Vous êtes sur la page 1sur 19

AIAA Guidance, Navigation, and Control Conference and Exhibit AIAA 2005-6256

15 - 18 August 2005, San Francisco, California

Modeling for Control of a Generic Airbreathing Hypersonic


Vehicle

Maj Mirmirani* and Chivey Wu*


California State University, Los Angeles
5151 State University Dr.
Los Angeles, CA 90032, United States

Andrew Clark and Sangbum Choi


Multidisciplinary Flight Dynamics & Control Lab.
California State University, Los Angeles
5151 State University Dr.
Los Angeles, CA 90032, United States

and

Richard Colgren
University of Kansas
1530 West 15th St.
Lawrence, KS 66045, United States

The unique airframe-engine configuration of airbreathing hypersonic flight vehicles


(AHFV) pose a significant challenge for design of controllers for these vehicles. The
Airframe-engine configuration, the wide range of speed and the extreme flight conditions
result in significant coupling among various dynamics and modeling uncertainties. There is
almost a complete absence of models that adequately include and quantify the unique
attributes for this class of vehicles. This paper describes a high-fidelity CFD-based model of
a full scale generic airbreathing hypersonic flight vehicle under development at the
Multidisciplinary Flight Dynamics and Control Laboratory (MFDCLab,
www.calstatela.edu/centers/mfdclab) at California State University, Los Angeles (CSULA).
The vehicle (CSULA-GHV), which has an integrated airframe-propulsion system
configuration, resembles an actual test vehicle. The vehicle is specifically designed to study
the challenges associated with modeling and control of airbreathing hypersonic vehicles and
to investigate and quantify the couplings between the aerodynamics, the propulsion system,
the structural dynamics, and the control system. The configuration of the vehicle and its
dimensions are developed based on 2-D compressible flow theory, and a set of mission
requirements broadly accepted for a hypersonic cruise vehicle intended for both space
access and military applications. Analytical aerodynamic calculations are conducted
assuming a cruising condition of Mach 10 at an altitude of 30 km. The 2-D oblique shock
theory is used to predict the shock wave angles, the pressure on the frontal surface, and the
Mach number at the engine inlet. The scramjet engine is simply modeled by a 1-D
compressible flow with heating. The exit flow is modeled using 2-D expansion wave theory to
predict the pressure on the rear surface. The unique aspect of this study is the use of coupled
simulations using multi-physic software in conjunction with theory enabling quantification
of the couplings which are broadly ignored in models used for control system design.
Simulation results developed to date are presented.

*
Professors, Mechanical Engineering Department, AIAA Member

Research Associates, MFDCLab, AIAA Member

Professor, Aerospace Engineering Department, AIAA Member

1
American Institute of Aeronautics and Astronautics

Copyright 2005 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
I. Introduction

THE dynamics of airbreathing Hypersonic Flight Vehicles (AHFVs), characterized by the tight integration of
airframe and the propulsion system make the modeling and control of these vehicles very challenging. The
couplings between the airframe, structure, and propulsion system, the thermal effects of hypersonic speeds, and the
wide range of speeds at which these aircraft fly pose significant control challenges. The longitudinal dynamics of
these aircraft is unstable in pitch mode and non-minimum phase. The aeroelastic effects for the full-scale vehicle
become significant effects due to the specific structural design requirements and the high aerothermodynamic
loading. Modeling and control techniques used for conventional aircraft are inadequate for AHFVs. Models that
incorporate the interactions and the salient features of these vehicles and takes account of the integrated airframe-
engine-control system are needed [4],[6]. The simplifying assumptions made in almost all existing literature must be
relaxed to obtain the accuracy required for flight test environment applications. To obtain such high-fidelity models
for this class of vehicles, for which little test data are available, will require full utilization of computational tools
combined with the underlying physics, or a hybrid numerical-analytic approach. An effort is underway at the
Multidisciplinary Flight Dynamics and Control Laboratory at California State University, Los Angeles in
collaboration with the University of Kansas and the University of Southern California to develop a high fidelity
simulation model for a full-scale generic airbreathing hypersonic vehicle, (CSULA-GHV), one resembling an actual
test vehicle such as NASAs X-43 and DARPAs FALCON. Such a unique high-fidelity model built around a set of
requirements identified for an actual future vehicle will have tremendous value for the hypersonic research
community in many ways.
This paper describes the preliminary design and modeling of a 2-D model of a generic vehicle, which incorporates
various dynamics of the vehicle and their interactions. The model is developed to investigate and quantify the
couplings between aerodynamics, propulsion, structure, and the control system. The configuration and dimensions
are developed based on 2-D compressible flow theory, and a set of mission requirements broadly accepted for a
hypersonic cruise vehicle intended for both space access and military applications.
To put our approach in context, the paper starts with a brief description of various modeling, and control
challenges of hypersonic flight and the state of knowledge in this field. It then proceeds to describe the design of a
2-D full-scale AHFV model based on oblique shock and expansion wave theories. Following the theoretical results,
the CFD Simulation data obtained to date at MFDCLab are presented. An extensive list of references is included in
the bibliography section.

II. History of Hypersonic Flight Research in the United States


Hypersonic airbreathing propulsion has been studied by NASA for more than 60 years, since the evolvement of the
hydrocarbon-fueled conventional ramjet (CRJ) engine concept [9],[10]. In the late 1940s, the feasibility of developing
a scramjet engine attracted the attention of the propulsion community. In the early 60s, the scramjets technical
hurdles of fuel injection and mixing, wall cooling and frictional losses, and nozzle performance, were outlined.
Meanwhile, comparisons of the performances of CRJ and scramjet engines determined that the scramjet engine
would outperform the CRJ somewhere in the speed range of Mach 6-8, and would be superior at higher speeds
[10],[11],[12]
. In 1964, NASA began the Hypersonic Research Engine (HRE) program, which aimed to flight test a
complete, regeneratively cooled, flight-weight scramjet research engine on the X-15 rocket-powered research plane
[13]-[16]
. In April 1965, the U.S. Air Force funded the Scramjet Incremental Flight Test Vehicle (IFTV) program [10].
The purpose of the program was to demonstrate vehicle acceleration from a boosted speed of 5400 ft/s to at least
6000 ft/s using four hydrogen-fueled scramjet modules located around the central vehicle body. In the 1970s, NASA
began focus on the rectangular airframe-integrated engine configuration [10],[13],[15],[17]-[20]. This concept utilizes the
inlet sidewalls to produce extra horizontal compression in addition to the vertical forebody compression (Figure 1),
and uses in-stream struts as housings for distributed fuel injectors. In 1986 the U.S. Air Force and NASA initiated
the National Aerospace Plane (NASP) program, a major hypersonic flight research program including flight tests.
The focus of the NASP program was to build an airbreathing single-stage-to-orbit (SSTO) experimental aircraft, the
X-30, which would be used for hypersonic flight testing and demonstration. NASP incorporated extensive
development of rectangular airframe-integrated scramjet technology, including a large number of newly developed
modular experimental engines which were tested at NASA Langley Research Center (LRC) in the Mach 4-7 regime
[10],[21],[22]
. The NASP was a full-scale operational prototype vehicle system development program rather than an
incremental technology program. In spite of the enormous amount of effort and achievements in many aspects, the
NASP program was terminated in January 1995 due to lack of funding, without having conducted flight tests.

2
American Institute of Aeronautics and Astronautics
In 1996, NASA started the Hyper-X program [9],[10],[30]-[33] as an initial stage of its current hypersonic plans. The
current plans focus on the development and flight testing of small-scale (X-43A, X-43B, X-43C, X-43D) and one
full-scale demonstrator vehicles [32]. X-43A is a 12-foot-long hydrogen-powered experimental vehicle with a five-
foot wingspan (Figure 2), smart scaled from a 200-foot operational concept. It has been used in three scramjet-
powered and un-powered flight tests at Mach 7 and Mach 10. Although the first trial in June 2001 was unsuccessful
due to a booster failure, the next two flight tests were successfully conducted at Mach 7 in March 2004 and at Mach
10 in November 2004.

Figure 1. The Flow Features on an AHFV Fore Body

Among the other three demonstrator concepts, which are yet to be flight-tested, X-43B is a 35'-45' reusable,
combined cycle demonstrator vehicle with combined turbojet and dual-mode scramjet power. X-43C is a 16-foot-
long hydrocarbon-powered vehicle utilizing a three-module engine which will accelerate the vehicle from Mach 5 to
Mach 7 during the flight test. X-43D is a concept for flight testing hydrogen-fueled scramjet engines at velocities of
Mach 15 or greater. Finally a large-scale reusable demonstrator (LSRD) vehicle, whose architecture is the same as
that of the operational vehicle and whose size is large enough to operate overall airbreathing propulsion speeds, is
planned to be built and flight tested.

Figure 2. X-43A (NASA LRC Archive)

The U.S Air Force is currently conducting another scramjet development program, the Air Force Hydrocarbon
Scramjet Engine Technology (HyTech) program [10],[32],[44]. Although HyTech is currently missile-oriented, it is
expected to be incorporated with hypersonic transportation and, in particular, in the X-43B and X-43C projects of the
Hyper-X program.

