CFD Simulation of Internal Flowfield of Dual-mode Scramjet

Similar documents
Dynamic Pressure Characterization of a Dual-Mode Scramjet

Shock-Boundary Layer Interaction (SBLI) CFD research and applications. IIT Bombay, 15th May 2012

Analysis of a Dual-Mode Scramjet Engine Isolator Operating From Mach 3.5 to Mach 6

ADVANCES in NATURAL and APPLIED SCIENCES

ASYMMETRY EFFECTS IN NUMERICAL SIMULATION OF SUPERSONIC FLOWS WITH UPSTREAM SEPARATED REGIONS

Numerical Simulation and Modeling of Combustion in Scramjets

ANALYSIS OF THE PREDICTION ABILITY OF REACTION MECHANISMS FOR CFD MODELING OF THE COMBUSTION IN HIGH VELOCITY ENGINES

DESIGN AND NUMERICAL ANALYSIS OF ASYMMETRIC NOZZLE OF SHCRAMJET ENGINE

Combustion and Mixing Analysis of a Scramjet Combustor Using CFD Pradeep Halder 1 Edla Franklin 2 Dr. P. Ravinder Reddy 3

Numerical investigation of swirl flow inside a supersonic nozzle

Numerical Simulation of Supersonic Expansion in Conical and Contour Nozzle

Numerical Investigation of Geometrical Influence On Isolator Performance

AEROSPACE ENGINEERING DEPARTMENT. Second Year - Second Term ( ) Fluid Mechanics & Gas Dynamics

Numerical and Experimental Investigations on Mach 2 and 4 Pseudo-Shock Waves in a Square Duct

WALL ROUGHNESS EFFECTS ON SHOCK BOUNDARY LAYER INTERACTION FLOWS

Optimization of Divergent Angle of a Rocket Engine Nozzle Using Computational Fluid Dynamics

SPC 407 Sheet 2 - Solution Compressible Flow - Governing Equations

ONE DIMENSIONAL ANALYSIS PROGRAM FOR SCRAMJET AND RAMJET FLOWPATHS

Shape Optimisation of Axisymmetric Scramjets

APPLICATION OF SPONTANEOUS RAMAN SCATTERING TO THE FLOWFIELD IN A SCRAMJET COMBUSTOR T. Sander and T. Sattelmayer Lehrstuhl für Thermodynamik,

DEVELOPMENT OF A ONE DIMENSIONAL ANALYSIS PROGRAM FOR SCRAMJET AND RAMJET FLOWPATHS

Introduction to Aerodynamics. Dr. Guven Aerospace Engineer (P.hD)

NUMERICAL SIMULATION OF HIGH-SPEED SEPARATION FLOW IN THE AEROSPACE PROPULSION SYSTEMS

Please welcome for any correction or misprint in the entire manuscript and your valuable suggestions kindly mail us

Review of Fundamentals - Fluid Mechanics

CFD ANALYSIS OF CD NOZZLE AND EFFECT OF NOZZLE PRESSURE RATIO ON PRESSURE AND VELOCITY FOR SUDDENLY EXPANDED FLOWS. Kuala Lumpur, Malaysia

Turbulent Compressible Flow in a Slender Tube

A New Method for Computing Performance of Choked Reacting Flows and Ram-to-Scram Transition

CFD MODELLING OF SUPERSONIC COMBUSTION

Design and Optimization of De Lavel Nozzle to Prevent Shock Induced Flow Separation

Investigation of Grid Adaptation to Reduce Computational Efforts for a 2-D Hydrogen-Fueled Dual-Mode Scramjet

Axisymmetric SCRamjet Engine Design and Performance Analysis

A Scramjet Engine Model Including Effects of Precombustion Shocks and Dissociation

J. F. Driscoll. University of Michigan

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Computational Analysis of Bell Nozzles

A Computational Study on the Thrust Performance of a Supersonic Pintle Nozzle

Theoretical Gas Flow through Gaps in Screw-type Machines

Numerical Investigation of Shock wave Turbulent Boundary Layer Interaction over a 2D Compression Ramp

Effect Of Inlet Performance And Starting Mach Number On The Design Of A Scramjet Engine

Chapter 1: Basic Concepts

THERMODYNAMIC ANALYSIS OF COMBUSTION PROCESSES FOR PROPULSION SYSTEMS

compression corner flows with high deflection angle, for example, the method cannot predict the location

