.. «a... .2 in. s 7?. r.. . 7.40:... . .. . .3:....).L.....n.u:: . .r . V I. :3. .. .2 15...! 5!). a“: :amwg.§amrp .43....5 la. .45.. that , 11...: fl (.5. c .3. .{zn .15 . cub-A m A .I (I . r .3. I, . .nn.».‘.e€v}.fiuhmn . 1.1.1 2 .323... :3 \ 3!. x1 K15 :tth..r§.v AL. up}! 2:3:10 I»? 33.353 ‘ :4 . L. 5...!) .x z. . 3% am. V . x. J .vqluvi; :11... .3. 3.8- ..,. , , Irwm 4...“...th an ramfixfl.” qfiih: i. .3. a 1.. u I... 10:11.3: E332! .35 1’. 23...»... . .. .. .1 12...-.. :1... it (do. a: .. t .9... .i . m... z 1.} 4 2-53:5: If! .5}... ..i?.....2.! : 1!...7912v 314.53,... 4.. 3...... . 31...... . .36??an A $2504 34.1: t 3m _ 43.4...» 533.512.! .lel-I..l . .151}: v ‘I I“ _ 3. 1:. 5.5.1 13...} imbaxdvddu 3.1.. .3. . a... a. m. 31%. . .. . 3 .5; :11}... .621. .4912.“ n , a. .9: {51:52.23 1... #1.: a l.‘ .. I .21!!!) .5..sz 4.1,! [.3 . It... ..‘ u (312..- LA u .) . 3.11....3 I; v «I...:J~» Z LIBRARY W11; Michigan State University This is to certify that the thesis entitled An Experimental Study of Microwave Plasma-Enhanced Combustion presented by Chandra Lynn Romel has been accepted towards fulfillment of the requirements for the MS. degree in Mechahical EnfleerinL MM. j Major Professor’s Signature Maj II) 200 6 O I #. Date MSU Is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:lClRC/DateDue.indd-p.1 AN EXPERIMENTAL STUDY OF MICROWAVE PLASMA-ENHANCED COMBUSTION By Chandra Lynn Romel A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Mechanical Engineering 2006 ABSTRACT AN EXPERIMENTAL STUDY OF MICROWAVE PLASMA-ENHANCED COMBUSTION By Chandra Lynn Romel Advantages of a hybrid flame, created by microwave power, include more efficient and stable combustion and a more concentrated, higher temperature flame/discharge, potentially useful for material synthesis, material cutting and welding, and various surface treatments. A compact torch has been designed and experimentally evaluated while operating in both plasma-only and plasma- assisted combustion modes. Operation of the torch in a hydrocarbon/oxygen combustion mode is investigated with microwave power applied to modify the combustion process. The objective of this investigation is to quantify the changes in the combustion process as microwave power is applied to intensify the discharge. Diagnostic measurements performed include (1) gas temperatures by optical emission spectroscopy, (2) flame/discharge power densities calculated from flame geometries, and (3) flame species to show how the presence of microwaves affects the flame. These measurements are made using a brass nozzle and a ceramic nozzle. ACKNOWLEDGMENTS I thank my mom for being my support for 26 years. I also thank the Man above for helping me through the tough times and for His love. I would also like to thank my advisor, Dr. lndrek Wichman, for challenging me and encouraging me. Special thanks goes to the Electrical Engineering professors, Dr. Grotjohn and Dr. Asmussen, and Materials Science professor, Dr. Case, all of whom advised me throughout this project. I also thank the electrical engineering students, Kadek Hemawan, Stanley Zuo, and Jeffri Narendra for their help and guidance. Thank you to Kadek who has given me permission to use select figures created by him. iii TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ vi LIST OF FIGURES ......................................................................................................... vii Chapter 1 ........................................................................................................................... 1 1.1 Introduction ................................................................................................................ 1 1.2 Literature on plasma-enhanced combustion ......................................................... 6 Chapter 2 ........................................................................................................................... 8 2.1 Nomenclature ............................................................................................................ 8 2.2 Single-step chemistry ............................................................................................... 9 2.3 Multi-step chemistry ............................................................................................... 27 Chapter 3 ......................................................................................................................... 43 3.1 Introduction to equipment ...................................................................................... 43 3.2 Overall experimental setup .................................................................................... 43 3.2.1 Torch .................................................................................................................. 43 3.2.1.1 Introduction to torch .................................................................................. 43 3.2.1.2 Torch design and dimensions ................................................................ 44 3.2.1.2.1 Ceramic nozzle design and fabrication ......................................... 47 3.2.2 Electrical network and cavity ......................................................................... 52 3.2.3 Gas system ....................................................................................................... 55 3.3 Diagnostics and measurement techniques ......................................................... 57 3.3.1 Optical emission spectroscopy (OES) ......................................................... 58 3.3.2 Infrared camera ............................................................................................... 60 3.3.3 Thermocouple ................................................................................................... 63 Chapter 4 ......................................................................................................................... 66 4.1 Introduction ............................................................................................................... 66 4.2 Rotational temperature theory .............................................................................. 66 4.3 Gas temperature of N2 ........................................................................................... 71 4.4 Gas temperature of C2 ............................................................................................ 74 Chapter 5 ......................................................................................................................... 78 5.1 Introduction ............................................................................................................... 78 5.2 Flame/discharge photographs .............................................................................. 78 5.3 Geometric analysis .................................................................................................. 88 5.4 Impact of microwaves on combustion .................................................................. 91 5.4.1 Optical emission spectroscopy (OES) .......................................................... 93 5.5 Flame/discharge temperatures and profiles ..................................................... 105 5.5.1 Data collected from infrared (IR) camera ................................................... 109 Chapter 6 ....................................................................................................................... 113 6.1 Summary of results .............................................................................................. 113 iv 6.2 Recommendations ............................................................................................... 115 APPENDICES ............................................................................................................... 117 Appendix A .................................................................................................................... 118 Appendix B .................................................................................................................... 119 BIBLIOGRAPHY ........................................................................................................... 121 LIST OF TABLES Table 3.1Torch part dimensions in millimeters and inches .................................. 46 Table 3.2 Nozzle flow limits. Maximum flow rates for each type of nozzle, in sccm and m/s. Note the ceramic nozzle allows a higher maximum flow rate than the brass nozzle with the same diameter. .......................................................... 57 Table 4.1 Rotational constants for the electronic states of nitrogen and carbon. 67 Table 4.2 Values used for the N2 rotational temperature calculations for the R branch of the (2,0) SPS. ...................................................................................... 73 Table 5.1 Gas flow limitations for three nozzles with units in sccm and m/s.......79 Table 5.2 Lengths of flames with varying flow rates, corresponding to Figure 5.1. The shaded cells highlight stoichiometric flow rates ............................................ 81 Table 5.3 Flame speeds calculated for flames shown in Figure 5.7 as power decreases from 39 W to 33 W. Calculations are based on Equation (5.1). ......... 88 Table A1 Values for thermocouple temperature measurements .................... 118 Table 8.1 Values used for flame extinction plot in Chapter 5. ........................... 119 Table 32 Values found in temperature plots in Chapter 5. ............................... 119 vi LIST OF FIGURES Figure 1.1 Schematic of experimental apparatus including the plasma/flame torch inside microwave cavity. The torch houses a gas line and water line, and microwave power is inputted midstream. A combustion flame is produced inside a cavity that creates an electric field pattern exciting a TM012 (transverse magnetic) mode. A window is present for observation and diagnostic measurements (optical emission spectroscopy (OES), for example) [3]. .............. 2 Figure 1.2 Species and radical concentration curves in regions 6,, 6n. and 6m of the combustion process. With the addition of microwaves, the electron/ion concentration is possibly greater than with combustion alone. ............................. 4 Figure 2.1 Temperature distribution in relation to flame front for the simple infinite-reaction-rate model described by Eq (2.3). Note the jump in temperature gradient across the flame front. 14 Figure 2.2 Schematic of reactant mass fraction distribution for the infinitesimally thin flame described by Eq (2.6) on either side of the flame, with f=0 on either side. Note, once again as in Figure 2.1, the discontinuity in the gradient of YR, [dYR/dX]=YRUUf/D. ............................................................................................... 16 Figure 2.3 Schematic of stretching transformation in nondimensional scale. The inner coordinate E is large and negative near 6 = 0 ', large and positive near 6 = 0*, and of the order unity in the flame sheet ........................................................ 21 Figure 2.4 Plot of anticipated solution after applying stretching transformation. .24 Figure 2.5 Curve representing a two-step combustion mechanism ..................... 29 Figure 2.6 Curve representing fuel species behavior for a two-step mechanism. ............................................................................................................................ 35 Figure 3.1 Left: photograph of disassembled PTB. Right: photograph of assembled torch demonstrating hand-held easiness. .............................................. 44 Figure 3.2 Internal structure schematic drawing of the PTB with the brass nozzle. This unit is also called the “torch applicator.” Note the center conductor and the water cooling system. The inner conductor is adjustable to create an optimal electric field that is concentrated at or near the tip of the nozzle [18]. .................. 45 Figure 3.3 Left: cross section of nozzle with gas and water lines. The gas line is soldered to the wall of the nozzle. Water cooling is necessary to keep the solder from melting. Water enters the inside of the inner conductor near the tip of the vii nozzle and circulates throughout, and exits near the base of the conductor. Right: cross-sectional drawing of the nozzle with dimensions in millimeters. .................. 47 Figure 3.4 Left: sketch of ceramic nozzle showing smoothly converging path from nozzle inlet to exit, including diameter dimensions. Right: photograph of ceramic nozzle attached to the inner conductor. Note the gold-sputtered coat on the ceramic. ........................................................................................................................... 48 Figure 3.5 Schematic of ceramic nozzle. Left: ceramic powder being pressed axially by two cylinders at 2000 psi inside a cylindrical metal die and shaped by a parabolic graphite mold. Right: nozzle after pressing and drilling of exit pathway. Note the outside corners of the nozzle are still square. Careful hand sanding rounds these corners (represented by dotted lines) so the flame does not sit on a radially-large surface. .................................................................................................... 50 Figure 3.6 Photographs of: (a) tumbler used for mixing powders and for shaping graphite, (b) axial press used for pressing powder in the die, and (c) digitally controlled furnace used for sintering specimens ....................................................... 51 Figure 3.7 Overall experimental setup, including electrical circuit, torch and cavity, gas system, and diagnostic setup [3]. ............................................................ 53 Figure 3.8 PTB and microwave cavity setup, including details of the cavity and the electrical field pattern and wave modes. The torch is placed vertically in the cavity and tuned so the flame interacts with the microwaves, creating a plasma- enhanced hybrid flame. The dimensions of the cavity-penetrated torch are given. ........................................................................................................................................... 55 Figure 3.9 Schematic of the gas setup. Gas flows from the compressed tank, through MFCs. It is then premixed before entering the torch. Methane (CH4) and oxygen (02) are the primary gases used in the experiments. The microwave power is fed through the side of the torch applicator (Pmicro) and the water cooling lines are also shown. ..................................................................................................... 56 Figure 3.10 Schematic of OES system. Light is emitted from discharge, analyzed in the monochromator, and collected and organized in the computer ................... 60 Figure 3.11 Left: photograph of the data acquisition system used to measure temperature with a Type K thermocouple. System is connected to a computer with a GPIP cable. Right: device built to sweep thermocouple through flame at a constant speed. Note the spherical thermocouple junction ..................................... 65 Figure 4.1 Energy level diagram for a band with P, Q, and R branches, relative to wavelength (A), rotational energy levels (J), and vibrational levels (v) [27].........68 viii Figure 4.2 Laboratory process for temperature calculations: light is emitted from the flame/discharge and focused by a lens into the spectrometer, where intensities and wavelengths are recorded. .......................................................... 71 Figure 4.3 Plot of intensity (or current since the units are arbitrary) versus wavelength for the N2 spectrum, commonly used for rotational temperature calculations. R20 - R30 list the order of the rotational bands. ................................ 72 Figure 4.4 Boltzmann plot for the bands R20 — R30 of the N2 spectrum. .............. 74 Figure 4.5 C2 spectrum of a Swan System, identifying the bands R25 — R45. ...... 75 Figure 4.6 Boltzmann plot for the bands R25 - R45 of the C2 spectrum, for the 180/60 sccm O2/CH4 pure flame. ........................................................................ 77 Figure 5.1 Photographs of pure combustion flames with varying flow rates, measured in sccm. The 0.4 mm-diameter brass nozzle is used. ..................... 80 Figure 5.2 Photographs of a pure plasma discharge, with a flow rate of 200 sccm argon and 30 W microwave power. Note that the torch is not confined within the cavity. .................................................................................................................. 82 Figure 5.3 Photographs of flame/discharges inside the cavity as 40-100 W microwave power is added. The 0.2 mm-diameter brass nozzle is used. The flow rate for the top set of photos is 40/0 sccm O2/CH4, i.e. no fuel is present. The flow rate for the bottom set is 40/20 sccm O2/CH4, a stoichiometric ratio. Scales are provided for flame dimensions. Camera speed and aperture are 1.6 and 3.5, respectively. ........................................................................................................ 83 Figure 5.4 Photographs of flame/discharges inside the cavity as microwave power is added, 20-80 W. The 0.4 mm-diameter ceramic nozzle is used. The flow rate is 200/100 sccm O2/CH4, a stoichiometric ratio. A scale is provided for flame dimensions. Camera shutter speed and aperture are 1.6 s and 3.5, respectively. ............................................................................................................................ 84 Figure 5.5 Photographs of flame/discharges inside the cavity with a ceramic nozzle comparing 0 W to 80 W of microwave power ........................................... 84 Figure 5.6 Photographs of flame/discharges inside the cavity as microwave power is added, 10-100 W. The 0.4 mm-diameter brass nozzle is used. The flow rate for the top set is 70/24 sccm O2/CH4. The flow rate for the bottom set is 153/70 sccm O2/CH4. A scale is provided for flame dimensions. Camera shutter speed and aperture are 1/30 s and 3.5, respectively. ......................................... 