III. Control Challenges of Hypersonic Flight


The NASP and the Hyper-X programs identified key technological obstacles to the feasibility of hypersonic
transportation. One primary challenge is flight controller design. Designing reliable and effective AHFV controllers
requires careful consideration of these vehicles unique dynamic characteristics, which differ in major ways from
those of typical aircraft. In hypersonic aircraft configurations with tightly integrated airframe scramjet engine
[4],[5],[45]
, the primary lift generating surface is the body itself. Besides the aerothermodynamic effects of hypersonic
speeds, the strong interactions between the aerodynamics, the propulsion system, and the elastic airframe make the
explicit characterization of flight dynamics of the AHFVs highly challenging [4],[45],[46]. For example, the optimum
operating condition will be based on the shockwave formed on the leading edge to be captured by the inlet lip
(Figure 1). However, aeroelastic effects and/or aircraft structure vibration will result in mass-flow spillage, because

3
American Institute of Aeronautics and Astronautics
the bow shock angle changes as the structure deforms, resulting in the engine operating in off-design conditions.
Therefore, these vehicles will require a variable geometry inlet to operate at or near optimum aerodynamic and
propulsion conditions at all times. As most high-performance aircraft AHFVs are non-minimum phase and unstable
in pitch. In addition, the dynamic characteristics of the hypersonic vehicle vary more significantly over the flight
envelope than other aircraft due to the extreme range of operating conditions and the rapid change of mass
distribution. Moreover, many aerodynamic and propulsion characteristics still remain uncertain and are hard to
predict due to the almost complete lack of flight test data and the inadequacy of ground test facilities. The following
sections list the major issues characterizing AHFV flight dynamics and control challenges in more detail.

IV. Effects of Hypersonic Speeds


At hypersonic speeds, the aerothermodynamic properties of the air deviate from the ideal gas behavior with more
significant impact at Mach 6 and faster [4], [46]-[48]. In particular, the temperatures behind the normal shock waves, the
loading on certain aircraft surfaces, and the pitching moment coefficient increase with the Mach number [4],[49],[50].
On the other hand, the high temperature gas effects of hypersonic flow regimes thicken the boundary layer around
the aerofoil, changing the effective aerodynamic surface significantly. This change as well as the flow separation
and reattachment phenomena lead to viscous behavior in terms of pressure distribution, shock waves, and drag,
differing significantly from the results of the inviscid characteristics of the aerofoil. High temperatures on the
control surfaces place constraints on control surface deflection limits. The variations due to wide speed ranges of a
complete flight and the sensitivity of the AHF dynamics to the flight conditions, requires a highly integrated
guidance scheme flight control system to provide a robust, stable high-performance flight. On the other hand,
hypersonic speeds cause the so-called path-attitude decoupling phenomenon, which is the resistance of the high
momentum of the AHFV to the changes in the desired flight path [51],[52],[53]. Hence the actual flight path significantly
lags the changes in the pitch attitude at hypersonic speeds.

V. Variations due to Speed Range


The dynamic characteristics, stability, and performance of AHFVs vary over the flight envelope more than other
aircraft due to their wide range of operating conditions and mass distributions. If the low-speed static stability
margins are desired to be kept in the conventional range, one must accept unstable configurations at hypersonic
speeds in pitch mode [4],[54]. The tightly integrated airframe propulsion system configuration needed for efficient
AHF also affects landing/take-off performance and increases transonic drag. Ground testing at NASA LRC that
examined subsonic AHFV behavior close to the ground plane demonstrated that turning the power off causes a sharp
increase in lift while turning the power on causes negative lift effect together with a large increase in the pitching
moment. These effects result from the diverging angle between the nozzle and the ground plane [55]. The external
nozzle configuration for good hypersonic performance causes flow separation at lower speeds and a steep increase
in drag in transonic regime.

VI. Scramjet Engine Dynamics


The Scramjet engines operate by supersonic combustion of fuel in an air stream compressed by the aircraft's
forward speed. Using hydrogen (or hydrocarbon) as the basic fuel for combustion, airbreathing scramjet engines burn
oxygen scooped from the atmosphere. Since the hydrogen and the scooped oxygen have less time to mix and react in
supersonic combustion, the combustor needs to be longer. Moreover, in order to produce sufficient thrust for
hypersonic flight, the engine inlet must capture as much of the airflow under the AHFV surface as it is possible. This is
done through the integration of the engine with the airframe so that the inlet area is contiguous with the vehicle
undersurface [4],[10],[13],[18]. The inlet sidewalls produce extra horizontal compression in addition to the vertical fore-body
compression.

VII. Airframe - Propulsion - Structural Dynamic Interactions


In AHFVs with airframe-integrated scramjet engines, the aerodynamics, propulsion system, and structural
dynamics are highly interactive; hence the control bandwidths for flight and propulsion systems cannot be separated
clearly. The engine-airframe integration causes significant interaction between the propulsion system and vehicle
aerodynamics [4],[23],[45],[55]. Considering the longitudinal dynamics, the inlet flow pressure acting on the fore-body
generates a nose-up pitching moment while the external nozzle flow generates a nose-down pitching moment. These
flows may also affect the lateral dynamics if they are laterally non-uniform.

4
American Institute of Aeronautics and Astronautics
Similarly, the aerodynamics affects the propulsion system in several ways. The capture and compression of the
flow through the inlet is determined by the properties of the bow shock wave under the vehicle forebody, which are
determined by the angle of attack (AOA) and the dynamic pressure as well as the free stream characteristics [4],[56].
The AOA and the dynamic pressure also affect the combustion kinetics and the exhaust flow/free stream shear layer.
The sensitivity of the performance to the AOA and the dynamic pressure is small at high Mach numbers and
increases as the speed decreases. At low speeds, the AOA and the dynamic pressure need to be greater than certain
values in order to provide acceptable performance. The frequency domain numerical analysis in Reference [4] shows
that the thrust and the pressure at the engine inlet are significantly affected by pitching control surface, e.g., elevon
defections over a wide frequency range. The analysis indicates a similar effect of the fuel flow rate and diffuser area
ratio changes on the AHFV pitch rate.
Another major source of dynamic coupling in AHFVs is the aeroelastic modes. Bending of the fore-body and aft-
body together with propagations throughout the entire airframe affect the flows through the inlet and the exhaust and
hence the aerodynamic performance. Elastic-rigid body interactions are also significant in AHFVs, since the low
structural vibration frequencies are close to those of the rigid body as a result of the requirement for very low
structural weight [4],[57]. Accurate determination of the structural elastic modes is critical for flight control, especially
for precise control of the AOA. However, the non-uniform aerodynamic heating in AHF, unconventional composite
materials used in building the airframe, and the shell-type structure of AHFVs cause significant variations and
uncertainties in the shapes and natural frequencies of the elastic modes [5],[58]-[63].

VIII. Databases and Uncertainties


Despite the unprecedented capabilities of the powerful CFD codes available today, the development of
conventional aircraft and the design of their control systems depend on the use of empirical estimates of
aerodynamic and engine data obtained from sources such as U.S. Air Force DATCOM, wind-tunnel data, and flight
test data. Design and development of hypersonic aircraft, however, must proceed without the benefit of a vast
statistical database of actual flight test data. The insufficiency of the flight test data and the inadequacy of the
ground test facilities lead to aerodynamic and propulsion uncertainties.
Unpredictable aerodynamic and thermodynamic behaviors due to hypersonic speed and aerodynamics-
propulsion-structural dynamics interactions constitute another uncertainty source [4],[7],[57]. The propulsion system
perturbations that indirectly impact the longitudinal dynamics, the elastic mode variations, the non-uniform pressure
distribution, and the uncertainties explained above comprise some of these unpredictable behaviors. The mass
property variations can also be considered to be a modeling uncertainty. In general, many of the critical aerodynamic
characteristics are hard to predict while the ability to experimentally determine them is limited. Hence, conventional
control methodologies that depend on relatively accurate models and a reliable aerodynamic database are not
suitable. Robust control design and compensation algorithms involving intelligent and adaptive flight control are
necessary.

