Você está na página 1de 102

A model and sounding rocket simulation tool with

Mathematica ®

Tiago Miguel Oliveira Pinto

Thesis to obtain the Master of Science Degree in

Aerospace Engineering

Supervisor: Prof. Doutor Paulo Jorge Soares Gil

Examination Committee
Chairperson: Prof. Doutor Filipe Szolnoky Ramos Pinto Cunha

Supervisor: Prof. Doutor Paulo Jorge Soares Gil

Member of the Committee: Prof. Doutor João Manuel Gonçalves de Sousa Oliveira

November 2015
ii
Acknowledgments

I would like to thank my supervisor, Professor Paulo Gil, for providing me the opportunity to develop
this work, for all the guidance and, especially, for sharing his knowledge throughout this journey.

To my fellow friends with whom I spent these years and stood beside me during the toughest and best
moments.

To my devoted father, mother and sister who helped me, in several ways, to reach my goals and gave all
of their heart.

iii
iv
Resumo

Durante o lançamento de foguetões de modelismo, a trajectória é consideravelmente influenciada pelo


vento já que estes são foguetões de controlo passivo. Este trabalho tem como objectivo o desenvolvi-
mento de um programa em Mathematica ® que simula a trajectória destes foguetões e de foguetes-sonda
sob a influência de um perfil de vento. É realizado um estudo da camada limite atmosférica e consideram-
se três perfis que levam em conta as condições atmosféricas e o tipo de superfı́cie no local do lançamento.
Mostrando a flexibilidade do programa desenvolvido, é realizada uma simulação Monte Carlo para se
inferir a dispersão do local de aterragem devida às incertezas nos parâmetros do vento. Com as poten-
®
cialidades do Mathematica , a ferramenta desenvolvida permite que seja definido um qualquer número
de estágios, com boosters externos ou internos no primeiro estágio, e que os modelos necessários para o
cálculo da trajectória sejam facilmente modificados ou acrescentados novos modelos na base de dados.
Foram efectuadas duas simulações com foguetões de potência distinta, ambos com dois estágios e
sob as mesmas condições. Também se determinou uma trajectória em que a direcção do perfil de vento
varia com a altitude e compararam-se trajectórias simuladas em estabilidades atmosféricas diferentes sob
a mesma intensidade de vento no lançamento. Os resultados obtidos mostram grande dependência das
trajectórias no vento, especialmente quando a estabilidade da atmosfera varia. Considerando as incertezas
definidas para a velocidade e direcção do vento, bem como para o tipo de terreno, estas mostram uma
forte influência na dispersão dos locais de aterragem.

Palavras-chave: Simulação de Trajectória, Modelismo de Foguetões, Camada Limite At-


mosférica, Perfil de Vento, Monte Carlo.

v
vi
Abstract

During the launch of model rockets, the trajectory is considerably influenced by the wind as these are
passive-guided rockets. This work aims to develop a program in Mathematica ® that simulates the tra-
jectory from model to sounding rockets under the influence of a wind profile. A study of the Atmospheric
Boundary Layer (ABL) is carried out and are considered three wind profiles that take into account the
atmospheric conditions and the terrain type in the launch site. Presenting the developed tool’s flexibility,
a Monte Carlo simulation is performed in order to deduce the dispersion of the landing sites due to the
®
wind parameters’ uncertainties. Owing to the Mathematica ’s potentialities, the developed tool allows
to define any number of stages, with external or internal boosters in the first stage, and to easily modify
the required models to compute the trajectory or to implement new ones in the database.
Two simulations using two-stage rockets provided with a distinct amount of power were predicted
under the same conditions. Also, a trajectory in which the wind direction changes with altitude was
determined and comparisons were performed between trajectories computed with different atmospheric
stabilities under the same wind speed at the launch. The results show a great dependence of the tra-
jectories on the wind profile, specially when the atmospheric stability changes. Considering the defined
uncertainties from the wind speed and direction, as well from the surface type, they present a strong
influence on the dispersion of the landing sites.

Keywords: Trajectory Simulation, Model Rocketry, Atmospheric Boundary Layer, Wind Pro-
file, Monte Carlo.

vii
viii
Contents

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
Resumo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv
List of Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii
List of Common Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix

1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Historical background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Model and high power rocketry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3.1 Associations and current events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Rocket trajectory simulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4.1 OpenRocket . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4.2 RockSim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Work goals and strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Model rocketry concepts 7


2.1 Flight profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Motors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.1 Coding system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Multi-staging and clustering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3 Wind and trajectory 13


3.1 Wind and atmospheric boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1.1 ABL profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.1.2 Atmospheric stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Kinematics and reference frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3 Rocket dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3.1 Dynamic stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

ix
4 Rocket design 25
4.1 Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2 Center of mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3 Moments of inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.4 Center of pressure and CNα . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.5 Drag coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.5.1 Skin friction drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.5.2 Compressibility effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.5.3 Recovery device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

5 Rocket trajectory simulator 39


5.1 Simulator description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2 Rocket assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.3 Atmospheric model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.4 Wind model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.4.1 Wind gusts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.5 Motors data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.6 Drag coefficient model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.7 Validation of the simulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

6 Simulation tests and results 53


6.1 Rockets description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.1.1 Model rocket specification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.1.2 Sounding rocket specification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.2 Rockets characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.2.1 Model rocket . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.2.2 Sounding rocket . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.3 Trajectory simulations conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
6.4 Results from the trajectory simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.4.1 Model rocket launch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.4.2 Sounding rocket launch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.4.3 The influence of atmospheric stability . . . . . . . . . . . . . . . . . . . . . . . . . 65

7 Stochastic simulations 67
7.1 Simulation rocket and conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
7.2 Landing site uncertainties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.3 Rocket optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

8 Conclusions 73
8.1 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

x
Bibliography 75

A Tangent ogive profile 79

B Connector’s Center of Mass (CM) and inertia 81

xi
xii
List of Tables

2.1 Rocketry motor total impulse classification system [23]. . . . . . . . . . . . . . . . . . . . 9

3.1 Surface roughness values for different land surfaces [1]. . . . . . . . . . . . . . . . . . . . . 15


3.2 Atmospheric stability classes according to intervals of Obukhov length, L [33]. . . . . . . . 17
3.3 Absolute and relative errors of H0 /(ρcp ) considering three heights. . . . . . . . . . . . . . 18

