You are on page 1of 21

Maximum Range Glide of a Supersonic Aircraft in the

Presence of Wind
Aseem S. Nevrekar1, Alfred G. Striz2, and Prakash Vedula3
School of Aerospace and Mechanical Engineering, The University of Oklahoma, Norman OK
73019-1052

A trajectory optimization that yields maximum range in glide for a

supersonic interceptor aircraft is presented in this paper, and the influence of wind

on the optimal trajectories is investigated. The optimization is performed using the

Chebyshev Pseudospectral Method (CPM) applied to equations governing the three-

dimensional motion of a point mass model of the aircraft. Comparisons are made

between cases without wind, an actual wind profile varying with altitude, and a

numerically determined constant average of this variable wind profile. Optimal

trajectories for the three cases are obtained for different initial heading angles

covering the first quadrant, along with the corresponding state and control variable

histories. Most of the results for the variable wind case and the corresponding

average wind case are found to be close. This observation suggests that the

assumption of a constant wind profile with the correct average wind velocity,

obtained from the actual variable wind profile, will lead to reasonably accurate

predictions of optimal glide trajectories for maximum range, as well as for the

associated state and control variables.

1
Graduate Student
2
Professor Emeritus
3
Associate Professor
1
Nomenclature

𝐶𝐷 = drag coefficient

𝐶𝐷0 = parasite/zero-lift drag coefficient

𝐶𝐿 = lift coefficient

𝐶𝐿𝛼 = lift curve slope

𝐷 = drag

𝒈 = algebraic path constraints

𝑔 = acceleration due to gravity

ℎ = altitude

𝐽 = objective function

𝑘 = induced drag factor

𝐿 = lift

𝑀 = Mach number

𝑚 = mass of aircraft

𝑞 = dynamic pressure

𝑅 = range

𝑆 = wing reference area

𝑣 = free stream velocity

𝑊 = weight of aircraft

𝑤𝑥 = x-component of wind velocity

𝑤𝑦 = y-component of wind velocity

𝑥 = x-direction in ground plane

2
𝑥𝑓 = distance travelled in x-direction along ground plane

𝑦 = y-direction in ground plane

𝑦𝑓 = distance travelled in y-direction along ground plane

𝛼 = angle of attack

𝛼𝑚𝑎𝑥 = maximum value of angle of attack

𝛾 = flight path angle

𝜂 = drag due to lift factor

𝜌 = atmospheric density

𝝉 = time in collocation interval [−1, 1]

𝜙 = bank angle

𝜒 = heading angle

𝜳 = initial and final conditions

I. Introduction

Extensive research has been conducted in the field of trajectory optimization since the 1950s

due to the availability of digital computers. Early work on an optimal control approach applied to

trajectory optimization was conducted by Bliss1, Pontryagin2, Bryson3 and Bellman4. Bellman

generalized the classical Hamilton-Jacobi theory to obtain optimal control functions and optimal

objectives for all possible initial conditions (based on the solution of a linear partial differential

equation). In the early 1960s, a simple Newton method was commonly used for trajectory

optimization, and the optimization was implemented parametrically5. Later on, in the 1970s, a

reduced gradient method6-8 began to be used for solving constrained optimization problems, for

3
nearly linear constraints and a small number of inequalities. One of the first applications of

trajectory optimization techniques was to optimize rocket thrust profiles in vacuum and in the

atmosphere. Bryson and Denham9,10 (1962) used a steepest-ascent method for solving optimum

programming problems. Bryson, Desai and Hoffman11 (1969) used the energy-state

approximation for performance optimization of supersonic aircraft, applying it to Minimum

Time-To-Climb (MTTC), Minimum Fuel-to-Climb (MFTC), and maximum range problems.

In more recent years, pseudospectral methods have increasingly been used to solve trajectory

optimization problems owing to their high degree of accuracy. Fahroo and Ross12 demonstrated

the usefulness of the Chebyshev Pseudospectral Method (CPM) in trajectory optimization by

considering the canonical Brachistochrone problem and the lunar landing problem. Dai and

Cochran13 applied CPM and the Gauss Pseudospectral Method to solve two-dimensional and

three-dimensional MTTC and MFTC problems for aircraft. Dekel and Ben-Asher14 used

pseudospectral techniques to solve optimal glide problems involving low altitude, low velocity

flight conditions.