86 Figure 5.7 Additional photographs from Figure 5.6. The flow rate is 70/24 sccm O2/CH4 and all other conditions are the same as listed in Figure 5.6. The top ix sequence shows the flame/discharge as microwave power increases; the bottom sequence is power decreasing. This intermediate behavior gives rise to a hysteretic affect. .................................................................................................. 87 Figure 5.8 Power density as a function of non-dimensional power, evaluated from flame/discharge geometries shown in Figure 5.6, with flow rates of 70/24 sccm O2/CH4 and 153/70 sccm O2/CH4. The powers of combustion are provided: 15 W for the first flow and 44 W for the second flow. ................................................... 90 Figure 5.9 Flame height as a function of non-dimensional power, evaluated from flame/discharge geometries, shown in Figure 5.4, with flow rates of 70/24 sccm O2/CH4 and 153/70 sccm O2/CH4 ........................................................................ 90 Figure 5.10 Flame extinction or blow-out plot of equivalence ratio versus total volumetric flow rate, VTOT. Each data point is provided with the combustion power of the flame at a particular flow rate, where fuel is present. Note that a discharge is maintained with no fuel when 20-100 W of microwave power is added. A spreadsheet of raw data can be seen in Appendix B. ......................................... 93 Figure 5.11 OES scan of C2 in a stoichiometric (200/100 sccm O2/CH4) combustion flame and hybrid (combustion plus 40 W microwave power) discharge using a 0.4 mm-diameter ceramic nozzle. .......................................... 95 Figure 5.12 OES scan of N2 in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 40, 50, 100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” is a flow rate of 160/54 sccm O2/CH4. The last set of data is from a discharge of only 160 sccm O2 and 50 W of microwave power (no fuel is present). Data from 30 W of microwave power and below show no N2 signal, similar to the combustion-only case, and therefore are excluded from the plot. .................................................. 96 Figure 5.13 Second set of data: OES scan of N2 in a hybrid (160/80 sccm O2/CH4 combustion plus 10-80 W microwave power) flame/discharge using the 0.4 mm- diameter brass nozzle. ........................................................................................ 97 Figure 5.14 OES scan of N2 in a hybrid flame/discharge with a flow rate of 160/80/10 sccm O2/CH4/N2 with varying microwave power, 30-80 W. Note that N2 is injected upstream into the flow of gases, CH4 and O2. .................................... 97 Figure 5.15 OES scan of C2 in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 10-100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” has a flow rate of 160/54 sccm O2/CH4. .......................................................................................... 98 Figure 5.16 Second set of data: OES scan of C2 in a combustion (160/80 sccm O2/CH4) and hybrid (combustion plus 10-80 W microwave power) flame/discharge using the 0.4 mm-diameter brass nozzle. ................................. 99 Figure 5.17 OES scan of O2 in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 40, 50, 100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” is a flow rate of 160/54 sccm O2/CH4. The last set of data is from a discharge of only 160 sccm O2 and 50 W of microwave power (no fuel is present). Data from 30 W of microwave power and below show no 02 signal, similar to the combustion-only case, and therefore are excluded from the plot. ................................................ 100 Figure 5.18 OES scan of OH in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 40, 50, 100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” is a flow rate of 160/54 sccm O2/CH4. Data from 30 W of microwave power and below show no OH signal, similar to the combustion-only case, and therefore are excluded from the plot. ............................................................................................................. 101 Figure 5.19 OES scan of CH in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 10-100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” is a flow rate of 160/54 sccm O2/CH4. ........................................................................................ 101 Figure 5.20 Relative intensities of N2 emissions as a function of absorbed microwave power, measured from flame/discharges with flow rates of 160/80 sccm, 160/54 sccm, and 160/0 sccm O2/CH4, and also a flow rate of 160/80/10 sccm O2/CH4/N2 with N2 injection. Intensities are recorded from four sets of experiments to show consistency, and values are recorded from peak intensities, at a wavelength of ~ 3769 A .............................................................................. 102 Figure 5.21 Relative intensities of C2 emission as a function of absorbed microwave power, measured from a stoichiometric flame/discharge with a flow rate of 160/80 sccm O2/CH4. Intensities are recorded from three sets of experiments to show consistency, and values are recorded from peak intensities, at a wavelength of ~ 5130 A .............................................................................. 103 Figure 5.22 Relative intensities of O2 emission as a function of absorbed microwave power, measured from flame/discharges with flow rates of 160/80 sccm O2/CH4 and 160/54 sccm O2/CH4. Intensities are recorded from one set of experiments, and values are recorded from peak intensities, at a wavelength of ~ 3340 A ............................................................................................................... 103 Figure 5.23 Relative intensities of CH emission as a function of absorbed microwave power, measured from flame/discharges with flow rates of 160/80 sccm O2/CH4 and 160/54 sccm O2/CH4. Intensities are recorded from one set of xi experiments, and values are recorded from peak intensities, at a wavelength of ~ 4270 A ............................................................................................................... 104 Figure 5.24 OES scan of combustion flame (160/80 sccm O2/CH4) and hybrid flame/discharge (addition of 40 W microwave power) for an entire range of visible emission spectra. The wavelength and relative intensities are given at the peak (A = 4280 A) for each set of data. Molecules and radicals are listed next to their visible spectra. A 0.4 mm-diameter ceramic nozzle is used for this data. ......... 105 Figure 5.25 Temperature profile of a stoichiometric CH4/O2 flame from a type K thermocouple. Photograph of the flame shows where the thermocouple scans the flame. The plot provides experimental data and also displays a qualitative example of a theoretical temperature distribution with an adiabatic flame temperature of 3000 K, taken as an average temperature. ............................... 106 Figure 5.26 N2 rotational temperature of a flame/discharge as a function of absorbed microwave power for data for three different flow rates: 160/80/10 sccm, 100/50/3.6 sccm, and 60/30/2 sccm O2/CH4/N2. ..................................... 107 Figure 5.27 N2 and C2 rotational temperatures of flame/discharges as a function of absorbed microwave power, for four sets of data (independent experiments). The flow rate for the first three sets is 160/80 sccm O2/CH4, and 160/80/10 O2/CH4/N2 for the fourth set. Tables of raw data are found in Appendix B. ....... 108 Figure 5.28Temperature profile images from the IR camera of a flame as it is modified by microwave power, 40 W and 60 W. The flame is not confined within the cavity. The temperature scale is provided, but does not provide an accurate temperature reading. The gas flow rate is 200/100 sccm O2/CH4. .................... 110 Figure 5.29 Temperature profiles sketched from the IR camera images (from Figure 5.28) for both a combustion flame (represented by the solid lines) and a hybrid flame/discharge with 40 W microwave power added (dashed lines). ..... 110 Figure 5.30 Temperature profile images from the IR camera for a plasma-only discharge: 200 sccrn argon excited by 40 W microwave power. Each image is recorded with a different filter. The image in the bottom right corner shows the temperature profiles of the hot gases beyond the plasma ................................. 111 Figure 5.31 Temperature profile images from the IR camera for a stoichiometric combustion flame formed by a ceramic nozzle. Different camera filters are used to highlight different sections of the flame. The temperature profiles of the nozzle are also illustrated. ............................................................................................ 112 xii Chapter 1 Introduction and Literature Review 1.1 Introduction Plasma-assisted combustion is an area of research that has been studied for several decades. Thoroughly understanding the physics of this coupled phenomenon and optimizing it have yet to be accomplished. The motivation behind the current research at MSU is to improve the combustion process, to conserve energy required for combustion, and to design and build an efficient, small-scale, plasma-enhanced, premixed-flame, combustion torch. Increased power, combustion efficiency, and flame stability are all advantages of combining a non-thermal plasma (NTP) with combustible gases [1]. Experimental data are gathered and theoretical calculations are performed in order to define the flame characteristics and behaviors;'flame temperature and flame dimensions (height, width) are measured, relative specie concentrations are observed, stoichiometric mass flow rates are calculated for three fuels, the premixed flame theory is reviewed, and a ceramic-flow nozzle is fabricated. The flame configuration employed in this project is, similar to other studies [1, 2], a circular premixed flame burner in which the premixed gases are combusted as they exit a nozzle into ambient air. The nozzle is enclosed in a microwave cavity, which houses the microwaves, producing the flame/plasma interaction. It is hoped that the experiments can demonstrate a quantitative difference between a pure fuel/oxygen flame and a flame that has been enhanced by microwave plasma. Varying levels of microwave power (measured in Watts) and premixed reactant flow rates (and mixture ratios) are examined. A schematic diagram of the experimental configuration is shown in Figure 1.1. “my top wall I plate E-field V f pattern flame/plasma observation window water[ . .I cooling gas leed Figure 1.1 Schematic of experimental apparatus including the plasma/flame torch inside microwave cavity. The torch houses a gas line and water line, and microwave power is inputted midstream. A combustion flame is produced inside a cavity that creates an electric field pattern exciting a TMo12 (transverse magnetic) mode. A window is present for observation and diagnostic measurements (optical emission spectroscopy (OES), for example) [3]. Flame stability, flame propagation speed, and combustion chemistry of premixed, laminar flames are all affected by inducing an electric current in the flow of gases. It is believed that the plasma assists in breaking down the combustion fuel thereby producing free radicals (e.g. H, OH, O), which are chemically unstable and therefore highly reactive. By using the plasma to convert gaseous fuels into reactive species, the ensuing combustion process does not rely on the self-generation of reactive species, and therefore hotter flames may be produced. Plasma also yields a potentially more concentrated flame. A possible mechanism for flame power augmentation is the acceleration of the reaction by enhanced radical concentration levels. Thus, if the microwave power is suitably and optimally “focused” it may produce, in the upstream gases (prior to the flame sheet), a larger concentration of free radicals, H, OH, O, that can accelerate the combustion reaction (see Figure 1.2). The accelerated reaction may produce a narrow, more spatially concentrated flame, thereby making a flame/plasma with a high power density. The practical goal is to employ this enhanced power density to perform some kinds of cutting, machining, material conditioning, localized heating, or other applications. I Reaction I l l Fuel (CH4) Zone, Products Oxygen 5' I l l | | Electrons/Ions I | With Plasma Intermediate I I Species (H2. CO. GIG) I I Electrons/Ions I Without Plasma Upstream, 1". ° €L Downstream, & an ............ ! .m... Figure 1.2 Species and radical concentration curves in regions 6,, 6”, and 6”, of the combustion process. With the addition of microwaves, the electron/ion concentration is possibly greater than with combustion alone. Microwaves are the type of electric field induced in the MSU torch. Microwave power produces a clean discharge and has been studied by researchers at MSU for many years [4]. Microwave plasmas are also referred to as non-thermal plasmas (NTPs), meaning these plasmas are not produced by extremely high temperatures but by some other means. NTPs have several advantages over thermal plasmas. High gas and ion temperatures are produced by thermal plasmas, whereas NTPs diminish the effects of high gas temperatures, while still producing energetic electrons and other reactive molecules (H, OH, 0, etc). As a result, with a microwave plasma it is possible to consider a gas mixture with multiple temperatures. When only a select number of molecules are excited, this results in little waste enthalpy associated with a gas stream [1]. The MSU apparatus currently operates with a non-thermal plasma. In addition to combustion and plasma considerations, there is the important issue of nozzle design, materials, and construction. The nozzle must channel the reactants into the flame; hence it must accommodate a controlled flow of the gases in the design range of flow rate. The nozzle must also allow the microwaves to concentrate in the appropriate locations so that the beneficial flame/plasma interaction will in fact occur. Whether this concentration (of microwaves) location is immediately in front of the flame or further upstream (in the nozzle, for example) is a matter of research and study. Finally, the nozzle material must allow the microwaves to concentrate without being absorbed, and thereby rendered ineffective. A plasma-enhanced torch is desirable for practical purposes. Industry is interested in such a torch for cutting and metal forming. The small, concentrated flame allows precision cutting. A plasma welder eliminates the mess that many combustion-based welders create. Applications may even be designed for outer space. Buoyancy does not play a role in the flame produced by the torch because it is so small and concentrated. In addition, the torch is easy to handle because of its small size. To further develop this ongoing plasma-enhanced combustion research, experiments are conducted and theoretical calculations are studied. A mini torch- was built that allows combustible gases to flow through and be activated by an electric field, producing free electrons, a current, and a plasma discharge. Temperatures are measured using an IR camera, an optical emission spectrometer (OES) for specific gas temperatures, and a thermocouple. Power density is calculated based on flame/discharge geometry and absorbed microwave power. In addition, a ceramic nozzle was designed and implemented for experiments. 