 and Figure 1.2. The longwall face is typically 300-400m long and 3m high and the active panel can be up to 2000m long. A shearer cuts along the coal face, moving back and forth along the face. Meanwhile, hydraulic roof supports hold up the strata (roof) and advance forward as the shearer cuts away at the face. When the roof supports advance forward, the strata that was once supported collapses creating the gob as shown in Figure 1.3. Figure 1.1 Schematic of an active panel of a longwall coal mine. Blue arrows represent airflow pattern. Figure courtesy of CSM research group. 3
Colorado School of Mines
tandem, Fig, Bogin, Brune, and Grubb (2016) has designed and built the experimental setups for investigation of high-speed methane deflagrations at both laboratory-scale (5cm diameter, 43cm length; 9cm diameter, 81cm length; 12cm diameter, 150cm length; 13.6cm diameter, 1.15m length; 30.5cm diameter 1.2m length) and large-scale (71cm diameter, 6.1m length) which provides guidance in developing the combustion models. Experimental results show that mine conditions such as humidity, temperature, and pressure have a significant impact on methane flame enhancement across all scales (Fig, Bogin, Brune, & Grubb, 2017). Additionally, Fig, Strebinger, Bogin, and Brune, 2018 have shown methane flame enhancement across a simulated rock gob. The research presented in this thesis aims at providing a more detailed description of which physical properties of the gob and mine environment impact methane flame enhancement; rock material, void size, void location, obstacle geometry, and porosity. Methane gas explosions in a longwall coal mine may occur behind the longwall shields such as the Buchanan mine fire in 2005 (Carico, 2005), in an entry/exit, or near the longwall face such as the Upper Big Branch explosion (Page, et al., 2011) which all have varying degrees of confinement. Thus, this research also investigates the impact of confinement on methane flame and pressure wave propagation. Finally, the ultimate goal of this project is to combine the CFD ventilation model and combustion model and perform large-scale simulations of methane gas explosions in a mine. 6
Colorado School of Mines
CHAPTER 2 BACKGROUND Combustion is integral to society and daily life, providing heat for living spaces, electricity for industrial processes, energy for transportation, and heat for food preparation. There are four essential pillars of combustion: heat, fuel, oxidizer, and chemical reactions. The combination of the four produces rapid oxidation of a fuel, producing heat and/or light (Turns, 2012). Since longwall coal mine explosions typically result in a devastating pressure wave or blast wave and an either subsonic or supersonic chemical reaction zone, this manuscript is mainly concern with understanding combustion flames and not autoignition events, though they may be possible. Flames are typically categorized by the method of fuel and oxidizer mixing. When fuel and oxidizer are mixed before the addition of heat, the resulting flame is called a premixed flame, such as spark ignition engines or gas fired furnaces. Non-premixed or diffusion flames occur when the fuel and oxidizer mix while chemical reactions occur, such as a cigarette lighter or candle flame. In longwall coal mining, typically the methane gas in the mine mixes with the fresh ventilation air and as such, any resulting combustion event and flame is either a partially premixed or premixed flame. The focus of this study is specifically on premixed flames since the resulting explosion can be more deadly than a non-premixed flame. There are two types of combustion waves: deflagrations and detonations. Deflagrations are combustion waves that propagate at subsonic velocities and are categorized as either laminar or turbulent. Under certain conditions deflagrations can transition to detonations (deflagration to detonation transition - DDT) which are combustion waves that propagation at supersonic velocities. The transition of deflagrations to detonations will be discussed in Section 2.3, but will be introduced here briefly. DDT can occur when a mixture is ignited in a confined space and sufficient run-up length and/or obstacles are present, resulting in a leading shock wave that is coupled to the combustion zone traveling at supersonic velocities. The upstream and downstream properties of deflagrations and detonations are significantly different: the ratio of the downstream temperature to the upstream temperature for a deflagration is almost 7.5 versus a detonation is 8-21 and the ratio of the downstream to the upstream density for a deflagration is 7
Colorado School of Mines
0.13 versus 1.7-2.6 for a detonation (Turns, 2012). Also, the pressure jump across a detonation wave is over 10 times larger than a deflagration (pressure jump across a deflagration is approximately 1) which is why detonations can be so devastating (Turns, 2012). For example, in longwall coal mining the Upper Big Branch explosion in 2010, which entrained coal dust leading to a more violent explosion, investigators estimated flames traveling upwards of 450m/s and pressure waves of 170kPa, which are in the DDT range (Page, et al., 2011). Therefore, many researchers in longwall coal mining are interested in understanding deflagrations and DDTs (Lee, Knystautas, & Chan, 1985; Oran & Gamezo, 2007; Oran, Gamezo, & Kessler, 2011) including the CSM research group. As mentioned, deflagrations are either laminar or turbulent flames which is dependent on the velocity of the flame, a characteristic length scale dependent on the surrounding geometry/environment, and fluid properties. Turbulent flows are dominated by fluid movement which is highly fluctuating and chaotic in both space and time and, thus are often difficult to define. The turbulent Reynolds number (Re) is used to help quantify a turbulent flow; it is the ratio of inertial forces to viscous forces and is defined below in Equation (2.1): (2.1) where v’ is the root mean square (RMS) fluctuating velocity of the flow (m/s), is the integral rms length scale representing the mean size of the large eddies (m), and υ is the kinematic viscosity (m2/s) (Irbin, Yetter, & Glumac, 2015; Turns, 2012). For turbulent Reynolds numbers equal to or less than one, the flow is considered laminar, otherwise the flow regime is turbulent. Turbulent flames have three main reaction regimes: wrinkled laminar flames (flamelets), flamelets in eddies, and distributed reaction. These regimes are defined by the relation of laminar flame thickness, δ , (i.e. the thickness of the reaction zone by molecular transport) to turbulent L length scales. The largest turbulent length scale is the macroscale, L, which is dependent on the surrounding geometry (e.g. in a cylindrical tube L would be the tube diameter). As mentioned, As mentioned, is the integral length scale and is always smaller than L. Next is the Taylor microscale, , which relates flow characteristics to the mean rate of strain (Irbin, Yetter, & λ lo Glumac, 2015; Turns, 2012). Finally, the smallest length scale is the Kolmogorov microscale, , κ 8
Colorado School of Mines
which is the smallest length scale and represents the scale at which turbulent kinetic energy is transferred to fluid internal energy (Irbin, Yetter, & Glumac, 2015; Turns, 2012). Knowing all this, the wrinkled laminar turbulent flame regime is when the laminar flame thickness is less than the smallest turbulent length scale, the Kolmogorov microscale. The flamelets in eddies regime is intermediate, when the laminar flame thickness is between the integral length scale and the Kolmogorov microscale (Irbin, Yetter, & Glumac, 2015; Turns, 2012). Then the distributed reaction regime is when the laminar flame thickness is greater than the integral scale. Figure 2.1 The expansion of unburned gas through a laminar, planar flame. Typically in combustion the three turbulent regimes are plotted in relation to the turbulent Reynolds number, Equation (2.1), and the Damkohler number (Da), Equation (2.2), which is the ratio of the characteristic flow time to the characteristic chemical time represented by Equation (2.3). (2.2) (2.3) where S is the laminar flame speed, which is defined as the “speed of an unstretched laminar L flame through a quiescent mixture” (Turns, 2012) as shown in Figure 2.1. The flame front moves perpendicular in the direction of the unburned fuel/oxidizer mixture (reactants) which has a 9
Colorado School of Mines
(2.7) Unfortunately, many of the turbulent flame speed models do not fully take into account flame stretching and stability (Turns, 2012) and so cannot be used for all flows, but are useful for estimation. For small Damkohler numbers less than 1 and small turbulent Reynolds numbers, the turbulent flame regime is a distributed reaction which is a zone between the burned and unburned gases which has small integral length scales of eddies, but large RMS velocities. At this point, it is unclear whether or not flames occur in this regime since they would be subject to large pressure differences and various length scales making the flame front inherently unstable (Turns, 2012). The last turbulent flame regime is the flamelets in eddies which is dominated by moderate Damkohler numbers and high turbulent Reynolds numbers. In this regime, pockets of burned and almost burned gas pockets are trapped in the flame zone and are transported to the burned gas by turbulent mixing. The combustion rate in this regime depends on the rate of unburned gas pockets. Premixed turbulent flames travel at subsonic velocities, but they can transition to detonations (DDT) depending on the confinement of the geometry, mixture ratio, and ignition source (Glassman, Yetter, & Glumac, 2015). For example, in a tube with both ends open or one end open, if a mixture is ignited from the open end it will always travel at subsonic velocities. However, if a mixture is ignited from the closed end of a tube that is either open-closed or closed-closed, then the combustion wave has the potential to reach the speed of sound as long as the tube is long enough (Glassman, Yetter, & Glumac, 2015). However, it is still unclear whether or not methane can transition to a detonation in coal mining conditions (Oran, Gamezo, & Kessler, 2011). Therefore, this research is mainly concerned with methane gas deflagrations and the potential of methane flames to transition to a detonation. All the flames presented in this proposal are turbulent premixed methane gas deflagrations in the wrinkled laminar regime or flamelets in eddies regime. 12
Colorado School of Mines
2.1 Parameters influencing laminar flames There are many factors that affect the laminar flame velocity of fuel/oxidizer mixtures including temperature, pressure, fuel type, and mixture stoichiometry which shall be discussed in more detail subsequently. The mixture stoichiometry refers to the amount of oxidizer required to completely burn an amount of fuel. Since methane is the predominant fuel in underground longwall coal mining, the global equilibrium reaction of methane and air is defined by Equation (2.8). In the global equilibrium reaction, methane reacts with air to form carbon dioxide, water, and nitrogen. (2.8) From this global reaction the stoichiometric air-fuel ratio is defined as the ratio of the mass of air to mass of fuel required for a stoichiometric reaction of methane and air, Equation (2.9). (2.9) And finally, the equivalence ratio is the ratio of the stoichiometric air-fuel ratio to the actual air- fuel ratio: (2.10) The equivalence ratio is typically used to define the stoichiometry of the mixture. If the equivalence ratio is equal to unity, then the fuel-air mixture is stoichiometric. If the mixture has excess air, meaning it is fuel lean, then the equivalence ratio is less than unity. If the mixture has excess fuel, it is fuel rich and the equivalence ratio is greater than unity. As previous discussed, the laminar flame speed of a fuel-oxidizer mixture depends on the density of unburned and burned gases, which is coupled to the stoichiometry of the mixture through the global reaction. Assuming a single global step chemistry, an estimate of the laminar 13
Colorado School of Mines
flame speed can be obtained with assumptions; these assumptions are outlined in Turns, 2012, following Spalding’s theory for laminar flame propagation (Spalding, 1979), and will not be discussed here (Turns, 2012). After several assumptions, including no pressure change across the flame, a general 1-D laminar flame theory is derived from conservation laws: (2.11) Equation (2.11) shows that the laminar flame speed is a function of thermal diffusivity (α), viscosity (υ), the average reaction rate, and density (ρ) which means it is also a function of temperature. As can be seen from this equation, the laminar flame speed of a methane-air mixture is highly sensitive to the mixture stoichiometry which has been studied experimentally by Andrews & Bradley (1972). The results of their study are shown in Figure 2.4 and Figure 2.5. As can be seen from these figures, the maximum flame speed is achieved at slightly fuel rich conditions, which also corresponds to the highest flame temperatures. Understanding how flame speed is affected by stoichiometry is extremely important for underground coal explosions since there is a widely varying distribution of methane in a mine. For example, in the longwall gob area, methane concentrations can be close to almost 100%, whereas near the edge of the gob, the methane is diluted from the ventilation air, resulting in more fuel lean mixtures. This research investigates the effects of mixture stoichiometry on methane flame propagation and pressure. In addition to mixture stoichiometry, increasing the unburned mixture temperature can greatly enhance laminar flame speed. Increasing the temperature of the unburned gas mixture promotes dissociation of minor species, which increases combustion rates and thus, flame speed. For methane-air mixtures, the laminar flame speed increases parabolically with the temperature of the unburned gas which was experimentally determined by Andrews & Bradley (1972): (2.12) 14
Colorado School of Mines
Although the laminar flame speed of flames has a positive dependence on the unburned gas temperature, it has a negative dependence on pressure as shown in Equation (2.13), which was determined by Andrews & Bradley (1972) for pressures above 5atm. This is due to the fact that an increase in pressure shifts the equilibrium and suppresses dissociation of minor species, resulting in slower flame speeds. However, this is not the case for turbulent flames where the increased pressure can help increase local temperatures and flame wrinkling leading to more violent explosions. This was introduced earlier in this Chapter and will be discussed in later sections. (2.13) Finally, since laminar flame speed is a property of a fuel/oxidizer mixture, each fuel has a different laminar flame speed. For example, the laminar flame speed of a stoichiometric mixture of methane and air at room temperature, 300K, and 1atm is approximately 40 cm/s, whereas hydrogen is significantly faster at 210 cm/s (Turns, 2012). A combination of factors result in hydrogen’s high flame speeds including a higher thermal diffusivity, higher mass diffusivity, and fast chemical kinetics. Thus far the properties of premixed, laminar methane flames have been described in terms of a propagating flame. However, there are many processes which can hinder the propagation of a flame, including flame quenching mechanisms, flammability limits, and minimum ignition energies. The quenching distance of a flame is defined as the critical diameter of a cylinder where a flame is extinguished due to heat losses to the cool walls of the cylinder (Irbin, Yetter, & Glumac, 2015; Turns, 2012). The analytical derivation of quenching criteria is a balance of heat of reaction and heat lost to the cool walls by conduction. However, typically quenching distances of mixtures are determined experimentally and depend on the flame thickness which is a function of mixture stoichiometry and initial conditions. For a stoichiometric mixture of methane and air at 300K and 1atm, the laminar quenching distance is typically on the order of 2mm (Turns, 2012). Understanding the quenching limits of methane flames is important for this research because typical coal mine explosions occur near or within the gob area such as in the Willow Creek explosion in 2000 and the Upper Big Branch explosion in 2010 (McKinney, et al., 2001; Page, et al., 2011). There has been anecdotal evidence over the 16
Colorado School of Mines
years from miners who have heard popping or seen flames deep within the gob, which never propagate to the longwall face (Brune, 2014). Since the gob area consists of different sizes of rock rubble and varying levels of compaction, perhaps some flames deep within the gob are quenched by the nearby rock rubble. Thus, it is one of the objectives of this project to model these types of phenomena and determine whether or not previous anecdotal evidence suggests methane flames can occur deep within the gob, as well as potentially propagate if originating near the longwall shields. Another property of a fuel-oxidizer mixture is the flammability limits of the mixture which are upper and lower limits in which a flame will propagate given a minimum amount of energy. The lower limit corresponds to the limit of propagation of a lean mixture and the upper limit corresponds to flame propagation in the richest mixture. Typically these limits are determined experimentally and are dependent on the type of experimental apparatus employed. In general for methane and air mixtures, the flammability limits are between 5-15% for upward propagation in a vertical cylinder (Coward & Jones, 1952). However, the flammability limits of a given mixture can vary depending on the apparatus (vertical or horizontal), gravity, and pressure. For example, Ronney and Wachman (1985) found that the lean flammability limits of methane and air can change almost 20% traveling upward versus downward. Taking into account the fuel-oxidizer mixture and environmental conditions, mixtures require a certain amount of minimum ignition energy for a flame to propagate. The minimum energy is typically determined by assuming that all heat from an electric spark is transferred to a critical volume of mixture (Turns, 2012). The critical volume is determined by balancing the amount of heat released from combustion and heat loss by conduction to the surrounding environment. Thus, for stoichiometric mixtures of methane and air at 300K, 1atm, the minimum ignition energy is approximately 0.5mJ which will be compared to experimentally determined minimum energies in Section 2.7 (Turns, 2012). This is important for longwall coal mining because there are many sources of ignition in a mine including rock-on-rock friction, machine- on-rock friction, hot smears, spontaneous combustion, and lightening among some. 2.2 Overview of flame dynamics in smooth cylindrical reactors Flame propagation has been studied for decades and the apparatus and methods by which researchers study flames vary depending on the purpose of the research. For example, typically when measuring the laminar flame speed of a mixture, researchers have used flat flame burners 17
Colorado School of Mines
or spherical bomb methods (Agnew & Graiff, 1961; Andrews & Bradley, 1972; Rallis & Garforth, 1980). However, in 1908 Henry Le Chatelier began studying flames in cylindrical reactors, where the flame traveled along the axis of the reactor (Le Chatelier, 1908). After this, many researchers began studying flame propagation in cylindrical or rectangular reactors because it allows researchers to easily vary experimental conditions. Additionally, many researchers focus their studies on industrial gas pipeline explosions which is also why using cylindrical reactors has become common in studying fundamental flame propagation. One of the main advantages of using a cylindrical reactor is that it easily lends to varying end conditions (i.e. open-open, open-closed, closed-closed) which is important because research has shown end conditions results in varying flame propagation speeds, overpressure, and flame propagation phenomena. In a longwall coal mine, methane gas explosions typically occur in rectangular, horizontal hallways or passageways. Additionally, methane gas explosions can occur in areas that are either completely confined or partially confined. The main difference with experimenting in a rectangular chamber is that there will be residual pockets of unburned gas in the corners of the chamber (Cooper, Fairweather, & Tite, 1986; Solberg, Pappas, & Skramstad, 1981). These pockets can continue to burn after the main flame front has propagated, which can increase pressure rise in the reactor (Cooper, Fairweather, & Tite, 1986). However, in a longwall coal mine explosion the main concern is the propagating flame brush and therefore, this research uses horizontal cylindrical reactors that are open on one end and closed on the other in order to more closely resemble flames propagating in a mine environment. Depending on the end conditions of the experimental setup, the flame propagation dynamics can be extremely different. For a mixture ignited at the open end of a horizontal cylinder, the flame first expands spherically as can be seen in Figure 2.6. As the flame comes closer to the edges of the reactor, the reaction front near the cool walls begins to lose heat due to the temperature gradient. At this point the flame front is slightly retarded and some hot exhaust gases flow out of the open end. During this time, some of the hot product gases also rise to the top of the reactor due to buoyancy. When this happens, the top of flame front is slightly tipped over at an angle as it continues propagating towards the closed end of the reactor. As the flame travels towards the closed end of the reactor it combusts all of the mixture. Near the closed end of the reactor the flame becomes unstable as it compresses and burns the residual mixture. This phenomena has been observed by many researchers (Ellis & Wheeler, 1928; Gerstein, Levine, & 18
