Вы находитесь на странице: 1из 10

THE AERONAUTICAL JOURNAL

NOVEMBER 2004

575

Effects of hydrogen-air non-equilibrium chemistry on the performance of a model scramjet thrust nozzle
R. J. Stalker, N. K. Truong, R. G. Morgan and A. Paull Division of Mechanical Engineering University of Queensland Brisbane, Australia

ABSTRACT
Two aspects of hydrogen-air non-equilibrium chemistry related to scramjets are nozzle freezing and a process called kinetic afterburning which involves continuation of combustion after expansion in the nozzle. These effects were investigated numerically and experimentally with a model scramjet combustion chamber and thrust nozzle combination. The overall model length was 05m, while precombustion Mach numbers of 3103 and precombustion temperatures ranging from 740K to 1,400K were involved. Nozzle freezing was investigated at precombustion pressures of 190kPa and higher, and it was found that the nozzle thrusts were within 6% of values obtained from finite rate numerical calculations, which were within 7% of equilibrium calculations. When precombustion pressures of 70kPa or less were used, kinetic afterburning was found to be partly responsible for thrust production, in both the numerical calculations and the experiments. Kinetic afterburning offers a means of extending the operating Mach number range of a fixed geometry scramjet.

Subscripts 1 s precombustion conditions reservoir conditions

1.0 INTRODUCTION
The theoretical promise of supersonic combustion ramjets (i.e. scramjets) as a means of high speed flight propulsion has been known for some decades(1,2) and has stimulated research into the operation of the components of scramjets. One such component, which is responsible for producing a major part of the overall engine thrust, is the thrust nozzle that follows the combustion chamber. In this nozzle, thrust is generated via the pressure distribution induced on the nozzle surfaces by expansion of the combustion gases. The combustion and expansion of these gases may take place so rapidly that they are unable to approach, or are unable to maintain, chemical equilibrium, with resulting effects on the pressure distribution in the thrust nozzle. This paper reports an investigation of such non-equilibrium effects in combusting hydrogen-air mixtures. The phenomenon known as freezing occurs when a chemically reacting gas undergoing a rapid expansion in a nozzle from an equilibrium state, is unable to adjust its composition rapidly enough to maintain equilibrium, and the composition of the gas freezes at some point in the nozzle expansion process. This implies that the energy release accompanying the formation of equilibrium product species does not continue, and the thrust generated by the nozzle flow is reduced. An early study(3) of freezing of hydrogen-air chemical recombination in a ramjet thrust nozzle, where the expansion takes place from a low subsonic Mach number, showed that if equilibrium flow is maintained past the low supersonic area ratios the loss in thrust caused by chemical freezing is not exceedingly high. In later numerical studies of scramjet thrust nozzles at high flight Mach numbers, where the nozzle expansion takes place from supersonic Mach

NOMENCLATURE
ci dci F Hs P Pc P1 Ps T x y mass fraction of species i mass fraction contour increments nozzle thrust (N) stagnation enthalpy (MJ/kg) pressure (kPa) maximum combustion pressure (kPa) precombustion pressure (kPa) reservoir pressure (kPa) temperature (K) streamwise distance (m.) cross stream distance (m.) ratio of specific heats hydrogen equivalence ratio

Paper No. 2780. Manuscript received 5 August 2002, revised version received 26 May 2004, accepted 29 September 2004.

576

THE AERONAUTICAL JOURNAL

NOVEMBER 2004

numbers, it was shown that although molecular vibration was in equilibrium throughout the nozzle, the chemical composition was not(4), that at precombustion pressures of the order of an atmosphere there was little recombination in the nozzle(5) and that, in spite of this, the lack of complete recombination yielded an overall thrust which was only 1% less than the value for an equilibrium expansion(6). Also, a study of thrust loss mechanisms in scramjets(7) showed that nozzle freezing represented only a small fraction of the overall losses. Thus the freezing phenomenon is not expected to have a substantial effect on nozzle performance. However, another non-equilibrium phenomenon can occur in a scramjet thrust nozzle, namely the process described here as kinetic afterburning. This takes place when the combustion reactions are unable to approach chemical equilibrium in the combustion chamber before they encounter the falling pressure and temperature in the nozzle. The resultant quenching effect may influence some of the combustion reactions more than others but, although the pressure distribution in the nozzle may be changed, energy release leading to thrust may still take place in the nozzle. Kinetic afterburning differs from conventional afterburning, where thrust is produced by adding fuel to the oxidizer-rich exhaust nozzle flow from a gas turbine. In kinetic afterburning, fueloxidizer mixing and the initial stages of the combustion reactions take place upstream of the exhaust nozzle, but the finite time required to complete the combustion reactions and produce energy release is exploited to allow energy release and thrust production to take place in the nozzle. This can be advantageous in avoiding high peak pressures in the combustion chamber. The possibility of kinetic afterburning was suggested by a study of the chemical kinetics of the ignition process in hydrogen/air mixtures(8), where it was pointed out that, after initial production of free radicals, the subsequent ignition process was less dependent on temperature than on the presence of those free radicals. This suggests that combustion chamber reactions may continue in the lower temperature regions created by a nozzle expansion downstream of the combustion chamber, leading to heat release in the nozzle. It is also believed that this effect may be responsible for a phenomenon observed in a previous paper (see Fig. 4 of Ref. 9) where, in a study of thrust production in a model scramjet, evidence of heat release in the nozzle was provided by noting that the difference between fuel-on and fuel-off thrust rose more rapidly than the pressure rise due to combustion in the combustion chamber. The present investigation was undertaken to obtain a better understanding of the non-equilibrium effects in a scramjet nozzle by comparing numerical models with experimental results. After describing and discussing the experimental arrangements and the numerical program employed, the paper goes on to separately consider nozzle freezing and kinetic afterburning. Considering each of these phenomena in turn then leads on to the conclusion.

