Aeronautics and Astronautics Part 10 docx

40 300 0
Aeronautics and Astronautics Part 10 docx

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

Plasma-Assisted Ignition and Combustion 349 combination of short high-voltage pulse and constant bias allow to provide selective and extremely nonequilibrium excitation of the gas. Critical high-voltage pulse duration depends on the gas parameters (density, composition) but for practically important range of parameters is restricted to few nanoseconds. Thus the possibility of selective excitation of the gas by electric discharge critically depends on the possibility of ultra-short high-voltage pulses generation. Figure 19 demonstrate recent progress of solid-state generators based on “turn-on” FID and “turn-off” DRD switches according to FID GmbH data [Efanov et al, 2011]. In modern pulsers the pulse rise time goes down to 80 ps, voltage rise rate reaches 1 MV/ns, maximal voltage 2-10 MV, and maximal current up to 100 kA. Wide range of possibilities proposed by current progress in solid-state electronics will lead to the increase of our abilities of nonequilibrium plasma generation with predicted properties. Fig. 19. Progress in ultra-short high-voltage pulse generators. A) typical nanosecond pulse shapes [Roupassov et al, 2008]; b) generators frequency-voltage map [Efanov, 2011]. 2.2 Non-equilibrium plasma recombination and energy relaxation For efficient production of large amount of active particles in the gas discharge it is necessary both efficient generation in the gas discharge plasma and slow recombination in collisions with major mixture components. 2.2.1 Rotational relaxation Due to fast rotational-translational (RT) relaxation, rotational degrees of freedom of the molecules are quenched rapidly. This process requires few collisions only. For example, for rotational relaxation in air O 2 (rot) + M → O 2 + M and N 2 (rot) + M → N 2 + M typical relaxation time is comparable with gas-kinetics time. This means that typical time of rotational states thermalization is ~ 0.5 ns under normal conditions. That is why rotationally-excited molecules cannot be considered as active particles for non-thermal acceleration of chemical reactions. Another important point is that the energy of excitation of rotational states is very small (roughly equal to translational temperature) and is significantly lower than typical chemical reaction’s thresholds. From the other hand, it is possible to heat the gas through the rotational degrees of freedom excitation. Aeronautics and Astronautics 350 2.2.2 Vibrational relaxation Opposite to rotational states relaxation, quenching of vibrationally excites states N 2 and O 2 (vibrational-translational (VT) relaxation) is very slow process. Time of VT relaxation usually is longer than typical time of plasma-assisted ignition (~10-100 s). These times become comparable when significant amount of H 2 or hydrocarbons is presented in te mixture. This means that the vibrationally-excited N 2 and O 2 molecules can be accumulated in the discharge with intermediate E/n values. VT relaxation leads to slow thermalization of vibrational energy of the molecules. This process becomes faster if the mixtures contain hydrocarbons. For example, VT relaxation of molecular oxygen on methane in stoichiometric methane-air mixture at T = 1000 K and pressure 1 atm has a characteristic time t ~ 1.3 s. Fast relaxation does not allow to maintain a significant deviation of vibrational temperature from translational on the long time scale. From the other hand, VT relaxation of oxygen in H 2 -air mixture lasts ten times longer and reaches t ~ 15 s for T = 1000 K and P = 1 atm (29% H 2 in the mixture). VT-relaxation of hydrogen in the same mixture takes approximately 380 s. Thus, vibrational excitation of hydrogen molecules can be very far from equilibrium during the ignition delay time and can effect significantly the radical’s production. Under uncompleted vibrational relaxation conditions chemical reactions between vibrationally excited molecules play an important role. There are several theoretical models for rate coefficients of reactions between excited reagents. Almost all these models were developed as an engineering substitution of time-consuming ab initio calculations [Kovach et al, 2010; Adamovich et al, 1996; Macheret et al, 1994; Park, 1988]. A model of vibrational energy usage was developed in [Losev et al, 1996]. The model assumes the decrease of the reaction threshold by E vib . The efficiency of vibrational excitation  can be estimated using activation energy and thermal effect of the reaction. A model proposed by Macheret [Macheret et al, 1994] allows to estimate the rate constant of simple