4.1 Solid textile parachutes’ projected to reference diameter (dproj /dref ) and respective drag
coefficient [52]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

6.1 Model rocket stages’ data (without connector and nose) [55] [56]. . . . . . . . . . . . . . . 53
6.2 Model rocket fins’ data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.3 Model rocket’s main characteristics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.4 Nike and Orion stages’ data (without connector and nose) [58] [59] [60]. . . . . . . . . . . 54
6.5 Sounding rocket fins’ data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.6 Sounding rocket’s main characteristics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.7 Launch site conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

7.1 High power rocket data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68


7.2 High power rocket fins’ data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
7.3 Mean (µ) and standard deviation (σ) of the landing site in the two directions (x – Southing;
y – Easting) changing each input within a distribution of 20 samples. . . . . . . . . . . . . 69

xiii
xiv
List of Figures

1.1 Launch of a V-2 rocket from Blizna in Poland in 1944 [3]. . . . . . . . . . . . . . . . . . . 2

2.1 Black powder motor design [21]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8


2.2 Grain sections showing different geometries and respective thrust profiles [22]. . . . . . . . 9
2.3 Rocket in stable conditions [24]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Air speed components. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.5 Direct staging method [26]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.6 Usual clustering arrangements [29]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3.1 Inertial (SXY Z) and topocentric-horizon (oxyz) reference frames [25]. . . . . . . . . . . . 18


3.2 Elevation-azimuth topocentric coordinate frame [37]. . . . . . . . . . . . . . . . . . . . . . 19
3.3 Additional angle of attack during rocket’s rotation [40]. . . . . . . . . . . . . . . . . . . . 22

4.1 Multi-stage rocket design. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25


4.2 Generalized fin geometry and respective division to find its centroid. . . . . . . . . . . . . 29
4.3 Generalized fin geometry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.4 Pressure drag coefficient of wedges, cones and similar shapes as a function of the half-vertex
angle (ε) [47]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

5.1 Simulator’s algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40


5.2 Rocket assembly list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.3 Motor list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.4 Body list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.5 Fins list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.6 Connector or nose list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.7 Delays list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.8 Committee on Space Research (COSPAR) International Reference Atmosphere 1986 (CIRA-
86) upgraded version structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.9 Temperature profiles for the CIRA-86 model. . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.10 Neutral wind speed profile for the ABL. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.11 Stable wind speed profile for the ABL. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.12 Unstable wind speed profile for the ABL. . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

xv
5.13 Wind gusts list structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.14 Angle of attack until the apogee with a wind gust perturbation. . . . . . . . . . . . . . . . 47
5.15 Thrust profile for the D12 motor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.16 Conical nose drag coefficient with Mach. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.17 Altitude’s relative error between analytical and numerical solution until the burnout. . . . 50
5.18 Descent velocity from a vertical launch in which the rocket loses mass at 25 s. . . . . . . . 50
5.19 Relative error between analytical and numerical solutions until landing from a projectile
launch with 70° of elevation, constant mass, no drag and no wind. . . . . . . . . . . . . . 51
5.20 Flight path angle from the projectile trajectory with the launch angle (elevation) of 70°. . 51

6.1 Model rocket properties’ functions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56


6.2 Sounding rocket properties’ functions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.3 Model rocket trajectory with a vertical launch and a neutral ABL; Parachute ejection at
the apogee. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.4 Vertical model rocket launch flight data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6.5 Model rocket trajectory in a 75° launch eastward with the downwind changing from north-
eastward at the ground to northward at the apogee. . . . . . . . . . . . . . . . . . . . . . 61
6.6 Sounding rocket trajectory with a vertical launch and wind blowing northeastward; Parachute
ejection at about 28 km of altitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.7 Sounding rocket’s dynamic pressure until the parachute ejection. . . . . . . . . . . . . . . 63
6.8 Vertical sounding rocket launch flight data. . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.9 Wind profiles in a stable atmosphere for a measured wind speed of 4.3 m s−1 at 6 m of
altitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.10 Trajectory profiles of the model rocket launch under three different wind profiles from a
stable atmosphere. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.11 Wind profiles in an unstable atmosphere for a measured wind speed of 4.3 m s−1 at 6 m of
altitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.12 Trajectory profiles of the model rocket launch under three different wind profiles from an
unstable atmosphere. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

7.1 Normal distribution N (0, 1) and respective probability density histogram from 1000 samples. 67
7.2 Landing positions and respective confidence ellipses from the Monte Carlo simulation. . . 71
7.3 Probability density function of the landing coordinates resulting from the Monte Carlo
simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
7.4 Altitude over time from nine launches with different rocket’s length and diameter config-
urations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

A.1 Ogive profile. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

B.1 Connector’s half section (thick lines). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

xvi
List of Acronyms

ABL Atmospheric Boundary Layer

CIRA-86 Committee on Space Research (COSPAR) International Reference Atmosphere 1986

CM Center of Mass

CP Center of Pressure

ISA International Standard Atmosphere

LDRS Large and Dangerous Rocket Ships

NAR National Association Rocketry

NARAM NAR Annual Meeting

NOAA National Oceanic and Atmospheric Administration

SEDS Students for the Exploration and Development of Space

SL Surface Layer

TARC Team America Rocketry Challenge

TRA Tripoli Rocketry Association

xvii
xviii
List of Common Symbols

Greek letters CDf Skin friction drag coefficient

α Angle of attack CDlam Skin friction drag coefficient in laminar

Γ Adiabatic lapse rate regime

γ Ratio of specific heats at constant pres- CDturb Skin friction drag coefficient in turbu-

sure and volume lent regime

CH Specific heat transfer coefficient


δ Latitude
CNα Normal force coefficient derivative
ε Nose half-vertex angle
c̄ Fin average chord
θ Potential temperature
cp Specific heat at constant pressure
θ0 Potential temperature at height z0
cr Root chord
κ Von Kármán constant
ct Tip chord
µ Dynamic viscosity
D Drag
µ⊕ Earth’s gravitational parameter
d Diameter
ξ Damping ratio
El Elevation
ρ Density
el Launch direction unit vector
ψ Stability correction function
f Coriolis parameter
Ω Earth’s angular velocity
g Gravitational acceleration
ωn Natural frequency
g0 Gravitational acceleration at sea level