Colorado School of Mines
Wong, 1951; Guenoche & Jouy, 1953). Additionally, Guenoche and Jouy (1953) found that if there is a small orifice on the closed end of the reactor it helps stabilize the flame. Further discussion on flame instabilities will be provided in Section 2.4. Figure 2.6 Images of a stoichiometric methane-air flame traveling from the open end of a 12cm diameter quartz reactor towards the closed end. Flame travels from left to right. CH = 4 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole at the ign closed end. If a mixture is ignited from the closed end of the reactor, it will also initially expand spherically as shown in Figure 2.7. During flame expansion, pressure waves emanate from the flame front causing turbulence to develop upstream of the flame front increasing the transport of unburned gases to the reaction zone, accelerating the flame. Then, due to the small volume and hot product gases expanding, the pressure behind the flame increases and further accelerates the flame front. Since there is no obstruction ahead of the flame, the flame is able to freely propagate towards the open end as shown in Figure 2.7. Due to the confinement and fluid movement ahead of the flame from the pressure generated in this type of explosion the flame accelerates much faster than a flame propagating from the open end of the reactor. This type of flame propagation has been observed and studied by many other researchers using reactors with open-closed end conditions and closed-closed end conditions (Clanet & Searby, 1996; Ellis & Wheeler, 1928; Xiaoping, Minggao, Wentao, Meng, & Juniie, 2015). 19
Colorado School of Mines
Figure 2.7 Images of a stoichiometric methane flame traveling from the closed end of a 12cm diameter quartz reactor towards the open end. Flame travels from right to left. CH = 9.5±0.3%. 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole at the closed end. ign The cases described thus far have been for flames propagating in smooth cylindrical reactors. However, there are many processes which can affect the propagation of a flame, some of which includes flame instabilities (both intrinsic and external), the degree of confinement, environmental conditions, obstacles, and ignition energies and sources. Understanding these effects are extremely important for developing a comprehensive combustion model of a longwall coal mine explosion. For example, in a real longwall coal mine, the passages and corridors are made out of irregular shaped rock, meaning the boundary conditions are not smooth. Also, there are many obstacles in a mine, including mine equipment, mine workers, and pillars which can either enhance or retard the flame. For brevity, the following sections will attempt to show some of the major effects of these different phenomena on flame propagation. 2.3 Transition of a deflagration to detonation As previously discussed, deflagrations, both laminar and turbulent flames, travel at subsonic velocities, but have the potential to transition to a detonation which travels at supersonic velocities. The shock wave, or pressure wave, ahead of a deflagration is supported by the constant expansion of combustion products as was discussed and shown in Figure 2.7. The combustion process of a deflagration is different than a detonation, where the shock wave produced by a detonation increases the temperature and pressures of the nearby gases such that chemical reactions occur in the form of a flame front. In a detonation, the chemical reaction front follows the shock wave and travels with the shock wave at an almost constant speed. The minimum speed that a detonation can travel at is called the Chapman-Jouguet (CJ) detonation 20
Colorado School of Mines
state (Ciccarelli & Dorofeev, 2008). The CJ velocity can be thought of as the sonic velocity, usually above 1000m/s, and depends on the fuel and mixture composition because it is a choked state that supports the minimum entropy increase across the detonation wave (Ciccarelli & Dorofeev, 2008). Detonations also have a complex structure and can be visualized many ways including soot tracking (Ciccarelli & Dorofeev, 2008; Kuznetsov, et al., 2002). Figure 2.8 shows soot tracking images of methane-air detonations with vary stoichiometries taken by Kuznetsov, et al. 2002. As demonstrated in these images, the detonation wave has a structure to it which is cellular; typically the size of the cell is determined by the distances between points where the pressure waves interact and is detonated by λ (Turns, 2012). The cell size changes depending on fuel, mixture compositon, and initial conditions, which can be seen in these images; the cell size changes as a function of mixture stoichiometry. The cell sizes are important because a detonation inside a cylindrical reactor will not be sustained if the cell size, λ, is larger than the diameter of reactor (Ciccarelli & Dorofeev, 2008). Figure 2.8 Soot tracking visualization of methane-air detonations with varying mixture stoichiometries. From the left: CH 8.5% by volume, 11%, and 12%. Image credit: (Kuznetsov, 4 et al., 2002). Detonations can also occur if the ignition energy source is large enough or a deflagration transitions to a detonation (Ciccarelli & Dorofeev, 2008; Kuznetsov, et al., 2002; Zipf, et al., 2013). However, in an underground longwall coal mine, there are not many ignition energy sources strong enough to immediately produce a detonation, which is why this research is mainly concerned with DDT. Ciccarelli and Dorofeev (2008) wrote an overview of flame acceleration and transition to detonation, noting that the first step to DDT is flame acceleration. To illustrate this consider the 21
Colorado School of Mines
discussion regarding a confined, closed-end ignition in Figure 2.7. In this scenario, the confinement forces the combustion products to expand in all directions and the pressure waves emmanating from the flame front creates movement in the upstream, unburned gases. Together these higher pressures and temperatures from the confinement accelerates the flame front down the reactor. The flame acceleration is highly dependent on the reactor size, the mixture composition, reactor wall/tube roughness, initial conditions, and presence of obstacles (Ciccarelli & Dorofeev, 2008; Kuznetsov, et al., 2002; Silvestrini, Genova, Parisi, & Trujillo, 2008). As discussed, the detonation cell size has to be smaller or on the order of the reactor diameter in order to sustain a detonation. The flame acceleraton process has a critical distance, the run-up length, at which the flame inside the reactor will transition to a detonation. Previous researchers have compared the length to diameter (L/D) ratios necessary for flames to transition in smooth reactors and for methane-air mixtures this is typically above an L/D ratio of 50 (Ciccarelli & Dorofeev, 2008; Lee J. , 1984). However, as the diameter of the reactor increases, the run-up distance decreases as shown in Figure 2.9. Figure 2.9 Run-up distance versus reactor diameter in a smooth reactor for different fuels. Figure credit: (Ciccarelli & Dorofeev, 2008). After achieving flame acceleration, the transiton to a detonation can happen in two main ways: 1) from shock reflection or focusing or 2) from instabilities and mixing (Ciccarelli & Dorofeev, 2008; Glassman, Yetter, & Glumac, 2015). In shock reflection or focusing, the pressure waves continually emmanating from the flame front eventually coalesce creating a 22
Colorado School of Mines
shock wave. As discussed in Section 2.1, the pressures and temperatures across a detonation wave are hugely different and can autoignite unburned pockets of fuel-air mixture creating a detonation. Shock focusing is when the shock waves are focused due to a concave wall and again, autoignites the unburned gas. The second method of transitioning an accelerating flame to a detonation is by instabilities and mixing processes. In general, one of the main ways the run-up distance can be decreased is by roughened walls or the presence of obstacles, which will change depending on the obstacle configuration (Ciccarelli & Dorofeev, 2008; Chapman & Wheeler, 1926; Glassman, Yetter, & Glumac, 2015; Kuznetsov, et al., 2002; Oran, Gamezo, & Kessler, 2011; Silvestrini, Genova, Parisi, & Trujillo, 2008; Zipf, et al., 2013). Significant research has gone into flame acceleration by roughened tubes and obstacles and this will be discussed in more depth in Section 2.6. In general, obstacles tend to increase mixing and fluid movement which increases unburned gases to the flame front in addition to stretching the flame front. Stretching of the flame front increases combustion rates leading to flame acceleration. This flame acceleration is important because in a longwall coal mine, the longwall face can be 300m long and the entryways and corridors in the mine can be kilometers long. Added to the fact that the mine walls are roughened rock, the run-up distance of flame acceleration in an underground coal mine will be significantly shorter than experiments in smooth reactors. There have been many researchers investigating methane flame acceleration to DDT for application to mine explosions including Zipf, et al. (2013) and Oran, Gamezo and Kessler (2011). Zipf, et al. (2013) performed experiments in a reactor 1.03m in diameter and 73m long with obstacles and was able to achieve detonation speeds near the CJ velocity. However, to get an immediate detonation, a pocket of methane and oxygen mixture was located near the spark and so it is difficult to determine whether or not this scenario is perfectly indicative of a coal mine explosion. Oran, Gamezo and Kessler (2011) developed a complex CFD model of methane flame acceleration to DDT and found transition to detonation by flame interaction with obstacles and boundary layers which helped to create hot spots for autoignition. The research presented in this manuscript is mainly concerned with high-speed deflagrations and determining which aspects of a longwall coal mine environment, environmental, physical, etc., impact methane flame propagation and the potential to transition to a detonation. Previous researchers on this project, Fig (2019), investigated the impact of 23
Colorado School of Mines
environmental conditions and flame propagation as a function of scale by performing experiments in cylindrical reactors at a variety of scales, 5cm diameter to 71cm diameter, with a fixed L/D = 8.5. Research showed that methane flame propagation velocities increases non- linearly as a function of scale (Fig, 2019) which agrees with DDT theory and other reseachers (Ciccarelli & Dorofeev, 2008; Chapman & Wheeler, 1926; Kuznetsov, et al., 2002; Silvestrini, Genova, Parisi, & Trujillo, 2008). The research in this manuscript is focused on how different obstacles, shapes, sizes, and configurations, can impact methane flame propagation as well as ignition location, energy, and confinement. 2.4 Impact of flame instabilities Although there may be instances in which premixed flames are planar as shown in Figure 2.1 (page 9), not all flames are planar. Premixed flames are subject to many different types of flame instabilities including instabilities which are inherent to the combustion process. Hydrodynamic instabilities are natural for premixed flames because the unburned gas expands across the flame front, which means the burned gas velocity is higher and density is lower. For turbulent premixed flames, there is also a larger pressure difference across the flame and it is this pressure difference eventually disturbs the gas flow such that the flame becomes wrinkled as shown in exaggeration in Figure 2.3 (page 11). For a wrinkled flame, the gas flow moves perpendicular to the flame front, which means that the flow converges and diverges so that the burned gas flow has areas where the flow is accelerated (Jarosinski & Veyssiere, 2009). This type of instability is called the Darrieus-Landau (D-L) instability and will most likely grow with time. Simultaneously, there are thermodynamic-diffusive effects which affect the flame in what is termed the diffusive zone (Jarosinski & Veyssiere, 2009). In this diffusive zone, which is the size of the flame thickness, heat is diffused away from the flame front towards the unburned gas. At the same time, mass is diffused and convected towards the flame front and always opposite of the heat flux. In areas of the wrinkled flame, where the flame front is concave towards the unburned gas, the heat flux is locally convergent, helping to stabilize the flame, and the mass or species flux is locally divergent (Jarosinski & Veyssiere, 2009). The balancing of hydrodynamic (D-L) instabilities and thermo-diffusive effects is typically evaluated in terms of the Lewis number (Le), which is a non-dimensional parameter and is the ratio of the thermal diffusivity and species diffusivity. For example, for a Le < 1 the species diffusion is greater than the stabilizing 24
Colorado School of Mines
thermal diffusion and a wrinkled flame front will continue to grow and become unstable, increasing flame speeds. However for Le > 1, the flame may become more stable, depending on other environmental and mixture conditions, decreasing flame wrinkling and flame propagation velocities. The D-L instability phenomena can cause other instabilities such as the Rayleigh-Taylor (R-T) instability. The R-T instability occurs when there are large density differences along an interface between two fluids pushing against each other. For a flame, this means that the wrinkling becomes exaggerated, which can also lead to Kelvin-Helmholtz (K-H) instabilities. As the flame becomes more wrinkled, the rises and troughs become larger which means that the velocity differences between the burned and unburned gases are greater. This larger difference in velocity gives way to the K-H instability which typically results in the shedding of vortices along the velocity interface. One example of the impact of the D-L and R-T instabilities on flame propagation is the tulip flame phenomena as shown in Figure 2.10. This phenomena has been realized for many years in different experimental setups and different fuel-oxidizer mixtures (Ellis & Wheeler, 1928; Guenoche & Jouy, 1953; Starke & Roth, 1989) and has been a subject of CFD modeling (Bychkov, Akkerman, Fru, Petchenko, & Eriksson, 2007; Gonzalez, 1996; Gutkowski, 2013). Figure 2.10 Images of the development of a tulip flame in a stoichiometric mixture of methane and air. Refer to Section 4.3 for more details. Flame moves from left to right. CH = 9.5±0.3%. 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole at the closed end. ign One of the most recent, comprehensive experimental studies of the development of the tulip flame was performed by Clanet and Searby (1996). They used vertical cylindrical Pyrex tubes with open-closed end conditions and propane-air mixtures. The reactors had varying 25
Colorado School of Mines
diameters and lengths: 2.5-5cm diameter and 0.6-6m in length. The mixtures were ignited from the closed end of the reactor such that the flame propagated from the bottom to the top of the vertical tube. They discussed four main stages to the tulip development, which have also been observed in certain experiments performed by the CSM research group as shown in Figure 2.10. The first stage is a hemispherical expansion of the flame which can also be seen at t=10ms in Figure 2.10. At this point in time the flame is unaffected by the walls of the reactor. As the flame develops, it starts to approach the walls of the vessel and creates a finger shape seen at t=30ms. When the flame interacts with the cool wall, it quenches and rapidly loses heat to the walls as seen at t=40ms. When the flame loses heat to the walls, the surface area and velocity of the leading flame front decreases, leading to large density gradients. The large differences in density and velocity lead to the R-T instability and the flame front becomes inverted, i.e. the tulip flame seen at t=50ms (Clanet & Searby, 1996). Finally, in the fourth stage of the process, acoustic effects dominate, K-H instabilities, and the tulip flame is further distorted, sometimes producing multiple inversions along the flame front as shown at t=60ms. It is important to note that acoustics play a large role in the stability of a flame. Other researchers studying detonations have found the development of tulip flames due to pressure waves or shock wave interacting with the flame front, inverting it and increasing flame acceleration (Markstein, 1957; Salamandra, Bazhenova, & Naboko, 1959). Understanding the effects of interactions between flames and acoustics is extremely important, especially for the transition of a flame from deflagration to detonation. As shown in Figure 2.10, acoustics can grossly disturb the flame front and the fluid velocity in the unburned and burned mixtures. This is important because stretching of the flame front leads to greater differences in local fluid velocities along the flame front and increases combustion rates, leading to faster flames. Understanding flame instabilities and the interaction of acoustic waves and flames is extremely important for developing a combustion model of methane gas explosions in longwall coal mines. In a coal mine, there are many walls for pressure waves to interact with and reverberate off, which means that in the event of an explosion the pressure waves can compress to form a shock wave and potentially a detonation. Additionally, many reports of mine explosions have shown that ventilation controls were destroyed (McKinney, et al., 2001; Page, et al., 2011), likely from the leading pressure wave. There has also been anecdotal evidence from mine workers who have seen flames moving back and forth within the gob area (McKinney, et 26
Colorado School of Mines
al., 2001). This could also be due to the acoustic oscillation of the flame as it continues to burn methane emanating from the gob area. In general, however, one of the main purposes of this research is to help explain under what conditions flame instabilities and acoustic instabilities may exacerbate the propagation of a methane flame in a mine. 2.5 Impact of confinement on flame propagation The environment of an underground coal mine is very confined, which is why mine fires and explosions can be so devastating. As previously discussed, ignitions in confined spaces can result in higher temperatures, pressures, and fluid motion, thereby increasing combustion and the speed of the propagating flame and can lead to large overpressures and/or blast waves. The effect of degree of confinement, or venting, on flame propagation has been the topic of research in a variety of industrial processes (e.g. oil refineries, chemical plants). In the 1970s, Bradley & Mitcheson developed a simplified theory of venting for spherical reactors and compared the theory to a significant amount of experimental data from other researchers (Bradley & Mitcheson, 1978). The theory assumes isentropic compression and expansion, the flame is symmetric and expands spherically, and there are no interactions between the flame and pressure waves. Despite the fact that their model is simplified, many of their findings agree with other researchers. For example, they find that central ignition produces the largest overpressure, consistent with Fairweather, Hargrave, Ibrahim and Walker (1999) as well as Kindracki, Kobiera, Rarata, and Wolanski (2007). This is due to the fact that the hot exhaust gases must travel significantly further to vent from the reactor, which increases local temperatures and pressures, accelerating the flame away from the vent. Additionally, they find that as the area of the vent on one end of the vessel increases, the max overpressure decreases similar to Cooper, Fairweather, and Tite (1986) and Zhang and Ma (2015). Finally, they also investigated how the pressure of the vent itself (or weight of the vent) affects the overpressure. They found that as the vent pressure increases, the overpressure of the explosion increases in addition to the velocity of fluid ahead of the flame (Bradley & Mitcheson, 1978). This has been studied by many other researchers who further quantified the effect of vent relief pressure or weight of the vent (Bao, et al., 2016; Cooper, Fairweather, & Tite, 1986). However, Bradley & Mitcheson (1978) also discuss how their assumptions are simplified; noting that pressure-flame interactions must be incorporated into future venting theory. 27
Colorado School of Mines