Figure 1. Model configurations.

2.2 The models A scaled outline of the basic model is shown in Fig. 1(a) and was used in this configuration for the test series B, outlined below. The model was two-dimensional, with an internal width throughout of 50 mm., and consisted of a constant area combustion chamber, 25mm high, which extended for 175mm downstream of the hydrogen injection station. At the combustor exit, one 50mm wide surface of the duct was deflected to form a simple thrust nozzle, and the deflected surface was instrumented with PCB piezoelectric pressure transducers located as shown. Two transducers were also located as shown to monitor the pressure rise in the combustion chamber, and a further two transducers were located 31mm and 68mm downstream of the intake to monitor the precombustion pressure. The upstream transducer was preferred for this role, as it was less susceptible to possible boundary layer effects. The model was supplied with test gas directly from the shock tunnel nozzle for the configuration of Fig. 1(a), implying that it was operated in the direct connect mode. The fuel injection strut which was 47mm thick, was located midway between the upper and lower surfaces of the duct, fully spanning the duct width, thus ensuring that an essentially two dimensional flow was produced. This was checked by sample pressure measurements across the duct. The leading edge of the strut was a symmetric wedge of 20 included angle, and was located sufficiently far upstream of the inlet that the leading edge wave system did not interfere with the inlet leading edge. Hydrogen was injected, at nominal Mach numbers ranging from 12 to 19, from a room temperature reservoir through a two-dimensional supersonic nozzle with a 16mm throat, which spanned the full width of the blunt trailing edge of the injection strut. Precalibration allowed the hydrogen mass flow to be determined by monitoring the hydrogen reservoir pressure. In order to increase the precombustion pressure for the test series A, as outlined below, the model configuration was altered by the addition of a simple intake, as shown in Fig. 1(b). This consisted of two opposing ramps, each set at an angle of 5, which were spaced to ensure that the reflected shock waves passed outside of the inlet, as shown in the figure. These shocks reduced the Mach number to 29 013. 2.3 Thrust nozzle The nozzle configuration was chosen to allow the effects of nonequilibrium chemistry to be adequately represented in the flow, whilst being easy to manufacture. For structural reasons, a scramjet nozzle should be as short as possible, whilst maintaining near optimum thrust performance, implying that nozzle expansion angles

2.0 EXPERIMENTAL APPARATUS


2.1 The shock tunnel The experiments were conducted in the free piston shock tunnel T3 at the Australian National University, in Canberra(10). This shock tunnel employed a shock tube 76mm in diameter and 6m long, with driver gas heating accomplished by a free piston in a tube 300mm. in diameter and 6m. long. It was operated at stagnation enthalpy levels and test times which ensured that the test flow was not affected by driver gas contamination(11). A contoured axisymmetric nozzle with a 25mm diameter throat and a 92mm exit diameter, yielding a nozzle area ratio of 135, was located at the downstream end of the shock tube, and produced a test section Mach number of 33 013.

STALKER ET AL EFFECTS OF HYDROGEN-AIR NON-EQULIBRIUM CHEMISTRY ON A MODEL SCRAMJET THRUST NOZZLE PERFORMANCE
2.5 Test conditions

577

Figure 2. Effect of nozzle angle on thrust, showing typical flow pattern.at 15 degrees, with pressure contours (dimensions shown are in metres).

will tend to be as large as this performance criterion will allow. In the present experiments, shock tunnel test section constraints limited the expansion angle to 15 with the chosen nozzle length. Using the finite rate chemistry computer code described below, the variation of thrust with expansion angle, for the chosen nozzle length, has been calculated with test conditions such that chemical equilibrium is achieved within the combustion chamber length. The reults are presented in Fig. 2, and indicate that a 15 expansion angle approaches the optimum angle for thrust production. The pressure field with an expansion angle of 15 is also shown in Fig. 2, with constant pressure contours displayed at equal intervals of pressure, and the ratio of pressure to the precombustion pressure, P1, displayed at the upstream and downstream contours of the nozzle pressure field. The contours of pressure were found to be essentially identical with those computed for an equilibrium expansion, indicating that approximate estimates of the thrust could be obtained by using a perfect gas expansion with = 125. This value of is not surprising when it is remembered that nitrogen, which does not take part in the reactions, constitutes the major mass component of the flow. Thus, using the ratio between the pressures on the upstream and the downstream nozzle pressure contours of Fig. 2 to obtain an effective nozzle area ratio of 38, it was found that this nozzle produced approximately 60% of the thrust that would be expected from a nozzle with an area ratio of 20. This latter value is typical of those used in theoretical scramjet studies (e.g. Ref. 7). Hence, as the effects of non-equilibrium chemistry are expected to occur at low nozzle area ratios (e.g. Ref. 3), this nozzle may be expected to provide a reasonable representation of both the nature and the magnitude of the effects of non-equilibrium chemistry. 2.4 Instrumentation The pressure transducers mounted in the model had natural frequencies ranging from 250kHz to 500kHz. The transducer outputs were multiplexed and recorded by a Biomation 2805 M transient digital data recorder with slave unit, recording a complete data set every sixteen microseconds. An Apple personal computer was then used to process and display the data after each test. The output from the hydrogen pressure monitoring transducer was displayed on a Tektronix CRO, and was constant for the duration of the test flow.