Latin letters H0 Sensible heat flux at the surface

Az Azimuth h Atmospheric Boundary Layer (ABL)

a Speed of sound depth

a Inertial acceleration I Transversal moment of inertia

a0 Linear acceleration of the local refer- I¯ Principal transversal moment of inertia

ence frame (non inertial) I Total impulse

b Fin span Isp Specific impulse

CD Drag coefficient kf Fin’s interference factor

xix
L Obukhov length u∗0 Local surface friction velocity

LM BL Length scale in the middle of the ABL V Air flow velocity

l Length v Velocity

M Mach number ve Exhaust velocity

Md a Aerodynamic damping moment vt True air speed vector


Md j Jet damping moment W Weight
Ms Stabilizing moment w Wind velocity
m Mass x Position relative to the nose tip
ṁ Mass burning rate xcm Center of Mass (CM) position relative
N Normal force intensity to the nose tip

p Pressure xcp Center of Pressure (CP) position rela-

R Specific gas constant of the air tive to the nose tip

R̄ Universal gas constant z0 Surface roughness length

R Position relative to the Earth’s center


Subscripts for d, l, m, ṁ, r, x or xcm
R0 Origin of the local reference frame rel-
b Body tube
ative to the Earth’s center
bt Boat tail
ReL Reynolds number
c Connector
Recr Critical Reynolds number
ca Connector aft opening
Retr Transition Reynolds number
cf Connector fore opening
r Radius

r Position d Delay charge

S Reference area e Ejection charge

s Sweep length ext External

T Temperature f Final motor

T Thrust g Gap

Tp Thrust profile int Internal

tb Burnout time m Motor

td Delay time from burnout to ejection n Nose

tf Fin thickness p Propellant

u Wind profile velocity s Structural

xx
Chapter 1

Introduction

The goal of the present work is to simulate trajectories from model to sounding rockets taking into account
the wind and the Atmospheric Boundary Layer (ABL). The simulation tool is also capable to forecast
the most probable landing region and perform rocket design optimization using Monte Carlo methods.

1.1 Motivation

Model rocket launches depend on its features and also on atmospheric conditions that may not be well
estimated and change the desired trajectory. Hence, the hobbyists’ goal has better chances to be fulfilled
if there is available a simulation tool provided with all the variables influencing the trajectory.
The scope of our work focuses on passively controlled rockets, therefore the wind plays an important
role during the flight due to its influence on the airflow that stabilizes the rocket. It is known that mean
wind velocity profiles change according the land surface (or the surrounding environment) [1] and time
of the day [2]. Thus, if the wind profile is foreseen wherever and whenever the launch is performed,
the simulations’ errors can be minimized and simulations in different conditions can provide to users
important information to help selecting the site and day of the launch to achieve successful flights.
Sounding rockets consist of one or more stages (propelled by solid or liquid fuel) carrying a scientific
payload in order to study the atmosphere and carry out microgravity research between 40 km to 2000 km
height [3]. From 40 km to 200 km, these rockets are specially interesting because neither balloons nor low-
earth satellites can reach these altitudes. Hence, allowing the developed tool to simulate trajectories of
fin-stabilized sounding rockets up to these heights, brings additional valuable applications to our project.

1.2 Historical background

Rockets were used for the first time in 1232 by the Chinese (who also used gunpowder for fireworks in
the first century A.D.) during the war with the Mongols [4]. In this conflict, rocket-like arrows, propelled
by gunpowder, were launched by the Chinese through a guiding stick. After the war, this knowledge was
also developed by the Mongols who presumably spread the use of rockets to Europe. In the following

1
centuries, rockets keep improving: they started to be launched through tubes (envisioning the modern
bazooka) and a surface-running rocket-powered torpedo was used to set ships on fire. Nevertheless, in the
16th century, they fell on disuse as weapons of war but remained as fireworks, which used the multi-stage
concept for the first time, employed by Johann Schmidlap, in order to reach higher altitudes. In the end
of the 18th century, rockets had a revival as weapons of war due to the Indian rocket barrages against
the British that inspired Colonel William Congreve to design rockets for the army [4]. However, they
had not still improved very much and the cannons continued to represent a better alternative considering
their accuracy and efficiency.
In 1898, Konstantin Tsiolkovsky proposed the use of liquid-propelled rockets regarding the space
exploration. Then, in 1926, Robert H. Goddard, who developed modern rocketry in the early 20th
century, launched the first liquid-fuel rocket [5]. This small rocket, weighing 2.7 kg empty and 4.7 kg
fueled, flew for 2.5 seconds reaching 12.5 m high and 56 m away from the launch site. Although it was
a small flight, this historical event was the precursor of the new technological advances that enabled
the comeback of warfare rockets in World War II. With the development of vehicles that allowed the
transportation of launchers, barrage rockets were an ideal battlefield complement due to its simplicity.
Therefore, the Nazis developed the Nebelwerfer (“Smoke Thrower”), consisting of six short, wide tubes
arranged in a circular cluster, that could launch six 150 mm rockets in less than ten seconds but had
a limited range and a poor accuracy [6]. The Soviet Union developed the Katyusha (roughly, “Little
Katie”) consisting of eight parallel guiding rails fixed on a steel frame that allowed to launch the rockets
in the desired direction. Each rail carried two rockets with 132 mm in diameter that could reach a 8 km
range. Although these rockets were also inaccurate, they were an effective weapon when fired in large
quantities. The most sophisticated rocket used during the war was the V-2 Vergeltungswaffe (“Vengeance
Weapon”), also developed by the Nazis under the directorship of Wernher von Braun [4]. This was the

Figure 1.1: Launch of a V-2 rocket from Blizna in Poland in 1944 [3].

world’s first operational ballistic missile [6] and the first to fly into space. It measured 14.2 m in length,
1.65 m in diameter, and had a range of 330 km [3]. Even during the World War II, this rocket, combined
with new flight conditions and technological advances, demanded further atmospheric and space research.
Hence, after the war, the V-2 was used as a sounding rocket which boosted the development of modern
rockets and new scientific researches in the following decades.