Soon after the studies by Bradley and Mitcheson (1978), McCann, Thomas, and Edwards (1985) performed a study in two, small-scale cubical vessels investigating the effects of Taylor instabilities and other acoustic instabilities (Helmholtz) and how they affect propagating flames and explosion overpressure. They found that as the vent relief pressure increased, the speed of sound also increased as well as the frequency of the oscillations (McCann, Thomas, & Edwards, 1985). This finding is important because as the frequency of pressure oscillations increase, they may disturb the flame front even more, creating cellular structures. Solberg, Pappas, and Skramstad (1981) performed studies investigating the effects of instabilities for large scale explosions and they found that high frequency pressure oscillations are much more important in large scale explosions and the onset of cellularity occurs much sooner, leading to faster combustion and flame speeds which has been observed by other researchers (Bauwens, Chaffee, & Corofeev, 2008). Because of this, there has been much research on the elimnation of these high frequency oscillations and vanWingerden and Zeeuwen (1983) found that lining a vessel with glass wool helped “damp the acoustic wave and prevent the coupling between the acoustic wave and combustion to occur”. Though this may be a useful technique for some, whether or not a similar technique can be employed in a mine environment is unlikely. Figure 2.11 Example of a typical pressure-time profile of an explosion in a near cubic vessel with a single vent. Figure modified from: (Cooper, Fairweather, & Tite, 1986). Finally, one of the most fundamental studies on pressure generated by vented explosions was by Cooper, Fairweather, and Tite (1986) who performed experiments in a near cubic vessel, the results of which are reproduced in Figure 2.11. The first peak pressure rise, P , is the result of 1 28
Colorado School of Mines
the initial kernel expansion when the production of combustion gases exceeds the volume removed by venting . The second major pressure peak, P , occurs after initial venting when the 2 flame front ignites unburned gases increasing the rate of combustion and pressure rise in the vessel. After the second peak, Helmholtz modes are excited and the flame front undergoes bulk motion (Cooper, Fairweather, & Tite, 1986; McCann, Thomas, & Edwards, 1985). Figure 2.11 also shows two additional pressure peaks, P and P . After the Helmholtz oscillations, the flame 3 4 expands, turbulence and combustion increases, and large density gradients are created in the vessel which increases the pressure, P . Finally, as the production of burned gases decreases, the 3 pressure in the vessel decreases and couples “with the acoustic modes of the vessel” sustaining pressure oscillations creating a high frequency fourth peak, P (Cooper, Fairweather, & Tite, 4 1986; vanWingerden & Zeeuwen, 1983). In addition to evaluating the physical mechanisms behind each peak, Cooper, Fairweather, and Tite (1986) also discusses the practical implications of each peak, noting that the presence of obstacles can produce pressure peaks similar to P . This 3 is important, especially for explosions in a mine environment because pressure waves can destroy mine structures and equipment and potentially harm workers. Figure 2.12 Experimental setup used by Guo, Wang, Liu, and Chen (2017) exploring the impact of multiple vents on explosion overpressure. PT refers to the location of the pressure transducer. Image credit: (Guo, Wang, Liu, & Chen, 2017). Though much of the research discussed has only been in regards to a single vent and investigating the factors leading to more deadly explosions, there has been some research into the effects of multiple vents as shown in Figure 2.12. Results from Guo, Wang, Liu, and Chen (2017) show that multiple vents can slightly help reduce explosion overpressure and they can largely reduce the flame length of the explosion extending out of the vent. 29
Colorado School of Mines
In general, previous research has shown that compartment geometry, ignition location, vent size, venting pressure, and reactor scale can greatly affect the max overpressure, flame propagation velocity, and acoustic interaction with the flame. These are all important considerations for explosions in underground coal mines. Depending on the ignition location, whether it is in a passageway, in the gob, or behind the shields, the resulting explosion can be significantly different. To this end, it is the goal of this research to determine which aspects of a longwall mine environment has the greatest affect pressure generation and flame propagation. Thus, experiments are performed with and without obstacles, changing the ignition location and degree of confinement. Additionally, pressure traces can be correlated to flame propagation velocities which allows for a thorough understanding of possible flame dynamics in a mine environment. 2.6 Impact of obstacles on flame propagation The environment of a longwall coal mine has a variety of obstacles (i.e. mine equipment, workers, rock piles) as shown in Figure 1.2 and Figure 1.3. The complex geometries make it difficult to determine how a flame interacts with the obstacle since different mechanisms can be at play simultaneously. Since many explosions occur in or around the gob area, the CFD model developed in this work must be able to model methane gas explosions in confined spaces, under different environmental conditions, and with rock rubble and mine equipment/structures. There has been a significant number of studies on flame interaction with obstacles and porous media, varying different parameters such as number of obstacles, obstacle location, and blockage ratio (BR = Area unobstructed/Total cross sectional area) among some (Chapman & Wheeler, 1926; Dong, Bi, & Zhou, 2012; Dunn-Rankin & McCann, 2000; Evans, Schoen, & Miller, 1948; Ibrahim & Masri, 2001; Kindracki, Kobiera, Rarata, & Wolanski, 2007; Masri, Ibrahim, Nehzat, & Green, 2000; Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982; Xiaoping, Minggao, Wentao, Meng, & Juniie, 2015; Yibin, Fuquan, Xiaoyan, Xin, & Hongbin, 2011). This discussion shall detail some of the more notable studies on flame interaction with obstacles and in following, will discuss how this manuscript investigates the impact of obstacle parameters on methane flame propagation and explosion overpressure and how it differs from previous researchers. One of the first, most fundamental studies on flame propagation across an obstacle was performed by Chapman and Wheeler (1926). They performed methane-gas experiments in a horizontal, 5cm diameter brass cylinder, 240cm long open at both ends. The obstacles were brass 30
Colorado School of Mines
orifice plates (or annuli) that were 1mm thin. In the first part of their study they placed the orifice plates 40cm from ignition and varied the orifice diameter (3.6, 2.5, 1.5cm). When they compared flame speed measurements, they found that the flame was slightly retarded upstream of the obstacle, and then accelerated across the obstacle and downstream. Different orifice diameters also produced different accelerations; the greatest acceleration was found to be when the orifice diameter was half the tube diameter, meaning a 50% BR (Chapman & Wheeler, 1926). They also varied the location of the orifice plate and found as they moved it further from ignition the velocity recorded at the end of the reactor increased, however this was for a closed-closed reactor end condition. Varying the thickness of the brass orifice plate decreased the downstream velocity, though this could be due to the fact that brass has a high thermal conductivity and thus absorbed heat from the flame, slowing it down. Additionally, they added two or more restrictions in the tube and found even greater flame speeds than a single restriction. Decreasing the distance between the restrictions from 70 to 30cm accelerated the flame to even greater speeds. They even attempted to reach detonation velocities by adding 12 restrictions, but shattered the glass viewing windows, recording pressures upwards of 3.9atm (Chapman & Wheeler, 1926). Later, Robinson and Wheeler (1933) performed similar experiments in a slightly larger, steel tube, 30.5cm in diameter and 32.3m long with 11 steel orifice plates. They were unable to reach detonation velocities, but they did note a central flame core traveling through the restrictions and residual burning towards the wall between the obstacles (Robinson & Wheeler, 1933). This is important and there has been a significant area of research investigating how unburned gas pockets trapped between obstacles impacts flame acceleration (Ciccarelli and Dorofeev, 2008 (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982)). In 1982, Moen, Lee, Hjertager, Fuhre, and Eckhoff performed a large-scale experimental study on the effect of obstacle BR on flame speed and pressure generation, also looking at the mechanisms of burning pockets of mixture between obstacles and downstream of obstacles. They ignited stoichiometric methane-air mixtures in a cylindrical steel tube, 2.5m in diameter, 10m long with one end open and one end attached to an ignition tube and the mixtures were ignited using planar ignition. Some of their results have been reproduced here shown in Figure 2.13 and Figure 2.14. In summary, they found that even a single plate with a low BR can enhance methane overpressure, the value of which highly depends on the location of the obstacle relative to ignition. In most circumstances, increasing the BR increased overpressure and flame 31
Colorado School of Mines
In their theoretical description of the processes leading to enhanced overpressure and flame speeds due to obstacles, they noted that as the flame travels across the obstacle it results in a “nonuniform flow field” which “increased burning rate” and overpressure (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). In their discussion of this finding, they noted, similar to Chapman and Wheeler (1926) that “the rate of burning increases due to the larger flame surface area, the flame induced flow velocity also increases, creating stronger flow field gradients” generating turbulence and a feedback between the unburned pockets downstream of the obstacles and the propagating flame brush (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). This feedback loop between the unburned recirculating gases trapped between obstacles and the main flame brush is one of the main mechanisms of flame acceleration and has been observed and studied by other researchers as well (Ciccarelli & Dorofeev, 2008). Figure 2.14 Flame time of arrive versus distance from ignition for various blockage ratios (BR) and varying number of orifice plates. Figure credit: (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). 33
Colorado School of Mines
Figure 2.15 Schematic of turbulent flame propagation in an obstacle filled tube. Figure credit: (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). In general, the experimental results from Moen, Lee,Hjertager, Fuhre, & Eckhoff, 1982 showed that BR, the number of obstacles, obstacle spacing, and obstacle location can have a huge effect on methane flame propagation and overpressure. Their results laid a strong foundation for other researchers including Lee, Knystautas, and Chan (1985), Phylaktou and Andrews (1991), and Fairweather, Hargrave, Ibrahim, and Walker (1999). Lee, Knystautas, and Chan (1985) performed experiments in different steel cylinders ranging 5-30cm in diameter and 11-17m in length using a variety of fuels at different stoichiometries (including methane and air). They determined there were four major propagation regimes. The first regime was the quenching regime, in which the obstacle BR were large enough to initially accelerate the flame, but eventually the flame was quenched due to heat loss to the cool unburned mixture entrained in the turbulent flame brush (Lee, Knystautas, & Chan, 1985). The second regime was the choking regime and the flame accelerates the entire length of the reactor, noting that eventually the flame may reach the local speed of sound (Lee, Knystautas, & Chan, 1985). The third regime is the quasi-detonation regime, where flame speeds are so fast and the orifice diameter is large enough to transition to detonation which is dependent on the detonation cell size and reactor/obstacle geometry (Lee, Knystautas, & Chan, 1985). Finally, the fourth regime is the C-J detonation regime, and again the size of the orifice to the detonation cell size must be greater than several times the cell size, which is dependent on the fuel mixture (Lee, Knystautas, & Chan, 1985). These findings and discussion are extremely important for understanding the potential for deflagration transition to detonation. Other researchers have also experimented with obstacles that are staggered, in the center of flow, and only on one side of the flow (Xiaoping, Minggao, Wentao, Meng, & Juniie, 2015). 34
Colorado School of Mines
They found that obstacles that are staggered or central to flow produce significantly higher overpressures and the flame propagation itself is more tortuous. Understanding the potential for a methane mixture to transition to detonation via obstacle induced turbulence is the subject of much research in the coal mining industry (Oran, Gamezo and Kessler, 2011 (Zipf, et al., 2013)). Although certain experimental setups and models have shown that methane gas mixtures can reach detonation velocities, it is still unclear from the studies discussed above, whether or not this is possible during a real mine explosion. For example, in the UBB mine explosion in 2010, the pressure waves entrained coal dust on the floor of the mine, which transitioned the methane explosion to a methane-coal dust explosion resulting in 29 deaths and significant damage to miles of mine structures and equipment (Page, et al., 2011). Because of this, there has also been some research on understanding how coal dust can exacerbate a methane gas explosion and to determine better techniques for depositing inert rock dust over coal dust (Bai, Gong, Liu, Chen, & Niu, 2011; Dong, Bi, & Zhou, 2012; Sapko, Weiss, Cashdollar, & Zlochower, 2000). In addition to using solid orifice plate type obstacles, researchers have also investigated the effects of a grid, different shapes (cylinders, rectangles, etc), and porous media. Evans, Schoen, and Miller (1948) looked at the effect of copper grids on propane-air flame phenomena in pyrex tubes, varying in cross sectional shape and lengths. They found that grids accelerated the flame and a “grid flame” consisted of “large cells or globules at the front, and fine-grained eddies behind the front” as compared to a flame propagating in an unobstructed tube (Evans, Schoen, & Miller, 1948). Stark and Roth (1989) investigated the development of a tulip flame in a tube obstructed with a grid. They found that as the grid was moved further from ignition, the tulip inversion continued to be affected by the obstacle until a certain distance and the inversion was suppressed (Starke & Roth, 1989). There have also been several studies investigating the effect of different solid shaped obstacles on flame propagation (Ibrahim & Masri, 2001; Masri, Ibrahim, Nehzat, & Green, 2000; Yibin, Fuquan, Xiaoyan, Xin, & Hongbin, 2011). Researchers looked at the effects of cylinders, squares, diamonds, triangles, and wall/plates on flame and pressure enhancement and found that changing the blockage ratio of the rectangular or plate type obstacles had the largest effect on flame propagation velocity and overpressure (Ibrahim & Masri, 2001; Masri, Ibrahim, Nehzat, & Green, 2000). Interestingly Masri, Ibrahim, Nehzat, and Green (2000) found significant 35
Colorado School of Mines
differences in the size and length of vortices produced downstream of the different obstacles as shown in Figure 2.16, though they were unable to draw any significant conclusions about the downstream vortices effects. In a second paper, however, results of the overpressure versus BR in Figure 2.17 seem to suggest that off-spherical objects (wall/plate) increases the overpressure of an explosion (Ibrahim & Masri, 2001). This is of significant interest for this research because some coal mine explosions have originated near or within the gob, e.g. the Willow Creek explosion in 2000, and so it will be important to understand how rock, which has a varying degrees of sphericity, impacts flame propagation. Up to this point, the CSM group has been unable to find other researchers investigating flame propagation across a rock pile. Due to the lack of experiments, the main focus of this research is to perform experiments with idealized rock and actual rock typically found in a mine to determine whether these shapes change flame acceleration mechanisms. Figure 2.16 Vortex pair behind a triangular obstacle. Image credit: (Masri, Ibrahim, Nehzat, & Green, 2000). Figure 2.17 Overpressure versus blockage ratio for different obstacle geometries. Figure credit: (Ibrahim & Masri, 2001). 36
Colorado School of Mines
Finally, it has been shown that obstacles, confinement, and ignition location can influence explosion overpressure and flame propagation. In a real longwall coal mine, the explosion can occur from the gob area (confined) and propagate towards the face or entries. Therefore, one of the main areas of concern of this research is flame propagation between obstacles and this particular research has only been able to find one paper were the ignition was between obstacles by vanWingerden and Zeeuwen, (1983). In their study, they had a square, wooden plate with cylindrical sticks mounted on it. The obstacle was located in the middle of a larger vessel such that the flame could propagate in all directions (similar to a spherical bomb). Experiments were performed with ignition located 1) on the top of a single plate, 2) centered between two plates, 3) on a single plate with sticks, and 4) between two plates with sticks supported between them. For all fuels tested, including methane, the fastest flame velocity was in configuration 4) and produced an overpressure twice that found for a stagnant mixture (van Wingerden & Zeeuwen, 1983). This is of no surprise since the flame was not only confined, but accelerated due to the obstacles. However, it must be noted that the flame propagated in all directions, there was no principle direction like would be the case for a mine explosion where the flame might tend to travel in the open passageways. Therefore, one of the objectives of this research is to further understand the differences with flame propagation between obstacles in addition to across obstacles. Although most of the obstacles typically found in a mine are solid obstacles, the gob area, walls, and rock on the belt has different levels of compaction and it is unclear whether they can be treated as a porous media. Thus, one of the major focuses of this research is whether or not the gob and other such obstacles act similar to a porous media. Because of this, it is important to understand some of the differences in flame propagation mechanisms between completely solid obstacles and porous media. Some of the most fundamental studies on porous media were performed by Babkin, Korzhavin, and Bunev (1991) and Howell, Hall, and Ellzey (1996). Babkin, Korzhavin, and Bunev (1991) performed experiments in square reactors filled with different porous media including steel polished balls, polyurethane foam, porous material made of foam, and aluminum glued combs. They show that methane flames are much faster at higher initial pressures and pore cavity size. However, it is unclear whether the thermal properties of the material had an effect. Additionally, they show that porous media can accelerate flames “as effectively as in rough tubes, or in tubes with periodical obstacles such as spirals” 37
Colorado School of Mines
(Babkin, Korzhavin, & Bunev, 1991). This is extremely important because the walls of a mine consist of rock rubble and as such will have a significant effect on flame propagation. Another notable finding of Babkin, Korzhavin, and Bunev (1991) and more fully explained by Howell, Hall, and Ellzey (1996) is the feedback mechanism between the flame and porous media. As the flame propagates through the porous media, pockets of unburned mixture are trapped and then heated by radiation and conduction. These pockets are entrained by the passing flame, thus accelerating the flame in a feed-back loop (Howell, Hall, & Ellzey, 1996). Babkin, Korzhavin, and Bunev (1991) also note that “the most probable stabilizing factor” is quenching of the flame (Babkin, Korzhavin, & Bunev, 1991). Further studies by Ciccarelli, Hlouschko, Johansen, Karnesky, and Shepherd (2009) experimented with a layer of 12.7mm diameter ceramic-oxide beads along the entire bottom of a horizontal, rectangular reactor investigating how bead layer affects the transition to DDT. They helped confirm that the interaction of the flame with the bead layer “drives the flame acceleration in the gap until DDT” (Ciccarelli, Hlouschko, Johansen, Karnesky, & Shepherd, 2009). Also they found reducing the height of the gap above the bead layer resulted in a faster acceleration to DDT because the trapped gases in the bead layer have a stronger coupling to the main flow and may act similar to a piston (Ciccarelli, Hlouschko, Johansen, Karnesky, & Shepherd, 2009). Fig (2019) performed similar experiments to Cicarelli, et al. (2009) in different diameter reactors using layers of rock. In these experiments, rock rubble was arranged in a non-reacting metal cage and inserted into a reactor at different locations. The length and height of the rock pile were explored and it was found that lining the entire length of the reactor resulted in the greatest flame acceleration for large vessels. This research aims at taking a step back and using spheres as an idealized rock rubble. The spheres take away any different in flame propagation due to surface topology so that this research can determine which property of the rock pile has the largest impact on flame acceleration. 2.7 Impact of spark energy on methane flame propagation There are many sources of ignition in a longwall coal mine such as rock-on-rock friction, metal-on-rock friction (from a carbide cutting tips on the longwall shearer), spontaneous combustion, welding components, or any electrical communication system/remote. Frictional ignition from rocks and metal-on-rock friction are two of the most common methods of ignition 38