The test conditions were obtained by using measurements of the shock speed and the shock tube filling pressure to obtain conditions immediately after shock reflection, and then assuming an isentropic expansion to the post shock nozzle reservoir pressure, measured 67mm from the end of the shock tube. A one dimensional equilibrium nozzle expansion from the reservoir condition to the measured pressure at the station 31mm downstream of the model inlet then yielded the test conditions. To ensure that these test conditions would apply to the same element of air, account was taken of the time required for passage of the air from the reservoir to the inlet measurement station, and from the inlet station to the scramjet nozzle. Two groups of tests were conducted, and are referred to as Series A and Series B. Series A was related to nozzle freezing and, as it was necessary to generate precombustion pressures which would allow essentially complete combustion in the combustion chamber, this series was conducted with the intake ramps attached to the model, as shown in Fig. 1(b). Series B was related to kinetic afterburning and the model was used in the basic configuration, as shown in Fig. 1(a), in order to generate the lower precombustion pressures which were necessary to avoid significant levels of combustion in the combustion chamber. The test period was defined by using records of the nozzle thrust, normalised with respect to the shock tunnel nozzle reservoir pressure. The nozzle thrust was obtained through integration of the measured pressures along the nozzle, using a linear interpolation for the pressure between transducer stations. Typical records of the normalised thrust are shown in Fig. 3, where the normalised thrust is seen to rise until it reaches a plateau level, where it displays relatively small variations with time. The test time is taken to occur during this plateau period. The normalised thrust for Series B during the plateau period is not as steady as that for Series A but, even in this case, the normalised thrust varies by only 8% during the test period. During the test period the flow passes a distance of at least three times the model length, which is sufficient to be assured of quasi-steady nozzle flow. Argon and helium, at various mixture ratios, were used for the shock tube driver gas as a means of delaying driver gas contamination of the test flow. The decay of shock tunnel nozzle reservoir pressure in Fig. 3(b) is associated with the use of helium alone as driver gas, while the nearly constant pressure of Fig. 3(a) is associated with the use of argon. Because the hydrogen injection pressure remained constant during a test, the decay in Fig. 3(b) represents an increase in the equivalence ratio over the test time, an increase which is exploited in matching numerical and experimental test conditions.

3.0 NUMERICAL ANALYSIS


Numerical analysis was undertaken using the program Supersonic Hydrogen Air Reaction Calculator (SHARC)(12). This program utilizes the two-dimensional parabolised Navier-Stokes equations to simulate supersonic mixing and combustion, solving these equations by employing a finite difference scheme developed by Patankar and Spalding(13). The solution is implemented through a space marching scheme, and the pressure field is calculated by the algorithm Semi Implicit Pressure Linked Equations (SIMPLE)(14). A two equation k model was used to simulate the turbulence in the flow, with constants as in Ref. 12. For the simulations involving non-equilibrium chemistry, the basic NASP scheme(15) for finite rate hydrogen-oxygen reaction chemistry was used, with nitrogen treated as an inert diluent. This reaction scheme is shown in Table 1. Strictly, to be used for hydrogen-air reactions, this scheme should include the effect of nitrogen oxides, but sample combustor calculations in which hydrogen was premixed with air showed no significant difference in the combustor pressure rise when the reactions involving nitrogen

578

THE AERONAUTICAL JOURNAL

NOVEMBER 2004
Table 1 Hydrogen-oxygen reaction scheme

OH + H2 H+ O2 O + H2 H + HO2 H + HO2 O + HO2 OH + HO2 H + O2 + M H + OH + M H+H+M H+O+M O+O+M OH + OH

H + H 2O O + OH H + OH O2 + H2 OH + OH OH + O2 H2 O + O 2 HO2 + M H2 O + M H2 + M OH + M O2 + M O + H 2O

A 216 108 191 1014 506 104 25 1013 15 1014 20 1013 20 1013 80 1017 862 1021 73 1017 26 1016 114 1017 15 109

n 151 0 267 0 0 0 0 08 20 10 06 10 114

1,726 8,273 3,166 349 505 0 0 0 0 0 0 0 0

Rate constant k = ATnexp(/T) Units are seconds, moles, cubic centimetres and kelvins Third body efficiences for termolecular reactions are: 25 for M = H2, 1625 for M = H2O and 20 for all other M.

Figure 3. Sample reservoir pressure and nozzle thrust records (a) Series A (b) Series B.

nozzle flow field, it was found necessary to limit the pressure gradient to 1Pa/Step in order to avoid a numerical instability. Solutions which were then generated for the combustor-nozzle combination with 460 transverse grid points were identical with those with 800 transverse grid points. Thus the solutions obtained were grid independent. Except where otherwise specified, the initial composition of the air-hydrogen mixture was taken to include one percent by molar concentration of the oxygen to be in the dissociated form, together with 001% each of the total mixture mass to be in the form of H and OH. This was a convenient method of starting the calculation, whilst avoiding detailed consideration of ignition phenomena, and is further discussed below.