exchange endothermic reaction. The model requires the fraction of energy release in the reverse reaction directed to vibrational excitation and it is applicable only to a certain type of reactions [Kovach et al, 2010]. It should be noted that almost all analytical models available estimate reaction rate constants using “vibrational temperature”. This assumes that we have Boltzmann distribution over vibrational levels. Such an approach cannot be used at non-equilibrium conditions when the population over vibrational levels has non-Boltzann shape [Capitelli, 1996]. State-to-state model was considered in [Starikovskii, 2003]. Reactions between excited hydrogen molecules H 2 (v) and radicals are extremely important for ignition and combustion. As an example of reaction rate dependence on the vibrational excitation of reagents let us consider the process H 2 (v) + O  H + OH(w) It is shown in [Light& Matsumoto, 1978] that ratio of specific constants of the reaction rates at v=1 and v=0 is k(v=1)/k(v=0) = 2600 at T=300 K. The process at v=1 leads to formation of radical OH in vibrationaly excited state [Light& Matsumoto, 1978] H 2 (v=1) + O( 3 P)  H + OH(w=1) (= ( 1.0 . . ) ∙10  cm 3 /s) H 2 (v=1) + O( 3 P)  H + OH(w=0) (≤4.7∙10  cm 3 /s) Experimental measurements show that the averaged factor of vibration energy usage in this reaction is = 0.31 [Rusanov&Fridman, 1985]. Figure 20,a shows results of calculation of the Plasma-Assisted Ignition and Combustion 351 reaction rate constant at translation temperature T tr = 300 K for various vibration temperatures with the Boltzmann distribution of molecules over vibrational levels. The dependence calculated using model [Starikovskii, 2003] is in good agreement with calculation by -model with experimentally found  = 0.31 at overheating degree T vib /T tr < 5 (Fig. 20,a). Model [Starikovskii, 2003] predicts the ratio k(v=1)/k(v=0) = 2795, which is in perfect agreement with experiments [Light& Matsumoto, 1978] (2600). Ratio of channels to OH(w=1) and OH(w=0) at T = 300 K estimated in [Starikovskii, 2003] is equal to k(w=1)/k(w=0) = 7.9, which also is in good agreement with experiments [Light& Matsumoto, 1978] (>2). The analysis of the reaction rate constant dependence on the vibrational excitation degree for reaction OH + H 2 (v)  H 2 O + H is shown on Figure 20,b. The predictions of model [Starikovskii, 2003] are in a good agreement with calculations based on experimentally measured value of  = 0.24 [Fridman&Rusanov, 1985]. Work [Light& Matsumoto, 1978] gives an experimental estimation for ratio of the rate constants of processes OH + H 2 (v=0)  H 2 O + H and OH + H 2 (v=1)  H 2 O + H: k v=1 /k v=0 ≤ 1000 at T=298 K, which is in good agreement with the estimation by model [Starikovskii, 2003] k v=0 = 2.910 -15 cm -3 s -1 , k v=1 = 1.810 -12 cm -3 s -1 (k v=1 /k v=0 = 620). Thus, vibrational excitation of reagents can significantly accelerate chemical reactions. The influence of vibrational excitation is limited by VT-relaxation of the molecules. This process becomes extremely fast in the presence of hydrocarbons. In mixtures with hydrogen the efficiency of vibrational excitation increases because of relatively slow vibrational relaxation of H 2 . Analysis of [Zatsepin et al, 2001] shows the oxidation rate increase in H 2 -air mixture at T = 300K in 3-5 times. Fig. 20. Dependence of the rate constant of reaction on non-equilibrium excitation degree T vib/ T tr at T tr = 300 K. 1 – model [Starikovskii, 2003]; 2 – -model [Macheret et al, 1994]. a) H 2 (v) + O  H + OH,  = 0.31; b) H 2 (v) + OH  H 2 O + H, = 0.24. As an example of possible applications of vibrational excitation of the flow we will mention the paper [Bezgin et al, 2006]. Peculiarities of an oblique detonation wave formation in a supersonic hydrogen–oxygen mixture flow over a plane wedge were numerically analyzed. K ( T, T ) / k ( T ) vib tr tr T/T vi b tr H (v)+O=H+OH 2 K(T , T )/k(T ) vib tr tr T/T vi b tr H(v)+OH=HO+H 22 Aeronautics and Astronautics 352 Preliminary excitation of molecular vibrations of H 2 was shown to lead to a noticeable decrease in the induction zone length and the distance at which the detonation wave was formed. It was demonstrated that the reason for these effects was an intensification of chain reactions in the H 2 –O 2 (air) mixture owing to the presence of vibrationally excited hydrogen molecules in the flow [Bezgin et al, 2006]. 