Colorado School of Mines
(Page, et al., 2011) and are difficult to quantify since the energy is highly dependent on the type of rock, the speed of the rock or machine, and contact point among some. This research aims to understand how these different types of energy and energy durations may affect methane gas deflagrations using the experimentally validated CFD model. In 1979, Maly and Vogel wrote a paper describing the three discharge modes of an electric spark, detailing the properties of each mode and where the majority of losses come from. The three discharge modes are the breakdown phase, arc phase, and glow discharge phase as shown in Figure 2.18. The energy supplied by the spark system during the breakdown phase ionizes the gas between the electrodes, creating a plasma. Approximately 94% of the energy during the breakdown phase is transferred to the plasma which helps provide the “conductive path between the electrodes necessary” for the arc and glow discharge phases (Maly & Vogel, 1979). During the arc and glow discharge phases, a significant amount of energy is lost through conduction to the electrodes and there is less dissociation of species (Maly & Vogel, 1979). Understanding these processes is important because it has helped researchers design more efficient methods of spark ignition, such as the plasma spark plug which is designed to transfer the majority of the energy during the breakdown phase. Figure 2.18 Schematic of voltage and current versus time of a typical spark ignition system. U refers to the energy in unit volts (V). Figure credit: (Maly & Vogel, 1979). 39
Colorado School of Mines
Most studies of premixed combustion employ spark ignition, especially for determining the minimum ignition energy required to ignite mixtures (Blanc, Guest, von Elbe, & Lewis, 1947; Shmelev, 2009). However, there are several parameters of the design of the spark ignition system which can affect measurements of minimum ignition energy, including the electrode material, distance between the electrodes, and the shape of the electrodes, whether they are plates, hemispherical, or to a point. In general, most researchers agree that for a stoichiometric mixture of methane and air at 1atm and 300K, the minimum ignition energy is approximately 0.3-0.5mJ (Shmelev, 2009; Turns, 2012). As discussed earlier, the minimum ignition energy is highly dependent on pressure, temperature, and mixture stoichiometry. In majority of the studies on the effects of spark ignition on flame propagation, researchers note that the ignition energy only affects the initial flame kernel development until a critical radius (Lintin & Wooding, 1958). To note, many of these researchers used small test chambers or spherical bombs, making it difficult to determine if ignition energy does not affect flame propagation in other apparatus (Blanc, Guest, von Elbe, & Lewis, 1947; Lintin & Wooding, 1958). However, some researchers have found that for lean limits and high spark energies, ignition may affect the subsequent flame propagation, suggesting that researchers “perform experiments with several ignition energies to determine conditions at which the flame is not affected” (Lawes, Sharpe, Tripathi, & Cracknell, 2016). Additionally, other researchers have found that a planar ignition of match heads produces overpressures and flame speeds faster than a single point ignition (Hjertager, Fuhre, & Bjorkhaug, 1988). Though the ignition source is a match head, it still lends to the point that the distribution of energy and quantity of energy could affect flame propagation. Therefore, it is the goal of this research to determine whether varying spark ignition energy and duration can affect methane gas deflagrations, especially for confined spaces. 2.8 Modeling methane flames There are three major ways to model complex fluid flow problems: direct numerical simulation (DNS), large eddy simulation (LES), and moment models. As the name implies, DNS is the most accurate method of simulating fluid flow because it directly solves the Naiver-Stokes equations. Unfortunately, the major drawback to DNS is that it requires a significant amount of computational time and energy due to the wide range of temporal and spatial scales of most flows, which makes it less desirable as compared to LES and moment models. LES models 40
Colorado School of Mines
resolve the large-scale turbulent structures directly by solving the filtered Navier-Stokes equations and models the smaller scales (sub-grid scales – SGS). In general, LES models are good for turbulent flows that are characterized by large-scale eddies, allowing for a much coarser grid. Finally, the third type of models, which are the most widely used, are moment models that solve the fluid flow equations using statistical analysis. The most common method of averaging the Naiver-Stokes equations is by using ensemble averages which solve for mean velocities producing a set of equations called Reynolds-Averaged Navier-Stokes equations (RANS) and uses 1st order closure models to obtain a complete set of equations. Many researchers over the decades have developed various models of premixed methane- air (or other mixtures) deflagrations and detonations using DNS (Gonzalez, 1996; Hawkes & Chen, 2004), LES (Bi, Dong, & Zhou, 2012; Chen, Luo, Sun, & Lv, 2017; Di Sarli, Di Benedetto, & Russo, 2012; Xiao, Makarov, Sun, & Mokov, 2012), RANS (Fairweather, Ibrahim, Jaggers, & Walker, 1996; Fairweather, Hargrave, Ibrahim, & Walker, 1999; Fig, Bogin, Brune, & Grubb, 2016; Jerome, Christophe, & Guillaume, 2017; Kozubkova, Krutil, & Nevrly, 2014; Wang, Ma, Shen, & Guo, 2013), and other reduced order models (Sezer, Kronz, Akkerman, & Rangwala, 2017). Some researchers have written in-house codes (Catlin, Fairweather, & Ibrahim, 1995; Fairweather, Ibrahim, Jaggers, & Walker, 1996) while most use commercial softwares such as Fluent (Gutkowski, 2013; Kozubkova, Krutil, & Nevrly, 2014) or AutoReaGas (Hong, Lin, & Zhu, 2016; Zhang & Ma, 2015). In general, most researchers are concerned with accurately solving the flame time of arrival and the overpressure of the explosions for gas explosions with and without obstacles. To accurately model these phenomena, the majority of researchers performed their own experiments to help validate their model (Dunn-Rankin & McCann, 2000; Fairweather, Hargrave, Ibrahim, & Walker, 1999; Jerome, Christophe, & Guillaume, 2017; Kozubkova, Krutil, & Nevrly, 2014; Yu, Sun, & Wu, 2002), but there are many researchers who rely on experiments from other researchers as points of model validation (Catlin, Fairweather, & Ibrahim, 1995; Di Sarli, Di Benedetto, & Russo, 2012; Hong, Lin, & Zhu, 2016; Li & Hao, 2017; Oran, Gamezo, & Kessler, 2011; Salvado, Tavares, Teixeira-Dias, & Cardoso, 2017; Sezer, Kronz, Akkerman, & Rangwala, 2017; Valiev, Bychkov, Akkerman, Law, & Eriksson, 2010). To build a comprehensive 2D & 3D combustion model of a longwall coal mine methane gas explosion, it is important to be able to capture the effects of obstacles on flame propagation 41
Colorado School of Mines
and pressure enhancement. Figure 2.19 is a summary of the numerical simulations of premixed combustion which model obstacles versus those which do not (Note this is not meant to be an entire review of all flame experiments ever performed, but a summary of those discussed in this manuscript). As can be seen, there are less researchers modeling obstacles, though it is becoming much more common in recent years, especially with the improvements made to commercial codes, allowing for much easier meshing around complex objects and coupling of more complex chemical reactions. However, most research groups modeling obstacles consider only a couple of obstacles of simple geometry as shown in Table 2.1. This research not only models a larger variety of obstacles, but validates the model across a much wider range of reactor volumes. Fairweather, et al. 1996 & 1999 have modeled premixed combustion and interaction with obstacles and validated the model across different reactor sizes (Fairweather, Ibrahim, Jaggers, & Walker, 1996; Fairweather, Hargrave, Ibrahim, & Walker, 1999). However, their code was developed in-house which makes it more difficult to replicate their settings. Additionally, although they have validated their model with a handful of reactors, the range is still limited when it is compared to all other experimental combustion researchers, Figure 2.20 and Figure 2.21.. This is extremely important because flame dynamics, especially across obstacles, do not scale linearly with reactor size (Fig, Bogin, Brune, & Grubb, 2016). This also helps demonstrate how novel and important this research is to the combustion community; modeling the effects of obstacles across a wide range of scales and validating the model with experiments is a small percentage of the total number of experimental and numerical combustion research. It is also important to note that this research was the first to couple a combustion model with a ventilation model of a longwall coal mine and model a methane gas explosion (Bogin, 2015). Although there are still improvements to be made, the coupled model developed in this research is not only applicable to the mining industry, but to other small- and large-scale industrial explosions or even to flame propagation through a building. 42
Colorado School of Mines
Health Administration (MSHA). Although the number of incidents and fatalities from coal mine disasters have declined over the years due to better safe practices and equipment enhancements, underground coal mine methane gas explosions still pose a risk to miners and equipment. Figure 2.22 Number of coal mine disasters and fatalities from 1900-2016. Figure credit: (Center for Disease Control and Prevention, 2017). Methane gas explosions have several mechanisms of blast injury which affect the human body including primary, secondary, tertiary, and quaternary effects. Primary injuries are injuries from the initial blast shock wave interacting with the human body. The most common primary injuries from coal mine explosions are blast lung, eardrum/middle ear rupture and concussion, but other injuries may include eye rupture or abdominal hemorrhage (Centers for Disease Control and Prevention, 2017; Institute of Medicine, 2014). As the blast wave travels through the body and blood, it can cause air to expand in the blood, creating air embolus which in many cases, end up in the lungs (blast lung) (Institute of Medicine, 2014). If the victims do not seek medical attention right away including a chest x-ray, they may not know they have blast lung until it is too late. Secondary injuries may result from airborne debris that is thrown from the explosion, which may result in blunt force trauma or perforation of the body (Centers for Disease 45
Colorado School of Mines
Control and Prevention, 2017; Institute of Medicine, 2014). Tertiary injuries are a result of individuals being through from the blast and include bone fractures and head injuries (Centers for Disease Control and Prevention, 2017; Institute of Medicine, 2014). For example, in the Willow Creek mine explosion of 2010, several miners close to the second explosion were thrown and one miner lost their cap lamp (McKinney, et al., 2001). Finally, quaternary injuries are those which are not a direct result of the blast, but are indirect including burns or inhalation of toxic fumes such as smoke, dust, or carbon monoxide (Centers for Disease Control and Prevention, 2017). Again, in the Willow Creek mine explosion one miner “was asphyxiated as a consequence of carbon monoxide poisoning” and another miner was “seriously burned and received a massive head injury” (McKinney, et al., 2001). However, in addition to injuries to miners, blast effects of methane gas explosions can damage/destroy mine structures, damage mine equipment, and reverse airflow in the mine (Brune, 2014; McKinney, et al., 2001; Page, et al., 2011). For example, pressures between 0.1- 5psi can shatter single-strength glass and pressures between 1-2psi can crack plaster walls or buckle sheet steel (Owen-Smith, 1981). Also, pressures between 2-3psi can crack cinder-block or concrete block walls and upwards of 8psi they can crack brick wall (Owen-Smith, 1981). These pressure examples are important because the demonstrate the overpressure effects on common mine materials which the pressures from the UBB explosion in 2010 easily exceed (Page, et al., 2011). Also important is the pressure wave fluctuations which may reverse airflow in the mine or possibly temporarily stop ventilation fans which can be dangerous as CO levels increase from the explosion. In summary, there are many different types of hazards from a methane gas explosion in an underground longwall coal mine. Although advancements throughout the years have helped reduce the number of fatalities from these explosions, they still pose a risk in active longwall coal mines. It is the goal of this research to better predict these hazards and understand the potential for human loss and mine damages in order to build stronger mitigation strategies for improved worker safety. 46
Colorado School of Mines
CHAPTER 3 DESIGN OF EXPERIMENTAL COMBUSTION REACTORS Due to safety hazards of performing methane-air explosions in a mine, this research performs experiments in both large- and small-scale reactors of varying diameter and length. The large-scale reactor is 71cm in diameter, 6.1m in length (L/D=8.5, Volume = 2.4m3) and is located at Edgar Experimental Mine in Idaho Springs, CO. The benefit of performing experiments in the large-scale reactor is that the resulting methane flame speeds and pressure rise are more indicative of those found in a longwall coal mine explosion. However, this reactor requires large amounts of gases which increases cost and requires significant amounts of time to set up and perform experiments. Therefore, small-scale experiments are also performed to provide additional insights across a wide range of conditions. There are several small-scale reactors including 1) 5cm diameter, 43cm long (L/D=8.5), 2) 5cm diameter, 1.5m long (L/D=30) steel reactor 3) 9.5cm diameter, 81cm long (L/D=8.5), steel reactor, a 4) 12cm diameter, 1.5m long (L/D=12) quartz reactor, 5) 13.6cm diameter, 1.5m long (L/D=11) quartz reactor, and a 6) 30.5cm diameter, 1.15m long (L/D=8.5) steel reactor. Additionally, a small experimental box, 51x34x15cm (LxWxH), was setup to explore the impact of reactor shape on methane flame propagation and interaction with a simulated gob. The main advantage of performing experiments at different scales is it allows researchers to understand the scalability of methane flame properties which is important for validating the combustion model before incorporation into the mine-scale ventilation model. The scalability of methane flame behavior is one of the major objectives of M. Fig’s research (Fig, 2019) and will be presented in part of this dissertation. This document will present the work performed in the 12cm diameter quartz flow reactor, 71cm diameter large-scale reactor, and laboratory-scale experimental box. The main advantage of the quartz reactor and experimental box is that they allow full visualization of the flame as it interacts with rock rubble and other obstacles. This is important because it provides further insight into methane flame dynamics and allows researchers to use imaging to validate the CFD models. 47
Colorado School of Mines
3.1 Laboratory-Scale Experimental Setup and Procedure The laboratory-scale experimental system consists of industrial-grade compressed methane and zero-grade air, mass flow controllers, a mixing tank, flame arrestors, a reactor, ion sensors, a pressure transducer, an ignition system, and data acquisition system as shown in Figure 3.1. Figure 3.1 Flow diagram of laboratory experimental setup. NI USB refers to a National Instruments USB used for recording data. Before any experiments are performed, compressed building air flows through all lines to check for any possible leaks. After all safety checks have been performed, the experimental process begins by flowing methane and air at specified rates for a desired stoichiometry. The flow is controlled by mass flow controllers (MFC); a Bronkhorst EL-FLOW® Select (0-50 SLM) controls the air and an Alicat Scientific MC Series (0-5 SLPM) controls the methane. The methane and air mix enter a mixing chamber, which is filled with turbulence inducing media, until homogeneity is reached. Before the mixture flows into the quartz reactor, an aluminum foil cover is placed over the open end of the reactor. A small perforation is made in the foil so that 48
Colorado School of Mines
while filling the reactor is purged of the ambient air by the premixed methane-air and provides an undiluted mixture without increasing the overall pressure in the reactor. Finally, the premix flows into the quartz reactors at 30psi for 5 minutes which is equivalent to approximately 2-3 volume fills depending on the reactor size. Homogeneity of the mixture inside the reactor is verified using an Infrared Industries IR-6000 gas analyzer and mixture stoichiometry is confirmed using a gas chromatograph with thermal conductivity detector (GC-TCD). At the closed end of the reactor, where the premix inlet is located, there is no stratification in the radial direction. Along the top of the reactor, the methane-air mixture did not vary in concentration; near the open end of the reactor the mixture varies in the radial direction by less than 0.2% (within error of the MFCs). After the reactor is completely filled with the methane-air mixture, the mixture is allowed to settle for 40 seconds to insure stagnant conditions. The mixture is then ignited using a capacitive spark ignition system as described in Section 3.3, providing approximately 60±5mJ of energy to the system. After each experiment, compressed building air flows through the reactor to 1) cool down the reactor to ambient conditions and 2) to flush out any remaining combustion products. During the shutdown procedure all lines are purged to the laboratory exhaust system and compressed building air flows through the lines to ensure no residual methane or premix is left in the system. Note all experiments are performed at 294±1K and 83±1kPa, which are ambient atmospheric conditions for Golden, CO – elevation approximately 1,730m above sea level. The DAQ used in this system consists of a National Instruments (NI) USB-6008 and USB-6009 DAQ board, both capable of sampling at 48,000 samples per second. The ion sensors are each wired to a single input on the USB-6008 board and a Kistler © 4260 piezoresistive pressure transducer is wired separately to the USB-6009 board. Additionally, high-speed imaging is taken using a GoPro Hero4 capable of sampling at 240 frames per second at 720 pixel resolution. In order to ensure repeatability, each experimental set consists of 4-5 experimental runs. The flame front propagation velocities reported are the average of all experimental runs and the error bars represent the standard deviation of the mean of the set. The overpressure traces shown are a single run from an experimental set and are meant to show the general trend of the explosion overpressure. The maximum overpressure reported is always the average of all the experimental runs and the error bars represent the standard deviation of the mean of the set. 49
Colorado School of Mines
