You are on page 1of 10

JOURNAL OF SPACECRAFT AND ROCKETS

Vol. 41, No. 5, September–October 2004

Optimal Conceptual Design of Two-Stage Reusable Rocket


Vehicles Including Trajectory Optimization
Takeshi Tsuchiya∗
University of Tokyo, Tokyo 113-8656, Japan
and
Takashige Mori†
Japan Aerospace Exploration Agency, Tokyo 182-8522, Japan

Many candidate concepts of reusable space transportation vehicles have been proposed around the world. This
paper applies an optimization method for conceptual designs of winged fully reusable rocket two-stage-to-orbit
(TSTO) vehicles in both horizontal takeoff and vertical launch styles. We first describe our methods for analyzing
vehicle design. Then we integrate those methods into the optimization problem, the solution of which yields the
minimized parameter, defined as the total gross weight of the first-stage booster and the second-stage orbiter. This
information allows us to determine the optimal vehicle configuration and flight trajectory for a highly feasible
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

TSTO vehicle. The optimal solutions show the necessity of lightening and miniaturizing components. Vertical
launch vehicles are lighter in total gross weight than horizontal takeoff vehicles. In addition to optimizing vehicle
configuration, this study also optimizes ascent and return trajectories. These optimizations enable the booster to
glide back to the launch site without propellant, despite the long downrange path from the staging point of the
ascent trajectory.

Introduction tion Investigation Programme4 and initiated the Future Launchers


Technologies Programme5 to identify a technically sound and cost-
T O support the advancement of space activities in the 21st cen-
tury, for example, scientific studies, commercial enterprises,
and manned missions, there is an urgent need to develop new types
effective reusable or semireusable launcher to succeed Ariane 5.
Conceptual designs of space launch vehicles cannot be conducted
of space transportation systems. Such systems must combine flexi- without the cooperation of numerous disciplines, including sev-
bility, reliability, low cost, and good operability for access to space, eral analysis technologies: aerodynamics analysis, heating analysis,
at levels that existing transports can hardly achieve. Reusable trans- structural analysis, propulsion analysis, trajectory analysis, controls
portation systems are identified as a promising technology. Indeed, analysis, cost analysis, operations analysis, and so on. Inasmuch as
many researchers recognize that a key to advancing space develop- the frontiers of the present technologies can make a reusable space
ment is the replacement of expendable rockets with reusable space transportation system feasible, it is necessary to make good use of the
transportation systems. analysis technologies in the various disciplines and to apply design
In Japan, a series of successes in launching the present mainstay optimization techniques. The application of optimization methods to
rocket, H-IIA, has inspired active public interest in next-generation multidisciplinary analysis is referred to as multidisciplinary design
technologies. Japan’s National Aerospace Laboratory has taken the optimization (MDO).6 Advances in computing have yielded many
lead nationally in the study of winged horizontal takeoff and landing practical MDO methods, which have been applied widely from com-
spaceplanes with new types of airbreathing engines running at hy- ponent design to overall design of aircraft and space transportation
personic speed.1 Various types of engines have also been studied.2 vehicle concepts.7 Design tools have been proposed, which has in-
Japan’s Institute of Space and Astronautical Science has flight tested tegrated computer software for preliminary designs of the space
a small vertical launch and landing test vehicle equipped with a transportation systems.8,9 Some researchers have made compara-
liquid-hydrogen rocket engine.3 In June 2000 the Committee on Fu- tive studies of the systems from their specialties, for example, Koelle
ture Space Transportation Systems established within Japan’s Sci- made a cost comparison of an SSTO vehicle concept with TSTO ve-
ence and Technology Agency, defined a new direction for Japan’s hicle concepts.10 Miele and Mancuso optimized ascent trajectories
space transportation systems in the 21st century: an incremental for single-stage-suborbit and TSTO rocket-powered vehicles. The
research and development program whose final goal is a single- maximum payload weight problem was studied for different val-
stage-to-orbit (SSTO) spaceplane, with an interim fully reusable ues of the engine specific impulse and spacecraft structural factor.11
two-stage-to-orbit (TSTO) spaceplane in a few decades.1 On the Mehta and Bowles made comparative studies of the TSTO vehicles
basis of that directive, Japan has begun conceptual studies of TSTO with the airbreathing engines on the first stage.12 The first author
systems. In other countries the United States is working on the Space in this paper also studied the simultaneous optimization of vehicle
Launch Initiative program to develop next-generation space launch- sizes and flight trajectories for SSTO spaceplanes.13
ers as a replacement for the space shuttle. The European Space The purpose of this paper is to integrate analysis models for the
Agency has established the Future European Space Transporta- conceptual design of winged TSTO vehicles with an optimization
method. The TSTO vehicles in this paper use only rocket propul-
Received 9 March 2003; revision received 26 August 2003; accepted for sion and are classified as either the horizontal takeoff and landing
publication 26 August 2003. Copyright  c 2003 by the American Institute of style or the vertical launch and horizontal landing style. These two
Aeronautics and Astronautics, Inc. All rights reserved. Copies of this paper styles of vehicle, with their different payload capabilities, are both
may be made for personal or internal use, on condition that the copier pay optimized in the same manner for the sake of comparison. Trajec-
the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rose- tory optimization is one of the characteristics of the optimization
wood Drive, Danvers, MA 01923; include the code 0022-4650/04 $10.00 in
correspondence with the CCC.
problems in this paper. Rather than basing ascent and return flight
∗ Lecturer, Department of Aeronautics and Astronautics; tsuchiya@ trajectories on supposed guidance laws, we optimize three-degree-
mail.ecc.u-tokyo.ac.jp. Member AIAA. of-freedom trajectories themselves, in addition to optimizing the
† Senior Researcher, Future Space Transportation Center, Chofu. Member other design parameters. The results show examples of the optimal
AIAA. winged fully reusable TSTO vehicles.
770
TSUCHIYA AND MORI 771