2
1.3 Model and high power rocketry

After the World War II, and during the Space Race, people became interested in amateur rocketry,
specially inventors and intellectuals, which led to several imprudent and unsupervised launches [7, 8].
George H. Stine reported these problems and, together with Orville Carlisle, started the first model
rocket company based on a solid-fuel motor to be safely used in rockets that could be recovered and
reused [8].
Model rockets are typically made of safe materials such as cardboard, plastic and balsa wood [7].
They are propelled by a replaceable, small and pre-packaged solid fuel motor that has an ejection charge
to deploy the recovery system [9]. This consists of a parachute (or a streamer) attached to the nose cone.
High power rockets differ from model rockets in the propulsion power and weight. According to the
American National Association Rocketry (NAR), a rocket is considered a high power rocket if it [10]:

ˆ Uses a motor with more than 160 N s of total impulse or multiple motors that all together exceed
320 N s;
ˆ Uses a motor with more than 80 N average thrust;
ˆ Exceeds 125 g of propellant;
ˆ Uses a hybrid motor or a motor designed to emit sparks;
ˆ Weighs more than 1500 g (including motor(s)) or uses any airframe parts of ductile metal.

These rockets are usually made from the same materials used in model rockets but require a construction
that ensures resistance to higher stress conditions. They usually carry electronic devices to record flight
data or to activate the recovery system. High power rockets fly under the safety code of the local governing
organizations and only a qualified user can purchase a motor provided with such power [10].

1.3.1 Associations and current events

The NAR is the oldest and largest sport rocketry organization in the world [11]. Together with the
American Tripoli Rocketry Association (TRA), which are the two major associations in the United States,
they provide regulations to other organizations around the globe. The NAR and the TRA provide safety
codes and are recognized to certify their members.
NAR clubs organize several NAR meetings where people launch their rockets and compete with
each other. These competitions are divided in four events [12]: altitude, duration, craftsmanship, and
miscellaneous events. The altitude and duration events consist in maximize the apogee and the flight
time, respectively, for given motor power limits. Within the craftsmanship event, scale models of sounding
rockets, missiles and space vehicles are launched with no restriction to the motor power. The miscellaneous
events contain other contests that do not fit under the previous events, such as landing in a chosen spot
and presenting a written research or engineering project.
Annually, more events take place in a week-long national meeting (NAR Annual Meeting (NARAM))
bringing together all the NAR members to decide the national champion. Similarly, the TRA is made up

3
of local groups (“Prefectures”) that schedule events and launches. It also organizes bigger competitions
like Large and Dangerous Rocket Ships (LDRS) which has a category to launch bizarre and original
rockets.
To younger enthusiasts, there is in the United States the Team America Rocketry Challenge (TARC)
for students from 7th to 12th grade in order to incite them following an aerospace career [13]. College
and university students are allowed to compete in the NASA Student Launch in which it is required to
fly a scientific payload up to a given altitude [14]. Teams from many American institutions participate
annually in this competition. A similar event is held by Students for the Exploration and Development
of Space (SEDS) opened to any university in the United States.

1.4 Rocket trajectory simulators

Before there was any rocket trajectory simulator, rocketeers had to rely on their experience to figure out
the probable trajectory under specific conditions. If the hobbyists were beginners or lacked such skills,
they could not find a way to predict their launches and reach their goal. Nowadays, these problems
are overcome due to the development of software capable to simulate rocket trajectories within several
different conditions. Also, these simulators allow to test what if trajectories in order to determine the
outcome from changing a given rocket feature or atmospheric condition and adapt their devices to match
the desired solution.
Two of the commonly used rocket flight simulators, as they include a rocket design tool, are Open-
Rocket and RockSim. The latter is a proprietary software which makes it impossible to validate its
methods. Also, it may represent a significant investment for students and rocket hobbyists. However,
open-source software (like OpenRocket) can be used to avoid these drawbacks.

1.4.1 OpenRocket

OpenRocket was developed by Sampo Niskanen [15]. In this program it is required to model the rocket
before the simulation. The mass is determined by the volume and material density but it can be manually
defined due to existent residual materials impossible to model in the program. The Center of Pressure
(CP) is found by the Barrowman Method (see Section 4.4) which is based on geometric dimensions of
the several parts of the rocket and the aerodynamic forces and moments. OpenRocket simulates from
subsonic to supersonic regimes but at transonic speeds interpolation functions are used as it is difficult
to determine the flow properties. The simulator also make the following assumptions [15]:

ˆ The drag comes from 2 sources: skin friction drag plus pressure distribution drag (body, parasitic
and base pressure drag). The shock wave drag is included in the body pressure drag;
ˆ All the boundary layer on the rocket’s surface is taken as turbulent since the error calculating the
apogee altitude is less than 5%;
ˆ The atmosphere model is the International Standard Atmosphere (ISA) and the user is only allowed
to specify the local temperature that is replaced in the first layer of the model;

4
ˆ The humidity effects are ignored because the difference in air density and speed of sound between
dry and saturated air is less than 1%;
ˆ The wind is determined by summing a constant speed along the altitude with a random, zero-mean
turbulence velocity;
ˆ A flat Earth model is used and the Coriolis effect is ignored;
ˆ During the recovery simulation all the drag comes from the deployed recovery devices. The default
parachute drag is 0.8 and it can be changed by the user.

This simulator can be used for multi-stage rockets and allows to define instants to ignite the motors
or eject the recovery system. Other required inputs are the launch site (longitude, latitude and altitude)
and the angles of the launch rod relative to the wind and the vertical direction.

1.4.2 RockSim

RockSim is the standard program used in model rocketry [15]. Like OpenRocket, it is provided with
features such as: modeling interface, multi-staging, automatic mass estimation (or defined by the user),
atmospheric conditions input at ground level and ejection delay. Some assumptions and features are
worth mentioning [16]:

ˆ When Mach is larger than 0.8, the drag coefficient is adjusted to match the expected CD increase
at transonic regime; it is not specified how it is done;
ˆ The low wind speed is defined by selecting a value from a list with standardized wind speed
ranges, which also sets the high wind speed from National Oceanic and Atmospheric Adminis-
tration (NOAA) standards; custom values can also be specified;
ˆ The wind is considered to move horizontally and may start only after a defined altitude;
ˆ Wind turbulence is optional and it is modeled by a sine wave oscillating between low and high
speeds;
ˆ The atmosphere is simulated by the 1976 US Standard Atmosphere model;
ˆ Thermals, consisting of convective circulations due to the warm surface in response to solar heating
[17], may be implemented specifying their position, size and strength;
ˆ Currently, this program ignores the lift produced by cross wind.