1.2 Literature on plasma-enhanced combustion There has been no systematic and prolonged effort to study plasma- enhanced combustion, but over the past several decades isolated attempts have been made to enhance combustion with an electric field. A short account follows. Plasma-induced combustion research was conducted at Los Alamos National Laboratory in the past five years or so in an attempt to improve the combustion process. More stable and efficient combustion was desired and thought to be achieved by introducing a supplemented electric field [1]. The experimental setup consisted of a coaxial silent discharge reactor, producing a dielectrical barrier discharge (DBD), or silent discharge. The propane and air mix just before ignition and no converging nozzle is used to form a flame. The gases are pre-treated by NTP prior to ignition. Rosocha et al. argue that NTPs have advantages over thermal plasmas, as discussed in Section 1.1. NTPs promote efficient combustion by increasing flame speed, decreasing flame length, reducing soot formation, and reducing NOx and other combustion by- product emissions. The Los Alamos team found that with an NTP system, plasma increases flame blow-out limits, but at a relatively low propane flow rate. The partial pressures of common hydrocarbon combustion products (water and carbon dioxide) increased while propane fragments decreased when the plasma was added, which suggests that the propane is being burned more completely. The flame propagation speed increased with an increase in electric field intensity. Experimental work done by Chintala et al. at Ohio State University shows that non-equilibrium RF plasma-assisted ignition and combustion achieves large- volume ignition in premixed flows. At higher reactant flow rates and lower temperatures, the RF discharge allowed ignition to occur, as compared with no discharge. Leaner combusting fuel flows were achieved with the RF discharge. Radical species such as CN, CH, C2, OH, and O atoms, measured by visible emission spectroscopy, were detected in the hydrocarbon air flow [2]. Much experimental work has been performed by Starikovskii et al. in Russia, quantifying pulsed nanosecond plasma-assisted combustion. They found that the plasma discharge significantly decreases the ignition temperature and that the flame’s blow-off velocity increased significantly for a low discharge energy input. An increase in the flame propagation velocity can be explained by the production of atomic oxygen in a discharge by the quenching of electronically excited N2 and the dissociation of molecular oxygen on electron impact (02+e'—* O+O+e'). A numerical model was developed to describe the influence of the pulsed discharges on the combustion process [5, 6]. Experiments that were performed by Whitehair et al. at Michigan State University use a high-temperature nozzle and discharge chamber materials for the development of a microwave electrothermal thruster. It was shown that regenerative cooling and flow stabilization improved thruster performance [4]. Chapter 2 Premixed Flame Theory 2.1 Nomenclature “ [Ole-s c=§35r~xrmmmoogw>m C "(Denma- N-<‘<><§C—I Integration constant Pre-exponential factor (cm3/gmol—s) Species parameter, cpT/Q + YR Constant-pressure specific heat (J/kg-K or J/kmol-K) Arbitrary coefficients Binary diffusion coefficient (m2/s) Damkohler number Activation energy (J/kmol) Fuel Thermal conductivity (W/m-K) or Rate constant (cm3/gmol-s) Damkdhler number Loss term (J/m3-s) Normalized heat loss term Lewis number Flame thickness (m) Third body Reaction order or integer in gamma function Products Heat release rate (Jls) Energy per unit mass (J/kg) Normalized heat release Reaction rate function Mass generation rate (1/s) Universal gas constant (J/kmoI-K) Normalized flame speed Flame speed (m/s) Time (3) Temperature (K) Velocity (m/s) Mean molecular weight (kg/moi) Distance in dimensional system (rn) Normalized mass fraction Mass fraction (kg/kg) Radical Greek Symbols Thermal diffusivity, k/pcp (m2/s) or exponential coefficient Enthalpy ratio, (Tb-Tu)/Tb Zeldovich number, -E(Tb-Tu)/(RuTb2) Laminar flame thickness (m) Normalized temperature, (T-Tu)/(Tb-Tu) Normalized temperature after stretching, 8(1-6) Burning rate eigenvalue Exponential coefficient Constant density (kg/m3) Nondimensional length scaling, x’/(a/uf) Normalized length after stretching, E/6 Normalized temperature Molar consumption rate (moI/ma-s) g «4 IIIW‘Dt >@a>ovm>o Subscripts Consumed Burned gas Burned Flame Fuel Homogeneous Laminar Unbumed gas Particular Reactant Burned reactant Unbumed reactant Radical NZJJUIJ‘DC r‘21:T|"“I'DCTO CCU 2.2 Single-step chemistry One-dimensional, premixed-flame theory for single-step chemistry has been studied for decades and is well understood [7]. Reactants combust and form products and heat energy, as shown in expression (2.0). Equation (2.1) is the governing energy equation for the transient propagation of a reaction front into a traveling gas. The first term represents the rate of change of temperature at position x; the second term represents convection in the gas moving with velocity, u; the diffusion and reaction terms follow. The source/sink term, in the simplified model of Equation (2.1), is replaced by boundary conditions on the temperature and is redefined in Equation (2.5) as pq. Assumptions to be made are: complete combustion; constant pressure and density; inviscid flow; and a laminar and infinitesimally thin flame front. The analysis derived in this chapter is referenced from [7] and also from notes of an independent study course taken from the advisor of this thesis, Dr. lndrek Wichman. Although the assumptions seem drastic, they provide simple mathematical models while proving remarkable success in laying the foundations for understanding the structure, properties, stability, and dynamics of many combustion phenomena. For example, the density changes can be neglected when considering the diffusive transport [1]. Explicit and asymptotic solutions are found by making difficult problems tractable to mathematical analysis [8,9,10,11,12,13,14,15]. provide examples of explorations of various types of flames with a single-step chemistry model. An asymptotic approach, based on high activation energy, is used in one-step chemistry analysis. The temperature far upstream is given by a constant unburned value, Tu, and the temperature far downstream is a constant burned value, Tb. Figure 2.1 schematically represents the temperature distribution for this idealized premixed flame (PF). Fuel + Oxidizer—-> Products + Heat (20) 10 2 fl+ufl=afl+sourcflsink (2-1) at 3x 8x2 Equation (2.1) is invariant under the application of a Galilean transformation according to the following definitions: f=x+ut f w=u+u f t'=t Essentially, the new x-coordinate follows the flame front since dx’/dt = dx/dt + Uf and dx/dt = 'Uf at the flame; therefore, dx’/dt = 0 at the flame. The velocity, u, is the flow speed entering the flame front from the upstream direction. Note also that it is assumed that the flame front propagates at a constant speed, Uf. Without this assumption a Galilean transformation is not useful. The transformation yields the governing equation in (2.2) for the gas temperature on either side of the flame. 3T 3T BZT _ '_= _ 2.2 at' “ ex “3va I ) The location of the origin is arbitrary since x’—> x’ + constant produces exactly the same equation as in (2.2); as a result, it is convenient to let the flame front sit at 11 x’= 0. The flame front is not affected by t = t’; therefore, in the coordinate system attached to the steadily propagating flame front, the unsteady term in Equation (2.2) vanishes (a(-)/at’ = 0). When the flow velocity is zero (u = 0), the transformed velocity is constant (u’= Ur), and Equation (2.2) reduces to a simple ODE, shown in Equation (2.3), in which the coefficients, Uf and a, are constant: . dim—012T f dx' (1er (2.3) The solution to Equation (2.3) is as follows: / / T=C1+C2exm “f) The physical domain is divided into two regions: one upstream of the flame sheet (- °° < x < 0); the other downstream of the flame sheet (0 < x < °°). In the upstream region, the two following boundary conditions are imposed: T(—oo)=Tu and lim T=Tb x—)0— These boundary conditions yield Equation (2.4), also seen in Figure 2.1: 12 x/(a/uf) T = Tu + (Tb — Tu)e (2.4) The following boundary conditions are imposed for the downstream region to find T=Tb as the downstream solution: T —E and —->—1 as E-—>—oo (2.14) —-)—E as z —)—oo (2.15) For the downstream solution, as § —+ 0", the following distributions are found: 23 Once again, applying 9 = 1 - O/B and 6 = E/B gives 9 = 0 and dO/dE = 0, so that, along with O = 90 + 91/[3 + the following boundary conditions are found and are displayed in Figure 2.4: dGO (90 = O and = O as d91 ('91 = 0 and = O as 9 90(5) ‘ q u s - s u. 5 Anticipated solution ‘- ‘Q - - ‘g ‘- ‘- .s -o -- ..... ---_ c .............. (2.16) (2.17) Figure 2.4 Plot of anticipated solution after applying stretching transformation. 24 The lowest-order problem consists of Equation (2.12) and BCs (2.14) and (2.16). The first-order problem consists of Equation (2.13) and BCs (2.15) and (2.17). Both problems are over-determined and thus each solution will also produce a result for the various eigenvalues, A0, A1, etc. To solve the lowest-order problem, first the eigenvalue is calculated by integrating Equation (2.12) from E = - °° to E = + °° and imposing the conditions given by Equations (2.14) and (2.16). Therefore, the following computes: =—..., e o 2 "if—W - .. ”I tigejzdefi _ 0f 1, d9, _.,, daz “ E:_oo2d(-9 d3 d3 " 90%2 d3 __1_[d€-)0]2 _ [[deOT 2 a 90—20 2 d: 90%” 1 1 =—0—1=—— 2[ I 2 O 3: oo - 0( + )A 9" —90d@ -A De" ‘GOde — A r 1 T i 008 0‘ oioe 0—-0(n+) 90(52—00) oo Thus, A0 = 1/2 l'(n+1) gives the lowest-order flame speed eigenvalue. The procedure above is replicated, but integrated from E = - °° only, to a finite (and unspecific) value, where 90 = OO(E). The result is as follows: 25 2 d9 n+1,9 (ca/"0] :1_°° : _Q(r( +1)O)EP( 1’90) :. n Isne—Sds 0 90 Isne—Sds Q 0 P +1,@ =—= SI 2.18 (n 0) F 0° ( ) [5"?st 0 Equation (2.20) is obtained below by taking the square root of Equation (2.18) and rewriting it as Equation (2.19). Equation (2.20) is solved numerically. dOO d3 =—\/P(n +1,€-)0) E—g(n,®o) (2.19) 90 . ds —E= azm (i g(n,S) The integration constant, a, is found by integrating Equation (2.20) while applying upstream boundary conditions on Oo(E). A transcendental equation forces a numerical solution for Oo(E), which is then used to deduce 6 and thus T, and finally YR from Equation (2.7). 26 The expressions for O1 and A1 can be found by introducing a variable change in the form of g, as follows: dOO d5 2 9 =—$ (221) l 2 ' By a similar procedure (by integrating the first-order problem from the far upstream (E = -°°) to the far downstream (E = +°°)). it is relatively straight forward to find, for the first-order eigenvalues correction, the following expression in Equation (2.22): A1 °° — = (n + 2)(n + 1) - 2 [(1— g)ds (2.22) A0 0 The first-order analysis will not be pursued any further. The interested reader may consult Chapter 5 of [16]. 2.3 Multi-step chemistry The analysis of a single-step chemical reaction does have limitations in describing combustion behavior. For example, the Zeldovich number (B) for a 27 satisfactory description of stable flame balls must be about three times the actual value found in typical hydrocarbon reactions [7]. Realistic Zeldovich numbers are invoked in the two-step mechanism shown in expression (2.23). This expression describes how a radical species Z attacks a fuel species F to produce more radical in the form of chain branching, until all the fuel particles are consumed. A third body, denoted by M, represents any type of molecule. In reaction 2 the radical collides with M to produce product P and leftover M. The rate constant, kg, in the first reaction varies with temperature; however, the rate constant, kg, for the second reaction does not. The chain-branching reaction requires energy from the flame to generate radicals, T3 = EB/Ru (heat release step). The second reaction has zero activation energy, and is therefore independent of T. Thus, To = 0 (heat absorption step). 1) F+Z-—>Z+Z : k =A e'Tb’T B B 2) Z+M—>P+M: kC=A (2.23) C schematically represents the first step in the above mechanism. 28 Figure 2.5 Curve representing a two-step combustion mechanism. The governing equations for the two-step mechanism include the fuel species equation, the radical species equation, and the energy equation, shown in Equations (2.24), (2.25), and (2.26), respectively. The primes imply differentiation with respect to x. pSY1; =pDFY}: —W a) F B pSYZ =pDZYZ —WZa)F 'szc pcpST =kT +QwC—r 29 (2.24) (2.25) (2.26) The laminar flame speed is denoted by S, W is the mean molecular weight, and we and we are defined as follows: In Equation (2.26), the quantity I refers to a heat loss term that can be used to represent the rate of bulk heat loss (through radiation, for example). For simplification, the equations are normalized using the following definitions for the independent (x —’ of) and dependent (Yz, YF, T) variables: 30 0 . a x = JCS: (xs rs chosen for convenience, xs = —) u - . W Z LeZ Y2 = yz Y ZS (YZS IS chosen for convenience, YZS = moi/7.1:: F - S . O C . s =— (u rs chosen for convenience later In the analyszs) u T — T 7: 0 (normalized temperature field) T f — T 0 Y F = y F Y F0 (normalized fuel mass fraction) The fuel, radical, and energy equations are now reduced to the following nondimensional equations: d d2 LeF‘ yF — yzF —Kr (2.27) d: d; dy dzy Le a Z = Z +Kr—y (2.28) 31 2 _ §fl=d T+ Q y —L (2.29) d6 6152 Le], Z Note the similarity of these equations to Equations (2.5) and (2.6). Select variables that are included in Equations (2.27), (2.28), and (2.29) are defined below: Lewis number: Le = 2 D x2 pYFO (0st —TB /T Damk6hler number: K S W A e f =——_ F B pDFYFO WF W2 £212 _ 7f T Reaction rate function: r = y F yZe QYFO cp (Tf — T0)WF Heat release: 6 = (x2 L = S k(Tf —TO) Heat loss term: The quantity r is redefined in terms of the Arrhenius rate expression k(T). as follows: 32 1 1 TB[T T +r(T —T )] k(T)=e f 0 f 0 The expression for k(T) is rearranged after some algebra to give the expression as shown below, where a = 1 — Tom and p = T3(Tf-To)/T;2. l—z' k(T) = e-flil‘MHJ To solve the fuel species equation, the reaction zone is assumed to be very thin. Upstream (f < 0') the flame, the reaction rate r is zero. Therefore, Equation (2.27) becomes Equation (2.30). Ler’yF =yF (2.30) Letting y;’ = exp(m§), it can be shown that the characteristic equation gives L9F§ m = m2, so that the two roots are m = 0 and m = LeFE for the homogeneous solution. The solution to Equation (2.27) is Written below: yF = Cleo: + CzeILeFE)‘: (2.31) The boundary conditions are as follows: 33 Far upstream: y F =1 as ,f —) —oo On upstream side of flame front: y F = 0 as f —> O— The values for C1 and C2 in Equation (2.31) are computed, and the fuel species upstream solution is given below: y F =1 — 19”ng): for 5 < 0‘ (2.32) The downstream (f > 0+) solution uses the same solution (Equation (2.31 )) as the upstream solution. Equation (2.33) and the boundary conditions that follow lead to the solution given below by Equation (2.34): y F = C3 + C4eILeFEI5 (2.33) Far downstream: y F <1 as 4‘ —> +00 At downstream side of flame front: y F = O as 6 —> O+ Solution: y F = 0 for .5 > 0+ (2.34) 34 A curve demonstrating this behavior for the fuel species is shown in Figure 2.6. Yr: 1 1 I a=o-I b;0* 2 Figure 2.6 Curve representing fuel species behavior for a two-step mechanism. Equation (2.27) can also be integrated, from a to b, to show that y): is continuous, or that the jump of y): across the flame is zero [7]. The radical species equation, given in Equation (2.28), is solved by implementing similar methods as the fuel equation. By integrating from a to b in the above Figure 2.6, when a = 0' and b = 0+ are f—locations that straddle the 6 reaction front, the result, IKrdf = —§LeF , is found. Then following expression a is deduced, which yields the jump condition on the gradient of yz: 35 Iy'zlay'z = —§LeF (2.35) 0“r yZIO‘ Letting y2 = exp{a§}, the characteristic equation obtained from Equation (2.28) with r = 0 is Le2§ a = 02 — 1, whose solution gives the two roots 612, seen below: a = LeZs 1,2 2 1 - 2 i§J(LeZS) +4 Solutions are found both upstream (§<0) and downstream (§>0) of the flame; the unknown coefficients are set equal to each other because yz is continuous across the flame front. The one unknown coefficient is then solved for using the jump condition from Equation (2.35). The solutions are written below: Le E yz - F ea]: for §0 “1‘0’2 Equation (2.29), the energy equation, is solved using a coupled homogeneous and non-homogeneous solution. Equation (2.29) is rearranged into Equation (2.36), where L=a2 7, assuming I is linear in temperature. 36 2 __ d T 'd7 azrz—Q yz (2.36) E7371?- LeF Using the results from the species solution, Equation (2.36) can be written as Equation (2.37): r e%,f<0 2 — — L - d 2' E617 a2 Q y Q er (152 d6 Le z LeF (a!1 —a2) < (2.37) Keaz, §>O The upstream solution (§<0) is found from a homogeneous and particular solution where r = r H + Tp. The homogenous equation for 1' is as follows: + \l4a2 +52 The following roots for a and p are redefined below, for convenience, noting that .1. 2 11(1) and M2) are always positive: 1 _ 2 Lez§ a(1)':§\/(Lezs) +4 +——2—-= a] 37 1/ 2 _2 f fl(2)=§ 4a +S —§=—fl2 These roots yield the homogenous solution for the normalized energy equation, as seen in Equation in (2.38): l1“); flag): TH =Cle +C2e (2.38) The particular solution is found from Equation (2.37) for the upstream (§<0) case: 2 —— _ i:_§£__027 = Q Ler can): (1,52 d5 P Le F (am + am) Letting To = Aexp{p(1)f), the following algebraic expression is found: 38 E Lops Le}: (a(1)+a(2)) — 2 AW?” ‘ 50(1) — a )= 1401(1) — #(1) )(0!(1) + #(2)) = '- Q Lela-S‘- where A = — . LeF (61(1) + “(2) )(0«'(1) —fl(1))(a(1)+/1(2)) The total solution for the upstream energy equation is written in Equation (2.39). T=T +1 = (2.39) _ _ a 6 41(2): _ Q LeF s e (I) “F (“(1) + “(2))(0‘01 ‘ ”(1) )(“m + ”(21) ,u e (1)5 Cl +C2e The coefficient C2 is zero because I < °° and 6 —+ - °°; therefore the second term goes to zero. As 6 —> 0', the following expressions are obtained: dr 2' _ =C -K a d — = C —a K 0 1 1 n [délo- #(1) 1 (1) 1 where K1 Q Ler _ Le]: (“(1) +a'(2))(a’(1) —fl(1))(a’(1) +l’(2)) Downstream of the flame ({>0), 7 < °° as c,‘ —> + °°; therefore, C1 must be zero and the first term in Equation (2.39) disappears. As 15 —> 0+, the following expressions are found: 39 dr 2' 2C -K and — =- C +a K 0+ 2 2 [615]“ 111(2) 2 (2) 2 ._ a LeF 5 LeF (0(1) + 0(2))(05(2) +fl(1))(0’(2) -#(2)) where K 2 Applying the jump condition leads to expressions for C1 and C2. The solutions are shown below: T : 5.? a1 —a2 e’ul‘f _ ea]; (6 < 0) 01-02 #1 "/12 (at ‘fllX-a’z +1111) (al-fll)(0!1—/12) a1 _ 072 81126 8026 6'; r = - 0’1 ‘02 {#1 ‘1“2 (-02 ‘/12 )(0'1 #12) (—a2 HQ )(-(12 +#1) 1.. The definition for K was seen earlier as the following: 2 xs W A pYFO pYZS e—TB/Tf K=——— pDFYFO F 8 WF W2 This expression can be simplified into the following: —TB /Tf 40 The asymptotic analysis for a second order reaction shows that K = 32 [7]. Therefore, Tf is defined by Equation (2.40). 2 2 A —T /T T T K=i£YFOe 3 f = .5. 1__(L (240) AC WF Tf Tf For various parameter choices the equations of this subsection can be solved (and have been solved [7]). The solutions plot the flame speedj , versus 1/2a2 the parameters a2/(Q--1) for fixed Lez and Lez for fixed Q , showing that for 1/282 each parameter value (a2/(—Q— -1) and Lez ) there are either two values of the flame speed 5 , or none (see Figure 3 of [7]). It is believed, consistent with several decades of premixed flame stability analysis, that the lower branch is unstable whereas the upper branch is stable. The influence of the radicals enters through the pre-exponential factor, A. As this increases (or decreases) the parameter containing A changes and the influence of the radicals is modified. The influence of the plasma, it is believed, can be examined by varying the reaction rate A, which is altered by the level of the applied electric field. This influence is not examined in this thesis. Another strategy, which also is not examined, is to postulate that in front of the main reaction zone a radical formation region is produced by the plasma. Thus, at a certain distance 6,, upstream of the reaction front the radical mass fraction takes 41 the fixed value yzp. The remainder of the solution is of course dictated by the conservation equations. This solution, however, in preliminary calculations becomes algebraically extremely complicated and it is not clear that a simpler approach, based on scaling analysis, would not be more feasible. 