Analysis Method and Optimization Problem


Figure 1 shows the shape of the TSTO vehicle in this study. The
vehicle is made up of a first-stage booster and a second-stage or-
biter. This figure also indicates parameters of vehicle component
dimensions. Table 1 shows the design variables that we have to op-
timize: flight performance parameters and flight trajectory variables
as well as the geometric parameters shown in Fig. 1. One of the fea-
tures of this study is that we optimize ascent and return trajectories.
This is the reason why the design variables contain state and control
variables of the flight trajectories.
Figure 2 shows the overall analysis process. If we give the weight
and volume of payloads to/from orbit and a set of values for the de-
sign variables, each analysis is conducted sequentially, as exhibited
in Fig. 2. Analysis in this paper refers not only to the calculation
of output values for the following analysis based on the values of
design variables and the input values obtained from the preceding
analysis, but it also refers also to the computation of the function
values of equality and inequality constraint conditions (C E and C I in
Fig. 2), which the variables must satisfy. We have integrated analysis
tools in more than one specialized field for automatic computation.
The values we find for design variables must satisfy all of the con-
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

Fig. 2 Analysis process.


straint conditions. If we cannot obtain unique values, we should find
the best values under given criteria. Thus, we give a performance
index and define an optimization problem. The following summa- Vehicle Definition
rizes the analysis methods, constraint conditions, and performance Both the vertical launch and horizontal takeoff styles of TSTO
index in the optimization problem. vehicle are composed of an orbiter, which can reach orbit and reenter
the atmosphere, and a booster, which is a rocket booster with wings
for either vertical or horizontal takeoff and fly back to the launch site
Table 1 Design variables of vehicle configuration after the orbiter has separated from it. The orbiter is derived from the
and flight trajectory Space Shuttle Orbiter of the United States and HOPE-X14 of Japan.
The orbiter’s nose is blunted to a 0.5 m radius in order to withstand
Parameter type Variable
aerodynamic heating during reentry. The fuselages have nonintegral
Booster’s geometry (see Fig. 1) l1 , l2 , l3 , D, Cr , Ct , b tanks for the cryogenic propellant: liquid oxygen (LOX) and liquid
Orbiter’s geometry (see Fig. 1) l0 , l1 , l2 , l3 , D, Cr , Ct , b hydrogen (LH2). Each tank is composed of a cylinder with dome-
Flight performance of booster Gross weight shaped ends. Between the tanks of the orbiter is a cubic payload bay
Maximum body normal acceleration area. The payload weight and volume requirements are specified in
Maximum body axial acceleration
the problem definition before the optimization is computed. Both the
Maximum dynamic pressure
Thrust (vacuum) orbiter and booster have rocket engines installed in the aft fuselage.
Flight performance of orbiter Gross weight Though the number of rocket engines should be one of the design
Maximum body normal acceleration variables, the optimization method in this paper cannot optimize
Maximum body axial acceleration the discrete number. Thus, we assign a minimum number on the
Maximum dynamic pressure condition that the thrust force per engine is less than 1.5 MN because
Thrust (vacuum) it is presumed in the weight analysis that a small number of engines
Flight trajectory Discretized state variables reduces the total engine weight. The wing sections of the booster
Discretized control variables and orbiter are NACA0005 and NACA0010, respectively. For lateral
Initial times of intervals
stability the orbiter has a vertical tail fin, but the booster has tip fins
Terminal times of intervals
instead because a tail fin would probably have little effect when
the orbiter and booster are attached to each other. Although the
vehicle component sizes are given by the design variables as shown
in Fig. 1, the sizes of the tail fin and tip fins are calculated from the
tail volume of HOPE-X. In addition, the orbiter has a body flap, the
effectiveness of which is equivalent to that of HOPE-X.
The vehicle definition has a surface mesh generator to make pan-
els for the surfaces of both the vehicles from the values of geometric
parameters in the design variables. The coordinate values of the pan-
els, as well as their attribution to either the fuselage or the wings, are
submitted to aerodynamics analysis. The vehicle element sizes that
are not included in the design variables are also calculated because
this information is needed for the weight analysis.
To attain the proper vehicle shape, the vehicle definition com-
putes inequality constraints for the values of geometric parameters.
We also constrain the flight performance parameters of the design
variables so that they are covered by the weight analysis. To be con-
crete, the maximum values of acceleration and dynamic pressure
are subject to the ranges of 2.5–4 G and 15–50 kPa, respectively.
At the same time the trajectory analysis gives the amount of pro-
pellant required by the booster and that required by the orbiter, and
the vehicle definition calculates the tank sizes accordingly. Some
constraint conditions demand that the booster fuselage should be
large enough to hold the tanks and the orbiter fuselage should hold
Fig. 1 Booster and orbiter layouts. the tanks and payload bay.
772 TSUCHIYA AND MORI

Aerodynamics Analysis Weight Analysis


The study involves estimating the vehicles’ aerodynamic proper- The weight analysis computes the component weights of the
ties in the different flowfield regimes that the craft encounter dur- booster and orbiter and their combined gross weight from the geo-
ing atmospheric flight, which ranges from subsonic to hypersonic metric and flight performance parameters in the design variables,
speeds. The aerodynamics analysis incorporates the surface panels the component sizes as given by the vehicle definition, and the
of each vehicle as provided by the vehicle definition and tabulates amount of the propellant as provided by the trajectory analysis. This
lift and drag coefficients for typical angles of attack and Mach num- study employs the Hypersonic Aerospace Sizing Analysis (HASA)
bers from subsonic to hypersonic speeds. The trajectory analysis program,18 a statistical weight estimation method. The HASA pro-
depends on the aerodynamics analysis, that is, the aerodynamics gram is improved in parts, as follows:
analysis sends to the trajectory analysis the results obtained from 1) The vertical launch boosters and the orbiters use their respective
interpolation of the values in the tables, and these results are used landing gear only for landing, not launching. The landing weight of
to compute the flight trajectories. Another important operation of a space transportation vehicle is generally much lighter than the
the aerodynamics analysis is to evaluate the longitudinal static sta- takeoff weight because the propellant weight is a high proportion
bilities from the pitching moments of the vehicles and find the aft of the takeoff weight. Thus, the weight of gear that is used only for
limits of the centers of gravity. We presume that the forward lim- landing is very different from that used for both takeoff and landing.
its, which result fundamentally from maneuverability, are locations We estimate the landing gear weight of the vertical launch booster
that are 10% of body length forward of the aft limits. The c.g. limits and that of the orbiter with the maximum landing weight, that is,
obtained are given to the weight analysis.
This study employs simple analysis methods. A panel method15 Wgear = 0.0101W01.124 (5)
in subsonic flight, the Prandtl–Meyer expansion flow theory, and
a tangent cone/wedge method in supersonic to hypersonic speed where Wgear (kg) is the landing gear weight. W0 (kg) is the total gross
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