1.5 Work goals and strategies

This work is mainly focused on the development of the trajectory simulator in order to provide a tool
that extends certain boundaries imposed by other simulators, although it may cost a larger processing
®
time due to the interpreted language of Mathematica . Therefore, the developed tool aims to: allow
the user to simulate trajectories from micro to sounding rockets (and even any type of missiles); give
flexibility to change easily a certain input in a given model and figure out its influence in the trajectory;
give freedom to edit or implement new functions and values into the simulator’s database; and process the

5
resulting data in order to obtain further conclusions. Besides these features, the developed simulator has
all the other important characteristics of the mentioned simulators like: multi-staging, boosters, launch
direction and atmospheric, wind and drag coefficient models.
Although the developed trajectory simulator does not include a modeling interface, it allows to define
a versatile rocket model that is functional to every type of rockets. This is accomplished by using a default
scheme of lists structured according the number of stages and boosters. Each major list corresponds to
each stage (not counting the first list that is stored for the boosters data); the internal sublists define the
respective components in each stage, and so on. Therefore, this structure of lists, allied to the simulator
code, allows the rocket model to have any number of stages. However, this scheme of lists must always
be respected so that the program runs properly.
The simulator is developed under a modular programming philosophy. The function that computes
the trajectory receives all the necessary inputs to the calculation and each of them is independent from
one another. Hence, all the required parameters can be easily modified without having to worry about
the other models. This allows to run several simulations under very different conditions and analyze their
impact on the trajectory. For example, compare the outcomes from changing the drag coefficient from a
complicated function to a constant value.
Two databases are already implemented as default, one for the rocket’s motors and another for the
atmospheric models. The databases are defined in a way that simplifies the access to the data and its
properties. For instance, a list of data can be totally changed to the respective parameters from another
model only by specifying the name of the new model. Also, other models can be joined in the databases
to expand the options to use during the simulations.
The developed simulator also allows the user to compute trajectories in loop changing a determined
group of parameters. These analysis may consist of calculations to determine a probable landing region
given a certain parameters’ uncertainties or a rocket design optimization evaluating the combination
of possible values for the components that achieve the desired goal in the trajectory. The mentioned
methods are explored in the present work.

6
Chapter 2

Model rocketry concepts

The developed tool is capable to simulate trajectories from model to sounding rockets, as already men-
tioned. However, throughout this work, we focus our attention mainly on model rockets. This is due to
the larger influence the wind portrays in the trajectory from smaller rockets, specially within the ABL.
Hence, simulations of these trajectories are the best approach to present the results of a rocket launch
combined with the wind profile estimation.
Model and high power rockets are an assembly of several pieces fitting together to produce the desired
shape of the rocket. Basic model rockets consist of a nose and fins attached to a cylindrical body tube.
The most common used noses are conical, parabolic or ogive shaped, and at their bottom exists the nose’s
shoulder to fit, internally, the nose to the body. To connect the stages of the rocket, or to change the
body tube diameter, as an option, conical transition sections can be used. If these parts increase the
rocket diameter, they are called shoulders, otherwise are reducers or boat-tails. The latter are used at
the back of the rockets to decrease the base drag. The more motors a rocket has, the more pieces it will
need to hold them in a fixed position and align the thrust force with the longitudinal axis of the rocket.
To ensure this, we must use engine mounts which consist of a mount tube, centering rings and a motor
block.

2.1 Flight profile

A model rocket’s flight is divided in five phases [18]:

1. Ignition and lift-off;


2. Engine burnout;
3. Coasting phase;
4. Apogee and ejection;
5. Recovery.

Each one of these phases must develop in harmony so the rocket achieves its goals and returns safely to
the ground.

7
Before the flight, the rocket is placed on a launch pad in order to be steady at the desired lift-off angle.
During the first instants of the launch, and because the motor is not steerable, the rod on the pad guides
the rocket while it gains the required speed to become aerodynamically stable. The ignition is usually
triggered electrically increasing the safety comparing to a fuse in case of any unexpected problem.
At the burnout the rocket has already reached its top speed and the motor no longer produces thrust.
Throughout the coasting phase the rocket gains altitude and loses speed until the recovery system is
deployed, safely, as the drag forces are significantly reduced. During this phase the delay charge burns
inside the motor producing a smoke trail that allows the rocket to be tracked at high altitudes.
The deployment of the recovery system, made by an ejection charge inside the motor that splits the
nose cone from the body, is desired to occur at the apogee where the rocket reaches its minimum speed.
This instant must be carefully predicted in order to choose a motor provided with an adequate delay
charge.
After the ejection, the rocket’s descending trajectory is very influenced by the wind. This may cause
the rocket to fly away from the launch site and, in worst cases, be lost. Usually, a parachute is used as
a recovery system but a streamer can be used in order to keep the rocket in sight, since the fall becomes
faster. It is also possible to cut a hole in the parachute to produce the same effect.

2.2 Motors

The two types of propellant used in model rockets are black powder and composite propellant [19]. Black
powder is a solid propellant using potassium nitrate as the oxidizer and carbon and sulfur as the fuel.
Composite propellant is made of ammonium perchlorate as the oxidizer, a powdered metal as the fuel
(usually aluminum or magnesium) and an elastomer as the binder. This propellant is more powerful than
black powder so it is normally used in motors from larger model rockets.
Black powder motors are the most common and consist of a paper case housing a clay nozzle, the
solid propellant, a time delay composition, an ejection charge and a clay cap (see Figure 2.1). Composite
motors have the same design although the nozzle and the case are made of high-temperature plastic [20].

Figure 2.1: Black powder motor design [21].