42 Chapter 3 Equipment and Experimental Methods 3.1 Introduction to equipment The test bed used to study the microwave/flame interaction consists of a miniature microwave plasma torch/burner (PTB) that operates inside a cylindrical cavity. The experimental conditions that occur can be divided into three distinct groups of flames/discharges: 1) pure microwave plasma discharge; 2) pure combustion; and 3) hybrid plasma/combustion discharge. Discussed in this chapter are the overall setup, including the torch, cavity, electrical network, and gas flow, along with the experimental methods employed for data collection and analysis. Images in this thesis are presented in color. 3.2 Overall experimental setup 3.2.1 Torch 3.2.1.1 Introduction to torch The PTB used for this research was designed and built by researchers in the MSU Electrical and Computer Engineering (ECE) department 1. It was originally designed for the formation of plasma (argon, helium) in order to study micro-scale plasmas (plasma diameter and height of 1mm and 3mm, respectively) [17, 18]. It operates under atmospheric pressure, but has the capabilities of operating in low pressures down to O(10‘1 atm) also. This pure microwave plasma discharge produces relatively low temperatures (1600 K-1800 K) which limits it to cutting paper and melting plastics. As a cutting tool, the torch’s discharge must cut plastics and metals with precision; therefore, 1 Under review for US. Patent in 2005 43 combustion is introduced to produce a high-temperature, hybrid flame/discharge that consists of a plasma discharge coupled with a flame [19]. 3.2.1.2 Torch design and dimensions The torch is considered “small-scale” because of the flame/discharge size it produces, which is of the several millimeter scale. Photographs of the brass PTB are shown in Figure 3.1. It is a simple hand-held torch comprising coaxial inner and outer conductors, with inlet/outlets for microwave power, gases, and water (used for cooling). The photograph to the left in Figure 3.1 shows the torch before assembly. It includes, from bottom to top, the outer conductor, the fine- tuning adjustor (which is placed inside the outer conductor and controls the placement of the inner conductor, with the help of a teflon spacer used for centering), the inner conductor, and the microwave coupling unit; the finger stock, allen wrench, and bolts are also shown. Figure 3.1 Left: photograph of disassembled PTB. Right: photograph of assembled torch demonstrating hand-held easiness. Figure 3.2 is a schematic of the torch with dimensions. The outer conductor has an 11.1 mm inner diameter, and the inner conductor has an outer diameter of 4.75 mm. All major dimensions of the torch are listed in Table 3.1. The torch is designed so the inner conductor can slide through the outer conductor in axial adjustment (useful for assembly and tuning of the discharge inside the cylindrical cavity). It is centered with two teflon spacers. Fine tuning is available at the base of the torch with an adjustable microwave tuning short. Water Water Input Microwave output power ‘Gases: fuel and oxygen Adjustable microwave tuning short Microwave coupling structure that allows inner tube to slide through Outer conductor ID: 11.1 mm Center conductor k1. / OD:4.75mm Nozzle exit diameter: 0.4 mm Figure 3.2 Internal structure schematic drawing of the PTB with the brass nozzle. This unit is also called the “torch applicator.” Note the center conductor and the water cooling system. The inner conductor is adjustable to create an optimal electric field that is concentrated at or near the tip of the nozzle [18]. A 1/16” gas line feeds premixed fuel and oxygen (mixed ~ 3 ft upstream of torch base) into the nozzle entrance. Because the brass nozzle is soldered to the inner conductor, flowing water prevents the solder from melting due to flame temperatures above 2000 K. The water inlet and outlet are at the base of the torch, and the water is allowed to circulate throughout the inner conductor. The 45 microwave current is fed into the torch by a 50-ohm coaxial cable with an N-type connection. The inner and outer conductors create an electrical field that ionizes the gases at or very near the tip exit of the nozzle. Table 3.1Torch part dimensions in millimeters and inches. Torch Dimension Units. mm Inches Inner Conductor Length 197.488 7.775 Inner Conductor OD 4.763 0.188 (3/16) Inner Conductor Thickness 0.356 0.014 Outer Conductor Length 148.590 5.850 Outer Conductor OD 12.70 0.5 (1/2) Outer Conductor Thickness 0.813 0.032 Nozzle Inlet Diameter 3.340 0.132 Nozzle Exit Diameter 0.40 0.016 Microwave @ut Coaxial Cable Diameter 6.350 0.25 (1/4) Gas Line Diameter 4.234 0.167 (1/16) Inlet Water Line Diameter 4.234 0.167 (1/16) Outlet Water Line Diameter 4.234 0.167 (1 Mg The nozzle, like the rest of the torch, is made of brass, and was designed for relatively easy machining. Drawings of the nozzle can be seen in Figure 3.3, which shows gas/water lines and dimensions. The exit diameter is 0.4 mm. A second nozzle was made with a diameter of 0.2 mm; however, most experiments presented in this thesis are performed with the 0.4 mm-diameter nozzle. 46 .__l 00 X '0 $4.— __f__ .0 .h ——+ 4——0.4 mm 1‘ 0.9 I / A ’ __J/ / / . 20 5 p 1 6 / / g g ,JSolder / 5 z; / r : Gas // § 1‘ ; fife—()4 line " 1 :Water : I j i Q d line / 3.3 1 4 mm 4 4'0 *— Figure 3.3 Left: cross section of nozzle with gas and water lines. The gas line is soldered to the wall of the nozzle. Water cooling is necessary to keep the solder from melting. Water enters the inside of the inner conductor near the tip of the nozzle and circulates throughout, and exits near the base of the conductor. Right: cross-sectional drawing of the nozzle with dimensions in millimeters. 3.2.1.2.1 Ceramic nozzle design and fabrication Innovative design and fabrication of a second, ceramic nozzle are explored and experimentally tested. The original nozzle of the torch cannot withstand temperatures that the combustion flame produces, and also reflects poor flow geometry. This nozzle is made of brass and is soldered to the inner 47 conductor. To prevent the solder from melting and detaching from the torch, the inner conductor is water cooled. Water circulates through the inner conductor to keep the nozzle at a relatively low temperature. A water-cooled nozzle causes the flame to stand off, forming a gap between the flame base and the nozzle exit. A detached flame is not desired because it is unstable and also loses heat energy to the cooled nozzle. In addition, the brass nozzle reflects poor fluid flow. The previous Figure 3.3 shows a drawing of the brass nozzle. The sharp corners induce flow separation and can cause considerable circulation patterns. To eliminate the necessity for water-cooling and to develop improved flow conditions, a smoothly- converging ceramic nozzle has been designed and fabricated. A sketch and photograph of a ceramic nozzle are shown in Figure 3.4. ——>l F—O.4 mm '1. 1' '1 1 1 | “C )4 | l l .1:— _L 3/16” or 0.5 cm Figure 3.4 Left: sketch of ceramic nozzle showing smoothly converging path from nozzle inlet to exit, including diameter dimensions. Right: photograph of ceramic nozzle attached to the inner conductor. Note the gold-sputtered coat on the ceramic. 48 The process for designing and fabricating a ceramic nozzle involved first shaping graphite “molds” to form the smoothly converging flow region, then preparing the powder used for pressing and sintering. The design of the nozzle is developed from shaping graphite rods; no specific equation was developed for the converging shape. The graphite (ordered from McMaster Carr) is used to form the converging shape of the nozzle. Rods of 1/4” diameter and lengths varying from 1 - 4 inches are tumbled in bottles lined with sand paper of various grits. The tumbler or rolling machine can be seen In Figure 3.5. Rods are first tumbled inside a bottle (3" diameter) lined with 40-grit paper for 24 - 48 hrs. The ends of the rods become rounded, closely resembling a parabolic curve. This curve is manually plotted and curve-fitted using Excel. Spotlight, a software program developed by NASA, is also useful for such analysis, thus eliminating the time spent manually plotting points along the rods edge. These procedures demonstrate that the rods are nearly parabolic at the ends. These rods are then tumbled with 400 grit paper, and finally with 1200 grit paper for 12 - 24 hrs at a time. The higher-grit paper creates a polished, reflecting surface on the rod. The rods are then finely polished with a diamond paste. The ends of the rods (noses) are cut off with a low-speed diamond saw. The parabolic noses are used to press the ceramic powder into its nozzle shape. The ceramic powder is made with 85% weight alumina (Al203) and 15% weight zirconia. Alumina is affordable and offers a 99.9% purity level. It is the close packing of aluminum and oxygen atoms within the structure that leads to its good mechanical and thermal properties [20]. Zirconia is added for increased 49 strength. The powder is tumbled/mixed for 24 hrs to ensure a homogeneous mixture. Approximately 10 g of powder is axially pressed under 2000 psi using a cylindrical die after inserting the graphite tip. Figure 3.5 provides schematics of the pressing process. Axial pressure 2000 psi Parabolic graphite mold Inner surface of nozzle \ Inner die cylinders Flow exit Figure 3.5 Schematic of ceramic nozzle. Left: ceramic powder being pressed axially by two cylinders at 2000 psi inside a cylindrical metal die and shaped by a parabolic graphite mold. Right: nozzle after pressing and drilling of exit pathway. Note the outside corners of the nozzle are still square. Careful hand sanding rounds these corners (represented by dotted lines) so the flame does not sit on a radially-large surface. The specimens are partially sintered at 800°C for 4 hrs using a heating rate and cooling rate of 5 °C min'1 in a digitally controlled Carbolite RHF 1500 furnace with SIC heating elements, shown in Figure 3.6. Partial sintering hardens the material but also allows an opportunity for machining the ceramic. The exit diameter (~ 0.4 mm) is drilled and the tip of the outer shell is sanded to form a 50 rounded tip. The specimens are then fully sintered at 1475°C for 4 hrs. During this process the specimens decrease in size, approximately 17% less than the original diameter. The ceramic particles exhibit “necking” or bridging between surfaces of adjacent particles, creating a denser overall nozzle. After full sintering no machining is possible for the ceramic nozzle. The final outer diameter of the nozzle is ~8.5 mm. (a) (b) (C) Figure 3.6 Photographs of: (a) tumbler used for mixing powders and for shaping graphite, (b) axial press used for pressing powder in the die, and (c) digitally controlled furnace used for sintering specimens. Several nozzles have been fabricated. The nature of the fabrication process does not produce Identical nozzles; each nozzle may be slightly different due to graphite shifting during pressing. Therefore, the best designs (axially aligned convergence path) are chosen for experimental analysis. The nozzle fits 51 snug onto the end of the inner conductor and epoxy glue is used as a sealer to prevent gas leakage. A photograph of a ceramic nozzle attached to the torch is shown in Figure 3.4. An optional step in the manufacturing process of the ceramic nozzle includes gold coating or sputtering of the ceramic’s outer surface. The ceramic seems to absorb the microwaves that the flame should absorb. Therefore, sputtering the ceramic nozzle with a metal (gold in this case because of its high temperature endurance) cause the microwaves to converge in the flame zone at the exit of the nozzle. An approximately 20 um-thick gold coating is sputtered onto the surface using a sputter machine. The ceramic nozzle shown in Figure 3.4 is sputtered with gold. 3.2.2 Electrical network and cavity Figure 3.7 provides a schematic of the overall experimental setup. The torch operates in plasma-only, combustion-only, and hybrid, plasma-enhanced combustion modes. For plasma production, power is supplied by 2.45 GHz microwave energy. The power absorbed by the system is measured by the difference between incident and reflected power, and is defined at the coaxial input line termination of the PTB. Both incident and reflected power are measured by 50-dB and 30-dB directional couplers, respectively, and are read using HP 435A power meters connected to HP 8481A thermistor power sensors. The reflected signal is fed into a 50-ohm matched load or dummy load (coaxial resistor model 8201 ). 52 Microwave I TOP Cavity Plate Incident Applicator - Camera/OES Microwave Power I GPIB Computer Reflected Power . Water Dumm Gas. ' . Loaii fuel 8. Cooling oxygen Figure 3.7 Overall experimental setup, including electrical circuit, torch and cavity, gas system, and diagnostic setup [3]. The microwave input power is side-fed into the torch applicator. The system is optimally tuned by repositioning the torch so the reflected power reads zero. Therefore, under these conditions, the incident power equals the power absorbed inside the cavity. The outer conductor and the inner conductor are independently adjustable. Tuning is also available by moving the top plate of the cavity in the vertical direction (see Figure 3.8). The coaxial inner and outer brass conductors allow an electric field to form in the pattern of a TM012 (transverse magnetic) mode inside the cavity (also shown in Figure 3.8) [18]. As a result of optimal tuning, the nozzle tip is positioned so the flame burns in a concentrated area of the electric field while absorbing microwaves. The cavity is sealed for safety to eliminate microwave radiation to the environment. The two viewing windows are covered with metal screening. 53 Figure 3.8 shows a schematic of the torch inside the cavity. The torch is placed vertically inside the cavity and held in place by finger stock. The cylindrical cavity is 7 inches (17.78 cm) in diameter, with a maximum height of 23 cm. The optimal operating height for this project’s experiments is 14.2 cm. Eleven millimeters of the outer conductor is inside the cavity and an additional 7 mm of the inner conductor is also inside the cavity, as shown in Figure 3.8. However, constant tuning of the torch/cavity is necessary for optimal interaction of microwaves and flame; as a result, these dimensions are not always exact for optimized interaction (i.e. zero reflected power). This inconsistency is caused by varying flame size, power input, and power-meter drifting. 54 \\\\ y ’/ \ 7 5f? / / - 5 Z / Cav1ty Z Z Z Top plate wall Z Z Z Z Z WWW. 7//////////////////.’< Z n 7 mm Z TM012 Z + 5 "‘09 Z 11 mm 2’ / / / . 6 % out3lde 5 LS Z . Z E-field Z cav1ty Z pattern Flame Z X 5 ’ , fl 5 1 Z g I 'I .4 % Zlmzxxxx’Z/zg . 7//.¢¢;¢7:¢¢2¢¢¢%|% f”? Input TORCH I I microwave 51 power Water .. C00"“9 Gas line Figure 3.8 PTB and microwave cavity setup, including details of the cavity and the electrical field pattern and wave modes. The torch is placed vertically in the cavity and tuned so the flame interacts with the microwaves, creating a plasma- enhanced hybrid flame. The dimensions of the cavity-penetrated torch are given. 3.2.3 Gas system The gases used for combustion are methane (CH4) and oxygen (02). Methane in 02 produces a relatively high theoretical adiabatic flame temperature (Tab = 3000 K), and yet is safer than hydrogen. It is readily available as “house gas” in the lab. Oxygen, as opposed to air, is used to produce a hotter, more 55 concentrated flame. The fuel and oxygen are premixed approximately 3 ft. before the nozzle exit plane. The gas flow is controlled by MKS digital mass flow controllers (MFC). The MFCs are calibrated for each type of gas, and the readout is 10.001 sccm accurate. A schematic of the gas system is shown in Figure 3.9. Gas flows from pressurized tanks to the MFCs through 1/4” flexible metal gas lines using VCR connectors. (1:331 _ Microwave cavity (V |Ml=c MFC NFC IMFC n I I ._.L_. A 03 C E) w O2 CH4. N2 H2 P micro Torch Water— coofing Premixed gas line Figure 3.9 Schematic of the gas setup. Gas flows from the compressed tank, through MFCs. It is then premixed before entering the torch. Methane (CH4) and oxygen (02) are the primary gases used in the experiments. The microwave power is fed through the side of the torch applicator (Pmicm) and the water cooling lines are also shown. 56 The gases are mixed downstream of the MFCs where a single 1/4" plastic tube replaces the multiple gas lines. This single gas line is fitted to a 1/16” plastic tube which is connected using epoxy glue to the torch gas inlet (1/16” brass tube) at the base. Gas then flows through the inner conductor to the nozzle and is ignited at the exit of the nozzle using a standard butane lighter. Flow rates from the nozzle range from 15 sccm to 309 sccm (2 m/s - 41 m/s) at the nozzle exit for the 0.4 mm-diameter brass nozzle. Reynolds numbers range from 1 to 50, using a constant 02 kinematic viscosity of 4 cm2/s (value interpolated at 2000 K). Table 3.2 lists the maximum flow rates in volumetric units (sccm) and length units (m/s) for each nozzle type. The ceramic nozzle allows a slightly higher maximum flow rate than the brass nozzle of the same diameter. This is most likely due to stand- off and temperature gradients between the flame and the nozzle. Table 3.2 Nozzle flow limits. Maximum flow rates for each type of nozzle, in sccm and m/s. Note the ceramic nozzle allows a higher maximum flow rate than the brass nozzle with the same diameter. Brass Brass Ceramic Nozzle Type 0.2mm 0.4mm 0.4mm diameter diameter diameter Max Total Flow Rates 75,10 309,41 321,43 (sccm or m/s) 3.3 Diagnostics and measurement techniques Experimental techniques used to measure temperature and diagnose the hybrid flame/discharge include: 1) optical emission spectroscopy (OES), used for identifying species and for calculating average rotational temperatures; 2) an infrared (IR) camera, used to examine relative temperature values and contours; 57 and 3) a type K thermocouple with a data acquisition system used to find local temperatures across the flame. The first two methods are most useful for the hybrid experiments because they do not perturb the flame since there is no physical contact. Even though a thermocouple is useful for obtaining point temperature measurements, it interferes with the electric field created by the microwave system. The cavity has two windows that are used for viewing and diagnostics. 