IX. Modeling Airbreathing Hypersonic Flight Dynamics


Despite The traditional development of equations of motion for flight vehicles is based on the Newtonian approach
that excludes the elastic degrees of freedom [64],[65]. At the other end of the spectrum, the purely structural dynamics
models for aeroelastic analysis exclude rigid-body modes [66],[67]. A common approach used in modeling for control
design assumes that the aircraft dynamics can be decomposed into two distinct parts with sufficiently separated
bandwidths: dynamics of the measured states, and dynamics of the unmeasured states. The dynamics with the
lower bandwidth includes the states that can be measured during the flight, such as vehicles velocity, altitude, angle of
attack, and pitch rate. The effects of structural dynamics, propulsion, and aeroelasticity, the unmeasured states, and
their contributions to the overall dynamics of the vehicle are in general hard to measure and model, and therefore have
been combined and treated as uncertainties in the control design formulation. In the case of AHFVs for example, with
few exceptions, the effect of propulsion and its interaction with aerodynamics of the vehicle has not been included in
the development of models and/or control algorithms.
The majority of the AHF dynamic models in the literature consider only the longitudinal motion. A widely used
longitudinal model in the literature is the rigid-body model of Reference [68], [69], for the winged-cone accelerator
configuration. The winged-cone configuration is of course significantly different from the rectangular airframe-
integrated engine configuration of AHFVs and the model is derived neglecting the coupling effects among
aerodynamics, propulsion, and structural dynamics.
The effects of couplings and the impact of the propulsive perturbations on the pitching moment are formulated in
Reference [57] as additive uncertainties and incorporated into a linear model of the vehicle. The coupling between

5
American Institute of Aeronautics and Astronautics
elastic and rigid structure modes is included in the frequency domain by defining a cover function, carrying information
about the flexible modes.
A more and realistic mathematical model for AHFV longitudinal dynamics can be found in Reference [5], [26]
where an analytical aero-propulsive/aeroelastic hypersonic-vehicle model derived for the rectangular AHFV
configuration with airframe-integrated scramjet. This model focuses on dynamic couplings and control system
integration.
An extensive computational model for a Generic Hypersonic Aerodynamic Model Example (GHAME) is presented
in Reference [45], [78]. GHAME is composed of five data sets: two aerodynamic models (one based on empirical data
sources and the other on analytical programs); two aero-thermodynamic models (a simple convection-radiation heat
flux and equilibrium model and a model obtained through analytical programs); and a simplified switching turbojet-
ramjet-scramjet engine model. The aerodynamic models are based on a rigid body configuration with geometry similar
to that of the conic accelerator [68],[69].

X. Airbreathing Hypersonic Flight Control


The available literature in the AHF control focuses on the longitudinal dynamics ignoring its lateral-directional
counterpart. The control design problem becomes significantly more complex when lateral-directional characteristics
are included.
Since robust AHF control addresses the modeling uncertainties and the trade-off between robustness and
performance, they are the most suitable approach. Robust AHF control approaches include linear model-based and
nonlinear model-based. In the linear model-based approach, the classical robust control [76],[77] techniques such as H
and -synthesis are applied to the AHFV models with structured and unstructured uncertainties. The resultant
controllers have the common property of being highly conservative with undesirable compromise in performance.
Included among the linear model-based approaches are the robust control design for the rigid conic accelerator
configuration in Reference [69] and [57]. The focus of the approach in Reference [81] is decoupling the phugoid and
short-period modes. The simulation results demonstrate the effectiveness of the H-based eigenstructure assignment
design compared to the traditional Shapiro design.
Nonlinear models are more suitable for representing the coupling effects. As an alternative to the classical
robustness theory developed for linear systems, the concept of stochastic robustness has been introduced in Reference
[73]-[75] for nonlinear systems. These techniques are applied to the longitudinal AHF in Reference [71] and [72]. An
LQ design is used in Reference [71], and is improved upon in Reference [72] using a triangular model and imposing
nonlinear dynamic inversion techniques [82].
A recent alternative to the nonlinear controller of Reference [72] is the adaptive sliding mode control approach of
Reference [83], [84]. In this work, a new adaptive sliding mode technique for a class of MIMO nonlinear systems
including the nonlinear models is presented. The adaptive scheme is based on the inverse dynamics of the original
system. The adaptive structure makes the controller efficient in dealing with the parametric uncertainties. The states
that are not available for measurement are estimated using a nonlinear sliding mode observer.
In all the above designs, the controller applies to a certain flight condition and hence a supervisor such as a gain
scheduling scheme is needed. Any supervisor design should take into account the significant variations in the AHF
system due to wide speed ranges. To this end, time- and/or parameter-varying approaches appear as a viable scheme or
as a companions to the gain scheduling schemes.
The literature includes a number of other efforts to control hypersonic flight. However, little or no attention is given
to the airframe/engine coupling. They merely concentrate on the application of specific techniques to the existing
hypersonic flight dynamic models. Main categories of this group of works are model reference adaptive control
approach for linear models [85], optimal and sub-optimal control approaches [86]-[88], genetic-based, algorithm-based
approaches [89], fuzzy control approaches [90], and neural network approaches [91],[92].

XI. NASA AHFV Programs Flight Control Designs


Although a significant number of AHFV control schemes have been designed and tested via simulations, none has
been completely implemented in a real AHFV primarily because the AHF technology is in its developmental and
testing stages. No operational vehicle yet exists, and the aim of the flight tests conducted or planned thus far has been to
analyze newly developed AHFV components and quantify the AHF dynamics rather than to perform a complete guided
flight. Nonetheless, the use of an efficient flight controller is one of the key technologies for flight demonstration of the
scramjet-powered hypersonic aircraft. In the Hyper-X program, such a controller achieved successful separation from
the booster rocket, maintaining the design condition during the engine test, and provided a controlled descent [33]. The
control laws to perform these tasks, developed as a Boeing/NASA partnership, have been tested via flight simulations

6
American Institute of Aeronautics and Astronautics
and stability margin analyses before being used in the flight tests. The basic goal of the flight control system in Hyper-
X was to maintain the desired AOA and bank angle to within 0.5 degrees during the test flights, and to follow steering
commands from the guidance system to maintain a desired descent trajectory after the tests are completed.
The Hyper-X flight control scheme design uses classical linear control techniques [93] rather than advanced
nonlinear laws [33]. The control design is based on a linear model including rigid-body modes and second-order
actuation modes. Lead-lag compensators are used to improve stability margins. Elastic mode effects are circumvented
in the controller design as a part of the gain margin requirements, avoiding the use of a separate structural filter. A
feedforward compensator suppresses the effects of the propulsion system and dynamic pressure variations on the
AHFV. The controller parameters are gain scheduled with AOA and Mach number.

XII. New Directions in Control Design


After performing a sufficient number of flight tests with demonstrator vehicles, the next step in AHF research will
be guided AHFs that follow certain prescribed paths. The final goal could be using AHFVs efficiently to follow
arbitrary paths similar to conventional aircraft in a wide speed span including hypersonic ranges. This goal could be
broadened to include the original single-state-to-orbit (SSTO) and triple-stage-to-orbit (TSTO) space access task of
NASP.
The airframe-integrated scramjet engine concept forms the basis of the technology of current and foreseen AHF
programs, and is the focal configuration of all major AHF research centers worldwide [10],[32],[96],[97]. Considering the
airframe-integrated engine configuration, an essential task is to enhance and optimize the propulsion systems developed
for this configuration. The two most dominant sources of uncertainty in AHF are the lack of flight test/wind tunnel data
and the commonly ignored coupling effects. The former can only be addressed in time, through performing ground and
flight tests. Until that time, reliance on high-fidelity simulation models such as the one presented in this paper is the
only viable alternative. Ongoing studies and discussions on data collection via experiments and numerical techniques
can be found in Reference [36], [37], [39]-[42], [58], [96].
In order to implement the control designs based on the integrated models, the aerodynamic and propulsive control
laws to be applied need to be interactive. Engine control effectors controlling fuel flow, the inlet/diffuser area ratio, and
the exit area of the internal nozzle should be in use along with the conventional control surfaces [16]. These control
effectors should be commanded via an integrated mechanism that processes the combined information from the
aerodynamic and propulsive measurements. The bandwidth of some of these dynamics may limit ones ability to
effectively control the vehicle.
To perform a fully guided AHF design, a supervisor to handle the variations in the complete flight envelope should
be designed. The supervisory scheme will be required to perform the transitions between these flight conditions. The
conventional methodology used in aerospace applications is gain scheduling, but in spite of its wide use and the recent
interest of the academic research community, there is still a lack of tests to guarantee global stability of gain-scheduled
systems in general [104]-[107]. Considering the speed of variations in the AHF system, wide operation ranges, and the lack
of broad flight data, recently developed time- and parameter-varying approaches can be used to fill the gaps in the gain
scheduling schemes.
Time- and parameter-varying methods have already been used in designing controllers for various aerospace
systems [108]-[110]. These control designs, however, are in general based on the assumption that the variations are slow in
a certain sense and they do not exploit the a priori information about the variation structures. A method to exploit the a
priori information about the system parameter in designing controllers for time-varying systems is presented in
Reference [111]. This method is used to design effective adaptive controllers for linear time-varying systems in
Reference [111]-[114]. Moreover, the control scheme of Reference [114] can be extended to a class of nonlinear
systems. This extended scheme is a potential candidate for AHF control.

