You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/258515051

Optimal Trajectory Analysis of Hypersonic Boost-Glide Waverider with Heat


load Constraint

Article  in  Aircraft Engineering and Aerospace Technology · November 2013

CITATION READS

1 275

3 authors, including:

S Tauqeer Ul Islam Rizvi


Beihang University (BUAA)
17 PUBLICATIONS   24 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

HBG Vehcile Trajectory Optimization View project

All content following this page was uploaded by S Tauqeer Ul Islam Rizvi on 09 October 2015.

The user has requested enhancement of the downloaded file.


Optimal trajectory analysis of hypersonic
boost-glide waverider with heat load
constraint
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun
School of Astronautics, Beihang University, Beijing, China

Abstract
Purpose – The purpose of the paper is to study the variation of optimal burnout angle at the end of the ascent phase and the optimal control
deflection during the glide phase, that would maximize the downrange performance of a hypersonic boost-glide waverider, with variation in heat
rate and integrated heat load limit.
Design/methodology/approach – The approach used is to model the boost phase so as to optimize the burnout conditions. The nonlinear,
multiphase, constraint optimal control problem is solved using an hp-adaptive pseudospectral method.
Findings – The constraint heat load results for the waverider configuration reveal that the integrated heat load can be reduced by more than half
with only 10 per cent penalty in the overall downrange of the hypersonic boost-glide vehicle, within a burnout speed range of 3.7 to 4.3 km/s. The
angle-of-attack trim control requirements increase with stringent heat rate and integrated heat load bounds. The normal acceleration remains within
limits.
Practical implications – The trajectory results imply lower thermal protection system weight because of reduced heat load trajectory profile and
therefore lower thermal protection system cost.
Originality/value – The research provides further study on the trajectory design to the hypersonic boost-glide vehicles for medium range
application.
Keywords Heat rate limit, Hypersonic boost-glide, Integrated heat load limit, Pseudospectral method, Trajectory optimization,
Waverider vehicle
Paper type Research paper

Nomenclature T ⫽ Thrust, N
Tg ⫽ Gas Temperature, K
C ⫽ proportionality constant, kg0.5m1.5/s0.3 Tw ⫽ Wall temperature, K
CD ⫽ coefficient of drag t ⫽ time, s
CL ⫽ coefficient of lift V ⫽ velocity, m/s
CL␣ ⫽ Lift coefficient derivative with respect to V3 ⫽ burnout velocity, m/s
angle-of-attack W/S ⫽ planform loading, kg/m2
D ⫽ drag force, N ␣ ⫽ angle-of-attack
g ⫽ acceleration due to gravity, m/s2 ␤ ⫽ inverse of density scale height, m
g0 ⫽ acceleration due to gravity at sea-level, m/s2 ␥ ⫽ flight path angle, deg or rad
L ⫽ lift force, N ␥3 ⫽ flight path angle at burnout point, deg
L/D ⫽ lift-to-drag ratio ⌿ ⫽ heading angle, deg or rad
m ⫽ mass, kg ⍜ ⫽ longitude, deg or rad
nz ⫽ normal acceleration, g ⌽ ⫽ latitude, deg or rad
Q ⫽ total heat load, J/m2 ␳ ⫽ density, kg/m3
Q̇, Q= ⫽ heat transfer rate, W/m2 ␳0 ⫽ sea-level density, kg/m3
q ⫽ dynamic pressure, Pa ␮ ⫽ gravitational parameter, m3/s2
Re ⫽ radius of earth, 6,378 ⫻ 103 m ␴ ⫽ bank angle, deg or rad
R ⫽ total missile range, m
S ⫽ reference area, m2
Introduction
The performance of the lifting reentry vehicle is linked to the
The current issue and full text archive of this journal is available on lift-to-drag ratio and a waverider design provides the best
Emerald Insight at: www.emeraldinsight.com/1748-8842.htm

The authors would like to thank the Chinese Scholarship Council for
funding the research.
Aircraft Engineering and Aerospace Technology: An International Journal
87/1 (2015) 67–78 Received 28 April 2013
© Emerald Group Publishing Limited [ISSN 1748-8842] Revised 10 August 2013
[DOI 10.1108/AEAT-04-2013-0079] Accepted 16 November 2013

67
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

lift-to-drag ratio in the hypersonic regime, making it ideally et al., 2013). It, therefore, becomes essential to model the
suitable for a hypersonic cruise, as well as hypersonic boost phase so as to optimize the burnout angle which would
boost-glide (HBG) vehicles. The sharp nose radius of the result in maximum downrange performance of the vehicle
waverider vehicle exerts an additional burden on the thermal under heat rate and integrated heat load constraints. The
protection system (TPS) because of high heat rates and higher current paper deliberates on computing the optimal burnout
total integrated heat load. NASA Ames and Sandia Slender conditions at the end of the ascent phase, and the optimal
Aerodynamic Research Probe (SHARP L1) has a nose control deflections during the reentry phase which would
leading-edge radius of the order of 3-6 mm (Whitmore and maximize the downrange performance of the small-size
Dunbar, 2003) and, therefore, experiences a very high heat waverider, with different heat-rate and integrated heat load
rate during the reentry phase. The waverider technology is still limits.
undergoing test and trial in the form of X-41 vehicle by The problem has been modeled as a four-phase optimal
Defense Advanced Research Projects Agency. The heat rate control problem. The numerical optimization study has been
limit relates to the maximum temperature a material can bear, carried out for two payload sizes on a launch vehicle, to show
whereas the integrated heat load directly relates to the validity of results within a burnout speed range of 3.7-4.3 km/s.
thickness of the TPS and hence the empty weight of the The approach, which encompasses modeling of mass,
vehicle. It is for this reason that the deterioration of propulsion, aerodynamic and thermodynamics, as well as the
downrange and cross-range performance with materials of flight dynamics of the launch and the glide vehicles, gives the best
varying degree of heat rate tolerance needs to be studied. burnout angle and the burnout altitude for the ascent vehicle, as
Trajectories with a limit on the total heat load provide the best well as the angle-of-attack control profile during the reentry
mechanism by which the TPS weight can be reduced. phase required for maximum downrange trajectories. The
In the past, numerous studies have been performed on non-linear optimal control problem is solved using an
lifting reentry vehicles for intercontinental missile hp-adaptive pseudospectral method implemented in Gauss
applications, orbital Space Planes and crew return vehicles. Pseudospectral Optimization Software, GPOPS® (Rao, 2011).
Trajectory optimization studies for a lifting crew return vehicle
was carried out by Whitmore et al. (2006). Optimal Physical model
trajectories for orbital transfer vehicles have been studied by Equations of motion
Hull and Speyer (1982), Zimmermann and Calise (1998), The following set of 3-df equations of motion is used to
Darby and Rao (2011). Gogu et al. (2009) solved a coupled simulate the flight dynamics of the launch, as well as the
trajectory optimization problem for an orbital transfer vehicle hypersonic glide vehicles. The equations have been extensively
with the objective to minimize the thermal protection system used in the study of re-entry vehicles and their guidance
mass. Wing-body configuration was studied by Parish II systems (Hull and Speyer, 1982; Gogu et al., 2009; Darby and
(1995) for Intercontinental Ballistic Missiles application. Rao, 2011).
The concept of HBG vehicle was introduced by Sanger and
Bred (1944), who envisioned a vehicle which reached dr
⫽ V sin ␥ (1)
hypervelocity using rocket motors and returned to earth along dt
a combined skip-glide trajectory. The vehicle utilized
aerodynamic lift to cancel the effect of gravity during the d⌰ V cos ␥ sin ␺
⫽ (2)
ascent, as well as the glide path. Further work was done by dt r cos ␸
Eggers et al. (1955) who also considered aero-thermodynamic
heating and advanced the concept of hypersonic boost-glide d⌽ V cos ␥ cos ␺
⫽ (3)
bomber. Li (2009) carried out trajectory optimization of a dt r
boost-glide hypersonic missile waverider configuration by
using the shooting technique and calculated the footprint of an dV T D
⫽ cos ␣ ⫺ ⫺ g sin ␥ (4)
HBG missile, boosted from a Minuteman III launch vehicle. dt m m
For most of these applications, reentry speeds greater than 6
km/s and reentry angle less than 5 degree have been
considered. Scant attention has been paid to the high
d␥
dt