CHAPTER 7 NUMERICAL MODELLING OF A SPIRAL HEAT EXCHANGER USING CFD TECHNIQUE

CFD in COMSOL Multiphysics

High Speed Aerodynamics. Copyright 2009 Narayanan Komerath

Carbon Science and Technology

Fluid Mechanics - Course 123 COMPRESSIBLE FLOW

1. Introduction Some Basic Concepts

TURBINE BURNERS: Engine Performance Improvements; Mixing, Ignition, and Flame-Holding in High Acceleration Flows

A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM

EVALUATION OF GEOMETRIC SCALE EFFECTS FOR SCRAMJET ISOLATORS

Shock/boundary layer interactions

Oblique Detonation Wave Engine: Preliminary Design and Performance Prediction

International Journal of Scientific & Engineering Research, Volume 5, Issue 7, July-2014 ISSN

Numerical investigation on the effect of inlet conditions on the oblique shock system in a high-speed wind tunnel

A Magnetohydrodynamic study af a inductive MHD generator

Introduction to Fluid Mechanics. Chapter 13 Compressible Flow. Fox, Pritchard, & McDonald

IX. COMPRESSIBLE FLOW. ρ = P

In the continuing effort to better understand hypersonic flow phenomena, mixing characteristics of various

Lecture1: Characteristics of Hypersonic Atmosphere

Application of COMSOL Multiphysics in Transport Phenomena Educational Processes

NAPC Numerical investigation of axisymmetric underexpanded supersonic jets. Pratikkumar Raje. Bijaylakshmi Saikia. Krishnendu Sinha 1

Computational Analysis of Scramjet Inlet

AME 436. Energy and Propulsion. Lecture 15 Propulsion 5: Hypersonic propulsion

Numerical study of the gas flow in nozzles and supersonic jets for different values of specific heats

Analysis of Shock Motion in STBLI Induced by a Compression Ramp Configuration Using DNS Data

Direct Skin Friction Measurements in High-Speed Flow Environments

Compressible Gas Flow

FEDSM COMPUTATIONAL AEROACOUSTIC ANALYSIS OF OVEREXPANDED SUPERSONIC JET IMPINGEMENT ON A FLAT PLATE WITH/WITHOUT HOLE

Numerical Study of Jet Plume Instability from an Overexpanded Nozzle

A multiscale framework for lubrication analysis of bearings with textured surface

Dynamics Combustion Characteristics in Scramjet Combustors with Transverse Fuel Injection

Fluctuating Pressure Inside/Outside the Flow Separation Region in High Speed Flowfield

Introduction and Basic Concepts

Numerical Investigation of a Transverse Jet in a Supersonic Crossflow Using Large Eddy Simulation

CFD ANALYSIS OF HYPERSONIC NOZZLE THROAT ANALYSIS

A finite-volume algorithm for all speed flows

THEORETICAL AND EXPERIMENTAL INVESTIGATIONS ON CHOKING PHENOMENA OF AXISYMMETRIC CONVERGENT NOZZLE FLOW

NUMERICAL ANALYSIS OF ETHYLENE INJECTION IN

NUMERICAL SIMULATION AND MODELING OF UNSTEADY FLOW AROUND AN AIRFOIL. (AERODYNAMIC FORM)

Flow Structure Investigations in a "Tornado" Combustor

Modelling Nozzle throat as Rocket exhaust

Tutorial: Premixed Flow in a Conical Chamber using the Finite-Rate Chemistry Model

NPC Abstract

Comparison of drag measurements of two axisymmetric scramjet models at Mach 6

Jet Aircraft Propulsion Prof. Bhaskar Roy Prof. A.M. Pradeep Department of Aerospace Engineering

Continued Hybrid LES / RANS Simulation of a Hypersonic Dual-Mode Scramjet Combustor

INFLUENCE OF AIR COOLING AND AIR-JET VORTEX GENERATOR ON FLOW STRUCTURE IN TURBINE PASSAGE

Scramjet Engine Model MASIV: Role of Mixing, Chemistry and Wave Interactions

The ramjet cycle. Chapter Ramjet flow field

Keywords - Gas Turbine, Exhaust Diffuser, Annular Diffuser, CFD, Numerical Simulations.

Aeroacoustics, Launcher Acoustics, Large-Eddy Simulation.