are applied to each vehicle. These methods use the surface panels weight to compute the gear weight of the horizontal takeoff booster
from the vehicle definition. The methods, however, cannot estimate and, in the case of the vertical launch booster and the orbiter, it is
parasite drag. Especially, base drag accounts for a large part of the the maximum landing weight.
total drag because a vehicle with rocket engines has a cutoff aft 2) The orbiter has a reaction control system and orbital maneu-
fuselage. According to Ref. 16, we estimate the base drag of each vering engines (RCS/OME) and their propellants. The system has
vehicle from its base area and the Mach number. In addition, we thrusters with a specific impulse of 310 s and a velocity increment
compute the skin-friction drag by the sum of the frictions from the capability of 340 m/s in orbit. The propellant is supposed to weigh
flow velocities on the panels. Transonic aerodynamic forces result 70% of the total RCS/OME weight.
from subsonic and supersonic forces, as explained in Ref. 16. 3) This study estimates the weight of the thermal protection sys-
The simple methods in this study are applicable to each vehicle tem (TPS) more precisely than the HASA method. We presume that
separately and not to the coupled vehicle. We presume that the cou- the nose edge and wing leading edges of the orbiter are provided
pled vehicle’s lift force is equal to that of the booster and that its drag with C/C materials (area density 40 kg/m2 ), and, on the other sur-
force is the sum of the total drag of the booster and the parasite drag faces of the orbiter and the booster, thermal protection materials
of the orbiter. Thus, this study does not consider the interference equivalent to 30-mm-thick ceramic tiles are bonded.
drag caused by the coupling of the two vehicles. 4) The propellant tanks are made of aluminum alloy, and each is
The methods used in this study are simpler and lower fidelity than composed of a cylinder whose ends are dome shaped. The working
the latest sophisticated computation methods, but they are useful in pressure of the tank is given by the sum of the ullage pressure
that the aerodynamics analysis, by which the optimization process (3 kgf/cm2 ) plus the head pressure. We compute the tank weight
computes the aerodynamic characteristics repeatedly, needs only from the tank wall thickness, which is obtained considering the
a short computation time. These methods are applied widely to working pressure with the membrane theory. For manufacturability,
aerodynamic preliminary analyses of candidate space transportation however, the thinnest wall is determined to be 3 mm thick in this
vehicles.6,17 paper. In addition, the weight of the isogrid structure that is generally
patterned on the inner surface of the tank, plus the weight of spot
Propulsion Analysis welds, sensor attachments, and so on, require the preceding tank
Given a flight altitude by the trajectory analysis, the propul- weight to be multiplied by a constant number 1.4 in order to avoid
sion analysis calculates a thrust force and specific impulse in the underestimation.
following manner. The rocket engine uses the cryogenic propel- 5) An equation to obtain rocket engine weights is quoted from
lant (LOX/LH2). This study considers a conventional rocket engine Ref. 19. The following equation holds for the thrust force Tinf (N)
given by historical perspective and not a specific engine. in vacuum and the engine weight Weng (kg):
A thrust force Tinf (N) per engine in vacuum is one of the design
variables. According to existing or developed rocket engines, we Weng = 0.0119Tinf
0.856
(6)
presume that a cross-section area of the rocket nozzle exit Ae (m2 )
of the booster and orbiter is proportional to Tinf , as expressed here: A thrust-to-weight ratio of a rocket engine with a maximum thrust
force of 1.5 MN amounts to approximately 65 using the preceding
Ae = 2.771 × 10−6 Tinf (1) equation.
The estimation of the component weights needs the gross weight,
This means that each rocket engine has a fixed nozzle. The atmo- which is the sum total of the component weights. For example, the
spheric pressure p (Pa) changes the thrust force to TRE (N) in the landing gear weight, as part of the gross weight, is expressed as
air: an equation with a variable of the gross weight. Thus, the method
analyzes the weights using the gross weight in the design variables to
obtain the gross weight. We define an equality constraint condition
TRE = K Tinf (2)
that both the weights correspond exactly.
K = 1 − Ae ( p/Tinf ) (3) Knowledge of the layout of the components in the vehicle makes
it possible to compute the center of gravity. The flight situations
define the constraint conditions by which the c.g. locations remain
Simultaneously, a specific impulse ISP (s) in the atmosphere is pro- within the limits that the aerodynamics analysis determined. There
vided by the specific impulse ISPinf (s) in vacuum (fixed to be 445 s is also an equality constraint condition that the c.g. location of the
in this paper) and the preceding parameter K : booster does not move when the orbiter is released from the booster.
This implies that the only c.g. location determines the attachment
ISP = K ISPinf (4) location of the orbiter in this paper.
TSUCHIYA AND MORI 773