1 T
V m 关共
sin ␣ ⫹
L
m 兲
cos ␴ ⫺ g ⫺
V2
r 共
cos ␥ 兲 兴 (5)
lift-to-drag configurations for medium and intermediate range
applications, i.e. downrange between 1,000 and 5,500 km
(NASIC, 2006). Moreover, the dependence of optimal
d␺
dt

1 T
V m 关共
sin ␣ ⫹
L sin ␴
m cos ␥
⫹兲V2 cos ␥ sin ␺ tan ␾
r
(6) 兴
burnout angle on the heat rate and integrated heat load bound
has not been shown in the previous studies. dm ⫺T
⫽ (7)
Reentry vehicles experience very high heating rates on dt Isp.g0
atmospheric reentry which is strongly dependent on burnout
conditions. Higher burnout angles result in high decent rates Where, r is the radial distance from the center of the earth to the
upon reentry, and the vehicle quickly approaches the heat rate reentry vehicle. The terms, ⍜ and ␸ are the longitude and
boundary, resulting in non-feasible trajectories because of latitude, respectively. The symbol V represents the total velocity
violation of heat rate bound. On the other hand, shallow of the vehicle, and ␥ and ␺ denote the flight path angle and the
burnout angles result in higher flight time within the azimuth angle, respectively. The terms, L and D are lift and drag
atmosphere, resulting in higher integrated heat load (Rizvi and are defined as L ⫽ 0.5CL␳V2S, and D ⫽ 0.5CD␳V2S,

68
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

respectively. The terms, ␣ and ␴ denote the angle-of-attack and The states at the end of the second and third phase, which
bank angle control, respectively. The angle-of-attack control include the burnout angle, burnout altitude, and the burnout
modulates the lift and drag forces which results in change in the speed are treated as free parameters and can be optimized.
velocity and the flight path angle. The bank angle control is used The fourth phase includes the free flight stage, as well as the
to tilt the lift vector which results in change in the heading angle. glide/re-entry stage (Figure 1).
The bank angle remains zero degree through out the flight for
optimal downrange performance. The term, “m”, represents Aerodynamic model
mass which reduces as fuel is consumed to produce thrust (T) The aerodynamic model for the waverider configuration is
during the boost flight. obtained using the experimental data by Crockrell et al.
(1995). The lift curve slope has been adjusted to that of a flat
Ballistic vehicle and reentry vehicle data plate at hypersonic speeds obtained using Newtonian flow
The vehicle data of the launch and the waverider vehicle data, as theory (Anderson, 1989). This is done, as the CD and
well as the physical constants, are summarized in Tables I and II, lift-to-drag ratio data provided by Crockrell et al. (1995) is
respectively. The boost vehicle is a two-stage vehicle with a valid only till Mach 4.5 and is provided as a function of CL till
burnout speed in vicinity of 3.7 km/s for a payload mass of 1,200 kg. a value of 0.25 only. It has been assumed that the CD trend
The burnout speed increases to 4.3 km/s for a payload mass of remains the same till a trim CL value of 0.7 which corresponds
900 kg. The burn time of each rocket motor is about 80 s. The to angle-of-attack of 50 degree. The assumption is valid in
boost vehicle data does not resemble any particular launch light of the Newtonian theory (Anderson, 1989). The
vehicle and is not an optimized vehicle for the purpose. aerodynamic model is assumed to be a function of the
angle-of-attack only. This is true in the hypersonic region in
Trajectory sequence which the aerodynamic coefficients do not vary with Mach
The optimization problem has been divided into four phases number (Bertin, 1994; Anderson, 1989).
as to handle discontinuities in control mode and mass at the The equations for a waverider configuration are:
end of each boost stage:
1 First 5 s of first-stage boost phase, during which pitch CL ⫽ ⫺0.04 ⫹ 0.8␣ (8)
maneuver does not take place.
2 The time duration after t ⫽ 5s till the burnout of the CD ⫽ 0.012 ⫺ 0.01␣ ⫹ 0.6␣2 (9)
first-stage rocket motor. It is during this phase that the
launch vehicle pitches down using angle-of-attack control. The aerodynamic model for the boost phase was used as
3 The second-stage boost phase during which the flight path provided by He (2002) (Figure 2). The lift and drag
angle changes to meet the burnout conditions. coefficients are a function of angle-of-attack, as well as the
4 The free flight and the reentry stage during which the Mach number during the boost phase as shown in Tables III
glide-vehicle is steered to optimal downrange with the and IV. The coefficients are non-dimensionalized with respect
help of angle-of-attack control. to the maximum frontal area of the launch vehicle.
Table I Launch vehicle data
Earth and the atmosphere
Quantity Launcher The earth is assumed to be a perfect, non-rotating sphere. The
Body diameter, [m] 1.4 acceleration due to gravity is given by Newton’s inverse square
Propulsion Solid law.
Booster 1 ␮
Propellant mass, [kg] 13,700 g⫽ (10)
r2
Empty mass, [kg] 2,600
Thrust, [N] 398,000
Isp, [s] 237 Figure 1 Schematics for boost-glide missile trajectory design
problem
Booster 2
Propellant mass, [kg] 3,250 Free Flight