2.2.3 Electronic levels excitation and relaxation At E/n ~ 100-500 Td the main channel of gas excitation is population of electronic degrees of freedom by electron impact and by energy exchange between vibrationally-excited states. An important exception from this rule is singlet state of molecular oxygen О 2 (a). This state has a low excitation threshold and the maximum efficiency of its population corresponds to E/n ~ 3-10 Td. There are number of different electronically-excited particles in low-temperature plasma. Unfortunately, reaction rate constants, quenching rates and products are known only for limited number of them. That is why we will mention here the most important levels from the point of view of plasma assisted combustion only. The efficiency of energy of electronic state usage for plasmachemistry depends on the ratio between channels of depopulation of the state. For example, if the radiative life-time of the state is too short we will have some photon flux but no reactions with this state. Collisional quenching efficiency depends on the products of the reaction. Quenching of triplet states of nitrogen molecules by molecular oxygen lead to oxygen dissociation and atomic oxygen production. Another example – quenching of singlet oxygen molecules by hydrogen or hydrocarbons mostly leads to heat release without formation of active radicals. The most important reactions with electronically-excited molecules from the point of view of plasma assisted combustion are channels which lead to radicals formation. There are four different ways to produce radicals through excitation of electronically-excited states: 1. Excitation of the molecular electronic state and radicals production in chemical chains: a. O 2 + e  O 2 (a 1  g )[0.98 eV] + H  O( 3 P) + OH b. O 2 + e  O 2 (b 1  g + ) [1.64 eV] + H 2  OH + OH c. N 2 +e  N 2 (A 3  u + ) [6.2 eV] + O 2  N 2 O + O( 3 P) 2. Excitation of the molecule to repulsive or pre-dissociative term leads to molecule dissociation and formation of two radicals: a. O 2 + e  O 2 (B 3    )[8.4 eV]  O( 3 P) + O( 1 D) + e b. O 2 + e  O 2 (C 3  u )[6.87 eV]  O( 3 P) + O( 3 P) + e c. H 2 + e  H 2 (a 3  g + )[11.8 eV]  H( 1 S) + H( 1 S) + e 3. Excitation of the molecule and dissociative quenching of excited state by another molecule: a. N 2 +e  N 2 (C 3  u )[11.02 eV] + O 2  N 2 + O( 3 P) + O( 1 D) b. N 2 +e  N 2 (C 3  u )[11.02 eV] + H 2  N 2 + H( 1 S) + H( 1 S) c. O 2 + e  O 2 (A 3  g + )[4.5 eV] + CH 4  O 2 + CH 3 + H( 1 S) 4. Excitation of the molecular electronic state with radiative depopulation, high-energy photon flux generation and dissociation (ionization) of gas molecules by this radiation: a. N 2 +e  N 2 (B 1  u ) [12.5 eV]  N 2 + h  O 2 + h  O 2 + + e b. N 2 +e  N 2 (B 1  u ) [12.5 eV]  N 2 + h  CH 4 + h  CH 3 + H c. H 2 + e  H 2 (a 3  g + )[11.8 eV]  H 2 (b 3  g ) + h  O 2 + h  O + O Plasma-Assisted Ignition and Combustion 353 Comprehensive detailed kinetic models were discussed, for example, in [Kossyi et al, 1992] for N 2 -O 2 mixtures, in [Zatsepin et al, 2001] for H 2 -O 2 -N 2 mixtures and in [Anikin et al, 2006] for C x H y -O 2 mixtures. It should be noted however, that channel branching, rate coefficients and even products of such reactions are not very well known. The first group of processes was investigated much better, than second and third. Simultaneous presence in the plasma of all sorts of excited particles and radicals makes detailed kinetic analysis an extremely challenging and resource-consuming task. As an example we just mention that mixture composition variation, very popular approach in combustion chemistry, will not work in plasma chemistry because simultaneously with afterglow kinetics variation we will change electron energy distribution function in the discharge phase and kinetics of gas excitation. Mechanism (I) requires very low electric field to increase the efficiency of the excitation process because of low energy threshold for oxygen singlet states population. On the contrary, mechanisms (II)-(IV) require high E/n value and high electron energy for upper electronic states excitation. 2.3 Low-energy electronic states excitation Singled oxygen molecules as a tool for ignition and combustion control were proposed by group of Starik [Smirnov et al, 2008]. The effect of the excitation of oxygen molecules to the O 2 (a 1  g ) and O 2 (b 1  g + ) electronic states in the electrical discharge on the velocity of laminar flame propagation in the H 2 –O 2 mixture was analyzed. The calculations showed that the excitation of O 2 molecules to the a 1  g and b 1  g + electronic states allows one to increase significantly (by a factor of 2.5) the velocity of flame propagation for the fuel lean hydrogen– oxygen mixture. For stoichiometric and fuel rich mixtures the increase in flame velocity due to an abundance of singlet oxygen molecules in the mixture was found to be significantly smaller (about a factor of 1.1). Later the same team proposed to use a laser radiation at λ = 762.346 nm for O 2 molecules