Finally, there are several safety design features in the laboratory-scale experimental setup. The compressed methane and zero-grade air are housed in an exhaust cabinet. High-pressure tubing is routed from the regulators to the electronic solenoid valves as shown in Figure 3.1. The solenoid valves are normally closed so that in the event of an emergency, all flows are stopped. Additionally, these valves are wired to an emergency stop button so that the entire system can be shut down. Downstream of the solenoid valves, stainless steel tubing is used and can withstand pressures up to 35MPa which is significantly greater than pressures used in this system. Upstream of the mixing chamber is a pressure relief valve that is set to relieve at pressures greater than 35kPa and evacuate the gases to the laboratory exhaust system in case of an accidental overpressure. Downstream of the mixing tank is a 3-way electronic solenoid valve, which is wired to the operator’s table and either fills the reactor or purges the premix into a high- pressure line vented in the exhaust cabinet. Two flame arrestors are employed; one flame arrestor is located in the reactor fill line, upstream of the premix inlet and a second flame arrestor is located in the purge line. Additionally, exhaust hoods connected to a point exhaust system (150 cfm) are located on either end of the quartz reactor and a methane sensor is located near the open end and can measure concentrations less than 25% of the lower explosive limit (LEL) of methane. Another safety design feature is the Plexiglas enclosure around the quartz reactor which 1) protects users in the case of cracking or breaking of the quartz and 2) keeps any leaking premix inside the system so that it can be exhausted from either the open or closed end of the system. Lastly, there are three fire extinguishers placed throughout the laboratory; one is near the operator’s table, a second near the exhaust cabinet housing the compressed methane and zero- grade air, and a third near the laboratory exit. Safety is the number one concern of this lab and the measures described have been evaluated by the Environmental Health and Safety office on campus and exceed requirements. Additionally, the Standard Operating Procedures (SOPs) for operating the gas flame tubes and spark ignitions systems have been created, signed, included in a binder in the lab, and taped to the operator’s table and gas cabinet. 3.1.1 Design of Quartz Flow Reactor The horizontal, cylindrical quartz flame rector used in this study has an inner diameter of 12cm, 0.25cm wall thickness, and length of 1.5m as shown in Figure 3.2 and Figure 3.4. One end of the quartz reactor is open to atmosphere and the other end is closed. The premix fuel-air inlet and pressure transducer are mounted on the closed end of the reactor as shown in Figure 3.3, 50
Colorado School of Mines
Figure 3.4 Image of the 12cm diameter quartz flow reactor laboratory setup. There are several techniques to measure the propagation velocity including IR/UV sensors, pressure sensors, ion sensors, and high-speed imaging and many researchers use a combination of sensors in order to check the reproducibility of an experiment. For example, Fairweather, et al. (1999) and Robinson and Wheeler (1933) used imaging to determine flame time of arrival. Many other researchers, including Hjertager, Fuhre, and Bjorkhaug (1988), Ciccarelli et al. (2009), and Moen, et al. (1982) used ion sensors to measure the flame speed. This research uses high-speed imaging and ion sensors made from copper wire as shown in Figure 3.5. The electrodes are housed in a ceramic tube and placed into the sensor ports such that the electrodes are flush with the top of the reactor. The sensor ports are 0.8cm in diameter and the ceramic tubes are 0.5cm in diameter. Therefore, silicone was wrapped around the ceramic tubes to 1) hold the sensors in place and 2) not allow any premix or combustion gases to escape out of the top of the tube. The IR gas analyzer was used to verify no gases escape from the tube via the sensor ports. The ion sensors were wired to a voltage source and resistor network as shown in Figure 3.6. As the flame passes across the electrodes, the ions trigger a voltage drop across the resistor which is recorded by the DAQ system. There are three main components to the design of the ion sensors: the wire, the resistor, the voltage source. A study was done using 18 gage copper wire of varying lengths and compared the results to using coaxial cable which has significantly more insulation. The additional insulation on the coaxial cable reduced the standard deviation of the recordings, but did not impact the results. 52
Colorado School of Mines
Figure 3.5 Image of custom-made copper ion sensors (left) and example of raw output from ion sensors (right). Figure 3.6 Ion sensor circuit. The vertical location of the ion sensor is extremely important and by performing experiments in the 12cm diameter quartz reactor researchers took high-speed imaging to determine the actual location of the flame front so that reported flame front propagation velocities are accurate. As can be seen in Figure 3.7, when ignition is 11cm from the open end the flame travels towards the closed end and the flame front is at the top of the reactor. Therefore, for all experiments when ignition is from the open end the ion sensors are placed at the top of the reactor. However, when ignition is 11cm from the closed end, the flame shape is much different as shown in Figure 3.8. Results of testing the location of the ion sensor in the vertical direction for ignition from the closed end are shown in Figure 3.9. As can be seen in this figure, when the sensors are placed at the top of the reactor, the flame front propagation 53
Colorado School of Mines
velocities recorded seem step-wise. This is because the flame expands more rapidly in the axial direction than in the radial direction. Thus, by placing the sensors further down into the reactor, at 1cm as shown in Figure 3.8, the recorded flame front propagation velocities are more accurate and the standard deviation of the mean is reduced significantly. In conclusion, for all closed-end ignition experiments the ion sensors are placed 1cm down from the top of the reactor. This study was extremely important because researchers are unable to visualize the flame shape in all the other explosion reactors, including the 71cm diameter large-scale reactor at Edgar Experimental Mine. Figure 3.7 Flame front shape when ignition is at the open end of the 12cm diameter quartz reactor. Flame propagates from the open end to the closed end (left to right). CH = 9.5±0.3%. 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign Figure 3.8 Flame front shape when ignition is at the closed end of the 12cm diameter quartz reactor. Flame propagates from the closed end to the open end (right to left). Yellow line represents the top of the quartz reactor. CH = 9.5±0.3%. Operating conditions 294±1K, 4 83±1kPa. E =60±5mJ. One (1) relief hole. ign 54
Colorado School of Mines
Figure 3.9 Impact of vertical location of ion sensor on methane flame front propagation velocity for a closed-end ignition. Ignition location: 1.39cm from the open end. Flame travels towards open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief 4 ign hole. 3.1.2 Design of Experimental Box Most of the experimental setups conducted for this research have been horizontal cylindrical reactors to understand the impact of mine conditions on the bulk flow of a high-speed methane gas explosion. However, in a real mine environment the entryways, crosscuts, longwall face, etc. are rectangular in shape. To begin to understand how the reactor shape and multiple pathways affect methane gas flame propagation velocities an experimental box was setup as shown in Figure 3.10. The experimental box is made of steel and has a Plexiglas covering on top in order to take high-speed imaging of the flame during experiments. The box is 51x34x15cm (LxWxH) and in the center of the box is a simulated porous gob (porous medium) consisting of 2 layers of lava rock, average size 6.1±0.2cm x 4.4±0.2cm x 3.2±0.2cm, and 2 upper layers of river rock, average size 6.1±0.3cm x 4.2±0.2cm x 2.7±0.1cm, arranged in a non-reacting metal cage that is 28x22x15cm (LxWxH). Although lava and river rock are not commonly found in a mine environment, they were chosen due to ease of availability and due to the fact they have similar thermal properties of rock in underground mines (thermal conductivity is between 1.5-5 W/m- K), along with irregular shapes meant to represent the complex pathways between the rocks in a gob. 55
Colorado School of Mines
Figure 3.10 Image of experimental box setup with a simulated gob (porous medium) consisting of lava rock and river rock. As can be seen in Figure 3.10, a small opening 7x12cm (HxW) was made in the bottom- left corner of the box which allowed for explosion relief. The experimental procedure is the same as for the quartz reactor described earlier in this section. In summary, the box is filled with premixed methane-air mixture and allowed to settle before ignition using a capacitive discharge electric spark providing 60±5mJ of energy. Note that the spark electrodes were located at a height of H=7.5cm from the bottom of the box (total height is H=15cm). In this experimental box setup, ignition was initiated near the opening (open-end ignition at x=0.06m, y=-0.14m) in the bottom-left corner of the box and conversely, opposite the opening (closed-end ignition at x=0.45m, y=0.14m) in the top-left corner of the box. High-speed imaging was taken using a GoPro Hero4 capable of sampling at 240 frames per second and 720 pixel resolution and average flame speeds and standard error of the mean were estimated using video recordings. All experiments were performed using a methane-air mixture stoichiometry of 9.5±0.3% by volume. Experiments are performed at 293±1K and 83±1kPa, which are ambient atmospheric conditions for Golden, CO. 3.2 Design of Large-Scale Experimental Reactor The large-scale experimental system consists of industrial-grade compressed methane, zero-grade air, industrial grade compressed nitrogen, mass flow controllers, NEMA 7 explosion proof solenoid valves, a mixing tank, flame arrestors, a reactor, ion sensors, pressure transducers, an ignition system, and data acquisition system as shown in Figure 3.11, Figure 3.12, and Figure 56
Colorado School of Mines
Figure 3.13: Image of the closed end of the 71cm diameter, 6.1m long steel reactor showing the mixing tank, pressure relief valve, premix access ports, flame arrestors, and sensor ports and wiring. Before beginning the experimental process, an aluminum foil cover is placed over the open end of the reactor as shown in Figure 3.13 and a steel grid is placed over the end to suppress the flame exiting the reactor. Additionally, the reactor is angled towards a rock burm for safety concerns and to help redirect noise away from the city of Idaho Springs. A small perforation is made in the foil so that while filling the premix is enclosed in the reactor without increasing the overall pressure in the reactor. Similar to the laboratory-scale reactor system, the experimental process begins by flowing methane and air at specified rates for a desired stoichiometry. The flow is controlled by mass flow controllers (MFC); a Bronkhorst EL- FLOW® Select (0-50 SLM) controls the methane and a MKS 1559A controls the air and nitrogen. The methane and air are delivered at 70psi and enter a mixing tank, which includes a pressure relief valve and a NEMA 7 explosion proof solenoid valve connected to a purge line for safety. The mixture then flows into two lines on which are flame arrestors are attached in case of flashback. Finally, the premix then flows into the reactors for at least 45 minutes which is equivalent to approximately 1.5-2 volume fills. Homogeneity of the mixture inside the reactor is 58
Colorado School of Mines
verified using an Infrared Industries IR-6000 gas analyzer on site. After the reactor is completely filled with the methane-air mixture, the mixture is allowed to settle for 30 seconds to insure stagnant conditions. During this settle time, air flows through all the lines and through the gas mixture tank, purging to the outside to ensure no premix is inside the system. The mixture inside the reactor is then ignited using a capacitive spark ignition system as described in Section 3.3, providing approximately 60±5mJ of energy to the system. Note that the spark electrodes are located in the first sensor port, which is 28.5cm from the closed end. During the shutdown procedure all lines are vented to the outside and compressed air flows through the lines to ensure no residual methane or premix is left in the system. Note all experiments are performed at 295±1K and 79±1kPa, which are ambient atmospheric conditions for Idaho Springs, CO. The DAQ used in this system consists of a NI CDAQ and NI USB-6009 DAQ board. The ion sensors are wired in parallel to the NI CDAQ and two Kistler © 4260 piezoresistive pressure transducers are wired separately to the USB-6009 board. The ion sensors are located in the center of the reactor and reported flame front propagation velocities are the average of all experimental runs and the error bars represent the standard error of the mean of the set. The overpressure traces shown are a single run from an experimental set and are meant to show the general trend of the explosion overpressure. The maximum overpressure reported is always the average of all the experimental runs and the error bars represent the standard deviation of the mean of the set. Finally, there are several safety design features in the large-scale experimental setup. The compressed gases and flow controllers are housed away from the reactor and are wired underground in a PVC pipe back to the control center. This allows the operator to run the entire system from the control center, in which a fire extinguisher is located. High-pressure tubing is routed from the regulators to the NEMA 7 rated electronic solenoid valves and the solenoid valves are normally closed so that in the event of an emergency, all flows are stopped. Downstream of the solenoid valves, stainless steel tubing is used and can withstand pressures up to 35MPa which is significantly greater than pressures used in this system. Upstream of the mixing tank is a pressure relief valve which vents to the surrounding atmosphere. Downstream of the mixing tank is a NEMA 7 rated electronic solenoid valve, which is wired to the operator’s table and either fills the reactor or purges the premix into a high-pressure line vented to the outside. Two flame arrestors are employed downstream of the mixing tank before the premix enters the reactor. Safety is the number one concern of this research and the measures described 59
Colorado School of Mines
have been evaluated by the Environmental Health and Safety office on campus and exceed requirements. 3.3 Design of Spark Ignition System For this manuscript, a spark ignition was be employed since it is one of the most common methods of igniting a combustible mixture due to its reliability, ease of setup, and ease of implementation (Turns, 2012). The current spark ignition system consists of a voltage source, a 1Ohm resistor, a 2μF capacitor, a manual switch, and single ignition coil as shown in Figure 3.14. To produce a spark, the manual switch is activated so that the current builds up in the capacitor, storing the spark energy. After the switch is deactivated, the energy in the capacitor is released and a magnetic field breakdown occurs between the capacitor and ignition coil. The ignition coil acts similar to a transformer and steps up the voltage across the spark electrodes. After reaching a sufficient voltage breakdown across the electrodes, which is dependent on the mixture composition and stoichiometry, a spark is produced and ignition is initiated. Figure 3.14 Schematic of spark ignition system. The primary side includes the 12V battery, 1Ohm resistor, fuse, capacitor, and manual switch. The secondary side includes the output from the ignition coil and spark electrodes. It is important to note that typically the battery, resistor, switch, and capacitor are called the primary side. This primary side is low voltage (12V), but has a high current (5A). The ignition coil has primary and secondary windings with a typical turns ratio of almost 100:1. The 60
Colorado School of Mines
low voltage and high current from the primary side flow through the primary windings and a magnetic field builds up between the primary and secondary windings, such that during the spark discharge process, the low voltage is stepped up to 10,000V or more and the current is reduced to milliamps. To estimate the amount of spark energy, it is assumed that the amount of energy stored in the capacitor is completely transferred to the spark. In reality, some of the stored energy is dissipated during the dielectric breakdown across the electrodes and other processes such as heat dissipation to the electrodes. The amount of energy stored in the capacitor can be estimated by the product of the primary current (i ) and voltage source (V), Equation (3.1): p (3.1) The measured primary spark current is 5.2 ± 0.2A and measured battery voltage is approximately 12 ± 0.5V which corresponds to spark energy of 62.4 ± 1.7mJ; however, over time this may fluctuate slightly and researchers have recorded energies 60 ± 5mJ. The spark current is measured using an oscilloscope and the voltage is measured using a voltmeter. In summary, a spark ignition system capable of handling currents upwards of 30A, secondary voltages in excess of 20,000V, and spark durations between 0.5-6.5ms has been designed. The configuration of the spark ignition system is shown in Figure 3.14 includes a 12V battery, 1Ohm resistor, 2μF capacitor, a 30A fuse upstream of the ignition coil, and an ignition coil, supplying approximately 60 ± 5mJ of energy through the spark. Safety features of the system include no. 10AWG multicore copper wiring, a 30A fuse, a separate plastic box for the ignition coil in case of overheating, a handlebar switch on the positive junction to the battery, a secondary switch to activate the system, a thermometer in case components overheat, and a large plastic container with a Plexiglas lid to allow visualization of components. After each experimental session, the circuit is disconnected from the battery and ignition coil and all components are cooled down to ambient. 61
Colorado School of Mines
CHAPTER 4 EXPERIMENTAL INVESTIGATION OF METHANE FLAME PROPAGATION AND PRESSURE GENERATION Previous related experimental research was performed with piles of rock rubble to help simulate a real longwall coal gob as shown in Figure 4.1 (Fig, Strebinger, Bogin, & Brune, 2018; Strebinger, et al., 2017). Results from these experiments have shown that for all reactors and across all scales the rock pile, meant to represent fallen rock rubble in the gob, accelerated the methane flame for all stoichiometries investigated as shown in Figure 4.2. Additionally, increasing the length of the rock pile further accelerated the flame and had the most significant increase for the largest reactor. However, as can be seen from Figure 4.1, the rock piles have varying rock material, rock orientation, rock size, rock pile porosity, void spacing between rocks, and void location. These variabilities make it difficult to determine what gob parameters had the greatest impact on increasing the flame front propagation velocities. Therefore, one of the main objectives of this research is to gain a strong, fundamental understanding of the impact of different gob parameters on methane flame propagation. Figure 4.1 Image of methane flame passing across rocks in the 13.6cm diameter quartz reactor (top) and 71cm diameter steel reactor at Edgar Mine (bottom). Image credit: (Strebinger, et al., 2017). 62
Colorado School of Mines
Figure 4.2 Impact of rock pile length in the 13.6, 30.5 and 71cm diameter reactors. The barrier lengths were varied from 3 to 7% of the overall reactor length. The impact of the length of the barrier is larger for larger reactors than for smaller reactors. Ignition from the open end of the reactor. Operating conditions 294K, 83kPa. Figure credit: (Fig, 2019). 4.1 Empty 12cm Diameter Reactor: Impact of mixture stoichiometry on open- and closed- end ignition Methane can emanate from various sources in the longwall coal mine, including the gob area, face, walls, etc. among some (Karacan, Ruiz, Cote, & Phipps, 2011). Previous research has shown that the ventilation air can leak into the gob area (Krog, Schatzel, & Dougherty, 2014), and can mix with methane, creating EGZs with different stoichiometries (Gilmore, et al., 2016; Juganda, Brune, Bogin, Grubb, & Lolon, 2017). Therefore, to understand the impact of stoichiometry on methane flame propagation for model validation, researchers investigated lean, stoichiometric, and rich mixtures of methane and air. To obtain a base case for the 12cm diameter quartz reactor, experiments were performed igniting different methane-air mixture volume fractions from either the open end of the reactor (open-end ignition) or the closed end of the reactor (closed-end ignition). In the case of open-end ignition, the mixture is ignited 11cm from the open end of the reactor, centered in the radial direction, and the flame travels from the open end to the closed end. For closed-end ignition, the mixture is ignited 1.39m from the open end of the reactor (11cm from the closed end of the reactor) and the flame travels from the closed end to the open end. The stoichiometry, or volume 63
Colorado School of Mines
fraction, of the methane-air mixture was confirmed using a GC-TCD and the volume fraction reported is within ±0.3%, which is limited by the accuracy of the mass flow controllers. Note that one of the relief holes (D,=1±0.2cm,A=1.13cm2) is open on the closed end of the reactor, which helps in flame stability. Figure 4.3 Impact of mixture stoichiometry on methane flame front propagation velocity versus distance for open-end ignition. Ignition location is 11cm from the open end. Reported mixture volume fractions are within ±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One ign (1) relief hole. Each data point is the average of 5 data points. Standard deviation range is between 1-15% of the mean. Figure 4.4 Impact of mixture stoichiometry on methane flame front propagation velocity versus time for open-end ignition. Ignition location is 11cm from the open end. Reported mixture volume fractions are within ±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One ign (1) relief hole. Each data point is the average of 5 data points. Standard deviation range is between 1-15% of the mean. 64