Trajectory Analysis station to monitor rockets launched from Japan. We also have a plan
This study implements a three-degree-of-freedom (3DOF) tra- to develop infrastructures for a future spaceport.
jectory analysis.20 State variables are altitude, longitude, latitude, This study considers two takeoff styles: the horizontal takeoff and
velocity, flight-path angle, heading angle, and mass. Control vari- the vertical launch. In the case of vertical launch, an initial condition
ables are angle of attack, bank angle, and thrust throttle. Generally, is that the flight-path angle is 90 deg. After the takeoff, having no
the trajectory analysis computes state variable profiles by integrating bank angle, the coupled vehicle rises and accelerates east by the
equations of motion with given control variable profiles and exam- power of the booster’s rocket engines; this paper does not employ
ines satisfaction ratings of constraint conditions during the flight. parallel burn by the orbiter or cross-feeding of the propellants.
One of the characteristics in this study is to optimize the flight tra- The horizontal takeoff vehicle has constraint conditions involving
jectory with the other design variables, as already described. The the angle of attack, throttle, dynamic pressure, thrust, body normal
trajectory analysis, however, must deal with dynamic variables de- acceleration, and body axial acceleration. The values for these are
pending on time, while the other design variables are static ones limited to the maximum values in the design variables, except for the
independent of time. Therefore, we discretize the state and control angle of attack and the throttle, which are below 20 deg and above
variables to transform them into static variables like the other de- 60%, respectively. The vertical launch vehicle has the same set of
sign variables.21 In the same manner we also discretize constraint constraint conditions, plus the constraint condition that the flight-
conditions and the motion equations expressed as differential equa- path angle must decrease steadily from the initial 90 deg. This is
tions, which are functions of the state and control variables. The because a free flight-path angle results in the optimal solution, in
number of discrete elements is fixed at 200 in total. Nevertheless, which the vehicle descends with the flight-path angle negative im-
if we were to discretize all 10 state and control variables with the mediately after liftoff to enhance acceleration just above the ground
200 elements, the more than 2000 resultant design variables would and then gains altitude by the lift force from its wings. This is an
make for a very heavy computational burden in the optimization. unrealistic flight path with respect to safety, though it has the advan-
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

Therefore, we classify the state variables into those of position and tage that the vehicle can accelerate even if powered by small rocket
lateral motion (longitude, latitude, and heading angle) and others engines. From this we constrain the flight-path angle of the vertical
(altitude, velocity, flight-path angle, and mass). The structures of launch vehicle to decrease monotonously.
the motion equations exhibit strong interaction among the latter and
a weak influence of variation of the former on the latter. It follows Orbiter Ascent Phase
that we discretize only the latter four of the seven state variables and In this phase the released orbiter accelerates by its rocket engines
the three control variables. When the values of the former three state and climbs to an altitude of 120 km. At that altitude the orbiter
variables are required, they are computed by numerical integration cuts off the rocket engines temporarily and zooms up with no thrust
of their motion equations. in an elliptical orbit whose apogee altitude is 200 km. Finally, an
The flight styles of the vehicles change the equations of motion. apogee kick puts it into a 200-km circular low Earth orbit (LEO).
This divides the flight trajectory into four phases, as described next. To estimate propellant consumption, we actually compute the flight
In addition, if each vehicle flies with multiple rocket engines the trajectory until the temporary engine cutoff and approximate the
engines are cut off one after another because the body axial ac- apogee kick by an impulse thrust.
celeration increases as the propellants are consumed. Because the At the beginning of this phase, the angle of attack is allowed to
equations of motion change discontinuously at the shutdown points, have values that are discontinuous with the terminal angle of attack
the phases must be subdivided into intervals, the number of which in the preceding phase. The bank angle remains 0 deg during the
corresponds to the number of engines. Initial and terminal times flight for the orbiter to fly east. The constraint conditions in this
in the intervals are contained in the optimized variables. We de- phase are the angle of attack, throttle, dynamic pressure, thrust,
fine equality constraint conditions to make the time and variables normal acceleration, and body axial acceleration. Their values are
equal between a series of intervals. The phases are explained as limited to the orbiter’s flight performance parameters in the design
follows. variables, except that the angle of attack is kept below 20 deg and
the throttle is kept in the 60–100% range.
Booster Ascent Phase
This phase indicates the flight from liftoff until the orbiter sep- Booster Fly-Back Phase
arates from the booster. We suppose that the launch site is Aeon After releasing the orbiter, the booster turns to fly back to the
airfield (longitude, 157.36◦ west and latitude, 1.78◦ north) on Christ- launch site, Aeon airfield. As the booster in this paper has no separate
mas Island of the Republic of Kiribati (Fig. 3). On the island, airbreathing engines for the return flight, the only rocket engines can
National Space Development Agency of Japan has a downrange propel the booster in this phase. In this paper they are not compared
with the vehicles with the airbreathing engines on the boosters. The
bank angle, which until now has been fixed at 0 deg, is allowed to be
free. The initial angle of attack must be equal to the terminal angle
in the booster ascent phase. As in the ascent phase, the angle of
attack, throttle, dynamic pressure, thrust, normal acceleration, and
body axial acceleration are constrained during the return flight. The
terminal constraint conditions are that the latitude and longitude of
the booster are identical with those of the launch site and the altitude
is zero.