The energy performance of an airflow window

EXPERIMENTAL INVESTIGATION OF MIXING AND IGNITION OF TRANSVERSE JETS IN SUPERSONIC CROSSFLOWS

Final 1. (25) 2. (10) 3. (10) 4. (10) 5. (10) 6. (10) TOTAL = HW = % MIDTERM = % FINAL = % COURSE GRADE =

Civil aeroengines for subsonic cruise have convergent nozzles (page 83):

Model for the Performance of Airbreathing Pulse-Detonation Engines

COMPUTATIONAL ANALYSIS OF CD NOZZLE FOR SOLID PROPELLANT ROCKET

Jet Aircraft Propulsion Prof. Bhaskar Roy Prof A M Pradeep Department of Aerospace Engineering Indian Institute of Technology, Bombay

Transcription:

CFD Simulation of Internal Flowfield of Dual-mode Scramjet C. Butcher, K. Yu Department of Aerospace Engineering, University of Maryland, College Park, MD, USA Abstract: The internal flowfield of a hypersonic airbreathing engine, as that which is found in a dual-mode scramjet engine, is very complex. The incredibly high temperatures seen in supersonic and hypersonic flows, transverse fuel injection into a supersonic cross-flow and sustaining supersonic combustion are only a couple of the reasons for such complexity. Heat addition downstream of the combustion zone causes a shock train to form upstream of the fuel injection location in order to match the pressure rise associated with said heat release. Properly modeling these phenomena is non-trivial. Several approaches taken by the authors are presented below. The results were found to have good agreement when compared to experimental data that was gathered from a model scramjet in the Hypersonic Center at the University of Maryland, College Park [1]. Keywords: CFD, scramjet, shock train, heat release, transverse fuel injection 1. Introduction The purpose of this research was to simulate the internal flowfield of a dual-mode scramjet combustor, including shock trains and thermal choking characteristics. The authors also aimed to quantitatively deduce the spatial distribution of heat release characteristics obtained in the experiments by calibrating the wall pressure profiles and boundary conditions. This paper presents the methods that were taken by the authors. For simplicity and to reduce computational cost, the simulations were split into two separate parts: the isolator and combustor. This allowed each part to be more accurately modeled with greatly reduced computational cost than to model the entire geometry at once. Each model was subjected to a number of different conditions and will be presented below. At each step, the results were compared to experimental data that were gathered from the model scramjet. This was done to validate the simulations and give them a basis for comparison. The results showed good agreement in most cases, even with several simplifications and assumptions that were taken by the authors. The main goal is to develop an approximate technique to extract heat release information from experimentally obtained pressure profile data when optical access or heat release data is unavailable. This will help gain additional insight into the performance and characteristics of the model scramjet. 2. Basic Theory and COMSOL Setup For all of the simulations that were performed, the fully compressible Navier-Stokes equations govern the fluid flow. These are found in [2] and [3]. The Navier-Stokes equations are a set of 3 individual equations derived using the conservation of mass, momentum, and energy. The conservation of mass, or continuity equation, is as follows: ρ t + (ρu) = (1) where bold quantities represent vectors in all equations. Equation 1 states that mass can neither be created nor destroyed: it must remain constant for a given control volume [3]. The second Navier-Stokes equation is derived from the conservation of momentum. ρ u t + ρ(u )u = [ pi + τ] + F (2) Equation 2 states that the time rate of change of momentum must be equivalent to the force exerted on the body [3]. The body force terms are on the right side of the equation. The third and final equation is derived using the conservation of energy. ρc p ( T + (u )T) t = ( q) + τ: S T ρ ρ T ( p p t + (u )p) + Q (3) Where τ is the viscous stress tensor, q is the heat flux vector, and Q is the heat source over a volumetric domain [2]. Equation 3 states that