Colorado School of Mines
Results show for open-end ignition the stoichiometric flame, 9.5% methane by volume, produced the fastest flame front propagation velocities, followed by the lean flame, 7.5%, and the rich flame, 11.5%. According to theory, the stoichiometric flame should produce the fastest burning velocities, followed by the rich and then lean flame (Turns, 2012). However, in this experimental setup, the rich flame produces a diffusion flame which burns on the open end of the reactor. This is important because the burning of the diffusion flame acts as a counterbalance to the main flame front traveling towards the closed end, thus retarding the rich flame. The lean flame does not produce a diffusion flame and is able to propagate freely down the reactor. Also to note, for all cases the flame front propagation velocity reaches a peak just over half way down the reactor. At this point in time, the pressure resistance due to compressed unburned gases on the closed end of the reactor begin to heat up and increase in pressure, pushing back against the propagating flame. Figure 4.5 Images of a stoichiometric methane flame traveling from the open end of the 12cm diameter quartz reactor towards the closed end. Flame travels from left to right. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. 4 E =60±5mJ. One (1) relief hole. ign Interestingly, the resulting methane flame from an open-end ignition has a unique shape as shown in Figure 4.5. Initially the flame expands spherically and after the onset of venting of exhaust gases out the open end of the reactor, the hot, buoyant exhaust gases near the reaction front rise to the top of the reactor. As the hot product gases rise, they push over the top of the flame and the resulting traveling flame has a unique angled shape. This has also been observed by other researchers (Ellis & Wheeler, 1928; Guenoche & Jouy, 1953). 65
Colorado School of Mines
Figure 4.6 Impact of mixture stoichiometry on methane flame front propagation velocity versus time for closed-end ignition. Ignition location is 1.39m from the open end. Reported mixture volume fractions are within ±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One ign (1) relief hole. Each data point is the average of 5 data points. Standard deviation range is between 0-11% of the mean. Figure 4.7 Impact of mixture stoichiometry on methane flame front propagation velocity versus time for closed-end ignition. Ignition location is 1.39m from the open end. Reported mixture volume fractions are within ±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One ign (1) relief hole. Each data point is the average of 5 data points. Standard deviation range is between 0-11% of the mean. When the ignition location is moved from the open end of the reactor to the closed end of the reactor, the stoichiometric mixture produced the fastest flame propagation velocities, followed by the rich and lean flame as shown in Figure 4.6. However, the magnitude of the velocities produced are significantly different with movement of the spark location. A closed-end 66
Colorado School of Mines
ignition results in flame propagation velocity magnitudes almost 50 times greater than an open- end ignition. This is because a closed-end ignition is an ignition from a confined space which means the hot exhaust gases are not able to flow out of the reactor as easily. The increase confinement increases the temperature and overpressure of the explosion in addition to increasing fluid motion ahead of the flame with little to no pressure resistance leading to faster flame propagation velocities. As can be seen in Figure 4.6, the flame propagation velocities accelerate fairly linearly, but near the open end of the reactor they begin to fall off. This roll-over effect was investigated with the 5cm diameter reactor and have observed that as the length of the reactor gets longer, the roll-over point moves further as well which helps show that the roll-over is affected by the open end condition (Fig, 2019). Figure 4.8: Images of a stoichiometric methane flame traveling from the closed end of the 12cm diameter quartz reactor towards the open end. Flame travels from right to left. Ignition location is 1.39m from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. 4 E =60±5mJ. One (1) relief hole. ign In addition to being quantitatively different then an open-end ignition flame, the closed- end ignition flame shape is also qualitatively different as shown in Figure 4.8. Because of the fast expansion of the flame, the buoyant exhaust gases do not rise to the top of the reactor as quickly as the flame expands axially. The shape of the flame is often referred to as “finger shape” and has been observed by many other researchers (Clanet & Searby, 1996; Ellis & Wheeler, 1928). 67
Colorado School of Mines
Figure 4.9 Pressure-time history of a closed-end ignition (CEI) versus open-end ignition (OEI) with no obstacle. P (OEI, 1 experiment) = 0.26kPa. P (CEI, 5 experiments) = 3.24±0.15kPa. max max CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. 4 ign Figure 4.10 Example of a typical pressure-time profile of an explosion in a near cubic vessel with a single vent. Modified from: (Cooper, Fairweather, & Tite, 1986). As has been discussed in Chapter 2, Section 2.5, the degree of confinement plays a major role in the propagation of methane flames, specifically in the amount of pressure rise of the explosion. Figure 4.9 helps to demonstrate the large overpressure produced by an explosion from the closed-end of the reactor. As can be seen by both pressure traces, there are two major pressure peaks followed by decaying oscillations. These researchers have not observed the third pressure peak described by Cooper, Fairweather, and Tite (1986) and have only observed the fourth pressure peak in some of the experiments presented in this manuscript. 68
Colorado School of Mines
Figure 4.11 Pressure-time history of closed-end ignition with varying mixture stoichiometry. P (CH = 9.5±0.3%) = 3.24±0.15kPa. P (CH = 7.5±0.3%) = 1.56±0.06kPa. P (CH = max 4 max 4 max 4 11.5±0.3%) = 1.39±0.38kPa. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) ign relief hole. Figure 4.11 shows the pressure-time history of a closed-end ignition for various methane- air mixture concentrations. As expected the 9.5% (black) stoichiometric mixture produced the largest pressure rise, P , which agrees with other researchers (Bao, et al., 2016; Kindracki, 2 Kobiera, Rarata, & Wolanski, 2007). In this case, peak P is approximately twice that of the lean 2 (red) or rich (blue) cases, respectively, which helps to show how dangerous explosions can be if the flammable mixture is stoichiometric. 4.2 Impact of gob factors on methane flame propagation in the 12cm diameter quartz reactor Methane gas explosions in longwall coal mines may occur in different areas of the mine and can interact with a variety of obstacles including rock rubble, mine structures, and mine workers. It is well known that many of these explosions occur in confined spaces and can potentially occur in the gob area directly behind the shields and propagate towards the working face. Therefore, it is extremely important to understand how various gob parameters such as rock material, void spacing and location, gob porosity, and rock orientation can affect methane flame propagation. To study these effects, several simulated gob inserts of varying geometry and materials have been made to explore the impact of these parameters as shown in Figure 4.12 and Figure 4.13. 69
Colorado School of Mines
Figure 4.12 Images of simulated gob materials. In order to study the effect of gob parameters on methane flame acceleration, this research uses solid, smooth spheres of varying material and size and compares the results to granite rock of similar size. Smooth spheres were used for several reasons: 1) spheres take away any differences in results due to surface topology 2) spheres are easy to model for combustion model validation and 3) spheres can be easily arranged in a variety of geometries. Glass was used since it has a low thermal conductivity of 1W/m-K similar to common rock types found in mines: sandstone (1.7W/m-K) and limestone (1.3W/m-K). The glass spheres will be compared to granite rock (2.85W/m-K) of averaged similar size. The effects of simulated gob material on methane flame propagation will be discussed in Section 4.3.2. Figure 4.13 Schematic of simulated gob-wall geometries. A non-reacting metal cage 12cm in diameter and 2.54cm in length was used to control the orientation of the spheres so that researchers may investigate different simulated gob-wall geometries shown in Figure 4.13. The cage geometry serves as a base case to understand the effect of the wall and checkerboard geometries. Spheres were arranged in a checkerboard geometry in order to study the effects of porosity on methane flame propagation. Adding or removing spheres from the checkerboard geometry is equivalent to changing the porosity of the 70
Colorado School of Mines
simulated gob geometry without changing the effective void spacing. The wall geometry was used to study the effects of void spacing on methane flame propagation. By changing the height of the wall, the effective void spacing changes. The wall geometry can also be oriented in different ways to study the effect of void spacing at the top of the reactor versus the bottom of the reactor. Although a void at the bottom of the reactor may not be perfectly analogous to a void in a longwall coal mine gob, it is still important to investigate for potential future modeling. Finally, these gob inserts were located at different distances from the open end of the reactor to test how the relative location of the obstacle to ignition affects methane flame front propagation velocity. All experimental results in subsequent sections are performed at 9.5±0.3%% methane by volume since Section 4.1 showed a stoichiometric mixture produces the most explosive flame of the stoichiometries investigated. 4.2.1 Impact of void spacing and void location As shown in Figure 1.3 on page 4, the rock rubble in the gob can be varying sizes, shapes, and have different void spacings. Therefore, the main goal of these experiments was to understand the impact of void spacing, or BR, on methane flame front propagation. To isolate the impact of the simulated gob geometries on methane flame propagation, the cage was inserted into the reactor 37cm from the open end, between the first two ion sensors. In Figure 4.14, the mixture was ignited from the open end of the reactor and results show flame acceleration across the cage. This acceleration is due to the fact that the cage induces some fluid movement in the nearby unburned gases which accelerates combustion as the flame passes across the cage. Downstream of the cage, the effect of induced fluid motion is lessened. In the next set of experiments, researchers investigated how void spacing of a single simulated gob wall impacts methane flame propagation for both an open-end ignition and closed- end ignition. Figure 4.15 shows the results of an open-end ignition when the simulated gob wall is located at 37cm from the open end and the void space decreases from 73% (H=3.8cm) to 13% (H=9.8cm). Results show that the simulated gob wall with 73% void space (H=3.8cm) had no significant effect on methane flame front propagation velocity across the obstacle. However, when the void spacing was decreased to 13% (H=9.8cm) there is an enhancement of methane flame front propagation velocity across the obstacle. This enhancement is due to flame stretching around the obstacle, increasing the surface area of the flame, thereby increasing combustion 71
Colorado School of Mines
rates. For the taller wall, H=9.8cm, the flame surface area increases significantly more to move around the wall, versus the smaller wall, H=3.8cm. The taller wall, H=9.8cm, also had more consistent flame propagation run-to-run compared to the wall with H=3.8cm which shows larger deviations from the mean and a slight slow down downstream of the wall. Figure 4.14 Impact of cage on methane flame front propagation velocity. Obstacle = Cage. Obstacle location=37cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data point is the ign average of 5 data points. Standard deviation range is between 2-8% of the mean. Figure 4.15 Impact of obstacle height (or amount of void space) on methane flame front propagation velocity for an open-end ignition. Obstacles = 6.35mm diameter glass spheres in a wall geometry, L=6.35mm, H = 3.8cm, 9.8cm. Obstacle location=37cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. 4 E =60±5mJ. One (1) relief hole. Each data point is the average of 5 data points. Standard ign deviation range is between 2-14% of the mean. 72
Colorado School of Mines
Although the 3.8cm tall simulated gob wall (73% void space) did not have as a significant impact on an open-end ignition as the 9.8cm tall simulated gob wall, it greatly accelerated the flame for a closed-end ignition. This is because a closed-end ignition results in a turbulent flame that when passing over an obstacle wall results in flow separation as shown in Figure 4.17. This flow separation is important because it forms eddies on the downstream side of the wall, which increases temperatures and fluid motion promoting flame acceleration, agreeing with observations made by other researchers (Chapman & Wheeler, 1926) (Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). For an open-end ignition the maximum overpressure rise is 0.26kPa as shown in Figure 4.9 and does not significantly affect the unburned mixture upstream of the flame. Thus, when the flame passes over the obstacle wall, the flame continues to burn radially and tangentially as shown by the arrows in Figure 4.17. Figure 4.16 Impact of amount of void space on methane flame front propagation velocity for a closed-end ignition. Obstacles = Cage and 6.35mm diameter glass spheres in a wall geometry, L=6.35mm, H = 3.8cm. Obstacle location=37cm. Ignition location is 1.39cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. 4 ign Each data point is the average of 5 data points. Standard deviation range is between 1-4% of the mean. In addition to accelerating the methane flame across the simulated gob wall, the 3.8cm high gob wall (73% void space) also resulted in a, likely, reflected pressure wave whose peak is greater than the empty reactor and cage obstacle as shown in Figure 4.18. Based on the speeds of the flame, it is most probable this reflected pressure wave occurred after the flame exited the reactor. What is also important is that after the reflected pressure wave from the obstacle wall, greater pressure oscillations were sustained while residual methane-air mixture is burned even 73
Colorado School of Mines
though the main flame front has exited the reactor. In summary, these results help demonstrate that obstacles can increase the maximum overpressure, sustain pressure oscillations, increases secondary burn duration and percentage of burned methane-air mixture, therefore increasing the total heat release during an explosion. This is important because large overpressures can knock out ventilations controls, damage mine equipment, and cause harm to workers (Brune, 2014; McKinney, et al., 2001; Page, et al., 2011). Additionally, the large positive and negative pressure oscillations induced by the obstacle wall may entrain air, which in a real mine situation, could sustain a methane fire or reverse ventilation airflow. Figure 4.17 Image of methane flame propagating across a simulated gob wall for an open-end ignition (left) and a closed-end ignition (right). Obstacle: 6.35mm diameter glass spheres in a wall geometry, L=6.35mm, H = 3.8cm. Obstacle location=37cm. CH = 9.5±0.3%. Operating 4 conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign Figure 4.18 Pressure-time history of closed-end ignition with and without an obstacle. Obstacle: Cage and 6.35mm diameter glass spheres in a wall geometry, L=6.35mm, H=3.8cm. Obstacle location = 37cm. P (Empty) = 3.24±0.15kPa. P (Cage) = 2.98±0.50kPa. P (Wall, max max max H=3.8cm) = 5.24±0.45kPa. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief ign hole. 74
Colorado School of Mines
Finally, after investigating the impact of void spacing on OEI an CEI, researchers were interested in examining the impact of a longer void space, meaning increasing the obstacle length. Figure 4.19 shows the results of increasing the simulated gob wall length from 6.35mm to 12.7mm. Results show that by increasing the length of the reduced void spacing further accelerates the flame. This is because the flame must travel through a longer void, which means the flame front is being stretched even more, increasing combustion rates and unburned gases to the flame front. Also to note, the standard deviation range for these experiments was a bit larger than previous experiments because of older spark ignition circuitry. Since then, improvements have been made to the circuitry. Figure 4.19 Impact of obstacle length on methane flame front propagation velocity for an open- end ignition. Obstacles = 6.35mm diameter glass spheres in a wall geometry, H = 7.62cm, L=6.35mm or 12.7mm. Obstacle location=37cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. 4 ign Each data point is the average of 4-8 data points. Standard deviation range is between 2-13% of the mean. Thus far all experiments considering the simulated gob wall have been performed with the void at the top of the reactor. However, in a reality, the void spacing can be distributed in different locations and although a bottom void is unlikely, a fundamental understanding is still important. Therefore in this next experiment, the simulated gob wall void was tested at both the top and bottom of the reactor. Results from Figure 4.20 show that the location of the void spacing can significantly affect methane flame propagation. In this experimental setup, the flame front is at the top of the reactor, so when the void location was at the bottom of the reactor, the 75
Colorado School of Mines
flame had to travel around the obstacle as shown in Figure 4.21. After the flame has propagated through the bottom void, the flame front propagation velocity is much slower than the top void. This is because, as can be seen in c) of Figure 4.21, the flame is more concave and has not fully developed into the elongated angled flame typically observed with a top void. Since the flame that has passed through the bottom void has less flame area, the downstream propagation velocities are less than the flame which traveled through the top void. As the bottom void flame continues to develop, it accelerates down the length of the reactor until reaching the same velocity as the flame which passed through the top void. Figure 4.20 Impact of void location on methane flame front propagation velocity for an open-end ignition. Obstacle: 6.35mm diameter glass spheres in a wall geometry L=12.7mm, H=9.8cm. Obstacle location=37cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data point is the ign average of 3-4 data points. Standard deviation range is between 1-18% of the mean. Figure 4.21 Images of methane flame propagation around an obstacle wall with a bottom void. Obstacle: 6.35mm diameter glass spheres in a wall geometry L=12.7mm, H=9.8cm. Obstacle location=37cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating 4 conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign 76
Colorado School of Mines
4.2.2 Impact of porosity A longwall coal mine gob consists of varying types of rock rubble which depends on the geological location and during the mining process, leftover rock rubble collapses and is compacted by the roof above. The gob itself is a confined space and there is no access to the majority of the gob and limited visual access to rock rubble near the gob fringes. Due to the limited access, it is difficult to exactly determine the size, shape, and porosity of the gob. Despite this challenge, there have been some researchers who have used tracer gas studies to determine airflow leakage into the gob (Krog, Schatzel, & Dougherty, 2014) and others have used photo- analysis to get a rock size distribution (Pappas & Mark, 1993a). Because of this challenge and lack of experimental evidence, researchers have turned to CFD modeling to determine porosities and permeabilities of the gob (Marts, et al., 2014; Ren & Edwards, 2000; Tanguturi, Balusu, & Bongani, 2017; Yuan, Smith, & Brune, 2000). In general, it has been found that near the edge of the gob, especially behind the longwall shields, the rock rubble is less compacted with porosities near 40%, versus the center of the gob which is much more compacted with porosities near 15% (Marts, et al., 2014). Since gob porosities near areas of potential explosion risks, i.e. fringe zones and behind the longwall shields, can be upwards of 40%, this research is interested in the impact of small changes in porosity on methane flame propagation. To explore this experimentally, simulated gob-wall checkerboard geometries were made using 6.35mm diameter glass spheres with 77% porosity and 67% porosity as shown in Figure 4.22. The simulated gobs were located 37cm from the open end of the reactor and results show that even a small decrease in porosity can significantly retard the methane flame across the gob (Figure 4.23). These results are important because in a real longwall coal mine there are different levels of compaction in different areas of the gob where explosions may occur. Understanding the effects of porosity of methane flames is key to developing a physically accurate combustion model. Although only a single obstacle was used in these experiments, after model validation, simulations can be run to investigate the impact of multiple checkerboard obstacles with less porosity, which would be more representative of the gob. 77