XIII. Development of a 2-D Longitudinal AHF Cruise Vehicle Model


An effort is underway at the Multidisciplinary Flight Dynamics and Control Laboratory at California State
University, Los Angeles in collaboration with the University of Kansas (UK) and the University of Southern
California (USC) to develop a high-fidelity simulation model for a full-scale generic airbreathing hypersonic
vehicle, one resembling an actual test vehicle. Based on a set of requirements broadly identified for an actual AHFV,
a full-scale generic airbreathing hypersonic vehicle (CSULA-GHV) is under development, with a high level of detail
over a wide speed range (Mach 0.3 to Mach 20), using available unclassified (open literature) information. The
CSULA-GHV is designed to the requirements of a global-reach vehicle that must travel to a target halfway around
the world in less than two hours, deliver a payload of 20,000 pounds, and return to base without refueling at altitudes

7
American Institute of Aeronautics and Astronautics
near the outer limit of the Earths atmosphere. The CSULA-GHV concept vehicle (Figure 4) has an integrated
airframe propulsion system configuration resembling that of the X-43.
As a first cut a 2-D version of the GHV has been designed for development of a longitudinal model for control
design application. The vehicle has an integrated airframe-propulsion system configuration, and is designed to study
the unique challenges associated with the modeling and control of airbreathing hypersonic vehicles. Specifically, the
model is developed to investigate and quantify the couplings between aerodynamics, propulsion, structure, and
control systems. The configuration and dimensions are developed based on 2-D compressible flow theory, and the
set of mission requirements described above. The vehicle is designed to the requirements broadly accepted for a
hypersonic cruise vehicle intended for both space access and military applications. The longitudinal equations of
motion ensued will include both an inverse-square-law gravitational model and a centripetal acceleration that results
from a curved flight path. Longitudinal control is affected by elevators and the engine thrust. Main aerodynamic
coefficients, CL , CD , and CM are obtained using analytical aerodynamic calculations assuming a cruising condition
of Mach 10 at an altitude of 30 km. The 2-D oblique shock theory is used to predict shock wave angles, pressure on
the frontal surface, and Mach number at the engine inlet as described in the following section. The scramjet engine
is simply modeled by a 1-D compressible flow with heating, which predicts the flow rate of hydrogen fuel required
for a chosen design Mach number at the engine exit. The exit flow is modeled by 2-D expansion wave theory, which
can be used to predict the pressure on the rear surface. Resultant aerodynamic forces, total lift and drag, and engine
thrust are then estimated by summation of these pressure forces and momentum change of the airflow.
The unique aspect of this study however is the coupled aerodynamics propulsion simulations of the vehicle
dynamics in conjunction with the theory which enable uncovering and quantifying the couplings which are broadly
ignored in models used for control system design in the past. The multi-Physic software FLUENT is used to
simulate the vehicle coupled aero-propulsion and aero-structural-propulsion simulations to develop a high fidelity
model.

A. 2-D GHV Configuration Design


The 2-D configuration for the hypersonic vehicle is designed based on inviscid compressible flow theory of a
perfect gas. As shown in Figure 3, the upper body of the vehicle is simply a flat surface, which is kept at zero angle
of attack for simplicity. The lower side consists of a frontal wedged surface, a scramjet engine with a constant
cross-section area and another trailing wedged surface. The frontal wedged surface serves as a diffuser for the flow
entering the scramjet, and the trailing surface acts as a propulsive surface. The leading edge angle is arbitrary
chosen to be = 50, the length of the engine 9.5 m, and the engine cross-section area is A = 0.6 m (height) by 1 m
(span).

Figure 3. Shock and Expansion Waves in a Generic Hypersonic Vehicle Configuration

A flight Mach number, M1 = 10 at an altitude of 30 km (where the standard atmospheric temperature and
pressure are T1 = 227 K and P1 = 1172 Pa, respectively) is considered, and a corrected specific heat ratio = 1.36 for
air at hypersonic speeds is assumed. The wave angle (1) of the oblique shock generated from the leading edge, the
Mach number (M2), pressure (P2) and temperature (T2) behind the shock can be determined by the oblique shock
relations:

( ) ( ) 2 1
2 cot 1 M1 2 sin 1
tan( )
M1 ( + cos ( 2 1) ) + 2
2
(1)

1
1+ ( ( ))
M1 sin 1
2

(
M2 sin 1 ) 2

( ( ) ) 2 2 1
M1 sin 1
(2)

8
American Institute of Aeronautics and Astronautics

P2 := P1 1 +

2
(
M1 sin 1
+1
2
1
( ))
(3)

P2 2 + ( 1) ( M sin( ) ) 2

1 1
T2 := T1
( 2
+ 1) ( M1 sin( 1) )
P1
(4)

The leading edge of the lower surface of the engine inlet intercepts the first oblique shock to capture the entire air
flow rate and to deflect the flow back by 50 when entering the scramjet, so that the shock reflection would terminate
at the upper edge of the engine inlet, and the flow through the scramjet would become one-dimensional. The wave
angle (2) of the oblique shock reflected from the lower leading edge of the engine inlet, the Mach number (M3),
pressure (P3) and temperature (T3) at the engine inlet are similarly determined by the oblique shock relations.
The combustion process in the scramjet is simply modeled by Rayleigh flow theory, that is, one-dimensional
compressible flow with heat addition. Selecting a Mach number (M4 = 5) at the exit of the scramjet, the total
temperature change, the air flow rate and the rate of heat added can be determined from the Rayleigh Flow relations:

2
1+ M M4 2 + ( 1) M4
2 2
T03
3
T04
2
2 + ( 1) M3
2 M3
1 + M4 (5)


m3 P3 A M3
R T3
(6)

Q (
m3 Cp T04 T03 ) (7)

where Cp = 1.084 kJ/kg-K is the corrected specific heat of air at hypersonic speed.
The required flow rate of hydrogen fuel (with lower heating value LHV = 120 MJ/kg) can then be estimated, and
the total mass flow rate exiting the scramjet can be found:

Q
mf m4 m3 + mf
LHV , (8)

The expansion waves angles (4 , 5) extending from the upper edge of the engine exit, the Mach number (M5)
and pressure (P5) along the trailing surface are simply determined by the two-dimensional expansion wave theory:

4 asin
1 +1
G
M4 , 1 (9)

M 2 1
4
4 G atan atan M4 2 1
G (10)

M 2 1
5
4 + 5 G atan atan M5 2 1
G (11)

5 asin
1
M5 (12)


1
1
P04 P4 1 + M4 2
2 (13)

9
American Institute of Aeronautics and Astronautics

1
1
P5 P04 1 + M5 2
2 (14)

where 4 is the Prandtl-Meyer function and P04 is the total pressure at the engine exit.
Numerical results of these computations are summarized in the following table:

Mach Wave Pressures Flow Rates Velocities


Numbers Angles (kPa) (kg/s) (m/s)

M2 = 8.42 1 = 9.450 P2 = 3.45 m2 = 128.66 V2 = 2945

M3 = 7.21 2 = 10.430 P3 = 8.83 m3 = 128.66 V3 = 2907

M4 = 5.0 4 = 11.530 P4 = 18.06 mfuel = 0.359 V4 = 2864

M5 = 6.3 5 = 9.130 P5 = 4.08 m4 = 129.02 V5 = 2965

With the wave angles known and the dimensions of the scramjet engine already specified, the geometry of the
entire vehicle can be determined. The overall length is 33.37 m and the total height is 2.42 m, including the height
of the engine (0.6 m). The slope of the trailing surface is found to be 5 = 110.
Applying the momentum equations results in a net axial force (ThrustDrag) = 2.16 kN, and a net normal force
(Lift) = 28 kN, which dictates the vehicle weight at the beginning of steady level flight. The location of center of
mass of the vehicle is determined by a balance of pitching moments for trimmed flight, and is found to be at 48% aft
from the leading edge.
It is seen that 2-D inviscid flow analysis shows that it is possible to generate positive excess thrust with this
simple configuration. The dimensions, estimated cruise weight and centre of gravity of the vehicle all seem to be
reasonable. The 2-D data will serve as an initial design for further modifications and design iterations into a 3-D
configuration (Figure 4).

Figure 4. Initial Configuration

B. 2-D GHV Preliminary Simulation Results

1. Comparison of CFD and Analytical Results


Parallel to the analytical results described above Computational Fluid Dynamics (CFD) simulations have been
conducted using Fluent, a multi-physics CFD code capable of simulating compressible flow coupled with
combustion to obtain the basic aerodynamic properties of the 2-D vehicle, the propulsion model and the coupling
between the two.
The GHV was modeled using a coupled solver with inviscid flow, the density calculated from ideal gas law,
specific heat constant from the kinetic theory. The model has been simulated over AoA between negative and
positive five degrees at intervals of one degree, elevon deflection angles between negative and positive five degree

10
American Institute of Aeronautics and Astronautics
at intervals of 2.5 degrees and over six different fuel flow rate (0 kg/s, 0.5 kg/s, 0.75 kg/s, 1.0 kg/s, 1.25 kg/s and 1.5
kg/s).
A major challenge in the CFD simulations is the inclusion of shock wave interactions and the combustion
coupling. To reduce the computational demand, the aircraft was split into three sections and simulated separately,
the forebody was run without combustion through the range of angles of attack, then the engine through the range of
angles of attack and fuel flow rates and finally the aftbody through the range of angles of attack, fuel flow rates and
elevon angles.

Figure 5. Mach Number Contours for Different Angles of Attack


Propulsion integration was implemented by using a one-step finite rate chemistry model and injecting hydrogen
fuel from the upper surface of the scramjet as a boundary condition. Although using multi-step chemistry models
would yield results with greater accuracy it has been shown in [126] that a one-step model can provide general
agreement with experimental data and is therefore adopted in this study.
The preliminary computer simulation results are shown in Figure 5. The figure shows the CFD results for the
contours of mach number, at the optimum design condition of zero-degree-AoA and at the off-designs of five-
degree-AoA and negative-five-degree-AoA.
On comparison with the data from the theory, the following comparison can be made with the shockwave from
the leading edge.
Table 1 Results form 2-D Shock Wave Theory
Angle of Attack Theoretical Wave Angle CFD Wave Angle
-5 -9.52 -10 1
0 9.52 10 1
5 14.42 14.5 1

2. CFD-Generated Aerodynamic data


Basic aerodynamic data, including coefficients of lift, drag, pitching moment as well as thrust data have been
generated using Fluent. A more complete aerodynamic data base is currently under development. The preliminary
results are shown below and are compared with published data when available.