energy must remain constant. It can change form, but it cannot be created nor destroyed [3]. All of the simulations presented herein were done using COMSOL 4.4 and the turbulent High Mach Number Flow branch of the CFD module. The first step in simulating the model scramjet in COMSOL was to generate the geometry. Due to its relatively simple design, a series of basic shapes were used to model the internal flowpath in 2D space. This was done using a rectangle for the isolator and a Bézier polygon consisting of line segments for the combustor and nozzle. The correct material properties were applied for the fluid and the inlet/outlet boundaries were added. For the isolator, the inlet conditions were set as shown in Table 1. The Reynolds number is calculated based on the isolator inlet height and the viscosity can be found in [4]. Table 1. Scramjet Inlet Conditions [1] Mach Number 2.419 Static Temperature 699.7 [K] Static Pressure 8334.2 [Pa] Total Pressure 683928.9 [Pa] Reynolds Number 154,119 Fuel Mass Flow Rate.377 [g/s] The isolator outlet was set to a hybrid flow condition with the static pressure equal to the approximate values seen at the zero location in the model scramjet. More information can be found in section 3 below. For both the isolator and combustor simulations, a series of three stationary studies were run each time in order to accurately model a no slip wall condition. The reasoning for this will be presented in section 3. In each study, the dynamic viscosity was decreased through a set range of values each time, artificially increasing the Reynolds number through the use of an auxiliary sweep. This method was adapted from the Transonic Flow in a Nozzle tutorial from COMSOL [5]. The first two stationary studies used COMSOL s k ε turbulence model and utilized a slip wall boundary condition and wall functions. The third study, a stationary with initialization study, used the Spalart-Allmaras turbulence model and the no slip wall condition. The second and third studies had the values of the dependent variables initialized using the solution from the previous study and were run through the auxiliary sweep. 3. Isolator Simulation The purpose of an isolator in a dual-mode scramjet engine is to facilitate a pressure rise caused from the combustion downstream without allowing the engine to unstart. If that were to happen, the flow inside the engine would become purely subsonic, greatly reducing performance or even preventing the engine from being able to operate altogether [7]. The pressure rise is facilitated by the subsonic boundary layer that exists in a flow with a no slip wall. That is where the difficulty of properly modeling the model scramjet arises. The solution procedure can be found in section 2 above. The no slip wall causes boundary layer separation and the formation of a shock train. Shock trains can have an oblique or a normal shock structure as shown in Fig. 1a. An oblique shock train produces a smooth, fairly uniform pressure rise as seen in Fig. 2. However, a normal shock train produces a pressure rise that resembles a stair step fashion and causes a greater loss of stagnation pressure a key airbreathing engine performance measure. The pressure rise from a normal shock train can be seen in Fig. 3. It is the difference in shock train structures that explains the difference between the pressure rises in Figs. 2 and 3. As will be seen later in Fig. 7b, the model scramjet had an oblique shock train but (a) Experimental schlieren photograph visualizing vertical density gradient [6] (b) COMSOL simulation showing vertical density gradient in isolator of dual-mode scramjet engine. Figure 1. Experimental (a) and simulated (b) shock trains in a constant area duct. Black and white indicate a negative and positive density gradient, respectively. Both images show a normal shock train structure.

P/P o.5.45.4.35.3.25.2.5-16 -14-12 -1-8 -6-4 -2 Figure 2. Normalized experimental static pressure measurements along top wall of isolator for all combustor equivalence ratios. Adapted from [1]. the simulations generated a normal shock structure as shown in Figs. 1b and 4. The isolator is a very simple.5 square by 8.5 long duct. There are static pressure ports along the top wall spaced at.5 increments. The time averaged data from these measurements was shown in Fig. 2. As expected, the higher the combustor equivalence ratio, Φ, the higher the pressure at the exit of the isolator/entrance to the combustor. As stated before, this is because the pressure increase due to the exothermic reaction forces the flow upstream to adjust. Simulations were run for all combustor equivalence ratios from the experiment. This not only allowed for the generation of the pressure data in Fig. 3, but also allowed for comparison of the shock train structures. This can be seen in Fig. 4. It is P/P o.5.45.4.35.3.25.2.5 =. =.4 =.8 = 2 = 6 =.2 =. =.4 =.8 = 2 = 6 =.2-16 -14-12 -1-8 -6-4 -2 Figure 3. Simulated static pressure measurements along top wall of isolator for all combustor equivalence ratios. expected that for the cold flow case, Φ =., there would be no shock structure because there is no pressure rise to account for. This is seen in Fig. 4a. As the back pressure is increased due to the higher combustor equivalence ratio, a shock train forms and propagates upstream. The increasing number of shocks allows for the higher static pressure rise that is needed. 4. Combustor Simulation Due to the boundary layer and shock train, all of the exit conditions from the isolator have a profile in the vertical direction. Because the geometry was split into two distinct sections, outlet values from the isolator were used as inlet values for the combustor. This caused an issue because some average or constant value had to be chosen for the Mach number, temperature, and pressure rather than the full profile. In order to find the best combination of combustor inlet parameters, a series of cold flow simulations with no heat release or fuel injection were run varying these values within the ranges seen at the exit of the isolator. This is seen in Fig. 5. In order to select which set of data to use for the input conditions, the area downstream of the combustion zone became the focus. The simulated data needed to closely match the experimental measurements in this area from x/h of 6 onward. Two of the simulated cases became the focus: both of the cases with an inlet pressure of 1.8 atm. The case with the higher temperature and Mach number matches the experimental data most closely, but it is known that adding heat will raise the pressure by a certain amount. If that particular set of values were to have been used, then the downstream pressure would have been too high after the fuel injection and heat addition were added into the model. Therefore, the authors selected the case with 1.8 atm, 95 K, and a Mach number of 1.32 to use for the inlet conditions for all future combustor simulations. The equivalence ratio of 2 was chosen in particular so that the simulations can help gain more insight into the physics of the model scramjet. Due to the difficulties of visualizing an enclosed supersonic flow over such a large length, it is unknown whether or not the flow transitions back to supersonic after it passes through the combustion zone. In order to determine what happens to the flow, there are several parameters that can be explored: determining the actual