Optimal glide problems are important, especially in cases involving instantaneous engine

failure. When the engine of an aircraft in flight is powered off, it becomes necessary to determine

the maximum range the aircraft can cover in this condition - subject to existing constraints - and

the corresponding trajectories so that the feasibility of an emergency landing within this distance

can be determined. It is reasonable to expect that the presence of wind affects the maximum

range travelled by the aircraft and the underlying optimal trajectory, in glide condition.

This paper studies the effects that different wind profiles have on the optimum glide trajectory

and on the state and control variables. CPM is employed to transform the maximum range

problem from its optimal control form to a Non-Linear Programming (NLP) problem. CPM has

4
an advantage over other collocation methods due to the high degree of accuracy attainable. This

is because the Chebyshev-Gauss-Lobatto (CGL) node points used in CPM result in interpolating

polynomials that are closest to the optimal polynomial in the max-norm approximation of a

function12,15. The derivatives of these interpolating polynomials are calculated at the CGL node

points exactly, using a differentiation matrix. CPM thus offers relatively rapid solutions to

trajectory optimization problems. The resulting NLP problem can be solved using the NPSOL16

5.0 or SNOPT17,18 6.2 solvers. Both these solvers utilize a Sequential Quadratic Programming

(SQP) algorithm for optimization.

II. Problem Formulation

The trajectory optimization problem can be considered as an Optimal Control Problem (OCP),

expressed in the following form:

The control functions 𝑢(𝑡) are chosen to optimize the cost function19

𝐽 = 𝛷[𝒙(𝑡𝑓 ), 𝑡𝑓 ],

subject to the equations governing the evolution of the state vector 𝒙,

𝒙̇ = 𝒇[𝒙(𝑡), 𝒖(𝑡), 𝒑, 𝑡],

the initial conditions,

𝜳0𝑙 ≤ 𝜳[𝒙(𝑡0 ), 𝒖(𝑡0 ), 𝒑, 𝑡0 ] ≤ 𝜳0𝑢 ,

the final conditions,

𝜳𝑓𝑙 ≤ 𝜳[𝒙(𝑡𝑓 ), 𝒖(𝑡𝑓 ), 𝒑, 𝑡𝑓 ] ≤ 𝜳𝑓𝑢 ,

the algebraic path constraints,

𝒈𝑙 ≤ 𝒈[𝒙(𝑡), 𝒖(𝑡), 𝒑, 𝑡] ≤ 𝒈𝑢 ,

5
the bounds on the state variables,

𝒙𝑙 ≤ 𝒙(𝑡) ≤ 𝒙𝑢 ,

and the bounds on the control variables,

𝒖𝑙 ≤ 𝒖(𝑡) ≤ 𝒖𝑢 ,

where 𝑡 is the independent variable (usually time), 𝒙 represents the vector of state variables, 𝒖 is

the vector of control variables and 𝒑 are the parameters not dependent on 𝑡. The subscripts "0"

and "𝑓" denote the initial and final conditions, respectively, while 𝑙 and 𝑢 denote the lower and

upper bounds, respectively.

a. Chebyshev Pseudospectral Method

The Chebyshev Pseudospectral Method (CPM) discretizes the continuous solution of an OCP

represented by state and control variables by using orthogonal polynomials, which are infinitely

differentiable global functions, for interpolation to satisfy the differential equations, transforming

the OCP to a Non-Linear Programming Problem (NLPP). Examples of such polynomials are

Legendre and Chebyshev polynomials, which are orthogonal over the interval 𝝉 ∈ [−1, 1], with

respect to an appropriate weight function: 𝑤(𝜏) = 1 for Legendre polynomials, and 𝑤(𝜏) =

1⁄√1 − 𝜏 2 for Chebyshev polynomials of the first kind20. Since the exact solution to the OCP is

given by infinitely many state and control variable values, CPM is an approximation. It uses