3.3.1 Optical emission spectroscopy (OES) OES is used to measure and analyze the photons or light emitted by the flame/discharge. Its primary function for this project is to calculate the gas temperature of a particular gas-phase species, nitrogen and carbon rotational temperatures in this case, which has been performed in a number of previous experiments [21,22,23,24]. The rotational temperature is essentially a measure of the energetic level of the molecule at which all rotational modes of the molecule are fully engaged. The theory of gas kinetic temperature is explained in Chapter 4. This temperature is a Iine-of-sight, average temperature. Flame species can also be identified using OES since different atoms and molecules emit unique wavelengths in the spectrum; however, a quantitative value (i.e. emission intensity) is not known from this spectrometer. The McPherson Model 216.5, 0.5 meter, f/8.7, plane grating monochromator is the OES system used. The operating wavelength range is 1050 A — 5000 A, from a grating of 2400 grooves/mm. A schematic of the setup is shown in Figure 3.10. Light is emitted from the flame/discharge and focused by 58 a biconvex lens (focal length of 5 cm) into the spectrometer. The adjustable entrance slit is set to a width of 50 pm. The monochromator diffracts the incoming light. Since the angle of diffraction varies with wavelength, only specific wavelengths exit and are measured. The monochromator is manually set to a specific wavelength by rotating the diffraction grating about the axis at a specific angle. The signal of species present (light based on wavelength) is amplified by a photomultiplier and read by the picoammeter to certify that ample signal is being captured. The system is covered by a black cloth to reduce any external light signal, thus enhancing the signal—to—noise ratio. The spectrometer is connected to a computer by a GPIB (general purpose interface bus) cable and data is collected from a Quick Basic program written by Jayakumaran Sivagnaname [22] to extract wavelengths of various flame/discharge species. 59 Cavity V Monochromator (:23, _. A ...... 1---- ............ Entraneex Focusing 3'” Flame/ I lens discharge L :' Grating High voltage Computer input . Exit J Plcoammeter """ slit """""" ' .='= EPIB Photo Collimating multiplier mirror Figure 3.10 Schematic of OES system. Light is emitted from discharge, analyzed in the monochromator, and collected and organized in the computer. The system (particularly the torch and lens) is aligned so a maximum signal is fed into the entrance slit. The entire image of the flame is focused through the lens, supplying data for an average overall flame temperature. Rotational temperatures calculated from the OES method come from C2 and N2 emissions. 3.3.2 Infrared camera Infrared (IR) cameras are useful for nonintrusively measuring temperatures without directly contacting the sample (i.e. flame), unlike intrusive thermocouples that can cause flame perturbations and conduct heat away from the flame. A major obstacle or limitation of the IR temperature analysis, however, is the inability to measure the flame while inside the cavity. The cavity's viewing window is protected with a metal screen, to prevent microwave leakage. The 60 camera cannot bypass this; because it measures the first object in sight, the cooler metal screen is detected rather than the flame inside the cavity. A metal stand or platform was built to prop the torch in either the vertical or horizontal direction, for operation outside the cavity. With the torch in an open, unprotected atmosphere, the disadvantage is it cannot be optimally tuned to couple the microwaves with the flame due to leakage. IR measurements were collected from the flame while the torch was outside the cavity. The camera measures the radiation an object emits, since an object, not at absolute zero temperature, will always emit or absorb energy In the form of electromagnetic radiation. A blackbody, the ideal object with an emissivity (a) of unity, absorbs and emits energy with no reflection. Examples include the sun, a candle flame, and even the flame of a welding torch! Even though torch flames have almost identical properties and characteristics as the flame used in this research, the process for measuring temperature with the IR camera is not so simple (1: #1). Because there are hot gases surrounding the flame, the emissivity is no longer unity; it is a function of temperature, 5(7), and in addition the temperature varies with location. Thus, IR measurements of gas temperatures are much more complicated than surface temperature measurements. The Stefan-Boltzmann law, defined in Equation (3.1) below, is used to measure the energy, R( T), emitted from a surface relative to a black body, where the energy is a function of temperature and proportionally related to the Stefan- Boltzmann constant, 033, defined in Equation (3.2). 61 R(T) = £O'SBT4 12 m (3.1) 2 5k 4 _ 053: ’r 332 =5.67x10 8 2W4 15h c m K (3.2) The Stefan-Boltzmann constant is defined as a combination of three other universal constants, the Boltzmann constant, k3, Planck’s constant, h, and the speed of light, 0 [25]. IR detectors are commonly used to measure temperature because for objects at temperatures of 10-5000 K, the radiation predominantly occurs between wavelengths of 0.2-500 pm in the IR spectrum. IR cameras monitor the temperature from an array of IR sensors, and each sensor represents a pixel in an acquired thermal image [26]. The IR camera used for this research comes from Mid-range Merlin from Indigo Systems Company, a division of Flear International. It has interchangeable 25 mm and 50 mm lenses, has focal length adapters for distant viewing, and contains filters used for higher temperatures. Windows 2000 and Therrnagram make up the computer’s operating system and software. Further specifications are available at http://www.indigosystems.com/product/merlin.html. The camera uses a cooled InSb detector array and is designed for use with 3-5 pm wavelengths, corresponding to temperatures of approximately 273- 1000 K. The temperatures of the flame/discharge studied in this project produce a temperature beyond that range (2000-3000 K); therefore, only relative 62 temperatures are extracted from the analysis, used to compare combustion-only and hybrid flames. Temperature contours throughout the flame body are also obtained. 3.3.3 Thermocouple A type K thermocouple is used to measure flame temperature (combustion only). Type K Chromel-Alumel thermocouples are rated for a long range capacity of 173-1643 K. As mentioned previously, combustion produces a much hotter flame than the thermocouple is capable of measuring. Therefore, the solution to the transient heat conduction equation, assuming the lumped heat capacity method (Biot number < 0.1), was used to calculate temperature based on time- step iterations, and can been seen in Equation (3.3) below. Here T(t) is the temperature as a function of time, To is the initial temperature (room temperature in this case), Ts is the surface temperature, h is the heat transfer coefficient, A is the cross-sectional area, p is the density, CD is the specific heat, and V is the volume, all with respect to the thermocouple material or junction. T(t)—TS —= — / V T0_TS exm (hA pcp )t (3.3) The thermocouple is welded together to form a spherical junction. It must be noted that the junction is not shielded (may account for inaccuracies in results). Because the sphere is so small (1.7 mm diameter), the lumped capacity method is implemented. The heat transfer coefficient is calculated based on the Whitaker Nusselt (Nu) number correlation for a sphere, seen in Equation (3.4) 63 and a general definition in Equation (3.5) below. This correlation if valid for Reynolds numbers between 3.5 and 7.6e4 and Prandtl numbers ranging from 0.71 to 380 [25]. The Reynolds number is denoted by Re, where Re = Ud/v and U is flow velocity, dis diameter, and v is kinematic viscosity. The Prandtl number, Pr, is defined as Pr = v/a where a is thermal diffusivity. Dynamic viscosity, (1, is the viscosity of the fluid at Too, and us is the dynamic viscosity of the fluid at the surface evaluated at the film temperature, T), the average temperature of the solid and fluid, L is the characteristic length, and k; is the thermal conductivity of the fluid (02 gas in this case). All parameters, except #3. are evaluated at Too. Nuavg = 2 + (0.4 Re0-5+ 0.06 Re” 3) Pr0'4(,u / 1190-25 (3 4) hL Nuavg = — kr (3.5) An Excel spreadsheet is created with known constants and used to calculate the temperature for a 0.5-second time step. Reynolds numbers ranged from 3-375, based on a constant 3.5 m/s flame speed (calculated from cone angle) and temperature-dependent O2 kinematic viscosities. This spreadsheet is found in Appendix A. 64 A motorized device built enables the thermocouple to sweep through the flame at a constant speed, while the data acquisition system reads the voltage difference and converts to temperature. A Hewlett Packard (model # 3852A) data acquisition system/control unit, connected to a computer with a GPIB cable, is used. The time delay for this system is 0.1 3. Photographs of the control unit and motorized device are shown in Figure 3.11. Figure 3.11 Left: photograph of the data acquisition system used to measure temperature with a Type K thermocouple. System is connected to a computer with a GPIP cable. Right: device built to sweep thermocouple through flame at a constant speed. Note the spherical thermocouple junction. 65 Chapter 4 Theory and Background of Gas Rotational Temperatures 4.1 Introduction Flame/discharge temperatures can be approximated by measuring gas temperatures. The gas temperatures are calculated by measuring the rotational energy levels of various molecules present in the flame/discharge. The homonuclear, diatomic molecules that are measured in this study are nitrogen (N2) and carbon (C2). Rotational temperatures of these gases are calculated based on signal (wavelength) measured by optical emission spectroscopy (OES). 4.2 Rotational temperature theory The total energy of a given state of a diatomic molecule is the summation of electronic energy (Te), translational energy (Tt), vibrational energy (G), and rotational energy (F), as written in Equation (4.1) in wave-number units. T=Te+T,+G+F (4.1) In general, rotational energy changes in a given vibrational and electronic state are small compared with the thermal translational energy, therefore making F a small number. Conversely, gas molecule collisions produce changes in the vibrational or electrical quantum numbers much less frequently than in rotational quantum numbers precisely because the rotational energies are lower than the others and changes are easier to excite [27]. The rotational temperature is a reliable measure of the gas kinetic temperature because the rotational population 66 distribution in a sufficiently long-lived vibrational state has a Boltzmann distribution [27,28,29,22]. The rotational energy listed in Equation (4.1) is defined in Equation (4.2) below in terms of J, the rotational quantum number which takes the values 1, 2, 3,..., By, the rigid rotator rotational spacing or rotation radius, and By, the first anharmonic correction to the rotational spacing. F = BvJ(J+l)—D,,J2(J+l)2 +... (4.2) In addition, there are nonrigid rotator corrections to both 3,, and DV. These corrections are defined in Equations (4.3) and (4.4) below, where v is the vibrational frequency, 8,, and 0., are constants that correspond to the equilibrium separation, and 0,9 and [3,, are the first anharmonic corrections. Constants for the first order rotational energy are listed in Table 4.1 below for N2 and C2 [27, 22]. B, = Be —ae(v+%)+... (4.3) D, = De + ,Be(v+%) + (4.4) Table 4.1 Rotational constants for the electronic states of nitrogen and carbon. State B. a. N2 (C3pu-B3pg) 1.8259 0.0197 02 (d3pg-a3pu) 1.7527 0.01808 67 Each molecule is denoted by a specific quantum state. For this research the Second Positive System (Capu—tBapg) is used for N2 values and the Swan System (danga3pu) is used for C2 values. These systems are now explained. The upper (u) and lower (9) states may have different electronic angular momenta, A. Therefore, two or three series of lines or branches may appear: the P, Q, and R branches. Shown in Figure 4.1 are these three branches, which result from transitions in vibrational and rotational energy levels. Q Branches v“ I A . Figure 4.1 Energy level diagram for a band with P, Q, and R branches, relative to wavelength (A), rotational energy levels (J), and vibrational levels (v) [27]. 68 For both N2 and C2, AA = 0 because the angular momentum at both the upper and lower electronic states is zero; therefore, the AJ = 0 transition is forbidden and it follows that AJ = :l: 1. The R branch is associated with AJ = +1 and the P branch is associated with AJ = -1. The Q branch is not present for N2 and C2 because AA = 0. It is important to note, however, that all branches may be present for a specific electronic transition with a different angular momentum, as shown in Figure 4.1 [27]. Generally, emission methods, as opposed to absorption methods, are suited for measuring the temperature in a Boltzmann distribution. The relative emission intensity (I) of rotational lines is described as Equation (4.5) by [30]. This intensity is not related to size. (4.5) B. ' '+ 1:16.48). J. exp[— VJ” 1W] kT, The intensity is defined by: K, the constant for all lines originating from the same electronic and vibrational level; v, the frequency of the radiation; 8,1211, the appropriate H6nI-London factor; 8.1, the molecular rotational constant for the upper vibrational level; J, the rotational quantum number (J’ represents the upper energy level and J” represents the lower level) ; h, Planck’s constant; c, the speed of light; k, the Boltzmann constant; and T,, the rotational temperature. It is seen from Equation (4.5) that the intensity is proportional to the frequency, v, to the fourth power and also to the exponential ratio of rotational temperature, T,. As the wavelength of an emitted species decreases, the 69 intensity increases to the fourth power; frequency is defined as the speed of light divided by wavelength, v = cl). As temperature increases, so does the intensity exponentially. The H6nl-London factor describes the line strength of rotational spectra, and it is dependent on J. The H6nI-London formula is defined in Equation (4.6) for emission associated with the R branch, where AA = 0, [27]. = (J'+A'+1)(J'—A'+I) S 4.6 J '+1 ( ) Since AA = 0, Equation (4.6) reduces to Equation (4.7) below. S = J '+1 (4.7) The experimental data is fitted to the expression / ~ SJuexp{- BVJ’(J’+1)hc/(kT,)}, using several (11 or more) R-branch emission lines to determine the rotational temperature, T,. The plot of In(IIS) is a linear function of the upper rotational energy for the diatomic molecules used in this study. Chapter 3 describes the experimental procedure for measuring the spectra of various flame/discharge species, which are used for the temperature calculation. Figure 4.2 pictorially shows how the light signal is emitted from the flame/discharge, and then focused through a lens into the spectrometer. 7O Flame/ Discharo e Spectrometer Figure 4.2 Laboratory process for temperature calculations: light is emitted from the flame/discharge and focused by a lens into the spectrometer, where intensities and wavelengths are recorded. 4.3 Gas temperature of N2 Rotational temperatures are first calculated based on the N2 spectrum, since this molecule has been extensively measured [21]. Nitrogen is measured in two different sources. In the first set of experiments, only methane and oxygen serve as the feed gases; N2 is detected from the atmosphere (air) as the flame/discharge reacts. In a second set of experiments, a small amount (~ 2%) of N2 is fed into the incoming methane and oxygen streams. Parameters that are varied in each experiment include microwave input power and gas flow rate. The system currently operates at atmospheric pressure. The Second Positive System (SPS) is used to determine the N2 rotational temperatures. The SPS describes the energy level transition from Capu to Bapg. Figure 4.3 is a plot of an N2 spectrum, measured in the laboratory, identifying 11 emission lines (R20 — R30) in the spectrum range 3758 A - 3783 A of the (2,0) transition in the SPS. 71 Intensity (a.u.) I r T 3753 3763 3773 3783 3793 Wavelength (A) Figure 4.3 Plot of intensity (or current since the units are arbitrary) versus wavelength for the N2 spectrum, commonly used for rotational temperature calculations. R20 - R30 list the order of the rotational bands. Table 4.2 provides values used in the rotational temperature calculations. These values include the rotational lines/bands (R20 — R30), wavelengths recorded from text (the wavelengths gathered from the experiments of this thesis are compared to book values to ensure accuracy), relative upper-level energies calculated from Equation (4.2), and the H6nI-London factor calculated from the corresponding rotational band. 72 Table 4.2 Values used for the N2 rotational temperature calculations for the R branch of the (2,0) SPS. Rotational Wavelength Relative upper-level S Line (A) energy (cm-1) R20 3780.44 837.76 19.80 R21 3778.58 917.42 20.81 R22 3776.66 1000.67 21 .82 R23 3774.68 1087.51 22.83 R24 3772.64 1177.95 23.83 R25 3770.53 1271 .97 24.84 R26 3768.37 1369.57 25.85 R27 3766.14 1470.76 26.85 R28 3763.86 1575.51 27.86 R29 3761.51 1683.84 28.86 R30 3759.1 1 1 795.74 29.87 The natural log of the ratio, l/S, is plotted against the relative upper-level energy, and the gradient of a linearly-fitted line is used to calculate the rotational temperature in Equation (4.5). Figure 4.4 displays an example of a Boltzmann plot for the R20 — R30 bands. This plot is created from a set of experimental flame/discharge data with a methane/oxygen flow rate of 80/160 sccm combined with 40 W microwave input power. Rotational excitation N2 temperatures are calculated even though N2 is not present in the gas flow; the flame reacts with N2 found in the surrounding air, similar to a diffusion flame. 