(a) Φ =. (b) Φ =.4 (c) Φ =.8 (d) Φ = 2 (e) Φ = 6 (f) Φ =.2 Figure 4. Simulated density gradients for all fuel flow conditions. The data in Fig. 3 was taken from these simulations. amount of heat that was given off during the experiments and the exact location of that release. The first step was to estimate the amount of heat release that would most likely be seen in the model scramjet. To do this, a heat release value was calculated based on the heat of reaction of the hydrogen fuel 119,954 kj/kg fuel [7]. Due to the effects of dissociation seen in high temperature gas dynamics and material properties limitations, an upper limit of 3, K was imposed. This limits the fuel to about one third of the maximum heat that could be released in an ideal stoichiometric reaction. That translates to a total power for the heat release of about 95 kw per unit width over the heat release domain. Based on P/P o.5.45.4.35.3.25.2.5 P=2[atm],T=1[K],M=1.5 P=2[atm],T=95[K],M=1.32141 P=1.9[atm],T=1[K],M=1.5 P=1.9[atm],T=95[K],M=1.32141 P=1.8[atm],T=1[K],M=1.5 P=1.8[atm],T=95[K],M=1.32141 Experimental, =2 2 4 6 8 1 12 14 Figure 5. Normalized simulated static pressure measurements along top wall of combustor as a function of combustor inlet conditions. Simulations were for cold flow with no fuel injection. Experimental data from [1]. these assumptions and properties, a range of simulations were run at or below this value. The results are shown in Fig. 6. As predicted before, the larger the amount of heat that is released, the higher the downstream pressure values. However, the simulation with 1/6 of the lower heating value of hydrogen fuel shows very good agreement downstream. In addition, it oscillates about the experimental values in the area above the cavity. The discrepancy is to be expected because of several issues that are discussed in section 5 below. 5. Issues It is difficult to compare these simulations number for number with the model scramjet. The geometry cross section is the same, but the simulations were completed in 2D a far cry from the 3D experiments. While COMSOL was able to generate the shock trains in the isolator, they are an inherently three dimensional, asymmetric structure in real life. The boundary layer would be on all four walls of the scramjet, whereas the simulations only have a boundary layer on the top and bottom surfaces. Another issue is that the model scramjet would have had a large amount of heat transfer into the stainless steel walls, especially during the first few seconds of each run while they warm up from room temperature. Last but not least, the experimental measurements are time averaged over runs of up to ten seconds. The COMSOL simulations are stationary solutions at one instant in time. This explains why the curves in Fig. 6 oscillate about the experimental value.