Colorado School of Mines
Figure 4.22 Schematic of checkerboard geometry with 77% porosity (left) and 67% porosity (right). Figure 4.23 Impact of obstacle porosity on methane flame front propagation velocity for an open-end ignition. Obstacle: 6.35mm diameter glass spheres in a checkerboard geometry L=6.35mm, Porosity=67% or 77%. Obstacle location=37cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief 4 ign hole. Each data point is the average of 4 data points. Standard deviation range is between 1-7% of the mean. 4.2.3 Impact of simulated gob location Experiments presented thus far have only shown results when the simulated gob is located 37cm from the open end. However, in a real methane gas explosion, the location of obstacles (and ignition) is unknown. Thus, to build a comprehensive model capable of predicting these explosions requires knowledge of the impact of obstacle location relative to ignition. To do this experimentally, the cage and wall obstacles were tested at 37cm, 62cm, and 87cm from the open end of the reactor. The results in Figure 4.24 and Figure 4.25 show that the location of the simulated gob-wall (or obstacle) relative to ignition location can significantly impact results. For all locations tested, the cage and wall resulted in acceleration across the obstacle. However, as the cage and wall were moved further from the ignition location the effect of acceleration across the obstacle was less pronounced. This is due to the fact that at 78
Colorado School of Mines
approximately 50cm from the open end the methane flame is almost fully developed. Additionally, the pressure resistance built up on the obstacle increases and suppresses the impact of local fluid motion enhanced by the obstacle. Because of these effects, the impact of acceleration across the obstacle decreases further from ignition. Figure 4.24: Impact of obstacle location on methane flame front propagation velocity for an open-end ignition. Obstacle: Cage. Obstacle location=37cm, 62cm, 87cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. 4 E =60±5mJ. One (1) relief hole. Each data point is the average of 4 data points. Standard ign deviation range is between 1-12% of the mean. Figure 4.25: Impact of obstacle location on methane flame front propagation velocity for an open-end ignition. Obstacle: 6.35mm diameter spheres in a wall geometry, H=9.8cm, L=6.35mm. Obstacle location=37cm, 62cm, 87cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. 4 ign Each data point is the average of 4-5 data points. Standard deviation range is between 0-8% of the mean. 79
Colorado School of Mines
Finally, comparing the results from Figure 4.24 with the empty cage versus the obstacle wall in Figure 4.25, the average velocities for a given experiment are almost the same. However, as has been shown, the cage has much more open void space than the wall, which begs the question as to why the obstacle wall does not produce faster average velocities than the cage. This difference is due to the fact that these obstacles are very thin, 6.35mm; because the main flame acceleration mechanism in these cases is flame stretching, the amount of distortion of the flame traveling across the thin cage is similar to traveling through the void space above the thin wall. 4.2.4 Impact of simulated gob bed There are many different types of obstacles in a longwall coal mine and thus far a single obstacle has been studied. This section aims to gain a better understanding of how a rock pile or porous pile can affect methane flame propagation and explosion overpressure. To this end, the obstacle used in these studies is a simulated gob bed consisting of 1cm diameter glass spheres as shown in Figure 4.26. Similar to the simulated gob wall, researchers varied the height and length of the simulated gob bed to simulate varying rock piles in a longwall coal mine. Figure 4.26 Images of glass spheres and an example of simulated gob bed. Both 15cm and 30cm lengths are tested with 1cm and 2cm heights. The experiments in Figure 4.27 use a simulated gob bed that is 30cm in length and 1cm in height (1 layer of 1cm diameter glass spheres) and has a void spacing of 96%. The simulated gob bed was located 11cm from the open end directly under the ignition location and then it was moved further down the reactor to investigate how location of the rock pile affects open-end ignition. Results show that when the simulated gob bed is closer to the point of ignition it accelerates the flame across the pile more than when it is located further from ignition. This is because during the initial kernel expansion of the flame there is a small pressure rise, as shown in 80
Colorado School of Mines
Figure 4.9 on page 68, which creates some movement in the nearby gases, promoting unburned gases to the flame front. The gob bed acts similarly to a porous media and the small movement in the gases is accentuated by the spheres which continues to enhance methane flame propagation as described by other researchers studying porous media (Babkin, Korzhavin, & Bunev, 1991). When the simulated gob bed is located further from ignition, at 44cm, there is little-to-no flame enhancement because the pressure resistance of the gob counteracts the fluid motion induced by the obstacle. Figure 4.27 Impact of simulated gob bed location on methane flame front propagation velocity for an open-end ignition. Obstacle: 1cm diameter glass sphere bed, L=30cm, H=1cm. Obstacle location=11cm, 44cm. Ignition location is 11cm from the open end. CH = 9.5±0.3%. Operating 4 conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data point is the average of ign 5 data points. Standard deviation range is between 1-6% of the mean. In contrast, when the ignition location is moved to the closed end of the reactor, the simulated gob bed located at 44cm enhances the flame propagation whereas the gob bed at 11cm had little-to-no impact (Figure 4.28). This acceleration is due to the fact that when the methane flame interacts with the simulated gob bed it creates a turbulent boundary layer as shown in Figure 4.29; this turbulent boundary layer was observed for all cases tested, but this figure is only showing images from an experiment with a L=15cm, H=2cm gob for visualization purposes. Fluid motion and combustion rates are increased outside of this boundary layer, which continue to enhance methane flame propagation such that the flame front across the simulated gob bed moves faster. For the simulated gob bed at 11cm, no flame enhancement was recorded, however this may mainly be due to the fact that there was only one ion sensor above the bed. 81
Colorado School of Mines
Figure 4.28 Impact of simulated gob bed location on methane flame front propagation velocity for a closed-end ignition. Obstacle: 1cm diameter glass sphere bed, L=30cm, H=1cm. Obstacle location=11cm and 44cm from open end. Ignition location is 1.39cm from the open end. CH = 4 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data ign point is the average of 5 data points. Standard deviation range is between 1-10% of the mean. Figure 4.29: Image of the methane flame before (t=42ms) it encounters a simulated gob bed 15cm in length and 2cm in height (Void space = 89%) and image of methane flame as it interacts (t=0.46ms) with the simulated gob bed during a closed-end ignition. Flame moves from right to left. Ignition location is 1.39cm from the open end. CH = 9.5±0.3%. Operating conditions 4 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign Additional experiments investigating how simulated gob bed length affect methane flame propagation have been performed and similar to the single obstacle, increasing the gob bed length greatly increases methane flame front propagation velocity and pressure rise as shown in Figure 4.30 and Table 4.1. Increasing the height of the simulated gob bed decreases the amount of void spacing from 96% to 89% and even this small change increases the maximum flame front propagation velocity from 79m/s to 82m/s (Figure 4.31,Table 4.1). Also, increasing the height of 82
Colorado School of Mines
the simulated gob bed increased the reflected pressure wave and sustained larger oscillations in the overpressure of the explosion as shown in Figure 4.32. This finding is important because larger pressure oscillations could possibly reverse airflow in a mine causing continual burning of an explosive gas mixture and/or damage ventilation controls. Figure 4.30 Impact of simulated gob bed length on methane flame front propagation velocity for a closed-end ignition. Obstacle: 1cm diameter glass sphere bed, L=15cm and 30cm, H=1cm. Obstacle location is 44cm from open end. Ignition location is 1.39cm from the open end. CH = 4 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data ign point is the average of 5 data points. Standard deviation range is between 1-12% of the mean. Figure 4.31 Impact of simulated gob bed height on methane flame front propagation velocity for a closed-end ignition. Obstacle: 1cm diameter glass sphere bed, L=15cm, H=1cm and 2cm. Obstacle location is44cm from open end. Ignition location is 1.39cm from the open end. CH = 4 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data ign point is the average of 5 data points. Standard deviation range is between 1-9% of the mean. 83
Colorado School of Mines
Figure 4.32: Impact of simulated gob bed height on the pressure-time history of a methane-gas closed-end ignition. Obstacle: 1cm diameter glass sphere bed, L=15cm, H=1cm and 2cm. Obstacle location is44cm from open end. Ignition location is 1.39cm from the open end. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign Table 4.1 Table summarizing the maximum flame front propagation velocity, maximum overpressure, and minimum overpressure recorded for the closed-end ignition experiments with and without different obstacles. Averages of 5 experimental runs. CH = 9.5±0.3%. Operating 4 conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole: ign Maximum Flame Maximum Minimum Obstacle Front Overpressure Overpressure Conditions Propagation (kPa) (kPa) Velocity (cm/s) Empty 6500 ± 270 3.24 ± 0.15 -2.45 ± 0.60 Wall L=6.35mm,H=3.8cm 8100 ± 360 5.24 ± 0.45 -2.56 ± 0.46 Bed L=15cm,H=1cm 7900 ± 340 4.56 ± 0.71 -3.13 ± 0.88 Bed L=30cm,H=1cm 8400 ± 720 4.79 ± 0.55 -3.61 ± 1.10 Bed L=15cm,H=2cm 8200 ± 650 6.23 ± 0.87 -5.74 ± 0.45 Bed L=30cm,H=2cm 8600 ± 630 5.52 ± 0.96 -5.50 ± 1.20 Finally, Table 4.1 summarizes the impact of a single simulated gob wall and simulated gob bed on methane flame propagation. In summary all simulated gob conditions increased methane flame propagation velocity and increased the overpressure of the explosion. Interesting to note is that a simulated gob bed with height 2cm, void spacing 89%, produced similar or greater flame propagation velocities and overpressures than a simulated gob wall of almost double the height. This is important because it shows that not only do mine structures and other 84
Colorado School of Mines
solid obstacles enhance mine explosions, but the overall mine environment (i.e. rock rubble in the gob, rock rubble on the belt, corridors made on rock) can generate sufficient turbulence with low blockage ratio and have a major impact on the explosion. 4.3 Ignition between simulated gobs in the 12 cm diameter quartz reactor All experiments performed have been considering methane flame propagation dynamics with a single obstacle. However, in a real longwall coal mine the EGZs typically are near the gob area which is composed of varying types of rock rubble. Explosions that originate nearby or from within the gob can be caused by static discharge of falling rock rubble, hot smears left by metal- on-rock friction, or a flame created by spon-com (Page, et al., 2011). The resulting flame then travels from the gob towards the working face. In order to understand how a flame may propagate in these areas, experiments were performed where the ignition electrodes were placed between different simulated gob geometries as shown in Figure 4.33. In these experiments, referred to as in-gob ignition experiments in this manuscript, the ignition location was in Port 1, 25cm from the open end of the reactor. Simulated gob cages, checkerboard geometries, or wall geometries (no spacing between the spheres) were centered on either side of the ignition electrodes either D=15cm or D=30cm apart. Figure 4.33 Image of in-gob ignition in Port 1 with obstacles on either side, a distance D apart. 4.3.1 Impact of simulated gob location The first set of experiments shown in Figure 4.34 compares the results of igniting a stoichiometric mixture of methane and air 25cm from the open end to results of in-gob ignition between empty cages located a distance D=15cm or D=30cm away from each other. Results show that igniting between empty cages enhances methane flame front propagation velocity across the obstacles due to enhanced fluid motion induced by the cage. Downstream of the 85
Colorado School of Mines
simulated gob, the flame front propagation velocities are unaffected since the empty cages have shown to induce movement only in nearby gases. When the cages are moved further from the ignition point they have less of an effect on the initial kernel development, inducing less motion in the overall gases resulting in slightly slower propagation velocities, though still faster than the open tube without a simulated gob. Figure 4.34 Impact of simulated gob location on methane flame front propagation velocity for an in-gob ignition. Obstacle: Cage. Obstacle location represented by bars, D=15cm, 30cm. Ignition location is 25cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. 4 E =60±5mJ. One (1) relief hole. Each data point is the average of 4-5 data points. Standard ign deviation range is between 1-12% of the mean. Replacing the in-gob simulated gob geometries with 6.35mm diameter glass spheres in a checkerboard geometry (77% porosity) results in slower flame propagation velocities than the empty cage cases (Figure 4.35). This is because the glass checkerboard geometry has a larger pressure resistance which allows less unburned gases to reach the flame front, resulting in slower downstream propagation velocities. Downstream of the obstacles, the compression of burned gases builds up, similar to a closed-end ignition, accelerating the flame towards the end of the reactor. Additionally, the increased pressure resistance of the glass checkerboard geometry also changes the flame speed trends; as the glass checkerboard geometries are moved further from ignition they do not inhibit the initial kernel expansion, but continues to induce turbulence in nearby gases thereby accelerating the flame. However, due to the increased pressure resistance to the flame, the in-gob experiments using the glass checkerboard geometries results in a higher peak overpressure than the open tube or empty cages (Figure 4.36). Also to note in this figure, 86
Colorado School of Mines
these overpressure rises are while the flame is expanding and escaping out of the in-gob arrangement and the flame continues to travel towards the closed end after 0.15s. Figure 4.35 Impact of simulated gob location on methane flame front propagation velocity for an in-gob ignition. Obstacle: 6.35mm diameter glass spheres in a checkerboard geometry (77% porosity). Obstacle location represented by bars, D=15cm, 30cm between obstacles. Ignition location is 25cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. 4 E =60±5mJ. One (1) relief hole. Each data point is the average of 5 data points. Standard ign deviation range is between 4-20% of the mean. Figure 4.36 Pressure-time history of in-gob ignition with and without obstacles. Obstacle: Cages, 6.35mm diameter glass spheres in a checkerboard geometry with 77% porosity. Obstacle location, D=15cm. Ignition location is 25cm from the open end. P (Empty) = 1.23±0.1kPa. max P (Cages) = 1.48±0.1kPa. P (Glass Checkerboard) = 1.97±0.1kPa. CH = 9.5±0.3%. max max 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign 87
Colorado School of Mines
Figure 4.37 compares the overpressure trace of the checkerboard obstacles spaced 15cm apart and 30cm apart. Although the peak pressures of these experiments are within the standard deviation, the initial pressure rise due to the kernel expansion occurs for a longer duration when the obstacles are spaced 30cm apart, which makes sense since there is more physical room for the flame to expand. Also, the peak pressure when the obstacles are spaced 15cm apart occurs earlier than when the obstacles are twice the distance away from ignition. Figure 4.37 Pressure-time history of in-gob ignition with glass checkerboard obstacles (77% porosity) spaced 15cm and 30cm apart, centered on Port 1. Obstacle: 6.35mm diameter glass spheres in a checkerboard geometry with 77% porosity. Obstacle location, D=15cm, 30cm. Ignition location is 25cm from the open end. P (15cm) = 1.97±0.1kPa. P (30cm) = max max 2.08±0.1kPa. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) 4 ign relief hole. Finally, the most interesting aspect of all the in-gob ignition experiments is that they produce a tulip flame as shown in Figure 4.38. This is important because a tulip flame is an inversion of the flame front due to hydrodynamic instabilities and acoustic interactions with the flame front (Clanet & Searby, 1996; Ellis & Wheeler, 1928). The tulip inversion stretches the flame front, which increases combustion rates and pressure resulting in typically faster flame speeds. Tulip flames have been studied over the past decades and the development of the tulip inversion has been identified by four major steps (Clanet & Searby, 1996). First, there is the initial kernel expansion of the flame as shown at t=0.01s in Figure 4.38. Next the flame travels towards the closed end of the reactor in a finger shape shown at t=0.03s. Third, the edges of the finger-shape flame reach the cool walls of the reactor and the flame front is flattened as shown at 88
Colorado School of Mines
t=0.04s. Because of this cooling, the flame front starts to become hydrodynamically unstable and the tulip inversion is initiated as shown at t=0.05s. Finally, after the onset of the tulip flame, acoustic-flame interactions dominate which in some cases produce multiple inversions as shown at t=0.06s. Understanding this phenomena is extremely important because stretching and inversion of the flame front can result in faster combustion rates and larger peak overpressures that could lead to more violent explosions. Additionally, matching up the location and shape of the flame with the overpressure traces is extremely useful for understanding this phenomena and is one of the major advantages of performing these experiments in the quartz flow reactor. Figure 4.38 Images of tulip flame resulting from in-gob ignition between two glass checkerboard geometries (77% porosity). Obstacle location, D=30cm. CH = 9.5±0.3%. Operating conditions 4 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign 4.3.2 Impact of simulated gob material A simplified experiment of an in-gob ignition was carried out with the ignition electrodes placed between two “rock walls” with a porosity of 77%, as shown in Figure 4.33. The main goal of these experiments is to better understand the major differences between solid, smooth spheres and rough, irregular rock of similar size and thermal conductivity. Initial results have shown that there is a competing effect of induced turbulence by an obstacle and the pressure restriction from the obstacle. Results in Figure 4.39 show that ignition between two empty cages enhances turbulence in nearby unburned gases which helps accelerate the flame within 25cm upstream of the obstacles. The glass spheres induce movement in nearby gases, but the pressure restriction from the obstacle slows the flame down as compared to the cage. Furthermore, granite pebbles, due to their surface roughness, induce more fluid movement in the nearby gases resulting in higher flame front propagation velocities in the first 25cm as compared to the glass spheres. 89
Colorado School of Mines