Chebyshev-Gauss-Lobatto (CGL) collocation to locate the nodes required for discretization. The

interpolation points are given by

𝜏𝑘 = cos(𝜋𝑘⁄𝑁), 𝑘 = 0, … , 𝑁 (1)

These nodes lie in the interval [-1,1] and are the extrema of the Nth order Chebyshev

polynomial TN(t). The jth order Chebyshev polynomial is given by

6
𝑇𝑗 (𝜏) = cos(𝑗 cos −1 𝜏), 𝑗 = 0, … , 𝑁 (2)

which results in

𝑇𝑗 (𝜏𝑘 ) = cos(𝜋𝑘𝑗⁄𝑁) (3)

This yields 𝑇0 (𝜏) = 1, 𝑇1 (𝜏) = 𝜏, 𝑇2 (𝜏) = 2𝜏 2 − 1, 𝑇3 (𝜏) = 4𝜏 3 − 3𝜏, and so on. These

nodes cluster around the end points of the interval. Interpolation at these nodes gives the results

closest to the optimal polynomial approximation to a given function21.

Using a linear transformation, the actual time 𝒕 can be expressed as a function of 𝝉 as

𝒕 = [(𝑡𝑓 − 𝑡0 )𝝉 + (𝑡𝑓 + 𝑡0 )]/2 (4)

where 𝑡0 is the initial time and 𝑡𝑓 the final time. The state variables 𝒙(𝝉) and control variables

𝒖(𝝉) can then be approximated using 𝑁𝑡ℎ order Lagrange interpolating polynomials as

𝒙𝑁 (𝝉) ≈ ∑𝑁
𝑗=0 𝒙𝑗 𝜙𝑗 (𝝉) (5)

𝒖𝑁 (𝝉) ≈ ∑𝑁
𝑗=0 𝒖𝑗 𝜙𝑗 (𝝉) (6)

where
𝝉−𝜏𝑗
𝜙𝑗 (𝝉) = ∏𝑁
𝑙=1 𝜏 (7)
−𝜏
𝑙≠𝑗 𝑗 𝑙

The time derivative of 𝒙𝑁 (𝜏𝑘 ) is expressed as the product of a (𝑁 + 1)×(𝑁 + 1)

differentiation matrix 𝑫 and 𝒙𝑁 (𝜏) as follows5,12.


𝑁 𝑁
𝑑 𝑑
𝒔𝑘 = 𝒙𝑁 (𝜏𝑘 ) = ∑ 𝒙𝑗 𝜙𝑗 (𝜏𝑘 ) = ∑ 𝐷𝑘𝑗 𝒙𝑗,
𝑑𝑡 𝑑𝑡
𝑗=0 𝑗=0

𝑘 = 0, … , 𝑁 (8)

where the elements of 𝑫 are given by

7
(−1)𝑗+𝑘
−(𝑐𝑘 ⁄𝑐𝑗 ) [ (𝜏 ], 𝑗 ≠ 𝑘
𝑘 −𝜏𝑗 )
(2𝑁 2 +1)
− , 𝑗=𝑘=0
𝐷𝑘𝑗 = 𝐷𝑘𝑗 = 6 (9)
(2𝑁 2 +1)
− , 𝑗=𝑘=𝑁
6
𝜏𝑘
{ , 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒
2(1−𝜏𝑘 2 )

with 𝑐0 = 𝑐𝑁 = 2 and 𝑐1 = … = 𝑐𝑁−1 = 1

In CPM, the constraints are applied directly at the CGL nodes

𝑑 (𝑡𝑓 −𝑡0 )
𝒅𝑘 = 𝒙𝑁 (𝜏𝑘 ) − f(𝒙𝑘 , 𝒖𝑘 , 𝑡𝑘 ) = 0 𝑘 = 0, … , 𝑁 (10)
𝑑𝑡 2

The derivatives of the state variables are thus calculated (with a high degree of accuracy)

using the function values at the CGL nodes and the differentiation matrix 𝑫.