0.06
0.006
CSULA, CFD Data
NASA, X-43A, Mach 7, Wind Tunnel Data CSULA, CFD Data
0.05
NASA, X43A, Mach 7, Wind Tunnel Data

0.005
0.04

0.03
Coefficient Of Pitching Moment

0.004

0.02
Coefficient of Lift

0.003
0.01

0
-5 -4 -3 -2 -1 0 1 2 3 4 5 0.002

-0.01

-0.02 0.001

-0.03

0
-0.04 -5 -4 -3 -2 -1 0 1 2 3 4 5

Angle of Attack (Degrees) Angle of Attack (Degrees)

Figure 6 - Lift and Pitching Moment Coefficient Comparison between CFD Data for CSULA-GHV (Power
Off) at Mach 10 and Wind Tunnel Data for X-43 at Mach 7 [36]

11
American Institute of Aeronautics and Astronautics
Lift and pitching moment coefficient data from CFD simulations of the CSULA-GHV at Mach 10 and wind
tunnel data for the X-43 at Mach 7 [36] are shown on the same graph in Figure 6. As it can be seen there is good
agreement in the range of values. However, the data comparison is not meant to be an exact match because the
CSULA-GHV and the X-43 are subtly different vehicles.

Coefficient of Drag L/D

0.009 15

0.008

10
0.007

0.006 5
Coefficient of Drag

0.005

L/D
0
0.004 -5 -4 -3 -2 -1 0 1 2 3 4 5

0.003
-5

0.002

-10
0.001

0
-5 -4 -3 -2 -1 0 1 2 3 4 5 -15
Angle of Attack (Degrees) Angle of Attack (Degrees)

Figure 7. Drag Coefficient and L/D for Power Off

Coefficient of Lift Coefficient of Drag


0.06 0.006
CSULA, CFD Data, 0.5 kg/s Fuel 0.5 kg/s Fuel
CSULA, CFD Data, 1.0 kg/s Fuel 1.0 kg/s Fuel
0.05
CSULA, CFD Data, 1.5 kg/s Fuel 1.5 kg/s Fuel
NASA, X-43A, Mach 7, Wind Tunnel Data 0.005
0.04

0.03
0.004
Coefficient of Drag

0.02
Coefficient of Lift

0.003
0.01

0
-5 -4 -3 -2 -1 0 1 2 3 4 5 0.002

-0.01 0.006

-0.02 0.001
0.004

-0.03
0.002
-0.25 0 0.25 0
-0.04 -5 -4 -3 -2 -1 0 1 2 3 4 5
Angle of Attack (Degrees) Angle of Attack (Degrees)

Lift Coefficient for Power On Drag Coefficient for Power On


Effect of Fuel Rate on Thrust
Excess Thrust Coefficient (CT - CD)
0.002 0.004
0.5 kg/s Fuel 0.5 kg/s Fuel
1.0 kg/s Fuel 1.0 kg/s Fuel
1.5 kg/s Fuel 1.5 kg/s Fuel
0.0015 0.002

0.001 0
-5 -4 -3 -2 -1 0 1 2 3 4 5
Excess Thrust Coefficient
Thrust Coefficient

0.0005 -0.002

0 -0.004
-5 -4 -3 -2 -1 0 1 2 3 4 5

-0.0005 -0.006

-0.001 -0.008

-0.0015 -0.01
Angle of Attack (Degrees) Angle of Attack (Degrees)

Thrust Coefficient for Power On Excess Thrust Coefficient for Power On

12
American Institute of Aeronautics and Astronautics
L/D Coefficient of Pitching Moment About 40% Aircraft Length
15
0.005
0.5 kg/s Fuel
1.0 kg/s Fuel 0.5 kg/s Fuel
1.5 kg/s Fuel 1.0 kg/s Fuel
1.5 kg/s Fuel
10
0.004

Coefficient Of Pitching Moment


5 0.003
L/D

0 0.002
-5 -4 -3 -2 -1 0 1 2 3 4 5

-5 0.001

-10 0
-5 -4 -3 -2 -1 0 1 2 3 4 5

-15 -0.001

Angle of Attack (Degrees) Angle of Attack (Degrees)

L/D for Power On Pitch Moment Coefficient for Power On


Figure 8. Preliminary Aerodynamic Data for Power On

Coefficient of Lift, dE = -5 degrees Coefficient of Pitching Moment, dE = -5 degrees

0.06 0.0035

0.05 0.003

0.04 0.0025

0.03 0.002
Coefficient Of Pitching Moment

0.02 0.0015
Coefficient of Lift

0.01 0.001

0 0.0005
-5 -4 -3 -2 -1 0 1 2 3 4 5
-0.01 0
-5 -4 -3 -2 -1 0 1 2 3 4 5
-0.02 -0.0005

-0.03 -0.001

-0.04 -0.0015

-0.05 -0.002
Angle of Attack (Degrees) Angle of Attack (Degrees)

L/D, dE = -5 degrees Excess Thrust Coefficient, dE= -5 degrees

15 0.003

0.002

10
0.001

5 0
Excess Thrust Coefficient

-5 -4 -3 -2 -1 0 1 2 3 4 5

-0.001
L/D

0
-5 -4 -3 -2 -1 0 1 2 3 4 5 -0.002

-5 -0.003

-0.004

-10

-0.005

-15 -0.006
Angle of Attack (Degrees) Angle of Attack (Degrees)

Figure 9. Aerodynamic Data for Trim Condition: Elevon Deflection = -5 degrees, Fuel Flow Rate = 1.5 kg/s

3. Coupling Effects Aerodynamic & Propulsion


Effects on Pitching Moment

An examination of the Power off and Power on graphs reveals uncovers the coupling between aerodynamic
properties and propulsion. Most notable is the effect of the scramjet propulsion on the pitching moment in causing a
substantial nose down pitching increment (shown in Figure 10). This is caused by the aftbody surface being
pressurized by the expanding scramjet exhaust flow. A simple analysis of the actual forces on airframe-engine
integrated configuration shown in Figure 11 can be used to explain this effect.

13
American Institute of Aeronautics and Astronautics
0.006
Power Off
Power On - 0.5 kg/s Fuel
Power On - 1.0 kg/s Fuel
0.005 Power On - 1.5 kg/s Fuel

0.004

Coefficient Of Pitching Moment


0.003

0.002

0.001

0
-5 -4 -3 -2 -1 0 1 2 3 4 5

-0.001
Angle of Attack (Degrees)

Figure 10 - Aerodynamics & Propulsion Coupling Effects on Pitching Moment


In power off condition, with aerodynamic forces acting only the lift force on the leading edge is significantly
greater than that on the trailing edge. In power on condition with combustion effects, the difference between these
forces are reduced to a great extent.

Figure 11 - Lift Forces for Leading and Trailing Edge for Aerodynamics Forces both With and Without
Propulsion Coupling

Effects of Angle of Attack on Thrust


It is well known that maximum thrust is generated when the bow shock created by the leading edge is impinged
on the lip of the engine inlet and the entire aerodynamic pressure is captured. In fact, the GHV configuration design
at AoA zero is on that basis. It can be seen from Figure 8, Thrust Coefficient for Power On that the maximum thrust
is indeed generated at zero degrees angle of attack.
Of more interesting note, is the different fall-off rate for positive and negative angles of attack on the thrust
curves, again a direct result of the engine airframe integrated configuration, which at the first glance may appear
paradoxical. It may appear that at high angles of attach there is a higher air flow rate through the engine and hence
more power. However, the fall-off rate is higher at positive angles of attack because the shock wave from the
leading edge misses the engine cowl whereas it penetrates deeper into the engine up to the negative five degrees
AoA when the leading edge is aligned with the flow and there is no shockwave.