73 R20 R30 -2.5 1 . 1 700 1050 1400 1750 Relative upper level energy (cm-1) Figure 4.4 Boltzmann plot for the bands R20 — R30 of the N2 spectrum. The accuracy of rotational temperature is within 1 100 K, as estimated from the reproducibility of the data obtained. 4.4 Gas temperature of C2 Rotational temperatures of C2 are calculated in a fashion similar to N2 rotational temperatures. The carbon source is methane, CH4, the fuel used for combustion in this thesis. The light emitted from the breakdown of CH4 eventually to C2 and the subsequent C2 excitation is processed in the spectrometer and the wavelengths and intensities of the C2 spectrum are extracted and plotted. The Swan System (d3nga3pu) is used for calculating the C2 rotational temperatures [24]. The emission intensity is measured using the (0,0) band head, with an intensity of 513 nm. Figure 4.5 is a plot of a C2 spectrum, Identifying 21 74 emission lines (R25 — R45) in the spectrum range 5028 A — 5089 A of the (0,0) transition in the Swan System. (0.0) i 3 R45 R25 '1 \ \ 4999 5019 5099 5059 5079 5099 5119 5139 Wavelength(A) Figure 4.5 C2 spectrum of a Swan System, identifying the bands R25 — R45. Values used to calculate C2 rotational temperatures are listed in Table 4.3 and are obtained as described in Section 4.3 for N2 rotational temperature. Note the difference in the H6nl-London factor, defined previously by Equation (4.6), takes a new definition for the C2 analysis [24], written in Equation (4.8) below. 5 = (J'+1)(J'—1) J' (4.8) 75 Table 4.3 Values used for C2 rotational temperature calculations for the R branch of the (0,0) Swan System. Note the change in the H6nl-London factor. RotLaiggnal Wav(eAl?ngth Reéitggyuafiglfvel S = (J'+11(J"11’J' R25 5089 564.40 24.96 R26 5086 609.55 25.96 R27 5084 656.44 26.96 R28 5081 705.07 27.96 R29 5079 755.43 28.97 R30 5076 807.53 29.97 R31 5074 861 .36 30.97 R32 5071 916.94 31.97 R33 5068 974.24 32.97 R34 5066 1 033.29 33.97 R35 5063 1 094.07 34.97 R36 5060 1 156.59 35.97 R37 5057 1220.84 36.97 R38 5054 1286.84 37.97 R39 5051 1354.56 38.97 R40 5047 1424.03 39.98 R41 5045 1495.23 40.98 R42 5042 1 568.1 7 41 .98 R43 5038 1642.84 42.98 R44 5035 1719.25 43.98 R45 5031 1 797.40 44.98 As with the N2 calculations, the natural logarithm of l/S is plotted against the relative upper-level energy and is linearly fit to extract a rotational temperature from the slope of the line. An example of this plot is shown in Figure 4.6, gathered from experimental data using the ceramic nozzle, having methane/oxygen flow rates of 60/180 sccm, and no microwave power present. This is a pure flame, not a flame/discharge. The uncertainty of the measured gas temperature is approximately 100 K based on reproducibility. 76 -2.5 .26 1 .27 . -2.8 - R25 -) R45 .29 - In(l/S) {3.05 (3.1 - -3.2 ~ -3.3 1 I f 1 1 r 5(1) 700 fl) 11(1) 111) 1500 17(1) 191) Mafive mper-Ievel energy (cm-1) Figure 4.6 Boltzmann plot for the bands R25 — R45 of the C2 spectrum, for the 180/60 sccm O2/CH4 pure flame. Results from OES measurements of the rotational temperatures for N2 and C2 molecules are tabulated and discussed in Chapter 5 of this thesis. 77 Chapter 5 Results and Discussion 5.1 Introduction The results in this thesis are collected primarily from experimental procedures. Much preliminary work was done building an expandable foundation for this research because plasma-enhanced combustion is not a fully developed and clearly understood area of research. The research takes two approaches: 1) studying the fundamental concepts of plasma-enhanced combustion phenomena and understanding the underlying effects of such a hybrid system; and 2) engineering a practical torch/burner to be used in industry. Results collected and evaluated include flame/discharge photographs, geometric measurements which account for flame speed and power density calculations and non-dimensional power plots, extinction (blow-out) data, species data from OES, and temperature measurements using different methods. Data are collected from both brass and ceramic nozzles; however, focus is on data from the 0.4 mm-diameter brass nozzle. Where the nozzle is not specified, it is safe to assume the 0.4 mm-diameter brass nozzle is used. Images in this thesis are presented in color. 5.2 Flame/discharge photographs The torch has gas flow limitations it cannot exceed to produce a flame, depending on which nozzle is used. Table 5.1 lists the maximum total flow rates for three nozzles: 1) 0.2 mm-diameter brass nozzle; 2) 0.4 mm-dlameter brass nozzle; and 3) 0.4 mm-diameter ceramic nozzle. These limitations are recorded 78 based on stoichiometric (2:1 sccm) methane/oxygen flow rates, stoichiometric referring to the ideal combustion process during which a fuel is burned completely. The minimum flow rates are not measurable due to limitations of the mass flow controllers (MFCs) i.e. the flame is maintained at a low flow rate of 18/9 sccm O2/CH4, and when the rate falls to 9 sccm and below the MFCs do not measure accurately. Table 5.1 Gas flow limitations for three nozzles with units in sccm and m/s. Brass Brass Ceramic Nozzle Type 0.2mm 0.4mm 0.4mm diameter diameter diameter Max Total Flow Rates 75 ’ 10 309 ’ 41 321 ’ 43 (sccm or mls) Several photographs are taken, using a digital Nikon D70 Outfit camera, of flame configurations using the three nozzles mentioned earlier. Figure 5.1 includes photographs of pure combustion flames with varying flow rates, using the 0.4 mm-diameter brass nozzle. As the O2 flow rate increases for a fixed fuel rate, the flame decreases in length. The length is increased with a higher overall CH4 flow rate, but still follows the trend that as 02 increases, flame length decreases. Table 5.2 lists the flame lengths for each flow rate. The stoichiometric flame parameters are highlighted in the table. As the overall stoichiometric flow rate increases, the flame length also increases from 3.7 cm, to 5.0 cm, and finally to 6.5 cm. As the total flow rate increases by 33% from 200 sccm to 300 sccm, the length increases by 43% from 3.7 cm to 6.5 cm. The flame length increases almost linearly with an increase in total stoichiometric flow rate, yielding a coefficient of determination (R2 value) of 0.9903. Notice that as the fuel rate 79 exceeds the 02 rate, the flame front increases in size. The flame front is the outline of the white inner core of the overall flame. The blue volume beyond the flame front is the heat released from the chemical reaction. O23 30 65 100 135 170 200 CH4: 5, Figure 5.1 Photographs of pure combustion flames with varying flow rates, measured in sccm. The 0.4 mm-diameter brass nozzle is used. 80 Table 5.2 Lengths of flames with varying flow rates, corresponding to Figure 5.1. The shaded cells highlight stoichiometric flow rates. Mass flow rate (sccm) Length of Flame Methane (C H.) Oxygen (02) (cm) 65 30 5.0 65 65 4.5 65 100 4.0 65 135 3.7 65 1 70 3.2 65 200 3.1 85 30 6.5 85 65 6.0 85 100 5.7 85 135 5.2 85 170 5.0 85 200 4.7 100 30 9.6 100 65 8.4 100 100 8.0 100 135 7.5 100 1 70 7.2 100 200 6.5 Figure 5.2 provides photographs of a pure plasma made from 200 sccm argon (Ar2). The plasma discharge is much smaller in length and diameter than a pure combustion flame. The temperature is also much lower than a flame temperature (~1800 K vs. ~3000 K), as will be discussed later in this chapter. Notice that the torch is not confined within the microwave cavity. Microwaves do not escape the torch when the inner conductor is kept inside the outer conductor and the microwave power is held at a relatively low value (10 — 40 W). The plasma discharge, when inside the cavity, can become very long (> 10 cm) when the torch is tuned and the microwave power is increased. 81 Figure 5.2 Photographs of a pure plasma discharge, with a flow rate of 200 sccm argon and 30 W microwave power. Note that the torch is not confined within the cavity. Photographs in Figure 5.3 show how a pure combustion flame is affected or altered by the presence of microwaves. A stoichiometric flow of 40/20 sccm O2/CH4 is burned as 40 - 100 W microwave power is added, shown in the bottom row of photographs. As 40 W microwave power is added to the flame, the torch is adjusted and tuned inside the microwave-sealed cavity (the metal screen used to prevent microwave leakage from the cavity’s window is visible in the photographs) to trigger optimal flame/microwave interaction. The flame transforms into a hybrid discharge as it changes color and increases in size (four times its size from pure combustion to a 100 W-hybrid flame/discharge). The position of this optimal tuning is referred to as the “sweet spot," where also the reflected power of microwaves is zero. The top row of photographs in Figure 5.3 shows flame/discharges with a flow rate of 40/0 sccm O2/CH4 and with 40 — 100 W microwave power. In other words, an oxygen plasma is present. Once the flame transforms from the pure combustion form Into a hybrid flame/discharge, the fuel is reduced to zero flow. 82 The fuel-free flame/discharge is smaller (5 mm shorter in length at 100 W) than that which burns fuel. 40/20 s 300 Figure 5.3 Photographs of flame/discharges inside the cavity as 40-100 W microwave power is added. The 0.2 mm-diameter brass nozzle is used. The flow rate for the top set of photos is 40/0 sccm O2/CH4, i.e. no fuel is present. The flow rate for the bottom set is 40/20 sccm O2/CH4, a stoichiometric ratio. Scales are provided for flame dimensions. Camera speed and aperture are 1.6 and 3.5, respectively. Photographs of a pure combustion flame and microwave-induced flame/discharges formed by a ceramic nozzle are shown in Figures 5.4 and 5.5. The flow rate of the flames in Figure 5.4 is 200/100 sccm O2/CH4, and 20 W, 40 W, and 80 W microwave power is added. The photographs do not show evidence of a plasma-enhanced combustion flame; the color, shape, and size of the original flame are not affected by microwaves even with fine adjustment (zero reflected microwave power), from what the human eye can detect. However, OES measurements show that the flame is indeed altered by microwaves and will be discussed later in this chapter. 83 The ceramic nozzle may absorb the majority of the microwaves, rather than the flame, or the gases might be excited by the microwaves upstream the nozzle. At this point an explanation is uncertain. Figure 5.5 shows that even though the flame Is not physically altered by microwave power, the flame increases in length as the total flow rate Increases. O W 20 W 40 W 80 W ZOO/100 S - l l l C 3.55 c m cm Figure 5.4 Photographs of flame/discharges inside the cavity as microwave power is added, 20-80 W. The 0.4 mm-diameter ceramic nozzle is used. The flow rate is 200/100 sccm O2/CH4, a stoichiometric ratio. A scale is provided for flame dimensions. Camera shutter speed and aperture are 1.6 s and 3.5, respectively. 80/40 80/40 140/70 140/70 200/100 200/100 Figure 5.5 Photographs of flame/discharges inside the cavity with a ceramic nozzle comparing 0 W to 80 W of microwave power. The sets of photographs in Figure 5.6 show very clearly how the microwaves alter a pure combustion flame, formed with the 0.4 mm-diameter brass nozzle. The flow rates are listed in the figure for each set, and a scale is provided to estimate flame/discharge dimensions. It appears that the microwaves start to dominate after 30 — 40 W microwave power is added to the flame. To further investigate, photographs are taken at 1-Watt increments from 31 W to 40 W, and then again from 39 W down to 33 W, shown in Figure 5.7. As power increases the flame transforms at 38 W; however, as power decreases, a major change does not occur until 35 W. A hysteresis is present. There appears to be a delay in the change of state (transition from hybrid form to combustion form) when the power is decreased. Further theoretical investigation for an explanation of this behavior is advised. 85 ANOJ-bmmV 10W20W 30w 40w 50w 60W 70w 80W 90w100w cm 70 sccm CH. 53 sccm O2 1 CW 10W 20W 30W 40W 50W 60W 70W 80W 90W 100W Figure 5.6 Photographs of flame/discharges inside the cavity as microwave power is added, 10-100 W. The 0.4 mm-diameter brass nozzle is used. The flow rate for the top set is 70/24 sccm O2/CH4. The flow rate for the bottom set is 153/70 sccm O2/CH4. A scale is provided for flame dimensions. Camera shutter speed and aperture are 1/30 s and 3.5, respectively. 86 31W 33W 35W 37W 38W 39W 40W 39W 38W 37W 36W 35W 34W 33W Figure 5.7 Additional photographs from Figure 5.6. The flow rate is 70/24 sccm O2/CH4 and all other conditions are the same as listed in Figure 5.6. The top sequence shows the flame/discharge as microwave power increases; the bottom sequence is power decreasing. This intermediate behavior gives rise to a hysteretic affect. Flame velocity, or flame propagation speed, is calculated from the geometry of a flame. The cone angle, a, is the angle between the center of the flame and an edge, when the flame is modeled as a two-dimensional cone and assumed axisymmetric. This angle is measured, and along with the flow rate of gases that is measured, u, these measurements are used to calculate flame 87 speed, 82, as defined in Equation (5.1). Flame speeds are calculated for the flames in Figure 5.7, as the microwave power decreases from 39 W to 33 W, and are listed in Table 5.3. As microwave power decreases, the flame speed also decreases. Significant change occurs at 35 W when the flame/discharge transforms from hybrid form to combustion form. Most probably microwaves excite the species in the combustion process at a higher level and accelerate the radicals/ions at higher rates. Typical book-value flame speeds for an O2/CH4 flame are ~ 100 cm/s. SL=usina (5.1) Table 5.3 Flame speeds calculated for flames shown in Figure 5.7 as power decreases from 39 W to 33 W. Calculations are based on Equation (5n. Cone Angle Microwave Power Flame Speed, SL 4099) (W) (cm/S) 12 39 259 12 38 259 9 37 195 9 36 195 3 35 65 3 34 65 2 33 44 5.3 Geometric analysis Along with flame speeds, deduced from flame geometry are quantities such as power density and non-dimensional power as a function of flame height. Power density is calculated first by modeling the flame/discharge as a cone to calculate a volume; it is defined here as the addition of microwave power and combustion power of a flame/discharge divided by the volume of the corresponding flame/discharge. Combustion power is approximated using the 88 higher heating value (HHV) tabulated in text for gaseous methane, based on stoichiometric combustion with air, a value of 55,528 kJ/kg, and a CH4 density of 0.68 kg/m3 [31]. Values of combustion power in this research range from 10 - 40 W. Power densities are plotted against non-dimensional power in Figure 5.8 for the two sets of data shown in Figure 5.6. Non-dimensional power is defined as PP/(Pp + PC), where Pp is microwave (plasma) power and PC is combustion power (PC is provided in the figure for each flow rate). With the absence of microwaves, power density is greater for the flame with a lower total flow rate, which is also slightly more fuel-lean. The flame volume is significantly larger for the second set of data (153/70 sccm O2/CH4). The power density is greatest just before the flame transforms in size and color, suggesting the flame/discharge of highest intensity is formed when ~ 30 W microwave power interacts with the flame. An intense and highly concentrated discharge is desired, but temperature must also be considered. As microwave power increases, temperature increases, as discussed in Section 5.5. Figure 5.9 shows how the power (both microwave and combustion) physically alters the flame in terms of geometry. Flame height is plotted against non-dimensional power. An exponential growth in flame height occurs as microwave power is added to the flame, for both sets of data. 89 —e— 70/24soom02/CH4 Flow 0.5 -~a~153/70$CUT102/CH4FIOW ‘7 0.0 0.2 0.4 0.6 0.8 1.0 PP/(PP'I'PC) Figure 5.8 Power density as a function of non-dimensional power, evaluated from flame/discharge geometries shown in Figure 5.6, with flow rates of 70/24 sccm O2/CH4 and 153/70 sccm O2/CH4. The powers of combustion are provided: 15 W for the first flow and 44 W for the second flow. F 8 7 1 o 70/24soanO2/CH4 Flow 0 6 l a 153/70soon02/CH4Flow ° 1 E 5 5 E] El El 0 E o o .m 4,, 0 2 n ° 3 I 0 ° 5 21 D U 0 1 9‘” 0 o o o 0 o E , 7, ., - , ---. E- . , , . 0.0 0.2 0.4 0.6 0.8 1.0 PP/(PPI'PC) Figure 5.9 Flame height as a function of non-dimensional power, evaluated from flame/discharge geometries, shown in Figure 5.4, with flow rates of 70/24 sccm O2/CH4 and 153/70 sccm O2/CH4. 