Empty mass, [kg] 650


Phase 3:
Thrust, [N] 97,000 2nd Stage Phase 4:
boost Free Flight + Reentry/ Glide + Terminal flight
Isp, [s] 250
Reentry/ Glide

Phase 2:
Table II Reentry vehicle data and physical constants 1st Stage boost
with change in
pitch angle
Configuration Quantity Numerical values
Terminal
Flight
Waverider Mass, [kg] 1200/900 Phase 1:

W/S, [kg/m2] 400 Vertical


ascent
Terminal Conditions:
Velocity: 0.72 km/s
RN, [m] 0.006 Flight Path Angle: − 80 deg

C, [kg0.5m1.5/s3] 1.1813 ⫻ 10-3 Earth Curvature

69
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

Figure 2 Lift-to-drag ratio variation with angle-of-attack for a Where, t0 and tf are the initial and final times in the fourth and
waverider configuration and comparison with experimental data of final flight phase, respectively.
reference Crockrell et al. (1995)
Normal acceleration
1
CL Aeromodel For a lifting vehicle, the normal acceleration is of greater
CD Aeromodel importance as compared to the total deceleration load. This
0.5
is because a lifting reentry vehicle typically reenters the
CL, CD

CD Data
atmosphere at shallow reentry angles and at high
0
angle-of-attack to reduce the descent rate. This results in a
higher normal-force as compared to the axial force during
−0.5
−10 0 10 20 30 40 50 the reentry, as well as the glide phase. The normal
Angle-of-Attack (deg) acceleration is computed using the relation given by
5 equation (13).

L cos ␣ ⫹ D sin ␣
0 nz ⫽ (13)
mg
L/D

-5 L/D Aeromodel
Controls
L/D Data
The waverider vehicle has angle-of-attack and bank angle
−10 control. The vehicles are assumed to have angle-of-attack
−10 0 10 20 30 40 50
Angle-of-Attack (deg) trim capability up from ⫺10 to ⫹50 degree. The
angle-of-attack control modulates the lift as well as the heat
rate. The bank angle control remains zero for optimal
Table III CL␣ variation with Mach number for a launch vehicle
downrange performance.
CL␣ Mach number range
2.8 0 ⱕ M ⱕ 0.25 Boundary conditions and path constraints
2.8 ⴙ 0.447 (Mⴚ0.25) 0.25 ⱕ M ⱕ 1.1 Initial conditions
3.18 ⴚ 0.66 (Mⴚ1.1) 1.1 ⱕ M ⱕ 1.6 The initial conditions correspond to the launch vehicle at
2.85 ⴙ 0.35 (Mⴚ1.6) 1.6 ⱕ M ⱕ 3.6 the launch pad, implies:
3.55 M ⱖ 3.6
r(i) ⫽ 6378 km; ⌰(i) ⫽ 0; ␾(i) ⫽ 0.0;V(i) ⫽ 0.0m/s; ␥(i)
⫽ 90 deg; ␺ ⫽ 90 deg; m ⫽ m0 (14)
Table IV CD variation with mach number for a launch vehicle
Final conditions
CD Mach number range For the current study, the maximum terminal speed is
0.29 0 ⱕ M ⱕ 0.8 considered to be 720 m/s at an impact angle of ⫺80 degree
Mⴚ0.51 0.8 ⱕ M ⱕ 1.068 for the glide vehicle. The requirement has been modeled as
0.091 ⴙ 0.5/M M ⱖ 1.068 a terminal boundary condition.

r(f) ⫽ 6378 km, V(f) ⫽ 720 m/s; ␥(f) ⫽ ⫺80 deg (15)

The atmosphere is modeled using 1976 ICAO atmospheric Dynamic pressure constraint
model. A dynamic pressure limit corresponds to the maximum
stagnation pressure which the vehicle structure can bear in
Stagnation point heat rate the presence of high acceleration loads. The terminal
The stagnation point heating rate is modeled by using the constraints imply that the vehicle shall sustain a dynamic
equation given by Scott et al. (1985). The convective heat pressure of approximately 320 KPa toward the end of the
transfer rate is given by equation (11). mission. The dynamic pressure constraint of 320 KPa has
been imposed as a path constraint during the entire flight.
Q= ⫽ C␳0.5V3.05 (11) The maximum dynamic pressure is kept equivalent to the
dynamic pressure experienced by the vehicle in the terminal
The heat rate, Q⬘ is directly proportion to the square root of phase.
density, ␳, and the cube of velocity, V. C is the proportionality
1 2
constant, which is dependent on the nose radius, RN, and the q⫽ ␳V ⬍ 320, 000 Pa (16)
catalytic behavior of the material and is defined in Table II. 2
The total heat load at the stagnation point is calculated using Heat rate constraint
the relationship below: The heat rate constraint is directly linked to the surface
tf temperature. The surface is assumed to be in radiative


equilibrium with the surroundings.
Q⫽ Q⬘dt (12)
t0 Q̇ ⫽ ␦␧th(T4g ⫺ T4w) (17)