4.0 NOZZLE FREEZING


oxides were eliminated, indicating that the reaction scheme of Table 1 was sufficient for the present analysis. The combustor involved injection of hydrogen by a central strut, and an attempt was made to model the ensuing mixing and combustion at the conditions of Fig. 2. It was found that the pressure rise within the combustor fell well below that observed experimentally, but the experiments were consistent with computations involving premixed hydrogen and air, indicating that mixing occurred more rapidly than predicted by the numerical calculations. This was also consistent with experiments reported in Ref. 16, in which shadowgraph images of the wake formed by hydrogen injection in a combustor with dimensions identical to the present one showed that the transverse dimension of hydrogen-air combustion wakes was much greater than that of hydrogen-nitrogen wakes, and almost filled the duct at the downstream station. Though no explanation for this phenomena was given, it was speculated that it was due to combustion enhanced hydrogen-air mixing. Thus the combustor was modelled with premixed hydrogen and air. As seen below, this was a good approximation only when pre-combustion pressures were high enough to produce vigorous combustion in the combustor, which was signified by matching of numerical and experimental pressures at the downstream end of the combustor. The numerical solutions used 110 steps along the combustor length, checks with a smaller step size showing that this yielded solutions which were independent of the step size. In computing the This section concerns flows at sufficiently high temperatures and pressures that combustion is essentially completed in the combustion chamber, and emphasis is placed on freezing of the combustion products in the nozzle expansion which follows. 4.1 Numerical modelling Figures 4 and 5 show results of the numerical analysis for equilibrium chemistry and finite rate chemistry at two precombustion temperatures. These two temperatures represent the lower and the higher temperatures in the range covered by Series A which are noted in Table 2. Contours of mass fraction of OH and H2O are shown, the latter because the amount of H2O formed represents the energy release in the flow, and the former because the distribution of the free radical OH represents the influence of chemical kinetics. For the low temperature case of Fig. 4, the mass fraction of H2O is almost the same for the equilibrium and finite rate cases, indicating that equilibrium is essentially achieved in the combustion chamber. The H2O mass fraction also remains essentially constant during the nozzle expansion. The mass fractions of OH are not high enough to significantly affect H2O mass fractions, but they do show non-equilibrium effects, in that the finite rate mass fraction is lower than the equilibrium value, and it does not decay as rapidly with distance along the nozzle as in the equilibrium case. The high temperature case of Fig. 5 exhibits the same general characteristics, with equilibrium in the combustion chamber and a more rapid nozzle decay of

STALKER ET AL EFFECTS OF HYDROGEN-AIR NON-EQULIBRIUM CHEMISTRY ON A MODEL SCRAMJET THRUST NOZZLE PERFORMANCE

579

Figure 4. Mass fraction contours low precombustion temperature (a) Finite rate (b) Equilibrium.

Figure 5. Mass fraction contours high precombustion temperature (a) Finite rate (b) Equilibrium.

the OH mass fraction in the equilibrium case than in the finite rate case. However, the nozzle expansion exhibits a larger increment of H2O formation than in the low temperature case. The relatively low OH concentrations lead to the expectation that nozzle freezing, indicated by the relatively higher values of OH mass fraction at the nozzle exit in the finite rate case, will not cause large thrust losses. Thus, for the low temperature case of Fig. 4, the ratio of finite rate thrust to equilibrium thrust was 0934 and, for the high temperature case of Fig. 5, the ratio was 0956. This is in accord with previous investigations(3-7), where nozzle freezing caused only a small loss in thrust. On the other hand, by using the equivalence ratio given for Figs 4 and 5, it is found that the largest possible mass fraction of H2O formed for complete reaction, is 0187 for Fig. 4 and 0190 for Fig. 5. Therefore, at the nozzle exit, only 86% of potential H2O formation has occurred in the equilibrium low temperature case, and 69% in

the equilibrium high temperature case. Estimates show that a further equilibrium expansion to an area ratio of approximately 20 would alter this figure by less than 2%. This indicates that the loss of thrust due to the higher precombustion temperature is more serious than that due to nozzle freezing described in the previous paragraph. This accords with the widely held appreciation of the desirability of avoiding high precombustion temperatures (e.g. Ref. 17). The pressure fields corresponding to Figs 4 and 5 were generally similar to that shown in Fig. 2, with an expansion fan centred on the combustion chamber-thrust nozzle corner, and subsequent reflection of that expansion onto the thrust surface. In Fig. 4(a), the OH field displays waves originating at the combustion chamber walls near the station where the reaction in the mainstream begins. These waves were more pronounced in the corresponding pressure field, but did not appear for the other cases of Figs 4 and 5. The waves may be due to early combustion and energy release in the higher temperature

580
Table 2 Experimental test conditions

THE AERONAUTICAL JOURNAL

NOVEMBER 2004
Table 3 Normalised thrust Series A

Quantity Units Exptl error Series A A1 A2 A3 A4 Series B B1 B2 B3 BA

Hs MJ/kg 4% 39 31 26 21 40 35 32 38

P1 kPa 8% 335 425 190 230 67 69 65 117

T1 K 50K 1,400 1,180 910 740 1,220 1,020 940 1,180

U1 ms1 2% 2,190 1,930 1,810 1,640 2,340 2,140 2,110 2,250

M1

Condition F/Pc (m2) Experiment Numerical (finite rate)