Effect of Scramjet on Lift

An examination of the list curves in Power on condition (Figure 8) suggests that the lift generated is insensitive
to the flow fuel flow rate or the thrust as the curves essentially collapse into a single graph for all angles of attack.
This is mainly due to the combustor being placed approximately near the C.G. of the vehicle. A review of the lift
forces generated in the engine compartment and the exhaust nozzle area (Figure 12) indicate that the lift force is also
coupled to the propulsion. However, since one drops with the increase in AoA almost at the same rate as the other is
increased, the coupling is not detectable from the overall lift curves. (Figure 12 depicts the lift force on both the
engine components and post engine surfaces)

14
American Institute of Aeronautics and Astronautics
100000
Post Engine - 0 kg/s Fuel
Engine Components - 0 kg/s Fuel
Post Engine - 0.5 kg/s Fuel 80000
Engine Components - 0.5 kg/s Fuel
Post Engine - 1.0 kg/s Fuel
Engine Components - 1.0 kg/s Fuel
60000
Post Engine - 1.5 kg/s
Engine Components - 1.5 kg/s Fuel

40000

Lift Force (N)


20000

0
-5 -4 -3 -2 -1 0 1 2 3 4 5

-20000

-40000

-60000

-80000
Angle of Attack (Degrees)

Figure 12 - The Effect of Propulsion on Lift Force both In the Engine and After the Engine at Different Fuel
Flow Rates

XIV. Conclusion
AHFV configurations with airframe integrated scramjets form the basis for the current and foreseen AHF
technology programs. On the other hand, almost all of the existing control designs in the open literature are based on
the rigid winged-cone accelerator configuration, which does not address all major dynamic characteristics of AHFVs
like X-30 or X-43A. Therefore, new designs based on realistic models that address all the major issues in AHF
dynamics are needed.
The hypersonic flight control law design is also associated with significant uncertainty. The two most dominant
sources of uncertainty in AHF are the lack of sufficient data and ignored coupling effects. The first will be handled
in time by flight tests. For the second it is necessary to develop accurate simulation models quantify and elaborate
the couplings and include them in the system models. The correct tradeoff between the comprehensiveness of a
model and complexity of designing/ implementing a controller based is the key in successful control system design.
A study has been initiated at Cal State LA in collaboration with USC and University of Kansas to develop in-
house a CFD-based high-fidelity model for a full-scale airbreathing generic hypersonic cruise vehicle, CSULA-
GHV., which based on a set of broadly accepted requirements for such a vehicle. The GHV resembles NASAs X-
43. As a first cut a 2-D model has been developed and presented in this paper. Coupled CFD data generated using
Fluent, a multi-physic code are presented and coupling effects are discussed. A comprehensive list of references is
included..

Acknowledgments
This project was supported by the United States Air Force under Grant No. F49620-01-1-0489 and by NASA
Dryden Flight Research Center under Grant No. NAG4-175.

References
[1] Scientific Advisory Board, United States Air Force, Washington, DC, New World Vistas: Air and Space Power for 21st Century,
1995
[2] National Research Council, National Academy Press, Washington, DC, Hypersonic Technology for Military Applications, 1989
[3] Naidu, D.S., Banda, S.S, and Buffington, J.L., Unified Approach to H2 and H Optimal Control of a Hypersonic Vehicle,
Proceedings of the American Control Conference, San Diego, California, June 1999.
[4] Schmidt, D. K., Mamich, H., and Chavez, F., Dynamics and control of hypersonic vehicles, The integration challenge for the
1990s, AIAA Paper 91-5057, Dec. 1991.
[5] Chavez, F. R. and Schmidt, D. K., Analytical aeropropulsive/aeroelastic hypersonic-vehicle model with dynamic analysis,
AIAA Journal of Guidance, Control, and Dynamics, Vol. 17, No. 6, Nov. 1994, pp. 1308-1319.
[6] Bilimoria, K. D. and Schmidt, D. K., Integrated Development of the Equations of Motion for Elastic Hypersonic Flight
Vehicles, AIAA Journal of Guidance, Control, and Dynamics, Vol. 18, No. 1, Jan. 1995, pp. 73-81.
[7] Chavez, F. R. and Schmidt, D. K., Uncertainty modeling for multivariable-control robustness analysis of elastic high-speed
vehicles, AIAA Journal of Guidance, Control, and Dynamics, Vol. 22, No. 1, Jan. 1999, pp. 87-95.

15
American Institute of Aeronautics and Astronautics
[8] Fidan, B., Mirmirani, M., and Ioannou, P. A., Flight Dynamics and Control of Air-Breathing Hypersonic Vehicles: Review and
New Directions, AIAA Paper 2003-7081, Dec. 2003.
[9] Freeman, D. C., Reubush, D. E., McClinton, C. R., Rausch, V. L., and Crawford, J. L., The NASA Hyper-X program, 48th
International Astronautical Congress, Oct. 1997.
[10] Curran, E. T., Scramjet Engines: The first forty years, AIAA Journal of Propulsion and Power, Vol. 17, No. 6, Nov. 2001, pp.
1138-1148.
[11] Weber, R. J. and Mackay, J. S., An Analysis of Ramjet Engines Using Supersonic Combustion, NACA TN 4386, Sep. 1958.
[12] Dugger, G. L., Comparison of hypersonic ramjet engines with subsonic and supersonic combustion, High Mach Number Air-
breathing Engines, Pergamon, Oxford, U.K., 1961.
[13] Heiser, W. H., Pratt, D. T., Daley, D. H., and Mehta, U. B., editors, Hypersonic Air-breathing Propulsion, AIAA, Washington,
DC, 1994.
[14] Hallion, R. P., editor, The Hypersonic Revolution, Vol. Vol. I: From Max Valier to Project PRIME (1924-1967), Air Force
History and Museums Program, Bolling AFB, DC, 1998.
[15] Hallion, R. P., editor, The Hypersonic Revolution, Vol. Vol. II: From Scramjet to National Aero-Space Plane, (1964-1986), Air
Force History and Museums Program, Bolling AFB, DC, 1998.
[16] Thompson, M. O., At the Edge of Space: The X-15 Flight Program, Smithsonian Institution Press, Washington, DC, 1992
[17] Henry, J. R. and Anderson, G. Y., Design considerations for the airframe-integrated scramjet, NASA TM X-2895, 1973.
[18] Weidner, J. P., Small, W. J., and Penland, J. A., Scramjet Integration on Hypersonic Airplane Concepts, AIAA Journal of
Aircraft, Vol. 14, No. 5, May 1977, pp. 460-466.
[19] Northam, G. B. and Anderson, G. Y., Supersonic combustion research at Langley, AIAA Paper 86-0159, Jan. 1986.
[20] Rogers, R. C., Capriotti, D. P., and Guy, R. W., Experimental supersonic combustion research at NASA Langley, AIAA Paper
98-2506, June 1998.
[21] Schweikart, L., editor, The Hypersonic Revolution, Vol. Volume III: The Quest for the Orbital Jet: The National Aero-Space
Plane Program (1983-1995), Air Force History and Museums Program, Bolling AFB, DC, 1998.
[22] Voland, R. and Rock, K., NASP concept demonstration engine and subscale parametric engine tests, AIAA Paper 95-6055,
April 1995
[23] McRuer, D., Design and modeling issues for integrated airframe/propulsion control of Hypersonic Flight vehicles, Proc.
American Control Conference, 1991, pp. 729-734.
[24] Schmidt, D. K., Dynamics and control of hypersonic aeropropulsive/aeroelastic vehicles, AIAA paper 92-4326, Aug. 1992.
[25] Schmidt, D. K., Integrated control of hypersonic vehicles, AIAA paper 93-5091, Dec. 1993.32
[26] Chavez, F. R. and Schmidt, D. K., An Integrated Analytical Aeropropulsive/Aeroelastic for Dynamic Analysis of Hypersonic
Vehicles, NASA ARC 92-2, June 1992.
[27] Schmidt, D. K. and Lovell, T. A., Mission performance and design sensitivities of air-breathing hypersonic launch vehicles,
AIAA Journal of Spacecraft and Rockets, Vol. 34, No. 2, March 1997, pp. 158-164.
[28] Schmidt, D. K., Optimum mission performance and multivariable flight guidance for Air-breathing launch vehicles, AIAA
Journal of Guidance, Control, and Dynamics, Vol. 20, No. 6, Nov. 1997, pp. 1157-1164.
[29] Schmidt, D. K. and Hermann, J. A., Use of energy-state analysis on a generic air-breathing hypersonic vehicle, AIAA Journal of
Guidance, Control, and Dynamics, Vol. 21, No. 1, Jan. 1998, pp. 71-76.
[30] Rausch, V. L., McClinton, C. R., and Crawford, J. L., Hyper-X: Flight validation of hypersonic air-breathing technology, XIII
ISABE, ISABE 97-7024, Sept. 1997.
[31] McClinton, C. R., Rausch, V. L., Sitz, J., and Reukauf, P., Hyper-X program status, 10th AIAA/NAL/NASD-ISAS International
Space Planes and Hypersonic Systems and Technologies Conference, AIAA 2001-1910, Apr. 2001.
[32] Hueter, U. and McClinton, C. R., NASA's advanced space transportation hypersonic program, 11th AIAA/AAAF International
Conference, AIAA 2002-5175, 2002.
[33] Davidson, J., Lallman, F. J., McMinn, J. D., Martin, J., Pahle, J., Stephenson, M., Selmon, J., and Bose, D., Flight control laws
for NASA's Hyper-X research vehicle, AIAA Paper 99-4124, 1999.
[34] Rock, K. E., Voland, R. T., Rogers, R. C., and Huebner, L. D., NASA's Hyper-X Scramjet Engine Ground Test Program, ISABE
99-7214, Sept. 1999.
[35] Engelund, W. C., Holland, S. D., Cockrell Jr., C. E., and Bittner, R. D., Propulsion System Air- frame Integration Issues and
Aerodynamic Database Development for the Hyper-X Flight Research Vehicle, ISOABE Paper 99-7215, 1999.
[36] Engelund, W. C., Hyper-X Aerodynamics: The X-43A Airframe- Integrated Scramjet Propulsion Flight-Test Experiments,
AIAA Journal of Spacecraft and Rockets, Vol. 38, No. 6, Nov. 2001, pp. 801-802.
[37] Engelund, W. C., Holland, S. D., Cockrell Jr., C. E., and Bittner, R. D., Aerodynamic Database Development for the Hyper-X
Airframe-Integrated Scramjet Propulsion Experiments, AIAA Journal of Spacecraft and Rockets, Vol. 38, No. 6, Nov. 2001, pp.
803-810.
[38] Woods, W. C., Holland, S. D., and DiFulvio, M., Hyper-X Stage Separation Wind-Tunnel Test Program, AIAA Journal of
Spacecraft and Rockets, Vol. 38, No. 6, Nov. 2001, pp. 811-819.
[39] Buning, P. G., Wong, T. C., Dilley, A. D., and Pao, J. L., Computational Fluid Dynamics Pre-diction of Hyper-X Stage
Separation Aerodynamics, AIAA Journal of Spacecraft and Rockets, Vol. 38, No. 6, Nov. 2001, pp. 820-827.
[40] Holland, S. D., Woods, W. C., and Engelund, W. C., Hyper-X Research Vehicle Experimental Aerodynamics Test Program
Overview, AIAA Journal of Spacecraft and Rockets, Vol. 38, No. 6, Nov. 2001, pp. 828-835. 33