70
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

If Tg ⬍⬍ Tw, it implies: Method used


A number of methods exist for solving a trajectory
冉 冊
1
Q̇ 4 optimization problem:
Tw ⫽ (18)
␴␧TH ● optimal control theory;
● shooting techniques; and
Where, ␦ is the Stefan – Boltzmann constant, which is equal to ● heuristic algorithms.
5.67 ⫻ 10⫺8 W/(m2.K4), and  is the surface emissivity which
The advantages and shortcomings of various techniques as
is generally close to 0.8.
applied to trajectory optimization techniques are discussed in
The Carbon – Carbon Composite (2014) material can
reference (Betts, 1998).
retain its properties till a temperature of 2,800 or 3,073°K
Heuristic techniques are simple to implement and are
(Darby and Rao, 2011). The temperature corresponds to the
widely used for problems where the variables are discrete.
heat rate limit of approximately 4.0 MW/m2 which can be
However, in case of trajectory optimization, the control and
computed using equation (19). Maximum temperature values
the state variables are not discrete but are continuous in time
corresponding to different heat rate values are tabulated in
domain, and therefore heuristic optimization is not suitable for
Table V.
this kind of problem. The number of trajectory points is very
large. In such a case, heuristic algorithms would take a very
Q̇ ⱕ Q̇max (19) large convergence times and the convergence would also not
be guaranteed (Betts, 1998).
The heat rate, Q̇ is directly proportion to the square root of Shooting techniques have also been very popular during the
density, ␳, and the cube of velocity, V. C is the proportionality past two decades to solve the nonlinear, constraint, trajectory
constant, which is dependent on the nose radius, RN, and the optimization problem. In case of direct shooting methods, the
catalytic behavior of the material and is defined in Table II. control vector is optimized by means of a sequential quadratic
programming (SQP) Algorithm. The objective function and
Objective function the constraints are evaluated by integrating the equations of
motion in time. The final outcome of shooting algorithm is
The objective is to find the angle-of-attack control deflections
sensitive to the initial guess. The control vector is also discrete,
that would maximize the downrange of the reentry vehicle
and the intermediate control deflections are computed
subject to dynamic pressure and heat rate/integrated heat load
through interpolation. It is because of this reason that the
bounds. Maximizing the downrange is similar to maximizing shooting techniques are only successful once the dynamics is
the longitude at the final time tf of Phase 4. This predicament slow, and the initial guess lies in the vicinity of the optimal
is a Mayer’s problem and can be expressed as: solution. Shooting techniques have been successfully used in
reentry plane change problems by Gogu et al. (2009), Hull
J ⫽ min[⫺xf(2)]4 (20)
and Speyer (1982), Zimmermann and Calise (1998). Li
(2009), Li et al. (2009) also used a shooting technique to solve
Where, min [⫺xf(2)]4 represents the maximum value of
coupled launch and reentry trajectory optimization problem
second state (i.e. longitude) at the final time tf of the fourth
with heat rate constraint.
phase. In pseudospectral method, the time interval of the optimal
It may be noted, that the objective function has only been control problem is split into a prescribed number of
optimized for the fourth and final phase. The boost phase subintervals. The end points are called nodes, and the entire
optimization is indirectly catered for, by optimization of the distribution of nodes can be referred to as a mesh within the
burnout parameters, which results in range maximization of time domain. The points within the nodes are called
the fourth phase. Moreover, the percentage of range covered collocation points and are governed by strict quadrature rules.
during the first three phases is approximately 10 per cent of The method is based on the theory of orthogonal collocation
the overall range; therefore, the results presented in the where the collocation points are the Lagrange – Gauss points.
analysis can be considered equivalent to overall range By using the properties of the Lagrange polynomials, the state
optimization. equations and the control constraints are transformed into
algebraic equations. In this way, the optimal control problem
Table V Temperature values corresponding to heat rate limits
is reduced to a non-linear programming problem.
In h method, low-order polynomials are used to
Q̇max Tmax approximate the state vector. The mesh is refined till the
(MW/m ) 2
(°C) required accuracy of the solution is obtained. In p method, a
3.5 2,690 single interval is used, and the accuracy is improved by
3.0 2,580 increasing the order p of the polynomial. In an hp adaptive
2.5 2,450 method, the required accuracy is achieved by increasing the
2.0 2,305 number of nodes, as well as the degree of polynomial.
1.5 2,125 Additional nodes are introduced where the curvature is sharp
1.0 1,895 and therefore the method is called the hp-adaptive
pseudospectral method (Darby and Rao, 2011). The

71
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

hp-adaptive pseudospectral method is implemented in Gauss Configuration is boosted to a burnout speed of approximately
Pseudospectral Optimization Software, GPOPS (Rao, 2011). 3.7 km/s while configuration is boosted to a speed of 4.3 km/s.
Pseudospectral methods have been extensively used in a The table provides the optimal values of the objective
variety of problems in trajectory optimization. Examples range function, the corresponding optimal burnout conditions, as
from low-thrust orbit transfer (Darby and Rao, 2011), pick well the time-of-flight, the maximum normal acceleration, and
and place maneuver of robots (Strizzi et al., 2002), solar sail the total heat load values corresponding to each trajectory.
trajectory optimization (Stevens and Ross, 2003), launch The table show that as the heat rate constraint is reduced, the
vehicle ascent guidance (Rea, 2003), reentry trajectory design down performance also decreases. The reduction in heat rate
(Josselyn and Ross, 2003), etc. is also accompanied by reduction in integrated heat load.
The optimal control combined ascent and reentry problem The per cent loss in the downrange with the reduction in the
for unconstraint heat load is expressed as to determine that heat rate limit is shown in Figure 3(a). The figure shows that
state vector 关r共t兲, ⌰共t兲, ⌽共t兲, V共t兲, ␥共t兲, ␺共t兲, m共t兲兴 and the near optimal performance with less than a 5 per cent loss in
control vector 关␣共t兲, ␴共t兲兴 that minimizes the cost function downrange can be obtained for both waverider configurations
subject to the dynamic constraints of all variables given in state with heat limit of 3 MW/m2. Heat rate bound of 2.5 MW/m2
vector, the initial and terminal flight constraints given in also results in acceptable performance with approximately 7
equations (14) and (15), respectively, and the path constraints per cent loss in downrange as compared to optimal
given by equations (16) and (19). The state vector changes to performance corresponding to heat rate bound of 4 MW/m2.
关r共t兲, ⌰共t兲, ⌽共t兲, V共t兲, ␥共t兲, ␺共t兲, m共t兲, Q共t兲兴 for the constraint This corresponds to the material temperature limit of 2,450°C
heat load problem, subject to an additional dynamic heat load as compared to that of 2,800°C that exists for state-of-art C-C
constraint, equation (11). The nonlinear optimal control material (Table III). The per cent reduction in the integrated
problem is solved using open-source GPOPS version 4.0. The heat load with reduction in heat rate bound is shown in
program utilizes hp-adaptive Radau Pseudospectral Method Figure 3(b). The graph shows that for the heat rate limit of
(RPM) for solving the optimal control problem. The RPM is 2.5 MW/m2, per cent decrease in the integrated heat load is
an orthogonal collocation method where the collocation only 6 per cent for both configurations. This can be
points are the Legendre – Gauss – Radau points. The number understood by observing the heat rate profile shown in
of mesh intervals and the number of collocation points are
Figures 4(c) and 5(c). The diagram shows the entire heat rate
determined iteratively. The error is computed by determining
profile, depicting exposure times to maximum heat
the accuracy by which the differential equations and the path
rate/temperature to which a material would be exposed at the
constraints are satisfied between the collocation points. The
nose stagnation point during such a flight. The difference in
mesh refinement continues till a solution is obtained which
the integrated heat load values for 4 and 3 MW/m2 stagnation
satisfies the error tolerance (Darby and Rao, 2011). The detail
point heat rate bound trajectories is low because, in the former
of the hp-adaptive algorithm is given by Garg et al. (2010). For
case, the maximum heat rate is encountered for a very small
the current research, the maximum number of mesh iterations
period of time. The optimal trajectory path at higher heat rate
was set to 12 with 4-10 nodes between the collocation points.
bound has a skipping motion, and the vehicle spends some
The error tolerance was set to 1e-3 which is also the default
time at higher altitudes, resulting in lesser convective heat
value. The non-linear programming derivatives were
transfer and therefore lower integrated heat load.
computed using a finite difference scheme.
Figures 4(a) and 5(a) shows the trajectory shapes with
different heat rate constraints on altitude vs velocity diagram.
Results The figures show that the constraint heat rate and the
Numerical results for heat rate constraint problem constraint dynamic pressure bound are not violated in all
Numerical results for the two boost-glide waverider cases. The figures also show that as the heat rate bound is
configurations with varying heat rate limits and the objective reduced, the burnout altitude reduces because of lowering of
to maximize the downrange are shown in Table VI. burnout angle. The optimum burnout angle for maximum