The propellant’s burning rate, which is proportional to the chamber pressure, is not constant and
changes according to the burning area [19]. Therefore, using different grain configurations, it is possible
to build several thrust profiles in order to get the desired propulsive properties. A black powder motor

8
usually uses an end burning configuration which burns from the nozzle toward the delay composition
keeping a constant burning area [19]. On the other hand, usually, a composite motor is a core burning
which has a central hole running lengthwise the grain. Thus, the burning occurs transversely to the
propellant from the interior to the exterior, increasing the burning area and the thrust. A core burning
grain is represented by the first geometry illustrated in Figure 2.2 where other configurations are given
as examples.

Figure 2.2: Grain sections showing different geometries and respective thrust profiles [22].

The delay charge does not produce thrust allowing the rocket to coast up to the apogee leaving a
smoke trail behind at a slower burning rate than the propellant. After the delay time is over, the ejection
charge fires which over-pressurizes the case and bursts the clay cap. This increases the pressure in the
rocket’s body and deploys the recovery system.

2.2.1 Coding system

Every motor is labeled with a code which provides information about its propulsive characteristics. This
code consists of a letter and two numbers (e.g., “B6-4”). The letter refers to the total impulse produced
by the motor and it was defined that 2.5 N s corresponds to an “A” motor [23]. In Table 2.1 are listed
the range of the total impulse for several motor classes. Usually, the motors are built to operate at the
top limit of these classes [20]. As showed in Table 2.1, the power of the motor doubles if we pick the
next letter, and so on. A motor from class G has a maximum total impulse of 160 N s, thus it is the most

Code Total impulse (N s)


1/4A 0 - 0.625
1/2A 0.626 - 1.25
A 1.26 - 2.50
B 2.51 - 5.00
C 5.01 - 10.00
D 10.01 - 20.00
E 20.01 - 40.00
F 40.01 - 80.00
G 80.01 - 160.00
H 160.01 - 320.00
I 320.01 - 640.00

Table 2.1: Rocketry motor total impulse classification system [23].

9
powerful to be used in model rocketry, as mentioned in Section 1.3. Consequently, motors from class H
to class O (which are not all tabulated) are considered high-power motors.
The first number after the letter gives us the average thrust (in Newton) produced by the motor.
Then, dividing the total impulse of the motor by the average thrust, we can find the burning time.
Regarding the features of the rocket or the goals of the launch, it is possible to choose the most suitable
motor for a given total impulse. For example, a high average thrust motor is convenient if the rocket
needs stability at launch, although it will decrease the apogee since the drag increases.
The second number tells us the coasting time delay (in seconds) from burnout until the ignition of
the ejection charge. Motors marked with “0” as the time delay number are boosters to be used in lower
stages of multi-stage rockets [20]. If the delay number is replaced with a “P”, the motor is also a booster
but it is plugged, which means without ejection charge to burst the top of the casing.

2.3 Stability

A model rocket is stable if its CP is behind the Center of Mass (CM) (seeing from the nose to the tail
of the rocket) [24]. The CP is the point where acts the resultant aerodynamic force produced by the air
pressure and moves toward the nose with increasing angle of attack (α). This force, when the angle of
attack is different from zero, may be decomposed in axial and normal (N ) components where the latter
creates a moment (M ) about the CM of the rocket (see Figure 2.3). In stable conditions, this moment

(a) Normal force (b) Stabilizing moment

Figure 2.3: Rocket in stable conditions [24].

produces a damped oscillating movement about the air flow direction until the rocket angle of attack
returns to zero and the normal force vanishes [24] (see Section 3.3.1).
The arm of the moment is the length between the CM and the CP, called static margin. Decreasing
the static margin until the CP coincides with the CM, makes the rocket unstable because the resulting
moment becomes smaller and smaller. If the CP overtakes the CM and a perturbation deviates the rocket
from the air flow direction, the aerodynamic moment will amplify the disturbance causing the rocket to
spin and crash.

10
With a larger static margin, however, the rocket may become overstable and reaches a lower apogee
because it turns sooner to the wind [24]. This effect, called weathercocking, is due to the contributions
of the rocket and wind’s velocities (v and w, respectively) that define the direction and intensity of the
true air speed as (see Figure 2.4)

vt = v − w . (2.1)

Considering windy conditions, the angle of attack is maximum when the rocket leaves the guiding rods
and is found by (see Figure 2.4)

v · vt
α = arccos . (2.2)
kvkkvt k

Note that v may describe other angles in respect to the ground, so that Figure 2.4 no longer depicts a
right triangle. In a broader sense, if there is a windless atmosphere, the rocket may also present an angle
of attack if its velocity is not aligned with the longitudinal axis of the rocket.

v
vt α

Figure 2.4: Air speed components.

Weathercocking is also enhanced by strong winds and low rocket’s velocity as it leaves the launch
platform. In order to reduce this effect, it is advised to have a static margin lower than one maximum
rocket diameter, fly in calm winds and use longer guiding rods or higher average thrust motors so that the
rocket gains more velocity to counterbalance the wind speed when leaving the launch pad [24]. Except
the former, these advices also contribute to guarantee the stability of the rocket at the moment it leaves
the guiding rods.

2.4 Multi-staging and clustering

Multi-staging enables the rocket to reach higher altitudes by discarding parts of structural mass made
useless by the spending of rocket propellant. If only one stage is burning at a given time and the previous
one is discarded right after its propellant is consumed, it is called serial staging [25]. Otherwise, if some
stages are burning at the same time, it is called parallel staging.
In model rocketry there are two ways of staging a rocket: direct and indirect staging [26]. With direct
staging the upper motor is ignited by the lower stage motor called the booster. These type of motors do
not carry delay composition and ejection charge in order to ignite the upper stage nearly instantaneously
after the burnout, as it is represented in Figure 2.5. As the motor burns, the remaining propellant wall
becomes thinner and bursts due to the pressure, which throws hot gases and burning particles directly to