16
American Institute of Aeronautics and Astronautics
[41] Cockrell Jr., C. E., Engelund, W. C., Bittner, R. D., Jentink, T. N., Dilley, A. D., and Frendi, A., Integrated Aeropropulsive
Computational Fluid Dynamics Methodology for the Hyper-X Flight Experiment, AIAA Journal of Spacecraft and Rockets, Vol.
38, No. 6, Nov. 2001, pp. 836-843.
[42] Huebner, L. D., Rock, K. E., Ruf, E. G., Witte, D. W., and Andrews Jr., E. H., Hyper-X Flight Engine Ground Testing for Flight
Risk Reduction, AIAA Journal of Spacecraft and Rockets, Vol. 38, No. 6, Nov. 2001, pp. 844-852.
[43] Reubush, D. E., Martin, J. G., Robinson, J. S., Bose, D. M., and Strovers, B. K., Hyper-X Stage Separation - Simulation
Development and Results, AIAA 2001-1802, April 2001.
[44] Faulkner, R. F. and Weber, J. W., Hydrocarbon Scramjet Propulsion System Development, Demonstration, and Application,
AIAA Paper 99-4922, 1999.
[45] White, D. A., Bowers, A., Iliff, K., and Menousek, J., Handbook of Intelligent Control: Neural, Fuzzy, and Adaptive Approaches,
chap. Flight, Propulsion, and Thermal Control of Advanced Aircraft and Hypersonic Vehicles, Van Nostrand Reinhold, New
York, NY, 1992, pp. 357-465.
[46] Bertin, J. J., Periaux, J., and Ballmann, J., editors, Advances in Hypersonics, Birkhuser, Boston, MA, 1992.
[47] Anderson, Jr., J. D., Hypersonic and High Temperature Gas Dynamics, McGraw-Hill, 1989.
[48] Bertin, J. J., Hypersonic Aerothermodynamics, AIAA, Washington, DC, 1994.
[49] Johnson, P. J., Whitehead, Jr., A. H., and Chapman, G. T., Fitting Aerodynamics and Propulsion in to the Puzzle, AIAA
Aerospace America, Sept. 1987, pp. 32-34.
[50] Maus, J. R., Giffith, B. J., Szema, K. Y., and Best, J. T., Hypersonic Mach Number and Real Gas Effects on Space Shuttle
Orbiter Aerodynamics, AIAA Journal of Spacecraft and Rockets, Vol. 21, No. 2, Nov. 1984, pp. 132-141.
[51] Raney, D. L., Phillips, M. R., and Person, L. H., Investigation of Piloting Aids for Manual Control of Hypersonic Maneuvers,
NASA Technical Paper 3525, October 1995.
[52] Chalk, C. R., Flying Qualities Criteria Review, Assessment and Recommendations for NASP, NASP CR-1065, NASP JPO,
Wright-Patterson AFB, 1989.
[53] McRuer, D. T. and Myers, T. T., Considerations for the Development of NASP Flying Qualities Specifications, WL-TR-92-
3042, U.S. Air Force, 1992.
[54] Penland, J. A., Dinlon, J. L., and Pittman, J. L., An Aero Dynamic Analysis of Several Hypersonic Research Airplane Concepts
from M=0.2 to 6.0, AIAA Journal of Aircraft, Vol. 15, No. 11, Nov. 1978, pp. 716-723.
[55] Whitehead, Jr., A. H., NASP Aerodynamics, AIAA Paper 89-5013, 1989.
[56] Walton, J. T., Performance Sensitivity of Hypersonic Vehicles to Change Angle of Attack and Dynamic Pressure, AIAA Paper
89-2463, 1989.
[57] Bushcek, H. and Calise, A. J., Uncertainty modeling and fixed-order controller design for a hypersonic vehicle model, AIAA
Journal of Guidance, Control, and Dynamics, Vol. 20, No. 1, Jan. 1997, pp. 42-48.
[58] Gupta, K. K., Voelker, L. S., Bach, C., Doyle, T., and Hahn, E., CFD-Based Aeroelastic Analysis of the X-43 Hypersonic Flight
Vehicle, AIAA Paper 01-0712, 2001.
[59] Cowan, J. C., Arena, Jr., A. S., and Gupta, K. K., Development of a Discrete-Time Aerodynamic Model for CFD-Based
Aeroelastic Analysis, AIAA Paper 99-0765, 1999.
[60] Gupta, K. K. and Petersen, K. L., Multidisciplinary Aeroelastic Analysis of a Generic Hypersonic Vehicle, AIAA Paper 93-
5028, Dec. 1993.
[61] Thuruthimattam, B., Friedmann, P., Powell, K., and McNamara, J., Aeroelasticity of a Generic Hypersonic Vehicle, AIAA
Paper 2002-1209, 2002.
[62] Raney, D. L., Pototzky, A. S., and McMinn, J. D., Impact of Aero-Propulsive-Elastic Interactions on Longitudinal Flight
Dynamics of a Hypersonic Vehicle, AIAA Journal of Aircraft, Vol. 32, March 1995.
[63] Lind, R., Buffington, J., and Sparks, A., Multi-Loop Aeroservoelastic Control of a Hypersonic Vehicle, AIAA Paper 99-4123,
August 1999.
[64] Etkin, B., Dynamics of Atmospheric Flight, Wiley, New York, 1972
[65] Meile, A., Flight Mechanics Vol. 1: Theory of Flight Paths, Addison-WESLEY, Reading, MA, 1962.
[66] Bisplighoff, R.L. and Ashley, H., Principles of Aeroelesticity, Wiley, New York, 1962.
[67] Mirmirani, M., Xu, H., and Choi, S., Linear Quadratic Gaussian Control of Coupled Aeroservoelastic of a Simple Wing Model,
Technical Report 101-2000, California State University, Los Angeles, December 2000.
[68] Shaughnessy, J. D., Pinckney, S. Z., McMinn, J. D., Cruz, C. I., and Kelley, M. L., Hypersonic Vehicle Simulation Model:
Winged-Cone Configuration, NASA TM-102610, 1990.
[69] Gregory, I. M., Chowdhry, R. S., McMinn, J. D., and Shaughnessy, J. D., Hypersonic vehicle model and control law
development using H and -synthesis, NASA TM-4562, 1994
[70] Brauer, G. L., Cornick, D. E., and Stevenson, R., Capabilities and Applications of the Program to Optimize Simulated
Trajectories(POST) - Program Summary Document, NASA CR-2770, 1977.
[71] Marrison, C. I. and Stengel, R. F., Design of robust control systems for a hypersonic aircraft, AIAA Journal of Guidance,
Control, and Dynamics, Vol. 21, No. 1, Jan. 1998, pp. 58-63.
[72] Wang, Q. and Stengel, R. F., Robust nonlinear control of a hypersonic aircraft, AIAA Journal of Guidance, Control, and
Dynamics, Vol. 23, No. 4, July 2000, pp. 577-585.
[73] Stengel, R. F., Optimal Control and Estimation, Dover, New York, NY, 1994.
[74] Ray, L. R. and Stengel, R. F., Stochastic Robustness of Linear-Time-Invariant Control Systems, IEEE Transactions on
Automatic Control, Vol. 36, No. 1, January 1991, pp. 82-87.