excitation to the b 1  g + electronic state. Experimental observation of the shortening of the induction zone length in a premixed mode of combustion in a subsonic H 2 –O 2 low pressure flow due to the presence of oxygen molecules excited to the singlet a 1  g electronic state was reported in [Smirnov et al, 2008]. The low pressure electric glow discharge was used to produce singlet oxygen molecules. The analysis showed that ~1% of O 2 (a 1  g ) molecules in the H 2 –O 2 mixture allows to noticeably reduce the ignition delay length and to ignite the mixture at a lower temperature. Authors conclude that the results obtained demonstrate the possibility to intensify the combustion of a hydrogen– oxygen mixture by means of excitation of O 2 molecules by electrical discharge at low pressure (P = 10–20 Torr). A numerical study of the plasma assisted ignition of hydrogen-oxygen mixtures at different E/n has been performed in [Wu et al, 2010]. Results at low E/n values are compared with experimental data [Smirnov et al, 2008] and good agreement between experimental and numerical data was demonstrated. It was shown that the efficiency of radicals production through the oxygen singlet states excitation is limited by collisional quenching of SDO molecules in oxygen-fuel mixtures; in oxygen-nitrogen mixtures main efficiency limitation comes from discharge energy flux alternation by vibrational excitation of nitrogen [Wu et al, 2010] (see also Figure 17,b). In paper [Wu et al, 2010] two different mechanisms of radical formation were analyzed: 1) at low E/n - through oxygen singlet states excitation with subsequent quenching and conversion into radicals in reactions with fuel molecules, and 2) at high E/n – through direct dissociation of molecular oxygen by electron impact and Aeronautics and Astronautics 354 quenching of nitrogen triplet states in collisions with molecular oxygen. It was shown that the first channel is more efficient in pure oxygen, while the second is much more efficient for mixtures containing more than 10% of nitrogen. 2.4 High electronic states excitation In papers [Kof&Starikovskii, 1996-1, 1996-2] authors proposed to use pulsed nanosecond discharges for plasma assisted ignition and flame stabilization. The idea was to maintain an extremely high electrical field for a short period of time. This approach allows to generate highly-excited nonequilibrium plasma with the energy distribution shifted to the electronic excitation and dissociation. Short pulse duration restricts the plasma conductivity increase and keeps the energy density in the gas on the relatively low level (equivalent gas heating is in the range of 10-100 K). Paper [Starikovskiy et al, 2011] summarizes the requirements to the pulse discharges to maintain the high efficiency of excitation: 1. High-voltage pulse amplitude is limited to set the value of the reduced electric field E/n > 200-300 Td in the discharge gap which provides optimal conditions for dissociation of molecular oxygen by electron impact and quenching of nitrogen excited states (in air and lean fuel-air mixtures). 2. High-voltage rise dU/dt > 3001000 kV/(nsatm) to obtain the field intensity sufficient for homogeneous ionization wave formation. This condition allows to achieve the homogeneous gas excitation in the gap and simplifies the analysis of the kinetic data. It shold be mentioned, however, that for practical applications inhomogeneous excitation may have specific advantages in some cases (for example, reduction of energy consumption). This type of the discharge was used in [Zatsepin et al, 2001] to investigate low-temperature kinetics in plasma of pulsed nanosecond discharge. Oxidation of molecular hydrogen in stoichiometric hydrogen-air mixture in the Fast Ionization Wave (FIW) was studied at total pressures p = 1-8 Torr, and the detailed kinetics of the process has been numerically investigated. The excitation of the gas in FIW and dynamics of molecular hydrogen concentration were monitored with the use of measurements of absolute H 2 radiation intensity (transition a 3  g +  b 3  u + ). Comparison of calculation and experimental results allows to make a conclusion that the gas is predominantly excited behind the FIW front in relatively low electric fields E/n ~ 300-600 Td at electron concentration n e ~ (1-2)10 12 cm -3 during approximately 10 ns and the excitation can be described with a good accuracy using the two-term approximation of Boltzmann's equation. In the subsequent processes the reactions including electron-excited particles play a dominant role for the time up to 100 ns, ion-molecular reactions – for the time of microsecond range, reactions including radicals mostly contribute for the time interval