b. NLPP Formulation

After discretization using CPM, the above OCP is converted to a NLPP 19 to optimize the

objective function 𝑱(𝒙) subject to the following:

constraints on individual state and control variables:

𝒙𝑙 ≤ 𝒙 ≤ 𝒙𝑢

constraints defined by linear combinations of state and control variables:

𝒃𝑙 ≤ 𝑨𝒙 ≤ 𝒃𝑢

constraints defined by non-linear combinations of state and control variables:

𝒄𝑙 ≤ 𝒄(𝒙) ≤ 𝒄𝑢

where 𝒙 represents the vector of unknown state and control variable values and the final time 𝑡𝑓

(if 𝑡𝑓 is unknown). If 𝑁 is the number of discretization nodes (CGL points) and 𝐾 is the total

number of state and control variables, 𝒙 contains 𝑁𝐾 elements if 𝑡𝑓 is known and 𝑁𝐾 + 1

elements if 𝑡𝑓 is unknown.

8
III. Aircraft Dynamic Model

The 3D equations of motion for an aircraft in glide are as follows13,14:

ℎ̇ = 𝑣 sin 𝛾 (11)

𝑣̇ = − (𝐷⁄𝑚) − (𝑔 sin 𝛾) (12)

𝛾̇ = 𝐿 (cos 𝜙)/ (𝑚𝑣) − 𝑔(cos 𝛾)/𝑣 (13)

𝜒̇ = 𝐿 (sin 𝜙)/(𝑚𝑣 cos 𝛾) (14)

𝑥̇ = 𝑣 cos 𝛾 cos 𝜒 + 𝑤𝑥 (15)

𝑦̇ = 𝑣 cos 𝛾 sin 𝜒 +𝑤𝑦 (16)

The parameters h, v, γ, χ, x and y are the state variables, and α and ϕ are the control variables.

Box/inequality constraints on the state and control variables, which also determine the upper and

lower bounds on these variables, are assumed to be:

0 ft ≤ ℎ ≤ 70,000 ft (17)

10 ft/s ≤ 𝑣 ≤ 3,000 ft/s (18)


40𝜋 40𝜋
− 180 rad ≤ 𝛾 ≤ rad (19)
180

−𝜋 rad ≤ 𝜒 ≤ 𝜋 rad (20)


20𝜋 20𝜋
− 180 rad ≤ 𝛼 ≤ rad (21)
180

50𝜋 50𝜋
− 180 rad ≤ 𝜙 ≤ rad (22)
180

Assuming the aircraft is flying at an altitude of 44,820 ft at Mach 1.35 when the engine shuts

down instantaneously and the aircraft begins to glide, the initial conditions for true airspeed,

flight path angle and position are given by v0 = 1,307.04 ft/s, γ0 = 0 rad, and (x0 , y0, h0) = (0, 0,

44,820) ft, respectively. The final altitude is assumed to be zero, i.e., hf = 0 ft. The subscripts "0"
9
and "𝑓" are used to denote the initial and final conditions, respectively. The optimization

problem will involve finding the maximum value of the objective function R = (𝑥𝑓2 + 𝑦𝑓2 )1/2 with

respect to equations (11) – (16) and inequalities (17) – (22).

The aircraft under consideration for glide optimization is a McDonnell Douglas F-4 Phantom

II supersonic interceptor, used for MTTC optimization in Reference 22. The aircraft weighs

42,000 lb, and has a wing reference area of 530 ft2.

Figure 1. McDonnell Douglas F-4 Phantom II supersonic interceptor


[Photo source: Robert G. Schmitt, United States Maine Corps]

The aerodynamic characteristics of the aircraft, including the effects of compressibility, are

shown in Figure 2. In particular, 𝐶𝐷0 , 𝐶𝐿𝛼 , 𝜂 and k are shown as a function of Mach number. A

piecewise cubic Hermite interpolation approach is used for the presentation of the data in Figure

2 and for use in the trajectory optimization problem.

Figure 2. Aerodynamic Characteristics of the Aircraft22


10
The variations of atmospheric density and speed of sound with altitude, based on the US

1976 Standard Atmosphere23, are shown in Figures 3 and 4. These functional dependencies will

also be incorporated in the trajectory optimization problem.