Table VI Numerical results for waverider performance with heat constraint


Q⬘max Q Down range Cross range
Configuration [MW/m2] [GJ/m2] [km] [km] ␥b(f)° Vb(f) [km] hb(f) [km] nz max [g] Time [min]
(a) Launch vehicle with 4.0 1.864 3,321 0 5.2 3.7 65.4 3.8 26.30
1,200 kg payload 3.5 1.886 3,268 0 3.9 3.7 61.6 3.2 26.02
3.0 1.913 3,188 0 1.6 3.7 55.7 2.3 25.73
2.5 1.758 3,096 0 2.2 3.7 55.7 1.8 25.10
2.0 1.517 2,956 0 2.2 3.7 54.1 1.3 24.55
1.5 1.158 2,647 0 2.6 3.7 54.0 1.3 22.96
(b) Launch vehicle with 4.0 2.325 4,517 0 4.8 4.3 68.0 3.8 31.1
900 kg payload 3.5 2.348 4,387 0 3.6 4.3 64.0 1.9 30.4
3.0 2.330 4,337 0 2.7 4.3 61.0 1.8 29.6
2.5 2.190 4,248 0 2.1 4.3 57.9 1.2 28.9
2.0 1.880 4,015 0 2.2 4.3 57.8 1.3 29.0

72
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

Figure 3 (a) Percent loss in downrange with reduction in heat rate Figure 5 Optimal trajectories for the boost-glide waverider
bound; (b) Percent reduction in integrated heat load with reduction vehicle with heat rate constraints and objective to maximize the
in heat rate bound downrange at burnout speed of 4.3 km/s
25 45 80 80
Vb = 3.7 km/s
40
Vb = 4.3 km/s, 60 60
Glide flight

Altitude (km)

Altitude (km)
20 35
40 40
30

percent heat load reduction


percent range reduction

,000
Heat rate bound 3,20

00
20 20 Boost flight

,0
15 25

35

2
4
Dynamic pressure bound
20 0 0
0 1 2 3 4 0 1,000 2,000 3,000 4,000
Velocity (km/s) Range (km)
10 15
(a) (b)
10
5

35 Q’ max = 4.0 MW/m2


5 5
4 Q’ max = 3.0 MW/m2
30

Angle-of-Attack (deg)
Heat Rate(MW/m 2)
0 25 Q’ max = 2.0 MW/m2
3
20
0 −5
1.5 2 2.5 3 3.5 4 1.5 2 2.5 3 3.5 4 2 15
2
Q’ (MW/m ) Q’ (MW/m 2) 10
1
5
(a) (b)
0
0
0 500 1,000 1,500 2,000 0 500 1,000 1,500
Time (s) Time (s)

Figure 4 Optimal trajectories for the boost-glide waverider (c) (d)


vehicle with heat rate constraints and objective to maximize the Notes: (a) Altitude vs velocity plot; (b) altitude vs range; (c) heat
downrange at burnout speed of 3.7 km/s rate vs time; (d) angle-of attack vs time
70

60 60 constant dynamic pressure line. For heat bounds of 3 MW/m2


50 50 Glide and lower, the vehicle first follows the heat rate bound followed
Altitude (km)
Altitude (km)

40 40 by the optimum dynamic pressure value of approximately 35


KPa, corresponding to trim CL at maximum lift-to-drag ratio
2

30 30
4

value and wing loading. A low value of heat rate constraint forces
0

20 20
00
35

Boost flight
10 Heat rate bound 10 the vehicle to maintain an equilibrium glide trajectory along the
0
Dynamic pressure bound 0 heat rate bound, resulting in a steeper decent angle as shown in
0 1 2 3 4 0 1,000 2,000 3,000
Velocity (km/s) Range (km) Figures 4(b) and 5(b) on an altitude vs range plot. Low dynamic
(a) (b) pressure at the burnout point causes the vehicle to trim at a high
40
angle-of-attack (Figures 4(d) and 5(d)) to arrest the descent rate
4 Q’ max = 4.0 MW/m2 and prevent the vehicle from violating the heat rate bound.
Higher angle-of-attack results in loss of energy to drag and, as a
Angle-of-Attack (deg)

30 Q’ max = 3.0 MW/m2


Heat Rate(MW/m 2)

3
Q’ max = 2.0 MW/m2 consequence, lower downrange performance. The
20
2
angle-of-attack demand reduces with an increase in dynamic
10 pressure as the vehicle descends along the constant heat-rate line.
1 The vehicle then flies at a constant dynamic pressure profile and
0
0
maintains an optimal angle-of-attack of approximately 10
0 500 1,000 1,500 500 1,000 1,500 degrees, corresponding to the maximum lift-to-drag ratio, to
Time (s) Time (s)
obtain maximum downrange before diving down to meet
(c) (d) terminal velocity and flight path constraints.
Notes: (a) Altitude vs velocity plot; (b) altitude vs range; (c) heat
rate vs time; (d) angle-of attack vs time Numerical results for the heat load constraint problem
Table VII provides the end trajectory optimization results for
downrange performance, corresponding to 4 MW/m2 bound is waverider configuration with integrated heat load constraint. The
approximately 5 degree which reduces to 2.2 degree for 2 MW/ deterioration of waverider downrange performance with heat
m2 heat rate bound (Table VI) for both configurations. The load is portrayed in Figure 6, which shows that for burnout speed
shallow entry angle reduces the trim angle-of-attack requirement of 3.7 km/s, near optimal performance with approximately 6.5
at the initiation of glide phase and therefore minimizes the energy per cent loss in the overall range can be obtained with trajectory
lost to drag force, resulting in maximum downrange. designs that result in 50 per cent reduced integrated heat. For
Figures 4(a) and 5(a) further show that for maximum heat higher burnout speed of 4.3 km/s, the loss in downrange is 8 per
rate bound of 4 MW/m2 the optimal trajectory has an unsteady cent for trajectory design with 50 per cent reduced heat load as
skipping phase followed by as steady-state descent along the compared to the unconstraint heat load trajectory profile.