of several milliseconds. The most critical processes have been separated for each time interval. The principal role of processes with formation of excited components that support the development of the chain mechanism of oxidation has been shown. Detailed state-to-state kinetic mechanism [Zatsepin et al, 2001] includes 750 chemical and 8700 vibrational exchange processes with participation of 254 particles including electron- excited and charged atoms and molecules, electrons, radicals, non-excited components, and vibrational-excited molecules H 2 , O 2 , N 2 , H 2 O and OH-radical. The most important processes in each time interval in plasma afterglow and radicals recombination were identified. Because the overall picture observed in [Zatsepin et al, 2001] is very typical for plasma assisted ignition by pulsed discharges, we will analyze it in more details. Plasma-Assisted Ignition and Combustion 355 2.5 Kinetics of plasma assisted combustion below self-ignition threshold The mixture compression in the engine before the ignition leads to temperature increase. For example, in IC engines initial temperature is close to 600 K, in GTEs – 600-700 K, in SCRAMjets 650-800 K. In these cases the initial temperature of the mixture is below or close to self-ignition threshold. That is why this range of parameters attracts in attention of researchers. From the other hand, this temperature interval is poor investigated from the point of view of chemical kinetic mechanisms. The problem is the lack of data for low- temperature mechanisms validation. As an example, methane combustion GRIMech-3.0 model was validated in the range 1250 – 2500 K. C1-C4 Konnov’s mechanism was validated down to ~910 K, hydrogen Popov’s mechanism [Popov, 2008] – to 880 K. Direct extrapolation of these models down to room temperature conditions or even to intermediate temperature range below self-ignition threshold, of course, is very questionable (see, for example, analysis in [Uddi et al, 2011]). Thus the task of kinetics investigations in low temperature region becomes extremely difficult and complex. We have to take into account kinetics in gas discharge and plasma afterglow and almost unknown mechanisms of chemical chains initiation under low temperature conditions. Another problem of investigations of kinetics in plasma is gas discharge inhomogeneity. Under low pressure conditions homogeneous gas ionization and excitation can be achieved even with rather slow voltage increase across the discharge gap. Pressure increase requires a sharp decrease of the voltage rise time (relations 1)-2) above suggest to keep the voltage rise rate on the level of ~ 1 MV/ns/atm for room temperature air to achieve homogeneous excitation). For low pressure conditions this leads to critical voltage rise time about 8 ns and correlate with homogeneous picture of plasma formation in the reaction chamber of 5 cm diameter. Uncontrollable inhomogeneous excitation significantly compromises the kinetic analysis. That is why some authors prefer to use controlled inhomogeneous excitation instead. For example, in papers [Bak et al, 2011; Stancu et al, 2010; Grisch et al, 2009; Wu et al, 2010, 2011] the point-to-point electrodes geometry was used. This geometry generates non-uniform streamer-like discharge but because of its high reproducibility allows to reconstruct the spatial distribution of excitation and kinetics in plasma. In [Grisch et al, 2009], detailed experimental investigation of a non-equilibrium nanosecond pulsed discharge in premixed CH 4 /air mixtures at atmospheric pressure has been carried out. The electron temperature and density properties were measured using laser Thomson scattering (LTS). Temperature measurements were performed using N 2 CARS thermometry to quantify the energy transfer in the gas mixture. Effect of the discharge on the local temperature shows the existence of the ignition of the gas mixture for equivalence ratio between 0.7 and 1.3. The experiments demonstrated significant reductions in ignition delay and increased lean burn capability relative to conventional spark ignition. Fast development of a flame kernel is then observed. OH and CH PLIF experiments were performed to confirm the large OH and CH streamer-induced production over the discharge volume. Papers [Bak et al, 2011; Stancu et al, 2010] discuss an important question on the channels of molecular oxygen dissociation in pulsed discharges. In [Bak et al, 2011] time-resolved emission measurements for N 2 (C-B) and N 2 (B-A) transitions were carried out in nanosecond pulsed discharges in air and pure nitrogen. 