A1 094 099

A2 075 (088) 093

A3 090 090

A4 085 086

15% 4% 074 055 094 070 115 097 124 114 29 28 30 30 33 33 34 33

Figure 6. Numerical and experimental pressure distributions Series A.

regions of the boundary layer, with the resulting increase in the boundary layer displacement thickness leading to the mainstream waves observed. 4.2 Comparison with experiment Series A Numerically predicted pressure distributions are compared with experiment in Fig. 6, for test cases covering precombustion temperatures ranging from 1,400K for A1 to 740K for A4, and precombusion pressures between 190kPa and 425kPa. Since

interest is centred on the scramjet nozzle performance, pressures are normalised with respect to the maximum combustion pressure, which is the pressure at the station immediately upstream of the entrance to the nozzle. The experimental precombustion conditions corresponding to each of the test cases are given in Table 2. As already noted, the equivalence ratio increased during the test time and therefore it was possible to match experimental and numerical equivalence ratios. The experimental error in equivalence ratio, quoted in Table 2, ensures that this is only a nominal match, but the normalised nozzle pressure distribution is found to be insensitive to equivalence ratio. For example, the time-wise rate of variance of the equivalence ratio is 103/sec for test case A1, which leads to an increase of 02 in the equivalence ratio during the time taken for the mainstream to pass twice the nozzle length, but the resulting changes in the normalised pressure distribution are such that they would be barely detectable if plotted in Fig. 6. The rate of variance in equivalence ratio is less for the other cases of series A, and therefore there is even less effect on the other normalised pressure distributions. The flow produced by the shock tunnel may include trace species, due to non-equilibrium effects in the shock tunnel nozzle flow. These trace species are represented in the numerical models by allowing the oxygen content of the test flow to include one percent of oxygen atoms. The influence of this assumption for initial flow composition was checked by performing calculations with the oxygen atom concentration, and the associated H and OH concentrations, reduced by a factor of ten. The result of these calculations for condition A3 are shown by the dotted curve in Fig. 6. The pressure distribution in the combustor is affected, but not in the nozzle. For higher precombustion temperatures it was found that the effects were negligible but, for the lower temperature, major effects took place, with no combustor pressure rise and the thrust reduced to 30% of the value with one percent of oxygen atoms. Figure 6 shows that the experimental nozzle pressure distributions closely match those predicted numerically for conditions A3 and A4, indicating that a sensibly uniform flow entered the scramjet nozzle and that combustion therefore was completed in the combustion chamber. For condition A1, the experiments show the length of the constant pressure region in the expansion to be somewhat shorter than that predicted by the numerical model. The pressure level at 0195m from the inlet indicates that, for this condition, the mixing and combustion process is less complete than for the other conditions, leading to possible transverse flow nonuniformities, which may cause compression and expansion waves to be incident on the thrust surface(18). For condition A2, the pressure levels in the constant pressure region fall below those predicted. For this case, it was necessary to obtain the experimental pressure distribution just before the test period commenced, in order to match the equivalence ratio, and therefore the pressure distribution does not represent quasi-steady flow. The normalised pressure distribution 250 micro sec later, which was well into the test period (and the equivalence ratio was 078), matched the numerical predictions satisfactorily. Thus the experimental results at condition A2 illustrate the asymptotic approach to the quasisteady flow condition. A figure of merit for the nozzle flow can be obtained as the thrust normalised by the maximum combustion pressure. Linear

STALKER ET AL EFFECTS OF HYDROGEN-AIR NON-EQULIBRIUM CHEMISTRY ON A MODEL SCRAMJET THRUST NOZZLE PERFORMANCE

581

Figure 8. Further development of nozzle combustion flow contours.

5.0 NOZZLE COMBUSTION KINETIC AFTERBURNING


Kinetic afterburning can be taken as occurring when the combustion reaction is interrupted by the nozzle expansion, which reduces the temperature and density of the combustion reactants, thus delaying completion of the reactions. However, the changes in temperature and density are not sufficient to completely quench the reactions, and they proceed to produce substantial heat release in the nozzle. 5.1 Numerical model kinetic afterburning with premixed hydrogen Numerical calculations were used to illustrate the development of kinetic afterburning, as shown in Figs 7 and 8. For Fig. 7, precombustion conditions were fixed and the initial reaction rate was varied by altering, by the same fraction, the mole fraction of oxygen atoms, and the mass fractions of H and OH. The occurrence of kinetic afterburning is apparent in the water mass fraction contours, which are presented in equal mass fraction increments. At the lowest oxygen atom mole fraction, water production takes place only near the downstream end of the combustion chamber and, on the thrust surface side of the flow, it is delayed by the centred Prandtl Meyer expansion from the thrust nozzle corner. On the straight surface side, the flow passes further downstream before meeting the leading edge of the expansion fan, and the spreading of the fan reduces the streamwise pressure gradients. As a result, the combustion reactions proceed further than on the thrust side, and more water production occurs before the flow enters the nozzle flow field. On the thrust surface side water production tends to be delayed until further downstream, causing the displacement of H2O mass fraction contours in the downstream direction. Closer to the thrust surface, these mass fraction contours reverse and are displaced upstream, an effect which is ascribed to boundary-layer heating. As the oxygen atom mole fraction is increased, thereby increasing the reaction rate, water production in the combustion chamber is increased, with a corresponding reduction in the production of water in the nozzle. At the highest oxygen atom mole fraction, it can be seen that almost all the water production takes place in the combustion chamber and, as in