Orbiter Reentry and Return Phase


In this phase the orbiter has completed its missions in orbit, which
in this case includes a payload pulled out of orbit and into the or-
biter’s payload bay. The payload-carrying orbiter reenters the atmo-
sphere and lands at the launch site. This study does not compute
the return trajectory of the orbiter by integrating the equations of
motion actually. Rather, we approximate the reentry guidance into
the atmosphere by using the analytic drag control guidance law22
employed for the Space Shuttle Orbiter. The constraint conditions
we use allow the orbiter to fly within the limits of the equilibrium
glide, heat rate, normal acceleration and dynamic pressure. The heat
Fig. 3 Launch site. rate is limited to 400 kW/m2 on the nose tip and the leading edge
774 TSUCHIYA AND MORI

of the wing, whereas the normal acceleration and dynamic pressure the vertical launch booster uses the gear for landing and not for both
limits are given by the design variables. In addition, the following takeoff and landing. For either booster the takeoff weight is about
constraint conditions are prepared: the lift-to-drag ratio of the orbiter five times the landing weight. This makes the landing gear of the
at subsonic speed must be greater than five, which is equivalent to horizontal takeoff booster six times heavier than that of the vertical
that of HOPE-X,23 and the orbiter can pull itself up at the maximum because of Eq. (5). The difference affects the total gross weight and
angle of attack during landing without violating the body normal the dry weight. This result shows the necessity of more accurate
acceleration limit. estimation of the landing gear weight. The difference in total gross
weight and dry weight between the vertical and horizontal takeoff
Optimization Problem vehicles demonstrates that the former seems more feasible than the
In addition to the design variables and constraint conditions just latter, as the definition of the minimized parameter of this study
described, the total gross weight of the booster and orbiter with a makes it clear that smaller vehicles are more likely to be realized.
payload weight is defined as a minimized parameter. We originally In the future, however, when the field of candidate space transporta-
minimized a dry weight of the vehicles, and the vehicles obtained tion systems is narrowed down, weight will not be the only criterion.
were very large. To restrict the vehicle size, we redefined the mini- Other criteria include cost, operability, reliability, and so on.
mized parameter to be the total gross weight including the propellant Figure 4 and Tables 2 and 3 show that, even if the orbiter has no
weight related to a vehicle volume. As a result, however, the sizes payload capability, the total gross weights of the orbiters and boost-
and weights of the optimal vehicles in both the case of the min- ers taking off vertically and horizontally are 126 and 146 Mg, re-
imized parameters differed by not more than 10%. Furthermore, spectively. An increase in the payload weight by 1 Mg increases the
qualitative trends of the solutions were same. For rocket vehicles total gross weight by 40–50 Mg. If fully reusable rocket TSTO sys-
with hydrogen–oxygen propellants, the dry weight and the gross tems are made to replace Japan’s expendable H-IIA rocket, which
weight are closely linked. This study obtains the optimal solutions has a payload capability of more than 10 Mg to LEO, the total
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

for both the vertical and horizontal takeoff vehicles, in which three gross weights of the vertical launch vehicles and the horizontal
sets of payload weights and volumes are delivered to/from orbit: takeoff vehicles will probably reach 550 and 600 Mg, respectively,
0 Mg and 0 m3 , 4 Mg and 40 m3 , and 8 Mg and 80 m3 . whereas the H-IIA rocket weighs about 290 Mg. Compared with
A sequential-quadratic-programming method is employed as an
optimization method. There are approximately 1450 optimized de-
sign variables. The reason for the large number is that we have not
only the geometric and flight performance parameters, but also the
seven state and control variables discretized to the 200 elements, as
explained in the preceding subsections.

Optimal Solutions
Optimal Configurations
Figure 4 shows the variation in weight configurations of the hor-
izontal and vertical takeoff vehicles with the payload weights and
volumes. Tables 2 and 3 detail the component weights as well as
the size and performance data for the vehicles. The results demon-
strate that the vertical launch vehicles are 10–15% lighter than the
horizontal takeoff vehicles in total gross weight. The vertical launch
vehicles are about 20% lighter in the total dry weight as well. The
difference is attributed to the landing gear weights of the respec-
tive boosters. As described in the weight analysis, the landing gear
weight of the vertical launch booster is estimated on the basis of the
maximum landing weight, rather than the takeoff weight, because Fig. 4 Component weight comparison for the two takeoff styles.

Table 2 Performance and weight data of optimal horizontal takeoff vehicles

Payload weight and volume


Characteristic 0 Mg and 0 m3 4 Mg and 40 m3 8 Mg and 80 m3
Vehicle type Booster Orbiter Booster Orbiter Booster Orbiter
Body length, m 47.57 17.61 63.49 25.95 70.78 29.91
Body diameter, m 3.56 2.16 4.40 4.47 4.94 5.59
Wing area, m2 185.1 31.8 326.9 157.0 444.0 247.7
Max. normal acceleration, g 2.8 2.5 2.6 2.5 2.6 2.5
Max. axial acceleration, g 4.0 4.0 3.5 3.5 3.3 3.3
Max. dynamic pressure, kPa 50.0 15.0 50.0 15.0 50.0 15.0
Total thrust (vacuum), MN 2.56 0.18 6.77 0.92 10.01 1.58
Total engines 2 1 5 1 7 2
Fuselage weight, Mg 7.60 1.28 13.39 3.97 17.24 5.78
Wing weight, Mg 2.98 0.25 6.46 1.36 9.63 2.33
Stabilizer weight, Mg 0.64 0.09 1.07 0.86 1.60 1.60
TPS weight, Mg 3.10 0.88 5.21 3.21 6.80 4.85
Landing gear weight, Mg 6.46 0.18 17.91 0.80 27.96 1.35
Tank weight, Mg 4.35 0.87 8.92 2.26 13.09 3.18
Rocket engine weight, Mg 4.02 0.38 10.55 1.52 15.48 2.67
Misc. weight, Mg 5.07 2.19 8.14 4.80 10.18 6.48
Dry weight, Mg 34.22 6.12 71.65 18.78 101.98 28.24
LH2 weight, Mg 13.33 1.72 29.26 8.66 41.71 14.92
LOX weight, Mg 79.98 10.30 175.56 51.94 250.24 89.53
RCS/OME propellant weight, Mg —— 0.73 —— 2.69 —— 4.29
Gross weight, Mg 127.53 18.87 276.47 86.07 393.93 144.98
Total gross weight, Mg 146.40 362.54 538.91
TSUCHIYA AND MORI 775