0-D kinetic simulations coupled with energy equation are conducted to predict quenching rate coefficients of quenching of N 2 * by N 2 and dissociative quenching of N 2 * by O 2 by matching the simulated emission curves to the Aeronautics and Astronautics 356 corresponding measurements. The dissociative quenching was found to be responsible for 82 % of O production whereas the electron-impact dissociation was ~5%. Papers [Pai et al, 2009, Stancu et al, 2010] reports the results of investigations of nanosecond repetitively pulsed discharge in atmospheric pressure discharge in air or nitrogen preheated at 1000 K. The ground state of atomic oxygen was measured by two-photon absorption laser induced fluorescence, the density of N 2 (A) was measured by cavity ring down spectroscopy and the densities of N 2 (B) and N 2 (C) were measured by optical emission spectroscopy. Measurements of O, N 2 (B) and N 2 (C) densities have confirmed that the formation of atomic oxygen occurs through the fast two-step mechanism through excitation and quenching of nitrogen triplet states [Stancu et al, 2009]. Papers [Wu et al, 2010, 2011] present measurements of time evolution of hydroxyl radicals in premixed hydrocarbon-air flow in the afterglow of a nanosecond pulsed discharge at atmospheric pressure. The temperature ranged from 300 to 800 K. The fuels were methane, ethane, propane and butane, at an equivalence ratio of 0.1. The plasma was generated by 20 kV pulses of 10 ns duration with < 1 ns rise time at repetition rate of 10 Hz. The tip electrode shape ensured a stable streamer discharge. The reactant flow rate was set at ~20 cm/s so that each discharge pulse occurred in a fresh gas mixture. Laser induced fluorescence was used to measure the concentration of OH radicals after the discharge. The energy of the excitation laser was adjusted to insure that the measurements were made under saturation conditions for all experiments. The time evolution of OH radicals was tracked by adjusting the delay time between the high-voltage pulse and the concentration measurement. It was shown that the OH concentration demonstrates three maxima: immediately after discharge, on time scale ̴ 100 µs, and the third ̴ 2-5 ms after the initiation. This behavior demonstrates relatively long chains development under low temperature conditions below self-ignition threshold. The important conclusion was made in [Uddi et al, 2011; Wu et al, 2010; Wu et al, 2011] that a new, validated mechanism for low temperature hydrocarbon combustion is required for qualitative description of plasma assisted combustion below self-ignition threshold. This problem is still unsolved at require a lot of new efforts. 3. Plasma assisted combustion above self-ignition threshold Kinetics above self-ignition threshold is relatively good understood for hydrogen and small hydrocarbons. Verified kinetic models exist for all saturated hydrocarbons from methane to n-decane at temperatures T > 1000-1200 K and pressures from several Torr to several atmospheres. Presence of detailed chemical models simplifies the analysis of plasma assisted combustion experiments in this range of parameters. The only difference between auto-ignition and plasma assisted ignition is a high concentration of radicals from the very beginning of the process and potential influence of non-equilibrium mechanisms with participation of vibrationally- and electronically- excited particles and ions. The challenge of high-temperature experiments is the controllable heating of the mixture in combination with the homogeneous non-equilibrium excitation by gas discharge. The problem was solved in [Kof&Starikovskiy, 1996-1; 1996-2] where the combined excitation of the combustible mixture by shock wave and fast ionization wave was proposed. Experimental installation was based on the shock tube coupled with the discharge section. Discharge was generated by Marks-type high-voltage pulse generator. The generator consisted of 10 steps and operated at U = 80-250 kV. Ferrite line with non-linear response Plasma-Assisted Ignition and Combustion 357 and an impedance of 40 Ohm allowed to decrease the pulse leading front down to 500 ps. The voltage increase rate on a high voltage electrode was up to 500 kV/ns and allowed a fast ionization wave formation in the discharge section. Ignition delay time was analyzed for oxygen-hydrogen mixtures and numerical analysis of chemical kinetics was performed for simultaneous mixture excitation by shock wave and high voltage ionization wave. Ionization wave influence (U ~ 250 kV, t pulse ~ 40 ns) on the ignition delay time of the mixture H 2 :O 2 :N 2 = 5:19:76 at p = 1 atm was investigated. High efficiency of the fast ionization wave for spatially-uniform excitation of the chemically- reacting systems has been found [Kof&Starikovskiy, 1996-1; 1996-2]. The