Figure 3. Atmospheric Density versus Altitude23

Figure 4. Speed of Sound versus Altitude23

This allows for the calculation of lift and drag. Here, the parameters 𝐶𝐷0 , 𝐶𝐿𝛼 , and η are

obtained from
fromFigure
Figure2,2,and
andρ is obtained
ρ is from
obtained Figure
from 3. The
Figure lift and
3. The lift drag
and coefficients are calcula
drag coefficients are

calculated as follows:
11
CD = 𝐶𝐷0 + η𝐶𝐿𝛼 α2

CL = 𝐶𝐿𝛼 α

q = 0.5 ρ v2

L = q S CL

D = q S CD

IV. Wind Profiles

The actual varying wind profile used in this paper is obtained from a nominal wind profile

defined by the centroids of the KSC 99% enveloping ellipse derived from monthly bivariate

normal statistics for the KSC Range Reference Atmosphere24,25.

An average velocity of this wind profile over altitude, is found to be 66.31 ft/s in the x-

direction and 6.62 ft/s in the y-directon, using the trapezoidal rule.

Figure 5. Actual Varying and Constant Average Wind Profiles

12
The values of wx and wy used in Equations (15) and (16) for both the actual varying and the

constant average wind cases, are determined from Figure 5.

V. Results and Discussion

Using initial heading angle values of 𝜒0 = 0o, 30o, 60o and 90o, maximum range glide

trajectories along with the time histories of the state and control variables are obtained for cases

(i) with no wind, (ii) with the actual varying wind, and (iii) with its constant average

approximation. The solver NPSOL 5.0 is used for optimization16. The results presented in this

section were obtained using 50 CGL nodes. At this grid resolution, the results were found to be

converged.

Figure 6. Altitude versus Distance when 𝜒0 = 0o

The variation of altitude with horizontal distance is shown in Figure 6, for cases (i) - (iii),

when the initial heading angle is set to zero. When the engine is powered off, the aircraft uses its

speed for an initial increase in altitude with an associated decrease in velocity, which reduces

drag and increases potential energy, resulting in a larger glide range. As observed in the

13
magnified inset, the no wind and actual wind profiles cross over at an altitude of 51,140 ft, and

the average wind and actual wind profiles cross over at 48,980 ft. The total distance covered in

the actual wind case is found to be larger (by about 8%) than the distance covered in the constant

average and no wind cases. The maximum distance covered in the actual wind case is very close

to that of the average wind case, with the difference being about 1.4%. These qualitative

observations made in Figure 6 are also found to be valid for other values of 𝜒0 , as seen in

Figures 7 and 8. For reference, the no wind case for 𝜒0 = 0o has also been shown.

Figure 7. Altitude versus Distance with Actual Wind

Figure 8. Altitude versus Distance with Constant Average Wind


14
The effects of initial heading angle for wind profiles (ii) and (iii) are presented in Figures 7

and 8, respectively. The total distance covered is observed to be greater for a lower value of 𝜒0 ,

with 𝜒0 = 0o yielding the greatest value and 𝜒0 = 90o the least value of the maximum range, as

seen in Figures 7 and 8.

It is observed that the wind modifies the trajectory of the aircraft. After starting in the

specified initial direction, the trajectory deflects so as to maximize the tailwind for maximum

range, and then settles on a linear path.

Figure 9. Ground Track for Different Initial Heading Angles with No Wind (left) and Actual

Wind (right)