73
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

Table VII Numerical results for waverider performance with heat constraint
Q⬘max Q Down range Cross range
Configuration [MW/m2] [GJ/m2] [km] [km] ␥b(f)° Vb(f) [km] hb(f) [km] nz max [g] Time [min]
a
(a) Launch vehicle with 4.0 1.864 3,321 0 5.2 3.7 65.4 3.8 26.30
1,200 kg payload 4.0 1.5 3,281 0 9.1 3.7 76.6 6.8 26.12
4.0 1.25 3,210 0 11.2 3.7 82.8 6.5 25.81
4.0 1.1 3,135 0 10.9 3.7 82.0 7.5 25.46
4.0 0.93 3,110 0 15.1 3.7 92.8 8.6 25.35
4.0 0.5 2,676 0 16.4 3.7 96.8 10.0 22.57
(b) Launch vehicle with 4.0 2.325a 4,517 0 4.8 4.3 68.0 3.8 31.1
900 kg payload 4.0 1.860 4,406 0 6.38 4.3 72.8 3.4 30.5
4.0 1.534 4,270 0 7.1 4.3 75.7 3.9 30.06
4.0 1.162 4,167 0 12.5 4.3 90.0 7.0 29.6
4.0 0.767 3,708 0 11.5 4.3 88.0 7.0 27.2
Note: a unconstraint heat load results

Figure 6 Percent loss in downrange with heat rate for the Figure 7 Boost-glide waverider trajectory for different heat
boost-glide waverider vehicle load constraints; altitude vs velocity diagram (Q̇ ⬍ 4 MW/m2,
Vb ⬇ 3.7 km/s)
20
Vb = 3.7 km/s 100
18
Vb = 4.3 km/s, q (Pa)
90 2
16 Q` constraint (MW/m )
2
80 Q = 1.86 GJ/m (unconstraint)
14 2
percent range reduction

Q = 1.25 GJ/m
12 70 2
Q = 0.93 GJ/m
60
Altitude (km)

10

8 50
Boost Flight
6 40
000 4
4 30 35,

2 20
glide/ reentry flight ,000
0
,00

0
3,2 Heat rate constraint
35

0 10
4

0 10 20 30 40 50 60 70 80
percent heat load reduction 0
0 0.5 1 1.5 2 2.5 3 3.5 4
Velocity (km/s)

The detailed trajectory shapes, path variable and control


deflections are shown in Figures 7 – 11 for burnout speed of
3.7 km/s. Trajectory shapes, path variables and angle-of-attack
Figure 8 Boost-glide waverider trajectory for different heat load
control profile for burnout speed of 4.3 km/s are shown in
constraints; altitude vs range (Q̇ ⬍ 4 MW/m2, Vb ⬇ 3.7 km/s)
Figures 12 – 16. Figure 7 and Figure 12 show the optimum
trajectory shapes, for the two burnout speeds, on altitude vs
Q = 1.86 GJ/m2 (unconstraint)
velocity diagram. The figures show that in both cases, the
140 Q = 1.25 GJ/m2
optimal trajectories with unconstraint heat load are smooth
Q = 0.93 GJ/m2
and the waverider vehicle glides along a equilibrium glide path 120
along constant dynamic pressure line (Figures 9(b) and
14(b)). As the integrated heat load constraint is reduced, the 100
Altitude (km)

burnout altitude increases, and the glide trajectory profile


results in unsteady, high amplitude motion. The figures also 80

show that the heat rate and dynamic pressure constraint lines
60
are not violated. Figures 8 and 13 show that by reducing the
total heat load, the burnout angle increases, which also results 40
in an increase in the free-flight range and decrease in the
reentry range, as well as the flight duration of the reentry 20
phase, resulting in a lesser integrated heat load. The burnout
angle increases from an optimum value of 5.2 degree to a value 0 500 1,000 1,500 2,000 2,500 3,000
of 15.1 degree corresponding to a burnout speed of 3.7 km/s, Range (km)

74
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

Figure 9 Boost-glide waverider trajectory for different heat load Figure 11 Boost-glide waverider trajectory for different heat
constraints load constraints; angle-of-attack vs time (Q̇ ⬍ 4 MW/m2,
Vb ⬇ 3.7 km/s)
6
Q = 1.86 GJ/m2 (unconstraint)
Heat Rate(MW/m 2)

60
4 Q = 1.25 GJ/m2 Q = 1.86 GJ/m 2 (unconstraint)
Q = 0.93 GJ/m2 50 Q = 1.25 GJ/m 2
2 Q = 0.93 GJ/m 2

40

Angle-of-Attack (deg)
0
0 200 400 600 800 1,000 1,200 1,400 1,600
Time (s) 30
(a)
400 20
Dynamic Pressure (kPa)

300
10
200

100 0

0
0 200 400 600 800 1,000 1,200 1,400 1,600 −10
0 200 400 600 800 1,000 1,200 1,400 1,600
Time (s)
Time (s)
(b)
Notes: (a) Heat rate vs time; (b) dynamic pressure vs time ( < Figure 12 Boost-glide waverider trajectory for different heat
4 MW/m2, Vb ≈ 3.7 km/s) load constraints; altitude vs velocity diagram (Q̇ ⬍ 4 MW/m2,
Vb ⬇ 4.3 km/s)
Figure 10 Boost-glide waverider trajectory for different heat 100
load constraints; normal acceleration vs time (Q̇ ⬍ 4 MW/m2, q (Pa)
90 Q` constraint (MW/m2)
Vb ⬇ 3.7 km/s)
80 Q = 2.325 GJ/m2 (unconstraint)
10 Q = 1.534 GJ/m2
Q = 1.86 GJ/m2 (unconstraint) 70 Q = 1.162 GJ/m2
8
Q = 1.25 GJ/m2
60
Altitude (km)

Q = 0.93 GJ/m2
6
50
4
40
000
2
30 35,
nz (g)

4
,000
0 20 3,20
0

Heat rate bound


,00

00
0,0
35

−2 10
3,2 Dynamic Pressure bound
4

−4 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Velocity (km/s)
−6

−8
0 200 400 600 800 1,000 1,200 1,400 1,600 to the higher induced drag which the vehicle experiences
Time (s)
because of higher angle-of-attack requirement to execute a
high amplitude skipping motion. It can therefore be seen
once the heat load is constrainted by 50 per cent (0.93 GJ/m2).
from Figures 11 and 16 that the angle-of-attack increases to
For burnout speed of 4.3 km/s, the optimal burnout angle
as much as 20 degree during the midcourse phase to ensure
increases from 4.8 degree to a maximum of 12.5 degree once
the integrated heat load is constraint by 50 per cent (1.162 low heat rate trajectory, while maximizing the downrange.
GJ/m2). Lowering the heat load constraint, the vehicle The angle-of-attack trend also shows that at re-entry, the
executes a higher amplitude motion, spending more time in trim angle-of-attack requirement is higher for sharper entry
the low density region, resulting in lesser convective heat angles. High angle-of-attack is necessary to initiate the
transfer and, therefore, lesser integrated heat load. The skipping motion and to avoid the maximum heat rate
convective heat transfer rate is shown in Figures 9(a) and boundary. The high angle-of-attack at reentry also results in
14(a) as a function of time. a loss of kinetic energy to drag-force which results in a lower
With each skip, the kinetic energy of the waverider cumulative heat load and a lower range. The high
vehicle reduces, and the vehicle quickly approaches the angle-of-attack in combination with high dynamic pressure
terminal velocity bound. The loss of kinetic energy is linked at initial skipping altitude of approximately 40 km causes