experimental works using this installation show the high efficiency of this methodology for high- temperature plasma assisted combustion investigation (see, for example, [Aleksandrov et al, 2009-1; 2009-2; Kosarev et al, 2009; 2008; 2008-2; Starikovskii et al, 2006; Starikovskii, 2005; Starikovskaia et al, 2004; Bozhenkov et al, 2003; 2002]). The kinetics of ignition in C n H 2n +2O 2 :Ar mixtures for n = 2 to 5 has been studied experimentally and numerically after a high-voltage nanosecond discharge [Kosarev et al, 2008]. The ignition delay time behind a reflected shock wave was measured with and without the discharge. It was shown that the initiation of the discharge with a specific deposited energy of 10–30 mJ/cm 3 leads to an order of magnitude decrease in the ignition delay time. Discharge processes and following chain chemical reactions with energy release were simulated. The generation of atoms, radicals and excited and charged particles was numerically simulated using the measured time-resolved discharge current and electric field in the discharge phase. The calculated densities of the active particles were used as input data to simulate plasma-assisted ignition. The sensitivity of the results to variation in electron cross sections, reaction rates and radical composition was investigated. Good agreement was obtained between the calculated ignition delay times and the experimental data. The analysis of the simulation results showed that the effect of nonequilibrium plasma on the ignition delay is associated with faster development of chain reactions, due to atoms and radicals produced by the electron impact dissociation of molecules in the discharge phase. Finally, we studied the role of various hydrocarbon radicals in the plasma-assisted ignition of the mixtures under consideration. Figures 21,a-d show the delay times measured and calculated in [Kosarev et al, 2008] in C 2 H 6 - to C 5 H 12 -containing stoichiometric mixtures with oxygen with 90% Ar dilution as a function of the gas temperature for autoignition and plasma-assisted ignition. The effect of gas discharge leads to a drastic decrease in the ignition delay and to ignition of the mixtures at noticeably lower temperatures and gas number densities. Good agreement between the measured and calculated ignition delay time after the discharge in most cases studied shows that the developed kinetic model adequately describes PAI under the conditions considered. Simulation of discharge processes was also validated by comparison between calculated and measured temporal evolution in the discharge current and in the specific energy deposited in the discharge phase. It should be mentioned that kinetic model of active particles formation in the discharge used in [Kosarev et al, 2009; Aleksandrov et al, 2009] is significantly simplified and rate coefficients of number of processes are not very well known. It is related to ions composition and in part to composition of hydrocarbon radicals. Fortunately under conditions of typical lean mixtures combustion the atomic oxygen always plays a major role. Atomic hydrogen and hydrocarbon radicals are less important but processes of their formation are also well investigated and could be modeled with rather high accuracy. Uncertainty in the radical’s Aeronautics and Astronautics 358 relative composition is not of critical importance under such conditions because the ignition delay time and rate of chemical energy release at high temperatures does not significantly depend on the radical’s nature [Aleksandrov et al, 2009]. Thus even the plasma-chemical systems are very complex and some processes are not investigated in details at the moment, plasma assisted ignition and combustion at high temperatures are controlled by rather simple and well-understood mechanisms and radicals. Fig. 21. The delay time for autoignition and ignition with the discharge as a function of temperature. Closed symbols correspond to measurements and open symbols correspond to calculations. a) C 2 H 6 :O 2 :Ar mixture; b) C 3 H 8 :O 2 :Ar mixture; c) C 4 H 10 :O 2 :Ar mixture; d) C 5 H 12 :O 2 :Ar mixture [Kosarev et al, 2008]. 3.1 Vacuum Ultraviolet Emission of the discharge It is well known that both equilibrium and non-equilibrium plasma are strong sources of vacuum ultraviolet radiation (VUV). Absorption of VUV radiation by oxygen leads to molecular oxygen dissociation with quantum efficiency close to one. Thus ultraviolet sources potentially can generate high concentration of active radicals in the gas. In [Berezhetskaya et al, 2005] the possibility to utilize the discharge VUV self-emission for ignition stimulation has been considered. The pulsed microwave radiation was generated by [...]