Table 3 Performance and weight data of optimal vertical launch vehicles

Payload weight and volume


Characteristic 0 Mg and 0 m3 4 Mg and 40 m3 8 Mg and 80 m3
Vehicle type Booster Orbiter Booster Orbiter Booster Orbiter
Body length, m 47.47 16.40 61.74 24.25 66.74 29.90
Body diameter, m 3.32 2.00 4.32 4.47 4.83 5.59
Wing area, m2 154.5 28.1 343.5 137.2 438.3 246.8
Max. normal acceleration, g 2.5 2.5 2.5 2.5 2.6 2.5
Max. axial acceleration, g 4.0 4.0 3.9 3.9 3.6 3.6
Max. dynamic pressure, kPa 50.0 15.0 50.0 15.0 50.0 15.0
Total thrust (vacuum), MN 2.65 0.14 7.66 0.75 11.25 1.48
Total engines 2 1 6 1 8 2
Fuselage weight, Mg 7.16 1.10 12.98 3.65 16.20 5.76
Wing weight, Mg 2.19 0.21 5.90 1.24 8.12 2.31
Stabilizer weight, Mg 0.48 0.07 0.97 0.96 1.25 1.58
TPS weight, Mg 2.77 0.77 5.09 2.97 6.26 4.83
Landing gear weight, Mg 0.95 0.16 2.18 0.74 2.98 1.33
Tank weight, Mg 4.14 0.74 8.47 1.98 11.41 3.19
Rocket engine weight, Mg 4.14 0.30 12.04 1.27 17.43 2.52
Misc. weight, Mg 4.80 1.99 7.93 4.53 9.82 6.45
Dry weight, Mg 26.63 5.34 55.56 17.34 73.47 27.97
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

LH2 weight, Mg 11.97 1.34 27.28 7.46 36.88 14.92


LOX weight, Mg 71.84 8.03 163.67 44.76 221.27 89.53
RCS/OME propellant weight, Mg —— 0.63 —— 2.52 —— 4.26
Gross weight, Mg 110.44 15.34 246.51 76.08 331.62 144.68
Total gross weight, Mg 125.78 322.59 476.30

Fig. 6 Optimized vertical launch vehicles.


Fig. 5 Optimized horizontal takeoff vehicles.
Because the booster (whether vertical or horizontal takeoff type)
existent airplanes, the horizontal takeoff vehicles are considerably must fly back, it has a fine fuselage and large wings, as these charac-
heavy. teristics extend the reachable range by reducing drag and increasing
Figures 5 and 6, which illustrate the horizontal and vertical take- the lift-to-drag ratio. The wing area of the horizontal takeoff booster
off vehicles, show little difference between them. An increase in coincides with the smallest wing area that allows the booster cou-
required payload capacity by 1 Mg and 10 m3 lengthens the boost- pled with the orbiter to take off at the maximum angle of attack. On
ers by about 4 m. Boosters that can carry the orbiters with a payload the other hand, the wing area of the vertical launch booster is not
capability of more than 8 Mg and 80 m3 would be more than 70 m so constrained and is designed solely to optimize fly-back perfor-
long, which would pose operational problems insofar as the boost- mance. Likewise, for either type of orbiter the wing area depends on
ers would be bigger than existing airplanes and rockets. Thus, to the flight conditions from reentry to landing, as the orbiter is subject
increase the feasibility of these vehicles, they must be reduced in to little aerodynamic force in its ascent flight after being released
both size and weight. from the booster.
776 TSUCHIYA AND MORI

Fig. 7 Trajectory plots of horizontal takeoff vehicle (4 Mg and 40 m3


payload).
Fig. 10 Acceleration and dynamic pressure history of horizontal take-
off vehicle (4 Mg and 40 m3 payload).
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

Fig. 11 Trajectory plots of vertical launch vehicle (4 Mg and 40 m3


Fig. 8 Altitude and velocity history of horizontal takeoff vehicle payload).
(4 Mg and 40 m3 payload).

Fig. 12 Altitude and velocity history of vertical launch vehicle (4 Mg


and 40 m3 payload).

Fig. 9 Control history of horizontal takeoff vehicle (4 Mg and 40 m3


payload). and that of the vertical launch vehicle is 1.7. Once the horizontal
takeoff vehicle is airborne, the flight-path angle increases to 54 deg
and then gradually decreases. As the dynamic pressure increases, it
Optimal Trajectories controls the throttle to prevent the dynamic pressure from exceeding
Figures 7–14 show optimal ascent and return trajectories of the the optimal limit. Additionally, with the increase in the body axial
vehicles with the 4 Mg and 40 m3 payload. Vehicles with differ- acceleration the horizontal takeoff booster controls the throttle in
ent payload capabilities exhibit about the same optimal trajectories. order to stay at the maximum acceleration limit; the vertical launch
The horizontal takeoff booster coupled with the orbiter takes off at vehicle stops one of its six rocket engines during the ascent. Mono-
the maximum angle of attack of 20 deg as it leaves the runway at tone acceleration and climb are common to both the vertical and
123 m/s. When taking off, the vehicle has a thrust-to-weight ratio horizontal takeoff vehicles. In both cases altitude and velocity reach
of 1.4, which is considerably larger than that of existing aircraft, approximately 50 km and 2.5 km/s, respectively, in about 130 s, at
TSUCHIYA AND MORI 777

of attack is set to that which maximizes the lift-to-drag ratio. This


extends the flight range and enables the boosters to glide back to the
launch site. The characteristics of the optimal return trajectories are
the nonlevel turn and the glide flight. Even though the boosters do
not use rocket engines for the return flight, they are able to return all
of the way back to the launch site by nonlevel turning and gliding.
Conclusions
In this paper the optimization method was applied to conceptual
studies of fully reusable rocket two-stage-to-orbit (TSTO) vehicles
in order to find the smallest vehicles that can fulfill the mission
requirements. The booster and orbiter vehicles were classified by
their manner of ascent: horizontal takeoff and vertical launch. Con-
ceptual studies of space transportation systems proceed by the col-
laboration of several different disciplines. First, we explained the
analysis methods in each discipline, that is, the vehicle definition,
the surface mesh generator, the low-fidelity aerodynamics analy-
sis, the propulsion analysis, the statistical weight estimation, and
the three-degree-of-freedom trajectory analysis, all of which were
Fig. 13 Control history of vertical launch vehicle (4 Mg and 40 m3 integrated into a conceptual analysis tool. On that basis, we spec-
payload).
ified the optimization problem in which the minimized parameter
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