Figure 10. Velocity versus Time when 𝜒0 = 0o


15
Figure 11. Lift-to-Drag Ratio versus Time when 𝜒0 = 0o

Figure 12. Angle of Attack versus Time when 𝜒0 = 0o

Figure 13. Flight Path Angle versus Time when 𝜒0 = 0o

16
The time evolutions of true air speed, lift-to-drag ratio, angle of attack and flight path angle

are shown in Figures 10, 11, 12 and 13, respectively, for wind cases (i)-(iii). It can be noted from

these figures that the curves for cases (ii) and (iii) are very close, indicating that the use of a

constant value of wind velocity matching the average value of the actual wind profile, gives good

estimates for these optimal trajectories. Some interesting additional observations can be made

regarding Figures 10-13. There is a rapid decrease in the velocity of the aircraft initially as it

increases its energy altitude. During this phase, kinetic energy is traded off for potential energy,

enabling the aircraft to obtain a greater glide range. This is followed by a constant velocity

phase, during which the lift-to-drag ratio increases to a maximum, with a simultaneous smooth

decrease in the angle of attack. Then comes a smooth, almost linear decrease in velocity, during

which time the lift-to-drag ratio remains at its maximum, and the angle of attack remains

constant. This maximum lift-to-drag ratio provides the smallest possible glide angle for

maximum possible range. During the landing phase, the velocity again rapidly decreases to a

final value observed to be close to the calculated stall velocity of 235.3 ft/s, determined using

classical aerodynamic theory26. The value of the angle of attack at this point is close to 0.35 rad,

the given value of αmax.. The above is observed to be true for all cases of 𝜒0 . There is no

significant variation in the flight path angle time history with wind profile or 𝜒0 , as seen in

Figure 14. These figures also show that there are durations in which true air speed, lift-to-drag

ratio, angle of attack and flight path angle remain nearly constant.

VI. Conclusion

The powered-off flight (glide) trajectories of a supersonic McDonnell Douglas F-4 Phantom

II were optimized for maximum range, using the Chebyshev Pseudospectral Method, in the
17
presence of different wind profiles and for different values of initial heading angle. The presence

of wind affects the path travelled by the aircraft and the total distance covered. The trajectories as

well as the state and controls variable time histories for the actual wind and constant average

wind profiles were found to match in most cases or were sufficiently close to each other. Hence,

it can be concluded that the assumption of a wind profile with constant wind velocity matching

the average wind velocity of the actual variable wind profile, will lead to reasonably accurate

predictions of optimal trajectories, state and control variables for maximum range glide

performance in the presence of wind.

It was also found that optimal glide trajectories of the McDonnell Douglas F-4 Phantom II

aircraft consist of four distinct phases, including (i) an initial climb phase, during which there is

an increase in altitude and a rapid decrease in velocity as trade-off between kinetic and potential

energy, as expected, followed by (ii) a constant velocity phase (consistent with predictions from

classical aerodynamics), (iii) a maximum lift-to-drag ratio phase, and (iv) a landing phase, during

which the aircraft approaches stall velocity.

VII. Future Work

Future work will include determination of optimal trajectories, state and control variables for

minimum time and minimum fuel for a supersonic aircraft in climb, in the presence of wind, with

realistic wind profiles. Optimum trajectories for various maneuvers will also be determined for

aircraft significantly different from the one considered in this paper, such as UAVs and MAVs.

18
References
1
Bliss G. A., Lectures on the Calculus of Variations, The Univ. of Chicago Press, 1946
2
Pontyragin L. S., The Mathematical Theory of Optimal Processes, New York, Intersciences,

1962
3
Bryson, A. E., and Ho, Y. C., Applied Optimal Control, Blaisdell Publishing Company,

1969, pp. 246


4
Bellman, R. E., Dynamic Programming, Princeton Univ. Press, Princeton, NJ, 1957
5
Betts, J. T., “Survey of Numerical Methods for Trajectory Optimization”, Journal of

Guidance, Control, and Dynamics, Vol. 21, No. 2, 1998, pp. 193-207
6
Betts, J. T., and Hallman, W. P., “NLP2 Optimization Algorithm Documentation,” The

Aerospace Corp., Technical Operating Rept. TOR-0089(4464-06)-1, reissue A, El Segundo, CA,

1997
7
Rosen, J. B., and Kreuser, J. L., “A Gradient Projection Algorithm for Nonlinear

Constraints,” Numerical Methods forNon-Linear Optimization, edited by F. A. Lootsma,