Further downstream of the obstacles, the induced turbulence diminishes and the flame front propagation velocities approach a similar value due to the pressure resistance experienced by the flame from the closed end of the reactor. Simulated gobs have a large impact on the pressure-time history inside the quartz reactor as shown in Figure 4.40 and Table 4.2. The granite pebbles produced the highest peak pressures for the longest duration, followed by the glass spheres, and the empty wire mesh. Though the experimental set-ups vary, these results agree with previous researchers who found obstacles can increase overpressure (Kindracki, Kobiera, Rarata, & Wolanski, 2007; Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982), enhance turbulence (Fairweather, Hargrave, Ibrahim, & Walker, 1999; Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982), and promote combustion and burning velocity (Kindracki, Kobiera, Rarata, & Wolanski, 2007; Fairweather, Hargrave, Ibrahim, & Walker, 1999; Moen, Lee, Hjertager, Fuhre, & Eckhoff, 1982). The rate of decay of the pressure waves after the second major peak is greatest for the glass spheres, followed by the cage and the granite rock. These results help to confirm that the granite rock increases mixing for a longer period of time in the unburned gases allowing for more complete combustion and faster flame front propagation. Figure 4.39 Impact of simulated gob material on methane flame front propagation velocity for an in-gob ignition. Obstacle: Cages, granite rock in a checkerboard geometry (77% porosity), and 6.35mm diameter glass spheres in a checkerboard geometry (77% porosity). Obstacle location represented by grey bars, D=15cm between obstacles. Ignition location is 25cm from the open end. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief 4 ign hole. Each data point is the average of 5 data points. Standard deviation range is between 1-20% of the mean. 90
Colorado School of Mines
Also, the recorded flame front propagation velocities of the glass checkerboard obstacles resulted in much larger standard deviations than the other materials used, a maximum of 20%. However, when exploring the overpressure traces, shown in Figure 4.41, the maximum overpressure and resulting pressure oscillations did not significantly change, within 10%. Recall that in these experiments there is always a balance between induced turbulence by the obstacle and pressure resistance from the obstacle. For the cages, there is little to no pressure resistance, so the induced fluid motion is more dominant, resulting in fast flame speeds and less error. For the granite checkerboard obstacles, there is pressure resistance, but the irregularity of the obstacle surface induces significant nearby fluid motion, thus dominating the flame acceleration process. The smooth glass spheres arranged in a checkerboard pattern had a competition between fluid motion and pressure resistance, resulting in larger standard deviations over 5 experiments. Figure 4.42 Images of entrained unburned gases and autoignition event resulting from in-gob ignition between two granite checkerboard geometries (77% porosity). Obstacle location, D=15cm. CH = 9.5±0.3%. Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief 4 ign hole. Another interesting discovery during these experiments was an autoignition event during an in-gob ignition between the granite checkerboards (Figure 4.42). As shown in these images, the flame begins propagating into a finger shape (t=25ms), after which the pressure oscillations entrain unburned gases (t=33.3ms). After t=33.3s, pressure oscillations continue to entrain unburned gases resulting in an autoignition of the gases between the granite checkerboard obstacles. Finally, after t=58.3ms, the flame continues to burn down the length of the reactor. To note, the resulting flame speeds from this experiment did not lie outside the standard deviation of the mean and thus, were included in the data presented here. Also, the likely autoignition did not present itself as a sudden spike on the overpressure traces, so unfortunately, there is only 92
Colorado School of Mines
photographic evidence of this event. However, what is most important about this experiment is that it shows how the pressure fluctuations from a gas explosion can continue to re-entrain unburned gases, sustaining combustion and burning leading to an autoignition. This is important for longwall coal mining because an ignition from within the gob may continue to be sustained if there is nearby, fresh air being entrained by an initial explosion. 4.4 Impact of ignition location in the 12cm diameter quartz reactor Since methane gas explosions can occur in a variety of locations with varying degrees of confinement, it is important to understand how methane flame propagation velocities and overpressures change depending on ignition location. Previous researchers have ignited mixtures at the open and closed ends of their combustion chambers (Solberg, Pappas, & Skramstad, 1981) (Kindracki, Kobiera, Rarata, & Wolanski, 2007), and some in the middle, but none have thoroughly investigated the impact of varying the ignition along the length of the reactor. Therefore, to gain a better understanding of the impact of ignition location, in the next set of experiments the ignition location was varied along the length of the reactor in Ports 1, 2, and 3 (ignition 25, 50, and 75cm from the open end respectively). Figure 4.43 demonstrates the influence of ignition location. As ignition is moved further from the open end, the maximum flame front propagation velocity towards both the open and closed end increases. Figure 4.44 shows that the peak overpressure in the explosion vessel also increases as the ignition point moves away from the open end towards the center of the vessel which agrees with previous researchers (Bradley & Mitcheson, 1978; Cooper, Fairweather, & Tite, 1986). Ignition within Port 3 was centered in the quartz reactor and produced the largest pressure rise and sustained high frequency pressure oscillations which were also observed in high-speed imaging shown in Figure 4.45. Researchers also found that ignition in Port 2, 50cm from the open end of the reactor, can cause a higher pressure rise in the reactor than a closed-end ignition, but has a slower flame front propagation velocity. The duration of the pressure-time history of ignition in Port 2 is longer than that for a closed-end ignition. These differences are mainly due to the large acoustical waves produced by explosion, which in the case of ignition in Port 2, continue to interact with the walls of the vessel and flame front thereby increasing the overall pressure rise. This result is important because in a real longwall coal mine, pressure waves may travel throughout the mine, reverberate off walls and interact with other mine structures and thus increase the overpressure. 93
Colorado School of Mines
Figure 4.43 Impact of ignition location on methane flame front propagation. Ignition location is 11cm from the open end, 25cm from the open end in Port 1, 50cm from the open end in Port 2, 75cm from the open end in Port 3, and 1.39m from the open end. Dotted lines indicate ignition location and arrows indicate propagation direction of recorded flame fronts. CH = 9.5±0.3%. 4 Operating conditions 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. Each data point is the ign average of 5 data points. Standard deviation range is between 1-22% of the mean. Figure 4.44 Pressure-time history of ignition within various ports. Ignition location is 25cm from the open end in Port 1, 50cm from the open end in Port 2, 75cm from the open end in Port 3, and 1.39cm from the open end (CEI). P (Port 1)=1.23±0.1kPa, P (Port 2)=5.69kPa, max max P (Port 3)=11.99±2kPa. P (CEI) = 3.24±0.15kPa CH = 9.5±0.3%. Operating conditions max max 4 294±1K, 83±1kPa. E =60±5mJ. One (1) relief hole. ign 94
Colorado School of Mines
Figure 4.45 Images of methane flame from ignition in Port 3. Ignition location is 75cm from the open end. P (Port 3)=11.99±2kPa, P (Port 3)=-12.1±2.6kPa. CH = 9.5±0.3%. Operating max min 4 conditions 294K, 83kPa. E =60mJ. One (1) relief hole. ign 4.5 Impact of relief holes in the 12cm diameter quartz reactor There have been many researchers who have investigated the impact of venting/relief on flame propagation and overpressure, finding, in general, that increased relief/venting decreases the overall pressure rise of the explosion (Bao, et al., 2016; Guo, Wang, Liu, & Chen, 2017; McCann, Thomas, & Edwards, 1985; van Wingerden & Zeeuwen, 1983). As discussed in Chapter 2 Section 2.5, much of the research into venting is for designing safe pipelines in oil/gas transport, but this research can also be applied to large-scale industrial explosions including methane gas explosions in longwall coal mines. In order to develop a strong understanding on the impact of venting/relief on the 12cm diameter quartz reactor, experiments were performed changing both ignition location and the number of relief holes (D=1.0±0.2cm) on the closed end of the reactor. As shown in Figure 4.46 and Figure 4.47, there was no measurable change in flame front propagation velocity or peak pressure during a closed-end ignition, CH = 9.5%, when the number of relief holes was changed 4 from 0-2 because the expansion of the reaction gases is greater than the volume of unburned gases vented from the closed end. 95
Colorado School of Mines
Figure 4.46 Impact of number of relief holes on methane flame front propagation for a closed- end ignition. Ignition location is 1.39cm from the open end. CH = 9.5±0.3%. Operating 4 conditions 294±1K, 83±1kPa. E =60±5mJ. Each data point is the average of 5 data points. ign Standard deviation range is between 0-8% of the mean. Figure 4.47 Impact of number relief holes on pressure-time history for a closed-end ignition. Ignition location is 1.39cm from the open end. P (0 holes)=3.29kPa, P (1 hole) max max =3.24±0.15kPa, P (2 holes)=3.33kPa .CH = 9.5±0.3%. Operating conditions 294±1K, max 4 83±1kPa. E =60±5mJ. ign The cases shown in Figure 4.48 through Figure 4.52 are for ignition in Port 2 with varying the number of relief holes on the closed end of the reactor. Similar to previous findings from other researchers, increasing the venting area can help reduce the overpressure (Bauwens, Chaffee, & Corofeev, 2008; Bradley & Mitcheson, 1978; Cooper, Fairweather, & Tite, 1986; Guo, Wang, Liu, & Chen, 2017; Solberg, Pappas, & Skramstad, 1981) and the flame front propagation velocity towards the closed end of the reactor (Bauwens, Chaffee, & Corofeev, 96
Colorado School of Mines
2008). This is because increasing the number of relief holes, or venting area, allows for venting of gases, lowering the pressures and temperatures, reducing combustion rates and the flame front propagation velocity. Also, due to the combination of these effects, the shape of the flame also changes as a function of relief holes. As can be seen in Figure 4.50, Figure 4.51, and Figure 4.52, the flame propagating towards the open end has the same shape as a closed-end ignition flame (Figure 2.7). However, as the number of relief holes increase, the shape of the flame propagating towards the closed end looks more similar to an open-end ignition flame (Figure 2.6). For the case of 2 relief holes, Figure 4.52, the increased venting was enough that the hot, buoyant exhaust gases had time to rise to the top of the reactor and push over the propagating flame. Again, this is due to increased venting, allowing for less pressure build up on the closed end of the reactor (which is also reflected in the overpressure traces in Figure 4.47). Additionally, it was found for this experimental setup that as venting area decreases, pressure oscillations are sustained at a greater magnitude and for longer durations which is important because pressure waves greater than 35kPa can severely harm human ear drums (Owen-Smith, 1981; Institute of Medicine, 2014). Sustained high pressures can also reverse airflow in a mine, destroy ventilation controls, and displace mine structures (Zhang & Ma, 2015). This research demonstrates that even small changes to confinement relief openings directly impact methane flame dynamics and overpressure. Figure 4.48 Impact of number of relief holes on methane flame front propagation for ignition in Port 2. Ignition location is 50cm from the open end. Dotted line indicates ignition location and arrows indicate propagation direction of recorded flame fronts. CH = 9.5±0.3%. Operating 4 conditions 294±1K, 83±1kPa. E =60±5mJ. Each data point is the average of 5 data points. ign Standard deviation range is between 0-37% of the mean. 97
Colorado School of Mines
discussed in Chapter 3, Section 3.1.2, the main purpose of the experimental box is to begin to understand how reactor shape and multiple pathways impacts methane flame propagation and interaction with a simulated gob. The experimental box used has dimensions 51x34x15cm (LxWxH) and a total volume of 18.4L which is only slightly larger in volume than the 12cm diameter quartz reactor (16.9L). Experiments in the box were performed with and without a simulated gob as described in Section 3.1.2, with ignition near the opening (unconfined) and ignition near the closed end (confined). Results in Table 4.3 show that for an ignition near the opening (open end), the simulated gob enhanced the flame front propagation velocities, which was further enhanced when ignition was located near the closed end in the top-right corner of the box. The methane flame front propagation result trends were expected since previous research in this manuscript has shown that 1) confined ignitions result in faster flames and 2) obstacles can enhance mixing and combustion rates. However, looking further at the shape of the flame and the flame propagation trends, video results from these deflagrations show that without a gob the flame expands in all directions as shown in Figure 4.53 and Figure 4.55. Also, the confined, closed-end ignition flame tends to travel faster towards the relief opening than towards the corners of the box which has been observed by other researchers experimenting in rectangular enclosures (Solberg, Pappas, & Skramstad, 1981). However, with a gob, the flame tends to travel through the gob faster than around the entries as shown in Figure 4.54 and Figure 4.56. This was interesting because one might hypothesize that the flame would tend to travel in the open spaces faster than through the gob since these are the path of least resistance. Based on these results it was seen that the enhanced turbulence by the gob increased the transport of unburned gases to the flame front, accelerating combustion rates and flame speed. Table 4.3 Average methane flame front propagation velocities and standard error of the mean with and without a simulated gob (porous medium) for ignition near the open end (bottom-left corner) and closed end (top-right corner). Averaged over 2 data sets. Ignition location Open end Closed end Flame Front Propagation Velocity with No Medium 0.9±0.1m/s 7.7±1m/s Flame Front Propagation Velocity with Porous Medium 1.8±0.2m/s 12±4m/s 100
Colorado School of Mines
Figure 4.55 Flame images of a closed-end ignition in the experimental box without a porous medium. Ignition location is in the top-right corner. CH = 9.5±0.3%. Operating conditions 4 294±1K, 83±1kPa. E =60±5mJ. ign Figure 4.56 Flame images of a closed-end ignition in the experimental box with a porous medium. Ignition location is in the top-right corner. CH = 9.5±0.3%. Operating conditions 4 294±1K, 83±1kPa. E =60±5mJ. ign 4.7 71cm Diameter Reactor Experiments As discussed in Section 3.1, this research performs experiments in both small-scale and large-scale reactors in order to investigate scaling of methane flame front propagation velocity and overpressure. The main purpose of the small-scale, laboratory experiments is to narrow- down the necessary experiments to perform at the large-scale. Based on experiments performed in the quartz reactor and experimental box, results show that methane gas explosions are sensitive to confined ignitions, the amount of void space or blockage ratio, and obstacle surface topology. Therefore, experiments performed at the large-scale in the 71cm diameter, 6.1m long reactor are as follows: 102
Colorado School of Mines
• Closed-end ignition with no obstacles, CH = 9.5% 4 • Closed-end ignition with a rock pile at the open end of the reactor, CH = 9.5%, H = 4 0.24m, L = 1.8m • Closed-end ignition with a rock pile at the closed end of the reactor, CH = 9.5%, H = 4 0.24m, L = 1.8m One of the main challenges with performing the proposed experiments in the 71cm diameter reactor is the weight of the rock pile and moving the rock pile towards the confined (closed end) of the reactor. In order to have control over containing and moving the rock pile, a winch and pulley system was set up and connected to half of a cut, metal barrel with diameter of 71cm as shown in Figure 4.57. This setup allows a single operator to change the height, length, and location of the rock pile with ease. Figure 4.57 Image of a rock pile being inserted into the large-scale 71cm diameter, 6.1m long steel reactor. Rocks are piled on top of a steel bed which is attached to a 2-ton winch and pulley system allowing the rock pile to be inserted at different locations along the length of the reactor. Experimental results with and without a rock pile in the 71cm diameter reactor are shown in Figure 4.58 and Figure 4.59. As can be seen, when the rock pile is close to ignition, the flame front propagation velocity is enhanced along the entire length of the reactor. When the rock pile is further from ignition, at the open end of the reactor, the initial flame development is unaffected by the rock pile. However, when the flame begins to interact with the rock pile, the pressure wave in front of the flame induces turbulent motion in the gases above and within the void spaces in the rock pile. This turbulence leads to increased flame speeds across the rock pile as 103
Colorado School of Mines
shown in Figure 4.58. Examining the pressure histories in Figure 4.59 shows that the rock pile at the closed end produced the greatest overpressure while the flame was in the reactor. However, when the rock pile was located at the open end of the reactor, a large, reflected pressure wave was produced after the flame exited the reactor. This is likely due to a pressure increase in the rock pile and density difference between products in the reactor and ambient air leading to a reflected pressure wave. In general, however, the rock pile greatly enhanced flame speed and overpressure which is important for understanding severity of explosions in an underground mine. For example, an ignition closer to the gob may result in increased flame speed and overpressure as compared to an ignition along the longwall face or near a cross-cut. Figure 4.58 Impact of rock pile location on methane flame front propagation velocity for a closed-end ignition in the 71cm diameter reactor. Obstacle: Rock pile, H=0.24m, L=1.8m. Ignition location is at the closed end of the reactor. CH = 9.5±1%. Operating conditions 4 295±1K, 79±1kPa. E =60±5mJ. Each data point is the average of 2-4 data points. Maximum ign standard error of the mean is 5.5m/s. Figure credit: (Fig, 2019). 104
Colorado School of Mines
CHAPTER 5 COMPUTATIONAL FLUID DYNAMICS MODEL SETUP To accurately solve high-speed methane gas deflagration physics in an underground longwall coal mine requires computational fluid dynamics (CFD) modeling and complimentary experiments to validate the model. Thus, the CSM research group has been developing and validating a mine-scale, CFD model of the ventilation and movement of explosive gas zones in and around the gob area (Gilmore, et al., 2016; Juganda, Brune, Bogin, Grubb, & Lolon, 2017; Marts, et al., 2014). Complimentary to the mine-scale ventilation model, the Fig and Strebinger have been performing methane-air combustion experiments and validating coupled, CFD combustion models using ANSYS Fluent (Fig, Bogin, Brune, & Grubb, 2016; Fig, Strebinger, Bogin, & Brune, 2018; Strebinger, Bogin, & Brune, 2019). Previous research focused on developing 2D combustion models of the laboratory-scale reactors for the 5cm, 9.5cm, 13.6cm and 30.5cm diameter reactors (Fig, 2019). These laboratory- scale 5cm and 9.5cm models investigated the impact of humidity, radiation, and presence of a rock pile on confined methane-gas deflagrations. Fig (2019) validated the models to accurately capture the relative trends of humidity slowing down flame propagation and concluded that including radiation into the model resulted in unrealistic flame shapes and propagation trends (Fig, Bogin, Brune, & Grubb, 2017; Fig, 2019). Fig also found flame acceleration across a modeled rock pile and trends matched experimental trends at a variety of scales (Fig, Strebinger, Bogin, & Brune, 2018; Fig, 2019). Additionally, researchers investigated the impact of chemical reaction models, 2-step, reduced mechanisms, and full mechanisms on methane gas deflagrations and found that the reduced and full chemistry mechanisms resulted in faster flames and required significantly more simulation time (Fig, 2019). Important to note in these models, researchers found that modeling the turbulence using the k-ε turbulence model resulted in more accurately flame front propagation velocities compared to the k-ω models. Also important to note, the flame was initiated using a constant heat flux from an aluminum circle meant to represent the spark electrodes. Although these modeling settings worked well in the 2D, small-scale CFD combustion models, these assumptions will be revisited for these larger-scaled reactors in this Chapter. 106