P/P o.5.45.4.35.3.25.2.5 1/3 LHV H 2 1/4 LHV H 2 1/6 LHV H 2 1/12 LHV H 2 Experimental, =2 order to best match the experimental measurements downstream. This was due to difficulties in implementing the full exit profiles. Once a reasonable set of conditions were found, the fuel injection and heat addition were added into the model. A range of heat release values were applied over the same domain area in order to find an estimated amount of heat that was released in the model scramjet. It was found that the authors assumptions were correct in that the heat release was found to be less than the upper limit that was proposed. 2 4 6 8 1 12 14 Figure 6. Normalized simulated static pressure along top wall of combustor as a function of combustor equivalence ratio. Experimental data from [1]. As the flame fluctuates in the cavity, the pressure will fluctuate as well. Time averaging the data evens out those oscillations and make it appear as a nice smooth curve. In order to compare results with the model scramjet, they must oscillate about one another and follow the same general trends. 6. Conclusions The model scramjet was simulated in two distinct sections: the isolator and combustor. This aided with convergence, greatly reduced time and computational cost, and was able to better generate flow features as seen in the model scramjet. All simulations were run in 2D with the turbulent High Mach Number Flow branch of the CFD module in COMSOL Multiphysics 4.4. Using a set of three studies, each simulation was run through a slip, wall functions, and no slip wall boundary condition. This aided in convergence and ensured that all simulations were subjected to the same process. The no slip wall condition allowed for the creation of a shock train in the isolator. Each combustor equivalence ratio generates a different pressure rise, all of which were able to be matched through the simulations. There were some discrepancies however, as COMSOL generated normal shock train structures, whereas an oblique structure was seen in the experiments [1]. The isolator exit conditions were applied to the combustor inlet and the simulations were run through several sets of boundary conditions in 7. Future Work Now that a basic scramjet model has been set up, the heat release can be tweaked to more closely match the experimental data. In order to do this, the heat release domain can be altered in shape or size, as well as intensity in order to gain additional insight into the combustion zone of the model scramjet. Another method will be to use a series of several smaller domains with differing heat release values rather than a single domain at one constant intensity. Once this is completed, focus can shift to modeling the model scramjet with the added fin upstream of the fuel injection location. The fin allows for better fuel penetration into the cross-flow and allows the combustion to take place further downstream. It is thought that this prevents the dual-mode scramjet from becoming thermally choked as is possible in the baseline case. This can be seen in Fig. 7. The change in shock structures can also be seen. In Fig. 7b there is an oblique shock train in the isolator and beginning of the combustor, whereas the fin case (Fig. 7c) has a much more simple shock structure originating mostly from the fin and fuel injection. It is these types of structures that hope to be modeled in the future. 8. References [1] Muñoz, Camilo A., Effect of Fin-Guided Fuel Injection on Supersonic Mixing and Combustion, PhD Dissertation, University of Maryland, College Park (214). [2] COMSOL, CFD Module User s Guide, CFD Module Documentation (213). [3] Anderson, John D., Modern Compressible Flow with Historical Perspective, pp. 239-26. McGraw Hill, New York (23).

(a) COMSOL simulation for baseline reacting case with Φ = 2 showing vertical density gradient in isolator (left) and combustor (right). The gap represents that the simulations were run separately and is equivalent to location in (b) and (c) below. The vertical lines in the right portion of the image represent the left and rightmost bundaries of the heat release domain. Far upstream and downstream have been cut off in this illustration. (b) Experimental schlieren image for baseline case and Φ = 2. The combustion zone can be seen in the cavity between locations 1 and 4 [1]. (c) Experimental schlieren image for fin case and Φ = 2. Due to the addition of the 3-D fin upstream of the cavity, most of the combustion takes place in the diverging area downstream of the combustor. [1]. Figure 7. Simulated and experimental schlieren images for baseline (a,b) and fin case (c). In (b) and (c), the F and IG represent the fuel and igniter locations, respectively. [4] Kadoya, K., Matsunaga, N., and Nagashima, A, Viscosity and Thermal Conductivity of Dry Air in the Gaseous Phase, J. Phys. Chem. Ref. Data, Vol. 14, No. 4, pp. 947-97 (1985). [5] COMSOL, Transonic Flow in a Nozzle, COMSOL Model Library Tutorial (212). [6] Geerts, Jonathan, PhD Candidate at University of Maryland, College Park (214). [7] Heiser, William H. and Pratt, David T., Hypersonic Airbreathing Propulsion, pp. 113,251-257. American Institute of Aeronautics and Astronautics, Washington, DC (1994). 9. Acknowledgments The authors gratefully acknowledge the support of the U.S. Air Force Office of Scientific Research with Dr. Michael Kendra as Program Manager, through the University of Maryland Center for Revitalization of the Hypersonic Testing and Evaluation Workforce program.