90 It is possible to deduce that the combustion-like flame transforms to the hybrid flame/discharge form as the microwave power exceeds the power produced by the combustion process. The combustion power of the flame for the set of data with a flow rate of 153/70 sccm O2/CH4 is calculated as 44 W. The photographs in Figure 5.6 show that between 30 W and 40 W microwave power, the flame dramatically changes in appearance, to what resembles more closely a plasma than a flame. However, the combustion power for the second set of data (70/24 sccm O2/CH4) is ~ 16 W, but the flame does not transform at a microwave input power near this value. It must also be noted that the torch is very sensitive to positioning/tuning. The “sweet spot” is dependent on several factors: 1) cavity height; 2) inner conductor position relative to outer conductor position; 3) outer conductor position inside cavity; 4) gas flow rate; and 5) microwave input power. Because all these factors must be considered, there are no exact measurements for factors 1, 2, and 3. The experimentalist has a feel and general idea of how each should be positioned; however, fine tuning is always required at the start of each experiment to find the sweet spot. 5.4 Impact of microwaves on combustion To further show how microwaves impact a combustion flame, a flame extinction (or blow-out) plot is created and several OES scans of the flame/discharge are made. Flame extinction plots are powerful for analyzing combustion efficiency in terms of fuel efficiency. An extinction plot is made to show that, with the addition 91 of microwaves, a flame is maintained burning less fuel than is required when no microwaves are present. Figure 5.10 is an extinction plot of experimental data showing this very effect. The figure plots the equivalence ratio versus total volumetric flow rate. The equivalence ratio, rp, is a combustion term, defined as rp = (VF/V0x)/(VF/Vo,.)s,o,-ch, where V is the volumetric flow rate in sccm; (p = 1 is the case for stoichiometric flow, (p > 1 represents fuel-rich flow, and rp < 1 represents fuel-lean flow, the range plotted in Figure 5.10. The figure includes values of combustion power for each gas flow rate. The data is collected by first creating a stoichiometric flame, lowering the fuel slowly (considering delay of MFC responses) until the flame blows out, and then recording the flow rates at the point of extinction. The data points in the figure are the values of total flow rates at the point where the fuel flow rate causes flame extinction, while holding O2 flow constant. The equivalence ratio is calculated based on the fuel flow rate when blow out happens. The experiment is run several times to ensure accuracy of results; accuracy ranges from :I: 1 sccm. Three extinction curves are plotted: 1) combustion-only data; 2) hybrid data with 10 W microwave power; and 3) hybrid data with 20 — 100 W microwave power. With the addition of a minimal amount of microwaves (10 W), the flame continues to burn at a lower fuel flow, a flow that in not possible without the presence of microwaves. The average fuel flow rate reduction is 9%, ranging from 5 - 12%, not including the first data point in the first two curves. The sweet spot cannot be (easily) located at such a low flow rate; as a result, the microwaves do not affect the blow-out curve. At 10 W microwave power, the 92 flame/discharge does not completely transform in size and color as seen in previous data. Although no physical alterations are evident, the flame is affected by microwaves. When 20 - 100 W of microwave power is added to the flame, a physical transformation takes place and a “discharge” is maintained with no fuel present. The third curve in Figure 5.10 reflects this behavior. 0.8 5 0.7 . 0.6 a 0.5 1 Ratio 0.4 a 0.3 - 0.2 4 90/o a9 + Chrbustim ody —--e—— 10 W 0-1 1 .13.. 20100 w Eq 1. 8 00 r1 1:1 13 _m n - w— 1: w a ‘El 70 12) 170 220 270 VTor (scam) Figure 5.10 Flame extinction or blow-out plot of equivalence ratio versus total volumetric flow rate, VTOT. Each data point is provided with the combustion power of the flame at a particular flow rate, where fuel is present. Note that a discharge is maintained with no fuel when 20-100 W of microwave power is added. A spreadsheet of raw data can be seen in Appendix B. I! 5.4.1 Optical emission spectroscopy (OES) OES is used to detect species present in the flame/discharge, leading to gas temperature calculations, and is used as a tool to gain perspective on how microwaves affect a flame. The molecules/radicals scanned in this research include N2, C2, 02, OH, and CH, all commonly found in hydrocarbon combustion 93 processes. The OES equipment used in this research is not capable of measuring quantitative values of intensity inferring the amount of CH, for example, present (arbitrary units, a.u., are labeled in plots; the scale could be labeled as current rather than intensity); however, relative intensities provide insight. Photographs in Figures 5.4 and 5.5 in 5.2 show that a combustion flame formed by a ceramic nozzle is not physically altered by the presence of microwaves. OES analysis provides evidence that the flame may in fact be altered. Figure 5.11 is a plot of C2 emission spectra, detected in a combustion- only flame and also in a hybrid flame/discharge with 40 W microwave power. All variables but microwave power are held constant. After 40 W microwave power is added to the flame, the C2 emission intensity increases. Figure 5.24 in this section supports these results. The plot provides wide-range spectra for a combustion flame and a hybrid flame/discharge. Again, the Intensity is greater with the presence of microwaves, suggesting the microwaves excite these molecules/radicals at higher levels. 94 6. 5% 2 5.E09« -———Corrbtslion -~--~Hybfid(40V\0 2i 4. E-09 ~ 3. E—09 1 Intensity (a.u.) 25095 , .; . ., A ' I '. I I It . 2,111,135 '3: 5 .1251 «5.1-05:54:39 ," : - lz';"=: , 19'". Eli" ‘- ‘33" - 15094 .‘ lr- ‘ 1511.54, '. . 0.3m T I T I I T I I 1 5070 am 509) 5100 5110 5120 5130 5140 5150 516) VibvelengflHA) Figure 5.11 OES scan of C2 in a stoichiometric (200/100 sccm O2/CH4) combustion flame and hybrid (combustion plus 40 W microwave power) discharge using a 0.4 mm-diameter ceramic nozzle. The next set of plots, Figures 5.12 — 5.14, show N2 emission spectra in combustion flames and hybrid flame/discharges. The general trend follows that as microwave power increases, N2 emission intensity increases. Three independent sets of data are presented to show consistency within experimental results. Figure 5.14 shows intensities resulting from an N2-injecti0n flame (N2 is flowed into the gases upstream ignition). Peak intensities vary in all three experiments; this is probably caused by alterations in the experimental setup, i.e. the distance between the lens and flame may differ, or the torch may not be tuned exactly the same each time. 95 2E—O71 2E—07 — htensity (a.u.) Fri «2 5.E-(B< 0.E+(I) *7 WU» Figure 5.12 OES scan of N2 in a stoichiometric (160/80 sccm 02/CH4) combustion flame and hybrid (combustion plus 40, 50, 100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean" is a flow rate of 160/54 sccm 02/CH4. The last set of data is from a discharge of only 160 sccm 02 and 50 W of microwave power (no fuel is present). Data from 30 W of microwave power and below show no N2 signal, similar to the combustion-only case, and therefore are excluded from the plot. 96 Figure 5.13 Second set of data: OES scan of N2 in a hybrid (160/80 sccm O2/CH4 combustion plus 10-80 W microwave power) flame/discharge using the 0.4 mm- diameter brass nozzle. 6.E-(B Intensity (a.u.) P: so .A w a U :3; $5 iIO" é Figure 5.14 OES scan of N2 in a hybrid flame/disfigé with a flow rate? W 160/80/10 sccm O2/CH4/N2 with varying microwave power, 30-80 W. Note that N2 is injected upstream into the flow of gases, CH4 and O2. ‘3’— : --‘1s—J .‘igr MW 97 «77 de —~—— l'M’Jidl‘U/V l: - WW {‘9} -~---|'deN it?“ “““ W070” ‘1 [15:55! ----Hytxidww .1“.’Rln."vlii"’l,fl.r:; ‘5‘ 115“” 5| ll‘flai}; ‘1‘ .."Vl' )1} 'I 523 ll, , ., ,. Ni" . , .1. 5 :5 ‘ I gfiéa‘f‘lhpml . “V“‘My‘l “dowry“ "“l‘ h . -.'.1 I . _ ’ ' I _ '5 :‘l 1N1 .‘. ’55 VJ‘f-‘i 1 m1? {‘3le' J Ix‘lhr~,_.\’,m',--:I;"' 1'..." w In (343% ,, WMM‘W *7 , v r 3775 37% Figures 5.15 and 5.16 show C2 emission spectra in combustion flames and hybrid flame/discharges. The general trend follows that as microwave power increases, C2 emission intensity increases. Two sets of data are recorded to show consistency in experimental results. 1.508— 1 ,il 1.50m l. 1 :1“ 1" 1 1‘ 11111171151 NW ‘1‘F4 115111r‘i1‘11‘1111l‘rl‘f‘1 11*”1‘1iiir1‘" 1% W515 i 54 A; aeoe- _.._.C‘arb(1m&)) ‘ma‘rw s; ' " mam ,3 SEE wmdaow g ....... wd4OW 9:3 4.509 “" de _ _._..wmd1oow 25(95 \w’l Mil-1W 1. m1 5:13:31” 15“,«!~.‘,‘2 ‘11‘1‘1‘ “1111‘1‘l‘h‘1‘11‘ :mamwfm/ymlr‘ 414.4. 'rP-‘lgti ’ 1W1” 1’ ‘11“? H I “‘ ' 5W Iri" o.E+oo . ' ‘ ‘ 1 1 1 4995 5015 was m 5075 5095 5115 5135 W609”! (A) Figure 5.15 OES scan of C2 in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 10-100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” has a flow rate of 160/54 sccm O2/CH4. 98 1.5001 ---~»7ow -———90w 9.5% 1 8.15419 7.E-(B< 6.5-(B 55(9- 4.E-(B‘ 3519‘ ”WW“ EINVWYVWW‘ 111111 i11111*'1‘11I‘1 “113 15:511’1'1‘51‘151 1551; 111W" 2509- 1191111515111”- ‘ 1f”: ‘NMIT ;:y {11‘ iMWI :‘i11t1111111141'7n1‘1‘11‘ “111111111 1111 1.1-1'11 01“] [I '1; L 44,“. s ‘- o.E+oo fir . , . -_f_ gz *5 z_ . 4990 5010 5090 5050 5070 5090 5110 5130 51 WW Figure 5. 16 Second set of data. OES scan of C2 In a combustion (160/80 sccm O2/CH4) and hybrid (combustion plus 10- 80 W microwave power) flame/discharge using the 0.4 mm-diameter brass nozzle. Intensity (a.u.) Emissions of 02, OH, and CH are also recorded in Figures 5.17 — 5.19. As seen for N2 and C2, a general trend follows that as microwave power is added to a stoichiometric flame, the emission intensity increases. 99 $507 bummedam __wm'dww ~ an'dSOW ’5 5507 —-— andKDW g: WbridSOW—fudlem Figure 5.17 OES scan of O2 in a stoichiometric (160/80 sccm O2/CH4) combustion flame and hybrid (combustion plus 40, 50. 100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean" is a flow rate of 160/54 sccm O2/CH4. The last set of data is from a discharge of only 160 sccm O2 and 50 W of microwave power (no fuel is present). Data from 30 W of microwave power and below show no 02 signal, similar to the combustion—only case, and therefore are excluded from the plot. 100 2507 — Comb (16080) "~de 2E—07~ ———I-ybn‘d1oow —~-—-wbddwN—mellean ;::-_ M Hill”) W H 0.E+(X) - Intensity (a.u. ) WWW Figure 5.18 OES scan of OH in a stoichiometric (160/80 sccm 02/CH4) combustion flame and hybrid (combustion plus 40, 50, 100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean” is a flow rate of 160/54 sccm 02/CH4. Data from 30 W of microwave power and below show no OH signal, similar to the combustion-only case, and therefore are excluded from theplot - - -erb(1w&)) fiE-w —--—me -4I-ymdaow SE43} ' _~de ‘ "1 —l-Mxid50N 3‘ i ii i' q i ——— H/Utd1ww d 4508] '1" 5 \ —|'M!id4OW—fuellem v I“ ’lflfh " {I ---- deV-fifilai ' 3508+ ’- =""“ 'i E IW ll“: ~~~~~~ d1(DW—fuellea1 l I. i f in W“ M‘W‘W _ 2543‘ .- ' 1 ; a- mum.“ kg?!“ W@ 1 1.508 0.500 of, fl W7— 43!) 450 43!) 4:50 W WW — Figure 5.19 OES scan of CH in a stoichiometric (160/80 sccrn OleH4) combustion flame and hybrid (combustion plus 10—100 W microwave power) discharge using the 0.4 mm-diameter brass nozzle. “Fuel lean" is a flow rate of 160/54 sccm 02/CH4. 101 The peak intensities from each scan (N2, C2, 02, and CH) are plotted against absorbed microwave power. Figures 5.20 - 5.23 contain data from multiple sets of experiments, i.e. independent experiments, to show consistency within data. Aside from a scattering of data points due to variations in each independent experiment, the trend follows that intensity increases as microwave power increases. j 25“ x3et15toich oSd1FudLm1 20, asazstoich <> -Set2NoFueI x ASdBStoioh 15~ +Set4Stoioh-Nerj 10~ N2Em'ssionlrMnsity(a.u.)x10-8 15 25 35 45 55 65 75 85 95 105 AbsorbendmwavePaweer) Figure 5.20 Relative intensities of N2 emissions as a function of absorbed microwave power, measured from flame/discharges with flow rates of 160/80 sccm, 160/54 sccm, and 160/0 sccm 02/CH4, and also a flow rate of 160/80/10 sccm 02/CH4/N2 with N2 injection. Intensities are recorded from four sets of experiments to show consistency, and values are recorded from peak intensities, at a wavelength of ~ 3769 A. 102 .s _s N A .5 0 Oz Em'ssion Intensity (a.u.) x10-9 O) on D A A 3 A 3 o A 4“ 0 ° 0 U 0861 OSetZ 2< A863 0 Y I 17 I V 1 0 Z) 40 a) a) 100 120 Absorbendrawavernrm Figure 5.21 Relative intensities of Oz emission as a function of absorbed microwave power, measured from a stoichiometric flame/discharge with a flow rate of 160/80 sccm 02/CH4. Intensities are recorded from three sets of experiments to show consistency, and values are recorded from peak intensities, at a wavelength of ~ 5130 A. oSd1Stdch 0861me jg: j 104 amunmns'ty (a.u.) x10-8 8 8 o o D Y i T T l T l l 20 30 40 5o 60 7D 80 90 100 AbsorbendrowavePcmaer) Figure 5.22 Relative intensities of 02 emission as a function of absorbed microwave power, measured from flame/discharges with flow rates of 160/80 sccm 02/CH4 and 160/54 sccm 02/CH4. Intensities are recorded from one set of experiments, and values are recorded fromfi‘ peak intensities, at a wavelength of ~ 3340 . 103 i l | CH Emission Intensity (a.u.) x10-8 w A o a . 0 0 0 0 0 0 20 4o 60 80 AbsorbendmwavePowerM oSd1Stdch nSet1FudLea'1 1CD 120 Figure 5.23 Relative intensities of CH emission as a function of absorbed microwave power, measured from flame/discharges with flow rates of 160/80 sccm O2/CH4 and 160/54 sccm O2/CH4. Intensities are recorded from one set of experiments, and values are recorded from peak intensities, at a wavelength of ~ A wide-range scan including spectra of all species mentioned in this section is shown in Figure 5.24. The plot provides a clear, overall summary of how microwaves alter a combustion flame at a micro-scale level, not detectable by the human eye. The spectra of N2 and C2 are used to calculate gas temperatures which approximate flame/discharge temperatures, discussed in the next section. 104 l 5.E-08- —-Wb"'d(4°W) CH (4280479508) 55% q ---»- Cbntxstion 4.508 ~ 4.E-08 l 3.E-08 J 3.5-08 J Intensity (a.u.) ZE-(B 4 2.5% - 1.E-08 ~ 5.E-09 ~ O.E+m‘ l 7' I T 2500 3000 3500 4000 4500 5000 5500 6000 6500 VlhvalengtlflA) Figure 5.24 OES scan of combustion flame (160/80 sccm O2/CH4) and hybrid flame/discharge (addition of 40 W microwave power) for an entire range of visible emission spectra. The wavelength and relative intensities are given at the peak (A = 4280 A) for each set of data. Molecules and radicals are listed next to their visible spectra. A 0.4 mm-diameter ceramic nozzle is used for this data. 5.5 Flame/discharge temperatures and profiles Data for flame/discharge temperature are gathered with three different diagnostic techniques using: 1) a thermocouple; 2) OES; and 3) an IR camera. The experimental method for each technique is provided in Chapter 3. As discussed previously, the thermocouple is capable of measuring a pure combustion flame temperature only because the metal interferes with the electrical field produced by microwaves. Figure 5.25 presents a temperature profile through the width of a stoichiometric flame, using a type K thermocouple. The photograph in the figure shows where the thermocouple collects data through the flame. From a time-step iterative algorithm, the maximum flame 105 temperature calculated is 2253 K. Adiabatic flame temperature for methane/oxygen flames, found in text [25], is approximately 3000 K. The theoretical curve drawn in Figure 5.25 is not plotted with theoretical data; it is a qualitative curve used for illustrative purposes. The 750 K difference in temperatures is most likely due, in majority, to heat loss. The thermocouple is not shielded, leading to radiation loss. Furthermore, the flame is not confined or insulated; much heat is lost to the environment. It is also important to mention that thermocouples are used to calculate point temperatures, whereas adiabatic flame temperatures tabulated are average temperatures of the entire flame. —E>perirrertdee —TleoretiwCuve Terrperature (K) CD 4.0 -3.0 -20 -1.0 0.0 1.0 2.0 3.0 4.0 Wclh of Flam (mu) Figure 5.25 Temperature profile of a stoichiometric CH4/O2 flame from a type K thermocouple. Photograph of the flame shows where the thermocouple scans the flame. The plot provides experimental data and also displays a qualitative example of a theoretical temperature distribution with an adiabatic flame temperature of 3000 K, taken as an average temperature. 106 Rotational temperatures of N2 and C2 are calculated for several sets of experiments, i.e. independent experiments. Figure 5.26 plots N2 rotational temperatures for three sets of stoichiometric flow rates. Note that N2 is added to the gas flow upstream the nozzle. It is shown that temperature remains constant for each stoichiometric flow rate, in agreement with theory. Possible outlier data in the plot is found at 70 W microwave power; the temperature spread is 652 K. Such a range may be caused by variations in the experimental setup, mainly tuning the torch to find the sweet spot. In addition, as the volume of the flame increases, so will the heat loss, causing a decrease in temperature. The temperature increases as microwave power increases, ~ 1200 K from 30 W to 80 W. NzRotationalTerrperatu'eM) fié 2&1) j 0 A o A 2600 — 0 1666610 U 2400 — o A n 1065036 2200 4 D A 6630/2 A m l T I T T T l 20 30 40 50 60 70 60 ed AbsorbendmwavePowarMb Figure 5.26 N2 rotational temperature of a flame/discharge as a function of absorbed microwave power for data for three different flow rates: 160/80/10 sccm, 100/50/3.6 sccm, and 60/30/2 sccm O2/CH4/N2. Figure 5.27 graphs rotational temperatures of both N2 and C2 gases. Four independent experiments present data in this plot, the last set defined differently 107 because N2 is added to the gas flow. Data show a general increase in temperature as microwave power increases. The greatest increase occurs for N2 rotational temperatures (Set 1), with a temperature increase of 1489 K from 20 W to 80 W microwave power. The smallest increase occurs for C2 rotational temperature (Set 1), with a temperature increase of 95 K from 0 W to 80 W microwave power. Rotational temperatures of N2 and C2 in disagreement with each other does not infer that the temperature is miscalculated. Further investigation and research is required to know which rotational temperature, if either, best represents the flame temperature. Raw data from Figure 5.27 is listed in Appendix B. WTerrperaue( E 3 C; .. Q‘DO <> . mu- 0 ><+ D: 0 J + 0 0 .. 200m . 0 1500~ oNZSd1 OQSd1 1000 .msa2 0 .Qsa2 I: 5001 xmsas 0C2863 g +Nery'ectionSet4 o T r . . . 