17
American Institute of Aeronautics and Astronautics
[75] Ray, L. R. and Stengel, R. F., A Monte Carlo Approach to the Analysis of Control System Robustness, Automatica, Vol. 29,
No. 1, January 1993, pp. 229-236.
[76] Maciejowski, J. M., Multivariable Feedback Design, Addison-Wesley, New York, NY, 1989.
[77] Zhou, K. and Doyle, J. C., Essentials of Robust Control, Prentice-Hall, Upper Saddle River, NJ, 1998.
[78] Bowers, A. H. and Iliff, K. W., Generic Hypersonic Aerodynamic Model Example (GHAME) for computer simulation, NASA
TM, NASA Ames-Dryden, Edwards, CA, 1987.
[79] Bushcek, H. and Calise, A. J., Controllers: Mixed and Fixed, AIAA Journal of Guidance, Control, and Dynamics, Vol. 20, No.
1, Jan. 1997, pp. 34-41.
[80] Young, P. M., Controller Design with Mixed Uncertainties, Proc. American Control Conference, 1994, pp. 2333-2337.
[81] Lohsoonthorn, P., Jonckheere, E., and Dalzell, S., Eigenstructure vs constrained H design for hypersonic winged cone, AIAA
Journal of Guidance, Control, and Dynamics, Vol. 24, No. 4, July 2001, pp. 648-658.
[82] Isidori, A., Nonlinear Control Systems, Springer-Verlag, 1997.
[83] Xu, H. and Mirmirani, M., Robust adaptive sliding control for a class of MIMO nonlinear systems, AIAA Paper 2001-4168,
2001.
[84] Xu, H., Leung, P., Mirmirani, M., Boussalis, H., and Ioannou, P., Adaptive sliding mode control of a hypersonic flight vehicle,
Advances in Astronautical Sciences 108, Part 2, 2001, pp. 1947-1962.
[85] Mooij, E., Numerical investigation of model reference adaptive control for hypersonic aircraft, AIAA Journal of Guidance,
Control, and Dynamics, Vol. 24, No. 2, March 2001, pp. 315-323.
[86] Dewell, L. D. and Speyer, J. L., Fuel-optimal periodic control and regulation in constrained hypersonic flight, AIAA Journal of
Guidance, Control, and Dynamics, Vol. 20, No. 5, Sep. 1997, pp. 923-932.
[87] Chuang, C. H. and Morimoto, H., Sub-optimal and optimal periodic solutions for hypersonic transport, Proc. American Control
Conference, June 1995, pp. 1186-1190.
[88] Naidu, D. S., Banda, S. S., and Buffington, J. L., Unified approach to H2 and H1 optimal control of a hypersonic vehicle, Proc.
American Control Conference, June 1999, pp. 2737-2741.
[89] Austin, K. J. and Jacobs, P. A., Application of genetic algorithms to hypersonic flight control, Proc. IFSA World Congress and
20th NAFIPS International Conference, Vol. 4, July 2001, pp. 2428-2433.
[90] Zhou, Z. and Lin, C. F., Fuzzy logic based flight control system for hypersonic transporter, Proc. Conference on Decision and
Control, Dec 1997, pp. 2730-2735.
[91] Cox, C., Neidhoefer, J., Saeks, R., and Lendaris, G., Neural adaptive control of the LoFLYTE, Proc. American Control
Conference, Vol. 4, 2001, pp. 2913-2917.
[92] Saeks, R., Neidhoefer, J., Cox, C., and Pap, R., Neural control of the LoFLYTE aircraft, Proc. IEEE Int. Conference on
Systems, Man, and Cybernetics, Vol. 4, Oct. 1998, pp. 3112-3117.
[93] Ogata, K., Modern Control Engineering, Prentice-Hall, Upper Saddle River, NJ, 2002.
[94] Gibson, C., Neidhoefer, J., Cooper, S., Carlton, L., Cox, C., and Jorgensen, C., Development and Flight Test of the X-43A-LS
Hypersonic Configuration UAV, AIAA Paper 2002-3462, 2002.
[95] Neidhoefer, J., Gibson, C., Saeks, R., Cox, C., Kocher, M., and Hunt, L., Accurate Automation Corporation's LoFLYTE
Program, AIAA Paper 2002-3502, 2002.
[96] Marren, D., Lewis, M., and Maurice, L. Q., Experimentation, test, and evaluation requirements for future air-breathing
hypersonic systems, AIAA Journal of Propulsion and Power, Vol. 17, No. 6, Nov. 2001, pp. 1361-1365.
[97] Bowcutt, K. G., Multidisciplinary optimization of air-breathing hypersonic vehicles, AIAA Journal of Propulsion and Power,
Vol. 17, No. 6, Nov. 2001, pp. 1184-1190.
[98] Krsti, M., Kanellakopoulos, I., and Kokotovi, P. V., Nonlinear and Adaptive Control Design, Wiley, New York, NY, 1995.
[99] Krsti, M. and Deng, H., Stabilization of Nonlinear Uncertain Systems, Springer-Verlag, 1998.
[100] Marino, R. and Tomei, P., Nonlinear Control Design - Geometric, Adaptive, and Robust, Prentice-Hall, London, UK, 1995.
[101] Qu, Z., Robust Control of Nonlinear Uncertain Systems, Wiley, New York, NY, 1998.
[102] Isidori, A., Lamnabhi-Lagarrigue, F., and Respondek, W., editors, Nonlinear Control in the Year 2000, Springer-Verlag, 2001.
[103] Fidan, B., Kosmatopoulos, E. B., and Ioannou, P. A., A Switching Controller for Multivariable LTI Systems with Known and
Unknown Parameters, Proc. 41st IEEE Conference on Decision and Control, Vol. 4, 2002, pp. 4688-4693.
[104] Hyde, R. A., editor, H1 Aerospace Control Design: A VSTOL Flight Application, Springer-Verlag, 1995.
[105] Shamma, J. S. and Athans, M., Analysis of Gain Scheduled Control for Nonlinear Plants, IEEE Transactions on Automatic
Control, Vol. 35, No. 8, August 1990, pp. 898-907.
[106] Shamma, J. S. and Athans, M., Gain Scheduling: Potential Hazards and Possible Remedies, IEEE Control Systems Magazine,
June 1992, pp. 101-107.
[107] Rugh, W. J. and Shamma, J. S., Research on Gain Sceduling, Automatica, Vol. 36, No. 10, October 2000, pp. 1401-1425.
[108] Winiewski, R., Linear Time-Varying Approach to Satellite Attitude Control Using Only Electro-magnetic Actuation, AIAA
Journal of Guidance, Control, and Dynamics, Vol. 23, No. 4, July-Aug.2000.
[109] Zhu, J. J. and Mickle, M. C., Missile Autopilot Design Using a New Linear Time-Varying Control Technique, AIAA Journal of
Guidance, Control, and Dynamics, Vol. 20, No. 1, Jan.-Feb. 1997.
[110] Biannic, J. M., Apkarian, P., and Garrard, W. L., Parameter Varying Control of a High- Performance Aircraft, AIAA Journal of
Guidance, Control, and Dynamics, Vol. 20, No. 2, March- April 1997.
[111] Tsakalis, K. S. and Ioannou, P. A., Linear Time Varying Systems: Control and Adaptation, Prentice-Hall, Englewood Cliffs, NJ,
1993.

18
American Institute of Aeronautics and Astronautics
[112] Limanond, S. and Tsakalis, K., Model reference adaptive and nonadaptive control of linear time-varying plants, IEEE Trans. on
Automatic Control, Vol. 45, 2000, pp. 1290-1300.
[113] Limanond, S. and Tsakalis, K. S., Adaptive and nonadaptive pole-placement control of multivariable linear time-varying plants,
Int. J. of Control, Vol. 74, 2001, pp. 507-523.
[114] Zhang, Y., Fidan, B., and Ioannou, P. A., Backstepping Control of Linear Time Varying Systems with Known and Unknown
Parameters, to appear in IEEE Trans. on Automatic Control, 2003.
[115] Nijmeijer, H. and Fossen, T. I., editors, New Directions in Nonlinear Observer Design, Springer-Verlag, 1999.
[116] Kokotovi, P. and Arcak, M., Constructive nonlinear control: A historical perspective, Automatica, Vol. 37, No. 5, 2001, pp.
637-662.
[117] Arcak, M. and Kokotovi, P., Nonlinear observers: A circle criterion design and robustness analysis, Automatica, Vol. 37, No.
12, December 2001, pp. 1923-1930.
[118] C. I. Marrison and R. F. Stengel, Design of Robust control system for a hypersonic aircraft, Journal of Guidance, Control, and
Dynamics, Vol. 21, No. 1, 1998.
[119] Q. Wang and R. F. Stengel, Robust Nonlinear Control of a Hypersonic Aircraft, J. of Guidance, Control and Dynamics, Vol. 23,
No. 4, 2000.
[120] J.-J. E. Slotine and W. Li, Applied nonlinear control, Prentice Hall, 1991
[121] P. A. Ioannou and J. Sun, Robust Adaptive Control, Prentice Hall, Upper Saddle River, 1996.
[122] J.-J. E. Slotine and J. A. Coetsee, Adaptive Controller Synthesis for Nonlinear Systems, Int. J. Control, Vol. 43, No. 6, 1986, pp.
1631-1651.
[123] Josue Cruz, CFD-based Aerodynamic Coefficients and Stability Derivatives for a Generic Hypersonic Air Vehicle, CSULA,
2003.
[124] Shahriar Keshmiri, Six-DOF modeling and simulation of a generic hypersonic vehicle for conceptual design studies,
CSULA, 2004.
[125] Fluent 6.2.16 User Guide, Fluent Inc. Lebanon, NH
[126] P. Hyslop, CFD Modeling of Supersonic Combustion in a Scramjet Engine Thesis, Aerophysics and Laser Diagnostics
Research Kaboratory, Australia National University, 1998

19
American Institute of Aeronautics and Astronautics

Vous aimerez peut-être aussi