Figure 7. Development of nozzle combustion flow contours.

regressions performed on the thrust and the maximum combustion pressure data showed that thrust values varied by less than 14% from the regression line over the test time, while the maximum combustion pressure varied by less than 4%. The value of normalised thrust was obtained from the two regression lines at the time corresponding to the equivalence ratios of Fig. 6, and yielded the values presented in Table 3. Since the value for A2 was obtained while the flow was still developing, the value obtained 250 sec later is also presented in brackets. If it is accepted that the increase in equivalence ratio does not significantly alter the normalised pressure distribution, the normalised thrust presented in brackets can be taken as that corresponding to condition A2. The normalised thrust values obtained numerically, assuming finite rate chemistry, are also presented. The experimental values are consistent with the numerical ones to within 6%. Thus, noting the ratios of finite rate thrust to equilibrium thrust obtained in Section 4.1, it can be concluded that, for the conditions of these experiments, the thrust was within 10% of the equilibrium value. This accords with results of previous numerical studies indicating that thrust was only weakly affected by nozzle freezing.

582

THE AERONAUTICAL JOURNAL

NOVEMBER 2004

Figure 9. Experimental pressure distributions with nozzle combustion Condition B2.

Figs 4 and 5, there is very little water production and corresponding heat release in the nozzle. As is evident in the pressure contours presented at the top of Fig. 7 (where the oxygen atom mole fraction is at its lowest value), the pressure field resulting from nozzle combustion is different from the typical high pressure case presented in Fig. 2. The Prandtl Meyer expansion exists in distorted form, while the nozzle flow exhibits two compression waves, which were not evident in Fig. 2. The upstream compression wave is due to the pressure rise at the thrust surface caused by the heat release associated with delayed water production on the thrust surface side. The downstream compression wave is due to reflection from the thrust surface of the compression wave caused by the pressure rise at the straight surface associated with water production at the upstream edge of the expansion wave. Figure 8 shows the effect of further reduction in the reaction rate, together with an increase in the equivalence ratio. The reduction in reaction rate is achieved through lowering of the precombustion pressure and temperature, and is seen to lead to less water formation on the thrust surface side upstream of the expansion, together with a more clearly defined region of delayed reaction, and greater water formation, downstream of the expansion. The greater heat release generates a stronger upstream compression wave than in Fig. 7. The heat release at the upstream edge of the expansion wave persists, producing the downstream compression wave as in Fig. 7. The effect of increased equivalence ratio is apparent in the larger mass fraction of water formation than in Fig. 7, leading to the stronger upstream compression wave. 5.2 Experiment Series B Experiments were done with the model configuration of Fig. 1(a) (i.e. without the intake ramps), and the zone of the combustion reactions was displaced downstream by reducing the precombustion

pressure. With vigorous combustion no longer occurring in the combustion chamber, the wake formed by hydrogen injection grew much less rapidly with downstream distance(16). Therefore, in contrast to series A, the flow could not be represented by a premixed model. The numerical model of Figs. 7 and 8 can now only be used to guide a qualitative discussion of the flow structure. The experiments involved precombustion temperatures ranging from 1,220K to 940K and, with the exception of condition BA, precombustion pressures of approximately 67kPa. Condition BA shows the effect of increasing this pressure by a factor of 175. Figure 9 presents a series of pressure distributions in which nozzle combustion occurs for condition B2. The pressure distributions are normalised with respect to the pressure immediately downstream of the inlet. To avoid fluctuations of 15% in this pressure which occurred during the test time, values were taken from a linear regression fitted over the test time. A number of mechanisms, including heat transfer to the walls of the shock tube, ensure that the shock tunnel reservoir conditions and therefore the test flow characteristics, are not perfectly steady. However, the time interval between pressure distributions is sufficient for the flow to pass almost one nozzle length (which is usually regarded as sufficient to establish steady inviscid flow), and the changes between pressure distributions are sufficiently small to regard each pressure distribution as representing a close approach to steady flow. On this basis, a tentative qualitative interpretation of the main features of the flow follows. At 624 microseconds, some combustion and heat release occurs in the combustion chamber, and this produces compression waves which raise the pressure on the thrust surface immediately downstream of the expansion corner. The small pressure peak at 031m from the inlet is caused by the interaction of the corner expansion wave with the low Mach number wake(18), and is followed by a larger pressure peak at 042m from the inlet. This second peak is ascribed to delayed water production and heat release on the thrust surface side of the wake. At 720 microseconds, the zone of combustion reactions has moved downstream, the combustion chamber pressure rise has disappeared, and the compression waves which raise the pressure on the thrust surface near the corner are somewhat displaced downstream. At 816 microseconds, the combustion reactions have moved further downstream and, because the wake flow on the thrust surface side meets the expansion at an early stage of the combustion reactions, the reactions are almost quenched, and the associated pressure peak is reduced. However, as in Figs 7 and 8, the combustion reactions on the straight surface side of the wake progress much further before reaching the expansion, causing a compression wave to be propagated towards the thrust surface, where it produces the pressure rise at 050m from the inlet. At 912 microseconds, further downstream movement of the combustion reactions produces some reduction in the pressures downstream of the expansion corner, together with a somewhat increased final pressure rise. The pressure distribution at 816 microseconds in Fig. 9 is accompanied by a broken curve, which shows the pressure distribution obtained with the numerical model. The numerical model does not incorporate the wake-like flow structure of the experiments, and therefore does not exhibit the compression waves due to heat release in the combustion chamber. The first peak in the numerical pressure distribution arises from the delayed combustion on the thrust surface side of the nozzle, and the second arises from combustion on the straight surface side. These two peaks appear in the experiments but, because of the wake-like structure of the heat release zone, are further downstream. It will be noted that there is a reduction in the pressure immediately upstream of the expansion corner at the three later times. This effect was also observed in fuel-off tests. Tracing of waves from the trailing edge of the injection strut indicates that this is probably due to the pattern of compression and expansion waves created by the injection strut and its wake. Experiments at conditions B1 and B3 yielded results which were qualitatively similar to Fig. 9, with some difference in the detail of