11
Figure 2.5: Direct staging method [26].

the next nozzle. Immediately, the joint between the motors is pressurized and the lower stage is released
[27]. Since the composite propellant is soft and rubbery, it cannot hold the chamber pressure in order to
burst and expel the particles. For this reason there are only black powder boosters, leaving composite
motors to be used in indirect staging. Direct staging is the simplest and cheapest method because it does
not require electronic devices to ignite the upper stage, unlike indirect staging [26].
With serial staging the lower stages must be equipped with fins (with increased area on each added
stage) to compensate the shift of the CM back to the tail [27]. These fins must be designed so that
the stages become aerodynamically unstable in order to tumble and decrease speed during recovery. A
larger fin area, however, over-stabilizes the rocket giving it tendency to weathercock. For this reason, a
multi-stage rocket should only fly with calm winds and never use more than three stages [28].
Parallel staging can be achieved if the rocket is provided with external boosters with a lower burnout
time than the central motor. In this case, when the core motor and the boosters are burning at the same
time, it is called the zeroth stage [25].
Several motors can be used together to provide the rocket with more thrust. This method is called
clustering in which, at most, four engines must be used since it makes the ignition less reliable [29]. The
cluster must be arranged in a way that the resultant thrust is applied through the longitudinal axis of the
rocket and all the motors at the same distance from this axis must develop equal thrust (see Figure 2.6).

Figure 2.6: Usual clustering arrangements [29].

In a direct multi-stage rocket, clustering should only be used in the first stage due to the difficulty
in igniting the upper motors at the same time [29]. In this situation, one motor must be placed at the
center (to start the next stage) and the others around it.

12
Chapter 3

Wind and trajectory

3.1 Wind and atmospheric boundary layer

The lowest layer of the troposphere is directly influenced by the ground surface characteristics. Therefore,
a wind speed profile is worth to be implemented since its effect can be relevant, mainly, in the trajectory
of small and medium model rockets. This layer is known as the ABL and its thickness may change from
about a hundred meters to a few kilometers varying in time and with geographic region [30]. The bottom
10% of the ABL is named the Surface Layer (SL) with, in average, 100 m height [31] which makes the
ABL about 1 km thick.
Above the ABL stays the free atmosphere, a more stable region where the frictional influences of the
surface can be ignored and the wind is nearly geostrophic [30]. The geostrophic wind, moving parallel
to the isobars, results from the balance between the Coriolis and the pressure gradient forces [17]. From
the momentum equations in steady state with no turbulence terms, the geostrophic wind’s components
in the eastward and northward directions are given, respectively, by [17]:

1 ∂p
ug = − , (3.1)
ρf ∂y
1 ∂p
vg = , (3.2)
ρf ∂x

where ρ is the air density, p is the air pressure and f is the Coriolis parameter (f = 2Ω sin δ) which
depends on the Earth’s angular velocity (Ω) and latitude (δ). With (3.1) and (3.2), the direction of the
geostrophic wind can be found if the horizontal pressure gradient is known for a given region. If there is
a temperature gradient, the geostrophic wind intensity changes with height and its direction may vary if
the isotherms are not parallel to the isobars [17]. This vertical variation is named thermal wind and can
be found using the equation of state [32]

p
ρ= , (3.3)
RT

where T is the air temperature and R = 287.0 J kg−1 K−1 is the specific gas constant of the air (which

13
is the quotient between the universal gas constant, R̄ = 8.314 J mol−1 K−1 , and the molecular weight of
air, Mair = 28.97 g mol−1 ), and the hydrostatic equation

dp = −ρgdz , (3.4)

in (3.1) and (3.2) resulting, approximately, in [17]:

∂ug g ∂T
'− , (3.5)
∂z f T ∂y
∂vg g ∂T
' . (3.6)
∂z f T ∂x

In this situation, when there are horizontal temperature gradients, the atmosphere is called baroclinic.
Considering only geostrophic balance, which is the case of a barotropic atmosphere, the wind is constant
with height and its intensity can be assumed as equal to the one at the top of the ABL.
As the behavior of the atmosphere depends on the weather conditions, the atmospheric stability is
classified in three different classes (neutral, stable and unstable) that help to determine the properties of
the atmosphere (including the wind) in different conditions. A neutral atmosphere implies an adiabatic
lapse rate and no convection which is the case of a partially or highly cloudy atmosphere that may reduce
the insolation at the surface [30]. Stable conditions occur mostly at night but can also appear when the
ground surface is colder than the surrounding air. An unstable atmosphere is formed in clear weather
during the day when there is high radiation from the sun causing ascending heat transfer. Knowing these
atmospheric conditions we can deduce the stability class. They also influence the wind profile and must
be taken into account due to convective effects.

3.1.1 ABL profiles

From measurements of the wind speed, the literature [33] presents the following wind profiles modeled
considering the entire ABL and its stability:
"   #
u∗0

z z z z
u(z) = ln + − , (3.7)
κ z0 LM BL h 2LM BL

for neutral conditions,


"   #
u∗0
  
z 4.7z z z z z
u(z) = ln + 1− + − , (3.8)
κ z0 L 2h LM BL h 2LM BL

for stable conditions, and


"   #
u∗0
  
z z z z z
u(z) = ln −ψ + − , (3.9)
κ z0 L LM BL h 2LM BL

for unstable conditions, where h is the ABL depth, z0 is the surface roughness length, κ is the von Kármán
constant, LM BL is the length scale in the middle of the ABL for its respective condition, u∗0 is the local

14
surface friction velocity near the ground, L is the Obukhov length and ψ is a stability correction for the
SL given by
!
1 + x + x2 √
   
z 3 1 + 2x π
ψ = ln − 3 arctan √ +√ , (3.10)
L 2 3 3 3

1/3
where x = 1 − 12z/L .
Above the ABL (z > h), (3.7), (3.8) and (3.9) are no longer valid since it is the region of the free
atmosphere. Within the ABL, unlike the free atmosphere, there are surface friction effects which cause
the wind direction to deviate from the geostrophic wind. Thus, the wind changes with height until it
becomes aligned with the direction of the geostrophic wind at the top of the ABL [17].
The last two terms in (3.7), (3.8) and (3.9) represent the contributions of the middle and upper layers
of the ABL [33] and, considering the neutral case when these terms are ignored, (3.7) reduces to

u∗
 
z
u(z) = 0 ln , (3.11)
κ z0

which is the simpler and most known log wind profile used in the SL.
The parameter h, if it is unknown, may be found from [33]

u∗
h=c 0, (3.12)
|f |

in a neutral and unstable ABL, and from [34]


s
u∗0 L
h = 0.5 , (3.13)
|f |

in stable conditions, where c is an empirical constant (ranging in the literature from 0.1 to 0.4 [35]). Note
that, when f → 0, the thickness of the boundary layer increases significantly and h no longer has physical
meaning. Therefore, the model is not valid for low latitudes [35].
The surface type influences z0 which is quantified and can be obtained from Table 3.1. The value of
the empirical constant κ is near 0.35 or 0.4 [30].