75
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

Figure 13 Boost-glide waverider trajectory for different heat load Figure 15 Boost-glide waverider trajectory for different heat
constraints; altitude vs range (Q̇ ⬍ 4 MW/m2, Vb ⬇ 4.3 km/s) load constraints; normal acceleration vs time (Q̇ ⬍ 4 MW/m2,
Vb ⬇ 4.3 km/s)
160
Q = 2.325 GJ/m2 (unconstraint)
Q = 2.325 GJ/m2 (unconstraint)
140 Q = 1.534 GJ/m2
8
Q = 1.534 GJ/m2
Q = 1.162 GJ/m2
120 Q = 1.162 GJ/m2
6

100
Altitude (km)

nz (g)
80
2
60
0
40

−2
20

−4
0
0 500 1,000 1,500 2,000 2,500 3,000 3,500 4,000 4,500 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000
Range (km) Time(s)

Figure 14 Boost-glide waverider trajectory for different heat load Figure 16 Boost-glide waverider trajectory for different heat
constraints load constraints; angle-of-attack vs time (Q̇ ⬍ 4 MW/m2,
Vb ⬇ 4.3 km/s)
6
Q = 2.325 GJ/m2 (unconstraint)
Heat Rate(MW/m 2)

60
4 Q = 1.534 GJ/m2 Q = 2.325 GJ/m2 (unconstraint)
Q = 1.162 GJ/m2 50 Q = 1.534 GJ/m2

2 Q = 1.162 GJ/m2

40
Angle-of-Attack (deg)

0
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000
Time (s) 30

(a)
20
400
Dynamic Pressure (kPa)

300 10

200
0
100
−10
0 0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000
Time (s)
Time (s)

(b)
The normal acceleration trend can be seen in Figures 10 and
Notes: (a) Heat rate vs time; (b) dynamic pressure vs time ( <
15, which shows that for lower heat load trajectories, the
4 MW/m2, Vb ≈ 4.3 km/s)
vehicle experiences higher normal acceleration peaks,
resulting in more energy loss to drag and, ultimately, lower
range. The normal acceleration experienced by the vehicle
higher lift and therefore a normal acceleration which reaches approximately 8 g value for 50 per cent reduced
initiates and sustains the skipping motion. For total heat integrated heat constraint cases of 0.93 and 1.162 GJ/m2
load constraint of 0.93 GJ/m2 (50 per cent heat load) at
corresponding to burnout speed of 3.7 and 4.3 km/s,
burnout speed of 3.7 km/s, the angle-of-attack saturated to respectively. The normal acceleration peaks are followed by
50-degree value for duration of approximately 10 s, regions where the load factor reduces below one, or even
resulting in loss of speed and initiation of the skipping reduces to zero. Zero load-factor regions imply free-flight
motion. The angle-of-attack trim control requirement conditions. The plot also shows that the vehicle experiences
reduces to 45 degree for the case where the heat load is up to ⫺6 g load during the terminal flight phase.
constrained to 1.25 GJ/m2 (67 per cent heat load), whereas
for the unconstraint heat load case, the angle-of-attack trim
control requirement is approximately 22 degree. Similar Discussion
trend of angle of attack exits at burnout speed of 4.3 km/s For a burnout speed range from 3.7 to 4.3 km/s, and a waverider
and is shown in Figure 16. design-based configurations, the optimal burnout angle is