STALKER ET AL EFFECTS OF HYDROGEN-AIR NON-EQULIBRIUM CHEMISTRY ON A MODEL SCRAMJET THRUST NOZZLE PERFORMANCE
Table 4 Thrust comparison-experiment

583

Test Condit F/P1 (m2)

A1 22

A2 22

A3 25

A4 24

B1 24

B2 24

B3 25

BA 24

Figure 10. Pressure distribution at an intermediate pressure level Condition BA.

the pressure distributions. At the higher temperatures of B1, the pressure peak at 042m. persisted throughout the test time, while for the lower temperatures of B3, it made a more transient appearance than in Fig. 9. Apparently the higher reaction rates associated with the higher temperatures favoured delaying, rather than quenching, the combustion reactions. For both B1 and B3, the pressure rise in the combustion chamber showed the same pattern of behaviour as in Fig. 9, in spite of the difference in precombustion temperatures. This may be an indication that the concentration of trace species of free radicals in the flow produced by the shock tunnel, or by the flame holding action of the blunt trailing edge of the hydrogen injection strut, is sufficient to bypass the initial, temperature dependent, part of the combustion process. It should be noted that, in both Figs. 8 and 9, kinetic afterburning is not the sole agency causing the combustion induced pressure increases in the nozzle, as these increases are partly due to combustion generated compression waves which may originate upstream of the thrust nozzle corner expansion. Figure 10 displays the effect of increasing the precombustion pressure to a level intermediate between series A and series B, at a temperature close to that of B1. Although the experimental model is retained in the configuration 1(a), the pressure distribution is similar to those in Series A, and there is no indication of kinetic afterburning. The pressure distribution resulting from a numerical calculation at 001%O, is also shown, and exhibits the same lack of any indication of kinetic afterburning. Clearly the kinetic afterburning effect is sensitive to the precombustion pressure, through the level of combustion in the combustion chamber. 5.3 Thrust levels Because there is a loss in thermodynamic efficiency associated with heat addition at a lower temperature, and because the effect of lower nozzle pressures on the chemical kinetics may result in reduced water formation, it would be expected that nozzle combustion would lead to a reduction in thrust. This is borne out by results from the numerical model. For the three conditions presented in Fig. 7, corresponding to oxygen atom molar concentrations of 001%, 01% and 1% the values of thrust normalised by the precombustion pressure are 235, 267 and 283 m2 respectively. Thus, as combustion in the nozzle is reduced, the thrust produced is increased by 17%, in accord with expectations. Normalised thrusts were also obtained experimentally for Series A and Series B. Linear regressions over the test times were performed for the thrust and the precombustion pressure, and values were taken from the regression lines, with allowance for the time of passage of

the flow from the precombustion pressure measurement station to the nozzle. The resultant normalised thrusts are presented in Table 4. In view of the differing equivalence ratios, it is difficult to make a strict quantitative comparison between series A and series B, but it is worth noting that the lower values of normalised thrust in series A are associated with high precombustion temperature or low equivalence ratio. However, the results in Table 4 suggest that there is not a large thrust loss associated with avoidance of combustion induced pressure rise in the combustion chamber, and this may have some practical application. A foreseeable difficulty with operating a fixed geometry scramjet over a range of Mach numbers is that an inlet contraction ratio which is suitable for the higher Mach numbers may not be suitable for the lower Mach numbers because the pressure rise in the combustion chamber then causes the boundary layer to separate, leading to choking of the engine flowpath (e.g. Ref. 9). If the straight wall in Fig. 1(a) were taken as the centreline of a symmetric combustion chamber and nozzle combination, then the pressure distributions of Fig. 9 indicate that nozzle thrust is possible without any pressure rise along the walls of the combustion chamber. Since this can be done without excessive loss of thrust, it suggests a means of extending the operating range of Mach numbers of a fixed geometry scramjet.