was defined as the total (combined) gross weight of the first-stage


booster and second-stage orbiter.
The results showed that TSTO vehicles with a practical payload
capability are larger and heavier than both existing airplanes and
comparable expendable rockets. Weight reduction and miniaturiza-
tion of components are necessary to increase the feasibility of the
vehicles. In addition, the vertical launch vehicles are lighter in total
gross weight than the horizontal takeoff vehicles because the land-
ing gears of the vertical launch vehicles are used only during landing
and therefore can be made far lighter than the gears of the horizon-
tal takeoff vehicles, which are used for takeoff as well as landing.
However, as we must evaluate cost, operability, reliability, and other
factors besides weight, it is not necessarily appropriate to suggest
that vertical rather than horizontal launch systems should be selected
for future development. Weight estimation must be made more ac-
curate in the future. This study is characterized by its simultaneous
optimization of the flight trajectories and the vehicle configurations.
The optimization demonstrated that the boosters could return to the
launch site in the down-ranges provided by the ascent trajectories,
Fig. 14 Acceleration and dynamic pressure history of vertical launch using only nonlevel turning and gliding without the use of the rocket
vehicle (4 Mg and 40 m3 payload). engines, that is, the boosters consume no fuel in their return flight.
Japan has recently developed airbreathing engines that can work
at hypersonic speed in future space transportation systems. In future
which time the orbiters are released. The dynamic pressure ranges studies we will apply the present method to conceptual designs
from 3 to 4 kPa at that moment. We should examine the possibility of of vehicles equipped with these new engines and compare their
separation in such an environment in the future. The engine thrusts performance with that of the all-rocket-propulsion TSTO vehicle as
of both types of orbiters right after the release are almost equal to described in this paper.
their respective vehicle weights. The orbiters rise and accelerate by
bang-bang control of the throttles. As we suppose that the thrust References
direction coincides with the body axis, the control of the angle of 1 Maita, M., and Shirouzu, M., “Japan’s Perspective for Future Space
attack is equivalent to the thrust direction control in a low-dynamic Transportation System,” AIAA Paper 2001-4543, Aug. 2001.
pressure environment. 2 Sato, T., Tanatsugu, N., and Naruo, Y., “Development Study on ATREX
For both vertical and horizontal takeoff vehicles, the booster’s op- Engine,” Acta Astronautica, Vol. 47, No. 11, 2000, pp. 799–808.
3 Inatani, Y., Naruo, Y., and Yonemoto, K., “Concept and Preliminary
timal return flight does not involve any use of rocket engines even
though the engine throttle is not constrained to be zero. That means Flight Testing of a Fully Reusable Rocket Vehicle,” Journal of Spacecraft
no propellants are consumed in the return. Downrange distances and Rockets, Vol. 38, No. 1, 2001, pp. 36–42.
4 Dujarric, C., Caporicci, M., Kuczera, H., and Sacher P., “Conceptual
from the launch site are about 100 km at the moment of separation.
Studies and Technology Requirements for a New Generation of European
However, the low dynamic pressure means that the no-thrust boost- Launchers,” Acta Astronautica, Vol. 41, No. 4–10, 1997, pp. 219–228.
ers cannot yet change their flight path. Instead the boosters continue 5 Bonnal, C., and Caporicci, M., “Future Reusable Launch Vehicles in
to climb by their own inertia; the horizontal takeoff booster and the Europe: the FLTP (Future Launchers Technologies Programme),” Acta As-
vertical launch booster reach altitudes of 68 and 78 km, respectively. tronautica, Vol. 47, No. 2–9, 2000, pp. 113–118.
Once the boosters begin their descent and the aerodynamic forces 6 Hammond, W. E., Design Methodologies for Space Transportation

work again, the downrange distances reach 400–500 km. The boost- Systems, AIAA, Reston, VA, 2001.
7 Sobieszczanski-Sobieski, J., and Haftka, R. T., “Multidisciplinary
ers start a turn with the bank angle that exceeds 70 deg. At a high
bank angle the boosters gain a short turning radius but lose altitude Aerospace Design Optimization: Survey of Recent Developments,” AIAA
and velocity. The optimization provides the best bank angle and Paper 96-0711, Jan. 1996.
8 Wolf, D. M., “TRANSYS-Space Transportation System Preliminary De-
angle of attack between the losses and gains. The boosters subtly sign Software,” Journal of Spacecraft and Rockets, Vol. 31, No. 6, 1994,
adjust the two angles in order to prevent the dynamic pressure and pp. 1067–1071.
body normal acceleration from exceeding the optimized limits. As 9 Rahn, M., Schottle, U. M., and Messerschmid, E., “Multidisciplinary
the ground projection of the flight path points to the launch site, Design Tool for System and Mission Optimization of Launch Vehicles,”
the boosters gradually lessen the bank angle to 0 deg, and the angle AIAA 96-4130, Sept. 1996.
778 TSUCHIYA AND MORI