Academic, London, 1972, pp. 297–300


8
Lasdon, L. S., and Waren, A. D., “Generalized Reduced Gradient Software for Linearly and

Nonlinearly Constrained Optimization,” Design and Implementation of Optimization Software,

edited by H. J. Greenberg, Sijthoff and Noordhoff, Alphen aan den Rijn, The Netherlands, 1978,

pp. 335–362
9
Bryson, A. E., and Denham, W. F., “A Steepest-Ascent Method for Solving Optimum

Programming Problems,” Journal of Applied Mechanics, Vol. 29, 1962, pp. 247–257
10
Denham, W. F., Steepest Ascent Solutions of Optimal Programming Problems, Ph.D

Dissertation, Harvard University, 1963

19
11
Bryson Jr., A. E., and Desai, M. N., “Energy-State Approximation in Performance

Optimization of Supersonic Aircraft”, Journal of Aircraft, Vol. 6, No. 6, 1969, pp. 481-488
12
Fahroo, F., and Ross I. M., “Direct Trajectory Optimization by a Chebyshev Pseudospectral

Method”, Journal of Guidance, Control, and Dynamics, Vol 25, No. 1, 2002, pp. 160-166
13
Dai, R., and Cochran Jr., J. E., “Three-Dimensional Trajectory Optimization in Constrained

Airspace”, Journal of Aircraft, Vol. 46, No. 2, 2009, pp. 627-634


14
Maya, D. K., and Ben-Asher, J. Z. “Pseudo-Spectral-Method Based Optimal Glide in the

Event of Engine Cut-off”, Technion - Israel Institute of Technology Haifa, Israel, AIAA

Guidance, Navigation, and Control Conference, 08-11 August 2011, Portland, Oregon, AIAA-

2011-6596
15
Trefethen, L. N., Spectral Methods in MATLAB, Society for Industrial and Applied

Mathematics, Philadelphia, 2000


16
Gill, P. E., Murray, W., Saunders, M. A., and Wright, M. A., “User’s Guide to NPSOl 5.0:

A FORTRAN Package for Nonlinear Programming”, Stanford Optimization Lab., Technical

Rept. SOL86 1, Stanford University, Stanford, CA, July 1998


17
Gano, S. E., Perez, V. M., and Renaud, J. E., “Development and Verification of a

MATLAB Driver for the SNOPT Optimization Software,” Proceedings of the 42nd

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 16-

19 April 2001, Seattle, WA, AIAA 2001-1620


18
Holmstrom, K., Goran, A. O., and Edvall, M. M., “User’s Guide For TOMLAB/SNOPT,”

Tomlab Optimization, Inc., 2005, http://tomlab.biz [retrieved 1 May 2008]


19
Boyd, J. P., Chebyshev and Fourier Spectral Methods, Springer-Verlag, Heidelberg,

Germany, 1989

20
20
Hargraves, C. R., and Paris, S.W., “Direct Trajectory Optimization Using Nonlinear

Programming and Collocation,” Journal of Guidance, Control, and Dynamics, Vol. 10, No. 4,

1987, pp. 338–342


21
Canuto, C., Hussaini, M. Y., Quarteroni, A., and Zang, T. A., Spectral Methods in Fluid

Dynamics, Springer-Verlag, New York, 1988


22
Betts, John T., Practical Methods for Optimal Control and Estimation Using Nonlinear

Programming, Society for Industrial & Applied Mathematics, 2nd Edition, 2009
23
U.S. Standard Atmosphere, 1976, U.S. Government Printing Office, Washington, D.C.,

1976
24
Range Reference Atmosphere Committee, “Range Reference Atmosphere 0-70 km, Cape

Canaveral, Florida,” Range Commanders Council, Doc. 361-83, White Sands, NM, Feb. 1983
25
Adelfang, S. I., Smith, O. E., and Batts, G. W., “Ascent Wind Model for Launch Vehicle

Design”, Journal of Spacecraft and Rockets, Vol 31, No. 3, 1994, pp. 502-508
26
Yechout, T. R., Morris, S. L., Bossert, D. E., and Hallgreen, W. F., Introduction to Aircraft

Flight Mechanics, AIAA Education Series, 2003

21

You might also like