6.0 CONCLUSION
Two effects of non-equilibrium chemistry on the thrust nozzle flow of a hydrogen fuelled scramjet have been investigated numerically and experimentally. Nozzle freezing is one of these effects and the continuation of combustion in the nozzle, which is referred to here as kinetic afterburning, is the other. Previous numerical studies have indicated that nozzle freezing does not have a substantial effect on thrust performance, and this is confirmed by the present study for nozzles of moderate area ratio. The numerical investigation was conducted with the hydrogen premixed at the entrance to the combustion chamber whereas the experiments involved hydrogen injection through a central strut and the formation of a combustion wake. In spite of this difference, experimental nozzle pressure distributions were consistent with the results of the numerical model at the higher inlet pressures, and the thrust, normalised by the peak combustion pressure in the combustion chamber, was within 6% of the numerical values, as obtained from finite rate calculations. The numerical finite rate values were within 7% of value obtained from equilibrium calculations. The investigation of kinetic afterburning was preliminary in nature, and was conducted by lowering the precombustion pressures to minimize combustion in the combustion chamber. This reduced the growth of the wake with downstream distance thereby ensuring that the flow entering the nozzle had a wake-like structure. Thus the premixed numerical model could only be used to provide a qualitative description of the flow. Nevertheless it illustrated the development of a flow with kinetic afterburning, revealing the presence of compression waves in the nozzle due to kinetic afterburning, and to combustion near the upstream edge of the nozzle expansion, and the resultant thrust production. These compression waves were evidenced in the experiments by sharp peaks in the surface pressure distributions. The nozzle thrust was only partly due to kinetic afterburning, but the experiments showed that it was possible to generate

584

THE AERONAUTICAL JOURNAL

NOVEMBER 2004

a flow in which the thrust, normalised by the precombustion pressure, was comparable with that obtained with complete combustion in the combustion chamber. Thus, useful levels of thrust were produced without producing a streamwise pressure gradient along the walls of the combustion chamber. This may have application in increasing the Mach number range over which a fixed geometry scramjet can operate.

ACKNOWLEDGEMENTS
The authors would like to express their thanks to Dr Craig Brescianini, and to Dr Peter Jacobs for assistance with the numerical analysis. They also with to thank the Physics Department at the Australian National University for experimental support, and to note their appreciation of the financial support provided by the Australian Research Council, and through NASA grant NAGW-674.

REFERENCES
1. 2. 3. 4. 5. FERRI, A. Review of problems in application of supersonic combustion, Aeronaut J, September 1964, 68, (645), pp 575-597. SWITHENBANK, J. Hypersonic Air-Breathing Propulsion, 1967, Pergamon Press, Oxford, Progress in the Aerospace Sciences, Kchemann, D. (Ed), 8, pp 229-294. FRANCISCUS, L.C. and LEXBERG, E.A. Effects of exhaust nozzle recombination on hypersonic ramjet performance: II Analytical investigation, AIAA J, September 1963, 1, (9), pp 2077-2083. RIZKALLA, O., CHINITZ, W. and ERDOS, J. Calculated chemical and vibrational nonequilibrium effects in hypersonic nozzles, J Propulsion and Power, January-February 1990, 6, (1), pp 50-57. HARRADINE, D.M., LYMAN, J.L., OLDENBORG, R.C., SCHOTT, G.L. and WATANABE, H.H. Hydrogen/air combustion calculations: the chemical basis of efficiency in hypersonic flows, AIAA J, October 1990, 28, (10), pp 1740-1744. SANGIOVANNI, J.J., BARBER, T.J. and SYED, S.A. Role of hydrogen/air chemistry in nozzle performance for a hypersonic propulsion system, J Propulsion and Power, January-February 1993, 9, (1), pp 134-138. RIGGINS, D.W. Thrust losses in hypersonic engines. Part 2: applications, J Propulsion and Power, March-April 1997, 13, (2), pp 288-295. NISHIOKA, M. and LAW, C.K. A numerical study of ignition in the supersonic hydrogen/air laminar mixing layer, Combustion and Flame, 1997, 108, pp 199-219. STALKER, R.J. and PAULL, A. Experiments on cruise propulsion with a hydrogen scramjet, Aeronaut J, January 1998, 102, (1011), pp 37-43. STALKER, R.J. Development of a hypervelocity wind tunnel, Aeronaut J, June 1972, 76, pp 374-384. STALKER, R.J. and CRANE, K.C.A. Driver gas contamination in a highenthalpy reflected shock tunnel, AIAA J, March 1978, 16, (3), pp 277279. BRESCIANINI, C.P. An Investigation of the Wall-injected Scramjet, PhD thesis, University of Queensland, Brisbane, 1993. PATANKAR, S.V. and SPALDING, D.B. Heat and Mass Transfer in Boundary Layers, Second Ed, 1970, International Textbook Co, London. ELGHOBASHI, S. and SPALDING, D.B. Equilibrium chemical reaction of supersonic hydrogen-air jets (the ALMA computer program), 1977, NASA CR-2725. OLDENBERG, R., CHINITZ, W., FRIEDMAN, M., JAFFE, R., JACHIMOWSKI, C., RABINOWITZ, M. and SCHOTT, G. Hypersonic combustion kinetics, 1990, National Aerospace Plane, NASP TM-1107, Wright-Patterson AFB, OH. MCINTYRE, T.J., HOUWING, A.F.P., PALMA, P.G., RABBATH, P.A.B. and FOX, J.S. Optical and Pressure measurements in shock tunnel testing of a model scramjet combustor, J Propulsion and Power, May-June 1997, 13, (3), pp 388-394. PAULL, A., STALKER, R.J. and MEE, D.J.Scramjet thrust measurement in a shock tunnel, Aeronaut J, May 1995, 99, (984), pp 161-163. STALKER, R.J., MORGAN, R.G. and NETTERFIELD, M.P. Wave processes in scramjet thrust generation, Combustion and Flame, 1988, 71, pp 63-77.

6. 7. 8. 9. 10. 11. 12. 13. 14. 15.

16.

17. 18.

Вам также может понравиться