0 20 4O 60 8) 1(1) i AbsorbendmwavePowerM Figure 5.27 N2 and C2 rotational temperatures of flame/discharges as a function of absorbed microwave power, for four sets of data (independent experiments). The flow rate for the first three sets is 160/80 sccm O2/CH4, and 160/80/10 O2/CH4/N2 for the fourth set. Tables of raw data are found in Appendix B. 108 Flame and hybrid temperature greatly exceed those of pure plasma temperatures. Plasma temperatures measure ~ 1700 K (measured from N2 gas temperatures) while combustion temperatures, measured by the same technique are close to double that. 5.5.1 Data collected from infrared (IR) camera Although quantitative temperatures cannot be measured with the IR camera at this time (another student’s ongoing Ph.D. research), qualitative deductions are made. The images shown in Figure 5.28 are temperature profiles, created by the IR camera, of a stoichiometric flame as microwave power increases from O W to 40 W to 60 W. The emissivity is unity for all the results shown. The flame analyzed in these images is not confined within the microwave cavity because the IR camera cannot bypass the metal screen in the cavity’s windows. Therefore, the torch is not optimally tuned due to microwave leakage. Results do show, however, that the microwaves do indeed alter the flame. The center of the flame, the white temperature profile in Figure 5.28, is the hottest part of the flame, relative to all the profiles shown. As microwave power is added, this center section expands while the outermost sections contract. This behavior suggests that microwaves create a more intense flame/discharge by decreasing volume and increasing temperature. Figure 5.29 is a sketch of the first two IR images in Figure 5.28 against each other, to clearly show the change in volume. 109 Fahrenheit 420.0 i 410.0 400.0 390.0 380.0 370.0 360.0 350.0 340.0 330.0 320.0 310.0 300.0 290.0 280.0 I (2:123 [54.0 Figure 5.28 Temperature profile images from the IR camera of a flame as it is modified by microwave power, 40 W and 60 W. The flame is not confined within the cavity. The temperature scale is provided, but does not provide an accurate temperature reading. The gas flow rate is 200/100 sccm O2/CH4. Figure 5.29 Temperature profiles sketched from the IR camera images (from Figure 5.28) for both a combustion flame (represented by the solid lines) and a hybrid flame/discharge with 40 W microwave power added (dashed lines). 110 IR images of a pure argon plasma are shown in Figure 5.30. Each image is recorded with a different filter, so as to highlight various sections of the plasma. For example, the bottom-right image highlights the heat produced downstream the plasma discharge, while the top-left image highlights the temperature profiles within the intense plasma itself. Figure 5.30 Temperature profile images from the IR camera for a plasma-only discharge: 200 sccm argon excited by 40 W microwave power. Each image is recorded with a different filter. The image in the bottom right corner shows the temperature profiles of the hot gases beyond the plasma. Figure 5.31 shows IR images of a flame produced by a ceramic nozzle. An interesting observation is the temperature profiles of the ceramic. The center of the nozzle, where the flame sits, appears much hotter than the outer edges. Further study of ceramic temperature profiles should be carried out. This IR camera is capable of accurately measuring temperatures of solid surfaces, and it is also capable of recording time-dependent temperatures. 111 Time-dependent temperature measurements of a gold-sputtered ceramic nozzle are made with the IR camera. Unfortunately visual results of this experiment are not reported in this thesis because the only current recorded format is an avi file (video file). In summary, the gold-sputtered ceramic generates higher temperature profiles than the ceramic with no gold coating and does not cool down as fast as the ceramic with no coating. Figure 5.31 Temperature profile images from the IR camera for a stoichiometric combustion flame formed by a ceramic nozzle. Different camera filters are used to highlight different sections of the flame. The temperature profiles of the nozzle are also illustrated. 112 Chapter 6 Summary and Recommendations 6.1 Summary of results An extensive experimental study of microwave plasma-enhanced combustion concludes that microwaves do in fact alter a combustion flame. Physical alterations of the flame are made in color and geometry, as color changes from a blue flame to a white/yellow hybrid discharge with the addition of microwaves, and flame lengths and diameters grow in size. Alterations at the micro-scale level are also made in flame species. Molecules and radicals detected in the flame/discharge become more intense, or more accelerated, by the presence of microwaves. In addition, temperatures increase with the addition of microwaves, as hypothesized. One question that remains asks what exactly is desired by adding microwaves to a flame. Is it desirable to create a completely transformed flame that changes in color and increases in size, thus raising temperature but lowing power density; or is it more desirable to add just enough microwave power (10 — 30 W) so the power density rises exponentially without physically changing the appearance of the flame. The trade-off is temperature. More microwave power yields higher flame temperatures. These questions may be more easily answered when considering the applications of such a hybrid technology. Industry may be interested in a small— scale, easily handled, clean torch used for welding, surface treatment, or localized heating. In these cases, a higher-temperature flame/discharge is 113 desired. When considering cutting comparable to laser technology, geometry is a major factor considered. The more concentrated and more intense the discharge is, the more precise the cut will be. Geometry also becomes a critical factor when considering medical applications, such as surgery. Other applications include engines, both jet engines and automotive engines. Results in this thesis show that, with the presence of microwaves, a flame is maintained with lower fuel consumption than would be possible without microwaves. Fuel-efficient engines are driving the automotive industry today. Both higher temperatures and larger geometries may be desired for this application. Motivation behind this research includes improving the combustion process by creating more efficient combustion with the addition of microwaves. Results support the hypothesis that microwaves do improve the combustion process. Flame stability and fuel-efficient combustion are enhanced in this hybrid system. It is yet to be shown in this research if cleaner combustion results from the addition of microwaves. Researchers at the Los Alamos National Laboratory have shown that a slightly different plasma-enhanced hybrid system increases combustion efficiency by reducing the amount of unburned hydrocarbon (propane) emissions and also reduces toxic NOx/SOx emissions [1]. It appears that the presence of microwaves accelerates the combustion reaction, reflected in flame velocities calculated from measured flame angles. Flame speed more than doubles with the addition of microwaves when the physical transformation takes place. 114 6.2 Recommendations Further investigation and study of the ceramic nozzle is encouraged. In terms of manufacturing, an improved method for attaching the ceramic nozzle to the brass inner conductor should be sought. Currently the nozzle is glued with epoxy to the inner conductor. This becomes a problem when microwaves are present. The electrical energy is drawn to this epoxy material, thus retracting from the flame itself. Additional temperature measurements made with the IR camera could provide insight as to whether the ceramic is or is not absorbing microwaves. An improved design of the microwave cavity would attract the torch to industry. The cavity is big and bulky. Designing a much smaller and compact cavity to allow an intense formation of an electric field is critical. Temperature measurements of the flame/discharge inside the cavity using the IR camera would also provide insight for what is happening to temperature as microwaves greatly transform the flame. If time and funds allow, more detailed diagnostic measurements should be made including cavity ring-down spectroscopy (CRDS) and laser-induced fluorescence (LIF) to determine the influence of plasma enhancement on the distributions of CH, C2 and OH. OES used for the research in this thesis provides profitable results in terms of qualitative analysis. For quantitative analysis, however, the suggested CRDS and LIF techniques are preferred. CRDS is an absorption technique where pulsed laser light is coupled into an optical cavity formed by highly reflecting mirrors. It is a sensitive 115 technique because the optical path length is made long by the highly reflecting mirrors, and it is convenient because the quantity being measured is the decay time of the light intensity, not the absolute intensity of the light as needed for direct absorption spectroscopy measurements. CRDS works at both atmospheric pressure and low pressures (5 — 30 Torr), suggested working conditions for future experimental work. LIF is a high-spatiaI-resolution technique but can be more difficult to calibrate and work with. One final suggestion considers a numerical study. A computational simulation can be created by factoring into a full flame chemistry simulation the electronic or plasma reaction component. Full flame chemistry simulations are developed and available as commercial software. The suggested hybrid simulation has yet to be fully developed. Such a simulation would provide validation for experimental results and provide a deeper understanding of the research. 116 APPENDICES 117 Appendix A Table A1 Values for therrnocou le temperature measurements. Width of Measured Measured Time Flame T(t) PfGViOUS psolld pr|d@2oc [.1 [1. Step (mm) 462g) T (degC) 459/1113) (JIkg-K) m-s/m’) (N-s/m’) 1 -3.158 24.4 8600 523 2.00E-05 2.00E-05 2 -2.842 26.5 24.4 8600 523 2.02E-05 2.02E-05 3 -2.526 28 26.5 8600 523 2225-05 2.03E-05 4 -2.211 31.1 28 8600 523 2.16E-05 2.05E-05 5 -1.895 51.6 31.1 8600 523 2.33E-05 2.18E-05 6 -1.579 101.2 51.6 8600 523 3.48E-05 2.45E-05 7 -1.263 167.4 101.2 8600 523 4.64E-05 2.75E-05 8 -0.947 274.2 167.4 8600 523 5.26E-05 3.23E-05 9 -0.316 382.9 274.2 8600 523 6.70E-05 3.62E-05 10 0.000 473.8 382.9 8600 523 6.62E-05 3.91 E-05 1 1 0.316 615.7 473.8 8600 523 6.46E-05 4.37E-05 13 0.947 1297.6 1285.9 8600 523 2.18E-04 6.40E-05 14 1.263 1196.2 1297.6 8600 523 6.51 E-05 6.11E-05 15 1.579 1096.2 1196.2 8600 523 2.81 E05 5.83E-05 16 1.895 985.2 1096.2 8600 523 3.93E-05 5.51 E-05 17 2.211 893.2 985.2 8600 523 4.03E-05 5.24E-05 18 2.526 818.5 893.2 8600 523 3.43E-05 5.03E-05 19 2.842 753.3 818.5 8600 523 2.93E-05 4.82E-05 20 3.158 712.5 753.3 8600 523 2.97E-05 4.68E-05 Calculat Calculat ed T- ed T- knmdm h flame flame pip, (WIm-K) Pr mm’ls) Re (WlmzK) (degC) (K) 375.15 1.000 0.02532 0.74036 1.61 E-05 5 170.31 27.00 300.00 1.002 0.02553 0.74000 1.74E-05 348.54 166.14 58.25 331.25 1.094 0.02807 0.73563 1.98E-05 306.07 175.16 49.47 322.47 1.056 0.02736 0.73685 1.91 E-05 316.92 172.25 76.25 349.25 1.071 0.02954 0.73311 2.12E-05 285.99 177.89 340.37 613.37 1.423 0.04968 0.70553 5.65E-05 107.1 1 207.29 697.37 970.37 1.688 0.06884 0.72788 1.21 E-04 50.01 225.12 897.52 1170.52 1.626 0.08008 0.73307 1.65E-04 36.62 234.90 1400.86 1673.86 1.852; 0.11066 0.71217 3.14E-04 19.29 270.49 1371.83 1644.83 1.691 0.10889 0.71338 3.05E-04 19.83 265.17 1318.27 1591.27 1.478 0.10564 0.71560 2.89E-04 20.91 256.70 1979.69 2252.69 1.312 0.14582 0.68814 4.85E-04 12.49 303.49 6685.03 6958.03 3.411 0.43163 0.49279 1.87E-03 3.23 716.22 1334.39 1607.39 1.065 0.10662 0.71493 2.94E-04 20.57 247.25 182.38 455.38 0.483 0.03806 0.71882 3.45E-05 175.70 159.48 480.65 753.65 0.714 0.05721 0.71431 8.19E-05 73.94 185.62 511.00 784.00 0.769 0.05884 0.71621 8.74E-05 69.30 189.02 323.78 596.78 0.682 0.04879 0.70449 5.35E-05 113.11 182.42 206.73 479.73 0.607 0.04002 0.71555 3.75E-05 161.38 168.80 216.29 489.29 0.634 0.04079 0.71426 3.87E-05 156.38 171.15 114.35 387.35 118 Table B.1 Values used for flame extinction plot in Chapter 5. Appendix B Combustion Only (0.4 mm brass) Combustion+Microwaves Blowout 10 W 20-100 W V02 (sccm) Vm‘ (sccm) ch (sccmL fiflsccm) °/o Fuel Reduction 180 72 64 0 11 160 56 53 0 5 140 44 41 0 7 120 35 32 0 9 100 26 23 0 12 80 19 19 0 0 Combustion Only (brass, 0.4) Combustion-HOW Combustion+20-100W Equivalence Equivalence VTOT Equivalence th Ratio vim (sccm) Ratio (sccm) Ratio (sccm) 0.80 252 0.71 244 0.00 180 0.70 216 0.66 213 0.00 160 0.63 184 0.59 181 0.00 140 0.58 155 0.53 152 0.00 120 0.52 126 0.46 123 0.00 100 0.48 99 0.48 99 0.00 80 Table B.2 Values found in temperature plots in Ctflter 5. Methane/0x1 en Mixture 160l80 sccm), Set 1 Microwave 1’0“" (W) o 10 20 30 4o 50 so 70 so Rotational Temperature (K) N2 N/A N/A 1717 2682 2368 2223 2609 2991 3206 Rotational Temperature (KLCZ 3124 2994 2935 2441 3111 3049 3057 3350 3318 Methane/Oxygen Mixture (160l80 sccm), Set 2 Microwave PW" (W1 0 1o 20 30 4o 50 100 Rotational Temperature (K) N2 N/A N/A N/A N/A 2524 2934 3612 Rotational Temperature (K) C2 3016 3115 2516 2584 2687 2886 3254 119 MethaneIOxygen Mixture (160l80 sccmL Set 3 Microwave Powerlwl o so 40 so Rotational Temperature (K) N2 N/A N/A N/A 2589 Rotational Temperature (K) CZ 2637 2352 2738 2849 Methane/Oxygen Mixture (160l80/10 sccm), Set 4 Microwave Poweriw) 30 40 so so 70 so Rotational Temperature 4KLN2 2368 2863 2674 3258 3329 3542 120 BIBLIOGRAPHY [1] L. A. Rosocha, Y. Kim, S. Stange, V. Ferreri, D. M. Coates, and D. Platts, “Plasma-enhanced combustion of propane using a silent discharge,” Plasma Physics Research Highlights, Los Alamos National Laboratory. [2] N. Chintala, R. Meyer, A. Hicks, B. Bystricky, J. W. Rich, W. R. Lempert, and I. V. Adamovich, “Non-thermal ignition of premixed hydrocarbon-air and CO-air flows by nonequilibrium RF plasma,” 42”” AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, 2004. [3] K. W. Hemawan, C. L. Romel, S. Zuo, I. S. Wichman, T. A. Grotjohn, J. Asmussen, “Microwave plasma-assisted premixed flame combustion,” International Workshop on Micro-plasma, Greifswald, Germany, 2005. [4] S. Whitehair, L. L. Frasch, and J. Asmussen, “Experimental performance of a microwave electrothermal thruster with high temperature nozzle materials,” 19’" AIAA/DGLR/JSASS International Electric Propulsion Conference, Colorado Springs, CO, 1987. [5] A. Yu. Starikovskii, “Plasma supported combustion,” 30th International Symposium on Combustion, Chicago, IL, 2004. [6] S. M. Starikovskaia, I. N. Kosarev, A. V. Krasnochub, E. I. Mintoussov, A. Yu. Starikovkii, “Control of combustion and ignition of hydrocarbon-containing mixtures by nanosecond pulsed discharges,” 43m AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, 2005. [7] J. W. Dold, R. W. Thatcher, A. Omon-Arancibia, and J. Redman, “From One- Step to Chain-Branching Premixed Flame Asymptotics,” Combustion Institute, 29:1519-1526 (2002). [8] N. Peters, Turbulent Combustion, Cambridge University Press, 2000. [9] A. Linan and F. A. Williams, Fundamental Aspects of Combustion, Oxford University Press, New York, 1993. [10] J. Buckmaster, Annu. Rev. Fluid Mech. 25:21 -53 (1993). [11] P. Pelce, Dynamics of Curved Fronts, Academic Press, London, 1988. [12] G. l. Sivashinsky, Annu. Rev. Fluid Mech. 15:179-199 (1983). [13] F. A. Williams, Combustion Theory, Benjamin/Cummings, San Francisco, 1985. 121 [14] J. D. Buckmaster and G. S. S. Ludford, Theaty of Laminar Flames, Cambridge University Press, 1982. [15] Ya. B. Zeldovich, G. I. Barrenblatt, V. B. Librovich, and G. M. Makhviladze, The Mathematical Theory of Combustion and Explosions, Consultants Bureau, New York, 1985. [16] I. K. Puri, Environmental Implications of Combustion Processes, CRC Press, Boca Raton, Florida, 1993, Chapter 5 by I. S. Wichman. [17] S. Zuo, K. Hemawan, J. J. Narendra, T. A. Grotjohn, and J. Asmussen, “Miniature microwave plasma torch applicator and its characteristics,” IEEE Conference on Plasma Science, 5A3, 2004. [18] K. W. Hemawan, Numerical A_nalvsis and Experimental Measurements of Material Loading in Cvlindricgl Microwave Cavitv Applicators, MS. Thesis, Michigan State University, 2003. [19] K. W. Hemawan, S. Zuo, C.L. Romel, T. A. Grotjohn, I. S. Wichman, E. Case, and J. Asmussen, “Exploring microwave plasma-assisted combustion,” International Conference on Plasma Systems (ICOPS), Monterey, CA, 2005. [20] J. B. Wachtman, Jr., Mechanical Prcyerties of Ceramics, John Wiley and Sons, Inc., New York, 1996. [21] J. Zhang, L. Liu, T. Ma, and X. Deng, “Rotational temperature of nitrogen glow discharge obtained by optical emission spectroscopy,” Spectrochism. Acta A, 58, 1915-1922, 2002. [22] J. J. Narendra, Characteristics and Modeling of Miniature Microwave Plasma Discharge Created with Microstrioline Technology, MS. Thesis, Michigan State University, 2004. [23] E. Stoffels, A. J. Flikweert, W. W. Stoffels, And G. M. W. Kroesen, “Plasma needle: a non-destructive atmospheric plasma source for fine surface treatment of (bio) materials,” Plasma Sources Sci. Technol., 11, 383-388, 2002. [24] W. Huang, Microwave Plasma Assisted Chemical Vapor Deposition of Ultra- nanocrfitalline Diamond Films, Ph.D. Dissertation, Michigan State University, 2004. [25] F. P. lncropera and D. P. Dewitt, Introduction to Heat Transfer, John Wiley and Sons, New York, 2002. 122 [26] J. L. Short, A New Svstem Developed to Characterize Thermoelectric Devices for Power Generation Applications, MS. Thesis, Michigan State University, 2005. [27] G. Herzberg, Molecular Spectra and Molecula_r Structure. I. Spectra of Diatomic Molecules, D. Van Nostrand Co., Inc., Princeton, 1950. [28] S. N. Suchard and J. E. Melzer, Spectroscopic Data 2: Homonuclear Diatomic Molecules, lFl/Plenum, New York, 1976. [29] G. L. King, Temperature and Concentration of Ionic and Neutral Species in Resonant Microwave Cavitv Plasma Discharges, Ph.D. Dissertation, Michigan State University, 1994. [30] B. C. Wadell, Transmission LinejDesign Handbook, Artech House, Inc., NonNood, 1991. [31] S. R. Turns, An Introduction to Combustion, The McGraw-Hill Companies, Inc., Boston, 2000. 123 |111:1][Elllljflljlfllljflf|