76
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

approximately 5 degrees corresponding to a 4 MW/m2, resulting which ensued lower flight time within the atmosphere along
in a downrange of 3,321 and 4,517 km, respectively. The with skipping motion, and therefore, lower integrated heat
stagnation point integrated heat load of the waverider loads. It has been shown that the integrated heat load can be
configuration is approximately 1.86 and 2.33 GJ/m2 at burnout reduced by 50 per cent with less than 10-per cent loss in the
speeds of 3.7 and 4.3 km/s, respectively, for unconstrained heat overall downrange of the vehicle in comparison to the
rate bound. unbounded heat load profile and a 4 MW/m2 heat rate bound
Trajectories with limits on heat rates show that heat rate for the launch and the waverider reentry vehicle under
bound of 2.5 MW/m2 results in trajectories with 6 per cent consideration.
loss in downrange, as compared to optimal trajectories,
corresponding to the stagnation point heat rate bound of References
4 MW/m2. The control strategy is also simple, i.e. an initial
Anderson, J.D. Jr. (1989), “Hypersonic and high temperature
heat rate control followed by a dynamic pressure control
gas dynamics”, Hypersonic and High Temperature Gas
which results in a steady optimal trajectory flight path. The
Dynamics, McGraw-Hill Book, New York, NY.
optimal burnout angle of approximately 5 degrees results in an
Bertin, J.J. (1994), “Hypersonic aerothermodynamics”,
endo-atmospheric trajectory. Constraining the heat rate
Hypersonic Aerothermodynamics, AIAA Education Series,
bound results in shallow reentry angles and, therefore, larger
Washington, DC.
flight time within the atmosphere. As a result, the integrated
Betts, J.T. (1998), “Survey of numerical methods for
heat load is not sensitive to reduction in heat rate bound.
trajectory optimization”, Journal of Guidance, Control, and
The optimal constrained heat load trajectories for waverider
Dynamics, Vol. 21 No. 2, pp. 193-207.
configuration resulted in trajectories with a higher burnout
Carbon-Carbon Composites (2014), “Carbon fibres, carbon-
angle up to a value of 16 degrees corresponding to burnout
polymer composites and carbon-carbon composites”,
speed of 3.7 km/s and, as a consequence, longer free-flight
available at: www.mse.mtu.edu/~drjohn/my4150/class14/class
range and smaller reentry range/flight time within the
14.html
atmosphere; high oscillatory angle-of-attack and, as a result,
Crockrell, C.E., Huebner, L.D. and Finley, D.B. (1995),
high amplitude oscillatory flight in the low density region of
“Aerodynamic performance and flow field characteristics of
the atmosphere and therefore lesser over all heat load. The
two wave-rider derived hypersonic cruise vehicles”, AIAA
higher angle-of-attack trim also resulted in loss of reentry
33rd Aerospace Science Meeting and Exhibit, AIAA Paper,
range because of loss of kinetic energy to drag force.
Reno, NV, pp. 95-0736.
Constraint heat load trajectory to the value of 0.93 GJ/m2 (or
Darby, C.L. and Rao, A.V (2011), “Minimum-fuel
50 per cent reduced heat load) at the nose stagnation point,
low-earth-orbit aeroassisted orbital transfer of small
corresponding to burnout speed of 3.7 km/s resulted in
spacecraft”, Journal of Spacecraft and Rockets, Vol. 48 No. 4.
optimal trajectory design with approximately 6.5 per cent loss
Eggers, A.J., Allen, J. and Neice, S. (1955), “A comparitive
in total downrange. The optimal burnout angle in this case is
analysis of the long range hyper velocity vehicles”, NACA
15.1 degree for a burnout speed of approximately 3.7 km/s.
Research Memorendum, National Advisory Comittee for
Trajectory with 50 per cent reduced heat load at 4.3 km/s
Aeronautics, Washington DC.
resulted in approximately 8 per cent loss in the maximum
Garg, D., Patterson, M.A., Hager, W.W., Rao, A.V.,
downrange at an optimal burnout angle of 12.5 degree. The
Benson, D.A. and Huntington, G.T. (2010), “A United
normal acceleration increases to approximately 8-g value as
framework for the numerical solution of optimal control
compared to a 4-g value for nominal trajectory, but stays
problems using pseudospectral methods”, Automatica,
within 15 g theoretical limit generally considered acceptable
Vol. 46 No. 11, pp. 1843-1851.
for missiles. It was observed that tighter bounds on heat rates
Gogu, C., Matsumura, T., Haftka, R.T. and Rao, A.V.
and integrated heat loads resulted in higher angle-of-attack
(2009), “Aeroassisted orbital transfer trajectory
control requirements.
optimization considering thermal protection system mass”,
Journal of Guidance Control and Dynamics, Vol. 32 No. 2a.
Conclusion He, L. (2002), Solid Ballistic Missile Design, BUAA Press.
The problem of optimal trajectory design for a boost-glide Hull, D.G. and Speyer, J.L. (1982), “Optimal reentry and
vehicle has been considered for burnout speed range of plane change trajectories”, The Journal of Astronautical
3.7-4.3 km/s with heat rate, and integrated heat load Sciences, Vol. 30 No. 2, pp. 117-130.
constraint. Trajectory design was optimized by numerically Josselyn, S. and Ross, I.M. (2003), “A rapid verification
solving a multiphase non-linear optimal control problem using method for the trajectory optimization of reentry vehicles”,
hp-adaptive pseudospectral method with the objective to Journal of Guidance, Control, and Dynamics, Vol. 26 No. 3,
compute the best burnout conditions, as well as optimal pp. 505-508.
control deflections during the reentry/glide phase for Li, Y. (2009), “Optimal attack trajectory for hypersonic boost
maximum downrange performance. glide missile in maximum reachable domain”, IEEE
It has been shown that for the vehicles under consideration, International Conference on Mechatronics and Automation,
and constraint heat rate trajectories resulted in low burnout IEEE, Changchun.
angles within the range of 2.2-5.2 degree and therefore higher Li, Y., Cui, N. and Rang, S. (2009), “Trajectory optimization
integrated heat load because of endo-atmospheric trajectories. for hypersonic boost glide missile considering aeroheating”,
The constraint heat load trajectories, on the other hand, Aircraft Engineering and Aerospace Technology, Vol. 81,
resulted in higher burnout angles of as high as 16 degree, pp. 3-13.

77
Hypersonic boost-glide waverider Aircraft Engineering and Aerospace Technology: An International Journal
S. Tauqeer ul Islam Rizvi, He Linshu and Xu Dajun Volume 87 · Number 1 · 2015 · 67–78

NASIC (2006), “Ballistic and cruise missile threat”, United Whitmore, S.A. and Dunbar, B.J. (2003), “Orbital space
States Air Force. plane: past, present, and future”, IAA/ICAS International
Parish-II, M.S. (1995), Optimality of Aeroassisted Orbital Plane Air and Space Symposium and Exposition, Dayton, OH.
Changes, Naval Post graduate School Monterey, CA. Whitmore, S.A., Banks, D.W., Andersen, B.M. and
Rao, A.V. (2011), “User’s manual for GPOPS version 4.0”. Jolley, P.R. (2006), “Direct-entry, aerobraking, and lifting
Rea, J. (2003), “Launch vehicle trajectory optimization using aerocapture for human-rated lunar return vehicles”, 44th
a legendre pseudospectral method”, AIAA Paper AIAA Aerospace Science Meeting and Exhibit, AIAA-1033,
2003-5640. Reno, NV.
Rizvi, S., Islam, T., He, H. and Nasimullah, M. (2013), Zimmermann, F. and Calise, A.J. (1998), “Numerical
“Vehicle performance tradeoff study of small size lifting optimization study of aeroassisted orbital transfer”, Journal
re-entry vehicle”, 10th International Bhurban Conference on
of Guidance Control and Dynamics, Vol. 21 No. 1,
Applied Sciences & Technology, National Centre for Physics,
pp. 127-133.
Quaid-i-Azam University, Islamabad.
Sanger, E. and Bred, J. (1944), “A rocket drive for long range
bombers”, Tech Information Branch, Navy Department.
Scott, C.D., Ried, R.C., Maraia, R.J., Li, C.P. and Further reading
Derry, S.M. (1985). “An AOTV aeroheating and thermal
Falcon Hypersonic Technology Vehicle (HTV-2) (2011),
protection study”, in Nelson, H.F. (Ed.), Thermal Design of
“Falcon Hypersonic Technology Vehicle (HTV-2)”,
Aeroassisted Orbital Transfer Vehicle, Progress in Astronautics
and Aeronautics, New York, p. 96. available at: www.globalsecurity.org/space/systems/x-41-
Stevens, R. and Ross, I.M. (2003), “Preliminary design of htv-2.html
earth-mars cyclers using solar sails”, American
Astronautical Society, AAS Paper 03-574.
Corresponding author
Strizzi, J., Ross, I.M. and Fahroo, F. (2002), “Towards
real-time computation of optimal controls for nonlinear S. Tauqeer ul Islam Rizvi can be contacted at:
systems”, AIAA Paper 2002-4945. rizvi.aeng@gmail.com

For instructions on how to order reprints of this article, please visit our website:
www.emeraldgrouppublishing.com/licensing/reprints.htm
Or contact us for further details: permissions@emeraldinsight.com

78

View publication stats

You might also like