... P=1atm, temperature Aeronautics and Astronautics O2/CH4 Kinetic Mechanisms for Aerospace Applications at Low Pressure and Temperature, Validity Ranges and Comparison Fig 6 Φ=1.0, P=1atm, temperature Fig 7 Φ=1.1, P=1atm, temperature Fig 8 Φ=1.3, P=1atm, temperature 381 382 Fig 9 Φ=1.5, P=1atm, temperature Fig 10 Φ=1.7, P=1atm, temperature Fig 11 Φ=1.9, P=1atm, temperature Aeronautics and Astronautics O2/CH4... which adopts 32 species and 177 reactions; the GRI-Mech 12 (32 species and 177 reactions) usually fits the reference values better than the other mechanisms; Ignition delay times at P=3atm and P=5atm are also compared and provided Here tables 12a-12b and 13a-13b, report the ignition delay times of the GRIMech3.0 detailed mechanism at P=3atm and P=5atm Reactants Temperature, K 100 0 1100 1200 1300 1400 1500... s, P=5atm Tables A1-A10 and A11-A20 (in the appendix) report, respectively, the percentage differences at P=3atm and at P=5atm Tables A1-A20 show that reaction mechanisms worse their accuracy increasing the operating pressure to P=3atm and P=5 atm, in fact: the 1-step and 2-step Westbrook and Dryer mechanisms start igniting only at Φ= 1.0 and Φ=1.1, respectively, both for P=3atm and P=5atm (at P=1atm... Westbrook and Dryer: 4 species and 1 reaction (Westbrook & Dryer, 1981); 2 Westbrook and Dryer: 5 species and 2 reactions (Westbrook & Dryer, 1981); 3 Minotti: 6 species and 2 reactions (Minotti et al., 2009); 4 Kee: 17 species and 58 reactions (Kee et al., 1985); 372 Aeronautics and Astronautics 5 GRI-Mech 12: 32 species and 177 reactions (Gri-Mech 1.2, 1994; Heffington et al., 1997); These mechanisms... semiglobal (multistep) models (Wesbrook & Dryer, 1981; Bowman, 1986), and finally to the inclusion of full detailed chemical kinetic mechanisms to better simulate chemistry interactions In addition, detailed mechanisms have been developed and validated for the 370 Aeronautics and Astronautics simplest fuel molecules (Westbrook and Dryer, 1981) and are not available for most practical fuels Finally there are... important at high values of E/n The experiments and calculations show that the fraction of the discharge power spent on “fast” gas heating increased from 10% at E/n = 100 Td to 30–55% at E/n = 100 0 Td This noticeably depended on gas pressure and only slightly on the electron density at  = d, nef The effect of pressure was negligible at E/n = 100 Td and became more profound at high E/n, at which most... temperature range 100 0K - 2000K and at pressure 1, 3 and 5atm The equivalence ratio (Φ) range tested was from Φ=0.3 to Φ=1.9 (ΔΦ=0.2), plus Φ=1 Table 1a and Table 1b provide the ignition delay times, tid, predicted by the reference detailed GRI-Mech 3.0 mechanism as function of temperature, for P=1atm, and at Φ previously indicated (Tables 12a-12b and 23a-23b report data respectively at P=3atm and P=5atm)... Boltzmann equation for electrons and of the balance equations for active particles Here, input data are electron-molecule cross sections and rate constants for reactions with excited and charged particles These data are available for simple molecules such as N2, O2, H2, and, to a smaller extent, for simple hydrocarbons However, little is known about cross sections and rates for complex hydrocarbon... K 100 0 1100 1200 1300 1400 1500 1600 1700 1800 1900 2000 Aeronautics and Astronautics Φ=0.3 0.125 0.0234 0.00577 0.0017 0.000578 0.000228 0.00 0101 0.0000519 0.0000303 0.0000202 0.0000154 Φ=0.5 Φ=0.7 Φ=0.9 Φ=1 0.134 0.153 0.17 0.172 0.0274 0.0322 0.0362 0.0382 0.00682 0.00827 0.00923 0.00982 0.00204 0.00242 0.00272 0.00282 0.000685 0.000738 0.000886 0.000921 0.000255 0.000285 0.000316 0.000331 0.00 0106 ... Meeting Orlando, Florida Jan 2011 Paper AIAA-2011-1271 Tachibana K., Phys Rev A 34 (1989) 100 7 101 5 Uddi M., Guo H., Sun W., Ju Y Studies of C2H6/air and C3H8/air Plasma assisted combustion kinetics in a nanosecond discharge 49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition 4 - 7 January 2011, Orlando, Florida AIAA 2011-970 Wu L., Fridman A.A and Starikovskiy . quenching and conversion into radicals in reactions with fuel molecules, and 2) at high E/n – through direct dissociation of molecular oxygen by electron impact and Aeronautics and Astronautics. chemical and 8700 vibrational exchange processes with participation of 254 particles including electron- excited and charged atoms and molecules, electrons, radicals, non-excited components, and. Horizons Forum and Aerospace Exposition 4 - 7 January 2 010, Orlando, Florida AIAA 2 010- 1593 Wu L., Lane J., Cernansky N., Miller D., Fridman A. and Starikovskiy A. Time resolved PLIF and CRD diagnostics

Ngày đăng: 19/06/2014, 12:20

Tài liệu cùng người dùng

Tài liệu liên quan