10 Koelle, D. E., “Economics of Small Fully Reusable Launch Systems 18 Harloff, G. J., and Berkowitz, B. M., “HASA—Hypersonic Aerospace
(SSTO vs. TSTO),” Acta Astronautica, Vol. 40, No. 2–8, 1997, pp. 535–544. Sizing Analysis for Preliminary Design of Aerospace Vehicles,” NASA CR-
11 Miele, A., and Mancuso, S., “Design Feasibility via Ascent Optimality 182226, Nov. 1988.
for Next-Generation Spacecraft,” Acta Astronautica, Vol. 45, No. 12, 1999, 19 Koelle, D. E., TRANSCOST Statistical-Analytical Model for Cost Esti-
pp. 705–715. mation and Economic Optimization of Space Transportation Systems, TCS-
12 Mehta, U. B., and Bowles, J. V., “Two-Stage-to-Orbit Spaceplane Con- Trans Cost Systems, Ottobrunn, Germany, 1991.
cept with Growth Potential,” Journal of Propulsion and Power, Vol. 17, 20 Zipfel, P. H., Modeling and Simulation of Aerospace Vehicle Dynamics,
No. 6, 2001, pp. 1149–1161. AIAA, Reston, VA, 2000.
13 Tsuchiya, T., and Suzuki, S., “A Simultaneous Optimization Technique 21 Tsuchiya, T., and Suzuki, S., “Spaceplane Trajectory Optimization with
for Spaceplane Shape and Trajectory,” AIAA Paper 2001-1847, April 2001. Vehicle Size Analysis,” 14th IFAC Symposium on Automatic Control in
14 Tsujimoto, T., Kouchiyama, J., and Shirouzu, M., “Current Status of the Aerospace, International Federation of Automatic Control, Seoul, Repub-
H-II Orbiting Plane Experimental (HOPE-X) Development,” Proceedings lic of Korea, 1998, pp. 444–449.
of 22nd International Symposium on Space Technology and Science, Japan 22 Harpold, J. C., and Graves, C. A., Jr., “Shuttle Entry Guidance,” Journal
Society for Aeronautical and Space Science, Motioka, Japan, 2000, Paper of the Astronautical Sciences, Vol. 27, No. 3, 1979, pp. 239–268.
00-g-06. 23 NAL/NASDA HOPE Joint Research Team, “Aerodynamic Data
15 Morino, L., Chen, L., and Suciu, E., “Steady and Oscillatory Subsonic Book—HOPE-X FY09,” National Space Development Agency, NASDA
and Supersonic Aerodynamics around Complex Configurations,” AIAA Jour- TMR-000001, Tokyo, March 2000 (in Japanese).
nal, Vol. 13, No. 3, 1975, pp. 368–374.
16 Raymer, D. P., Aircraft Design: A Conceptual Approach, AIAA, Reston,
VA, 1999.
17 Sova, G., and Divan, P., “Aerodynamic Preliminary Analysis System II. J. Martin
Part II: User’s Manual,” NASA CR-182077, April 1991. Associate Editor
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082
This article has been cited by:

1. Mathieu BalesdentNicolas BérendPhilippe Dépincé. 2012. Stagewise Multidisciplinary Design Optimization


Formulation for Optimal Design of Expendable Launch Vehicles. Journal of Spacecraft and Rockets 49:4, 720-730.
[Citation] [PDF] [PDF Plus]
2. Mathieu Balesdent, Nicolas Bérend, Philippe Dépincé, Abdelhamid Chriette. 2012. A survey of multidisciplinary
design optimization methods in launch vehicle design. Structural and Multidisciplinary Optimization 45:5, 619-642.
[CrossRef]
3. Jonathan Black, ; J. Simmons, Verifying Launch Vehicle Conceptual Designs Using First Principles and Historical
Trends . [Citation] [PDF] [PDF Plus]
4. Amer Farhan RAFIQUE, Linshu HE, Ali KAMRAN, Qasim ZEESHAN. 2011. Hyper Heuristic Approach for Design
and Optimization of Satellite Launch Vehicle. Chinese Journal of Aeronautics 24:2, 150-163. [CrossRef]
5. A. F. Rafique, L. S. He, Q. Zeeshan, A. Kamran, K. Nisar. 2011. Multidisciplinary design and optimization of an air
launched satellite launch vehicle using a hybrid heuristic search algorithm. Engineering Optimization 43:3, 305-328.
[CrossRef]
Downloaded by Stanford University on October 7, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.1082

6. Mathieu Balesdent, Nicolas Bérend, Philippe Dépincé, Abdelhamid Chriette. 2010. Multidisciplinary Design
Optimization of Multi-Stage Launch Vehicle using Flight Phases Decomposition. International Journal for Simulation
and Multidisciplinary Design Optimization 4:3-4, 117-125. [CrossRef]
7. Yunjun Xu. 2009. Trajectory Analysis for Vertical Takeoff and Vertical Landing Reusable Launch Vehicle’s Upper
Stage. Journal of Aerospace Engineering 22:1, 58-66. [CrossRef]
8. Mateen-Ud-Din QaziLinshu HePermoon Mateen. 2007. Hammersley Sampling and Support-Vector-Regression-
Driven Launch Vehicle Design. Journal of Spacecraft and Rockets 44:5, 1094-1106. [Citation] [PDF] [PDF Plus]
9. Xiao Huang, ; Bernd Chudoba, Overview of a HTHL Hands-On SAV Design Synthesis Methodology . [Citation]
[PDF] [PDF Plus]
10. Vernon T. Bechel, Mathew Negilski, Joshua James. 2006. Limiting the permeability of composites for cryogenic
applications. Composites Science and Technology 66:13, 2284-2295. [CrossRef]
11. Mateen-ud-Din Qazi, He Linshu. 2006. Nearly-orthogonal sampling and neural network metamodel driven conceptual
design of multistage space launch vehicle. Computer-Aided Design 38:6, 595-607. [CrossRef]
12. Takeshi TSUCHIYA. 2006. Near-Optimal Guidance Method for Maximizing the Reachable Domain of Gliding
Aircraft. TRANSACTIONS OF THE JAPAN SOCIETY FOR AERONAUTICAL AND SPACE SCIENCES 49:165,
137-145. [CrossRef]
13. Takeshi TsuchiyaTakashige Mori. 2005. Optimal Design of Two-Stage-To-Orbit Space Planes with Airbreathing
Engines. Journal of Spacecraft and Rockets 42:1, 90-97. [Citation] [PDF] [PDF Plus]

You might also like