Surface type z0 (cm)


Sand 0.01...0.1
Cut grass (∼ 0.01 m) 0.1...1
Small grass, steppe 1...4
Uncultivated land 2...3
High grass 4...10
Coniferous forest 90...100
Uptown, suburbs 20...40
Downtown 35...45
Big towns 60...80

Table 3.1: Surface roughness values for different land surfaces [1].

15
The length scale, LM BL , which is deduced by the Rossby similarity theory (that relates the geostrophic
wind and u∗0 ) applied to (3.7), (3.8) and (3.9) at the top of the ABL (z = h), is found from [33]:
 1/2 −1
 ∗  !2  
h u0 h 
LM BL =  ln −B + A2  − ln  , (3.14)
2 |f | z0 z0

for neutral conditions,


 1/2 −1
 ∗  !2  
h u0 h 4.7h 
LM BL =  ln −B + A2  − ln −  , (3.15)
2 |f | z0 z0 2L

for stable conditions, and


 1/2 −1
 ∗  !2    
h u0 h h 
LM BL =  ln −B + A2  − ln +ψ  , (3.16)
2 |f | z0 z0 L

for unstable conditions, where A ≈ 4.9 and B ≈ 1.9 in a neutral atmosphere [33]. Many empirical
functions have been developed in the literature to find A and B for the remaining conditions [34]. Show-
ing good agreement between theoretically and empirically derived functions, these parameters may be
approximated by [34]:

h h
A = ln − 2.2 + 2.9 , (3.17)
L L
h
B = 3.5 , (3.18)
L

for stable conditions, and

h |f | h
A = ln + ln ∗ + 1.5 , (3.19)
|L| u0
u∗0 fh
B=κ + 1.8 ∗ e0.2h/L , (3.20)
fh u0

for unstable conditions.


1
The velocity u∗0 is defined as u∗0 = (τ/ρ ) /2 , where τ is the surface stress and can be found solving the
equations in order to u∗0 using the measured wind speed at the launch site with the respective height of
the measurement.
The Obukhov length (L) represents the thickness near the surface in which the shear stress dominates
over the buoyancy effects in generating turbulence. It is defined by [17]

u∗0 3
L=− , (3.21)
κ Tg0 ρc
H0
p

where T0 is the temperature at the reference pressure (p0 = 1000 mbar), cp is the specific heat at constant

16
pressure and H0 is the sensible heat flux at the surface which is given by

H0 = −ρcp CH U (θ − θ0 ) , (3.22)

where CH is the heat transfer coefficient (ranging from 0 to 0.01 for land surfaces [17]), U and θ are,
respectively, the wind speed and potential temperature at the height of their measurements and θ0 is the
potential temperature at z0 . Potential temperature is defined as

 R/cp
p0
θ=T , (3.23)
p

and the subtraction in (3.22) may be approximated by [17]

θ − θ0 ' T − T |z=z0 + Γ(z − z0 ) , (3.24)

where Γ = 9.8 K km−1 is the adiabatic lapse rate in a dry atmosphere.

3.1.2 Atmospheric stability

The atmospheric stability can be determined from Table 3.2 if the Obukhov length is known. To determine

Obukhov length interval (m) Atmospheric stability class


10 ≤ L ≤ 50 Very stable
50 ≤ L ≤ 200 Stable
200 ≤ L ≤ 500 Near stable
|L| ≥ 500 Neutral
−500 ≤ L ≤ −200 Near unstable
−200 ≤ L ≤ −100 Unstable
−100 ≤ L ≤ −50 Very unstable

Table 3.2: Atmospheric stability classes according to intervals of Obukhov length, L [33].

this parameter we can use (3.21) with measurements taken from the atmosphere. In a range of a few meters
the temperature is practically constant, therefore, the height of the measurements should not be close to
the surface in order to reduce errors. Since to get L we do not need to find H0 but H0 /(ρcp ) (surface
kinematic heat flux), which depends of five measured parameters (xi=1,2...5 ), its maximum absolute error
is determined by [36]

5
 X ∂H0 /(ρcp )
∆ H0 /(ρcp ) = ∆xi
i=1
∂xi

= CH U (∆T + ∆T |z=z0 + Γ(∆z + ∆z0 ) + CH T − T |z=z0 + Γ(z − z0 ) ∆U , (3.25)

where ∆xi is the absolute deviation of each measurement.


Considering three measured heights with CH = 0.01, ∆z = ∆z0 = 1 mm, ∆T = ∆T |z=z0 = 0.1 K,
∆U = 0.1 m s−1 , using a temperature profile from the atmospheric model (developed in Section 5.3) to

17
get the temperatures and (3.11) to estimate U (with z0 = 7 cm, κ = 0.4 and u∗0 = 0.35 m s−1 ), the
absolute and relative errors from (3.25) are listed in Table 3.3. We can see that, even for 700 m above

z (m) (above z0 ) Temperature (K) Absolute error (K m/s) Relative error (%)
300 289.34 1.6651 × 10−2 39.1
500 287.85 1.8105 × 10−2 22.1
700 286.43 1.9286 × 10−2 15.1

Table 3.3: Absolute and relative errors of H0 /(ρcp ) considering three heights.

the ground, the relative error is still reasonably high. We could get lower errors taking measurements
at higher altitudes but the tabulated ones are already impractical to reach. Therefore, and since the
absolute deviations of temperature may be higher, it is not reliable to find L from the previous method.
Nevertheless, if the atmospheric stability is deduced qualitatively from the weather conditions and time
of day (as described in the last paragraph of Section 3.1), we get to know which wind profile to use and
a value for L can be estimated from Table 3.2.
Summarizing Section 3.1, there are three ABL profiles that we aim to determine for each atmospheric
stability. The inputs relevant to mention are: the roughness length (z0 ), the Obukhov length (L) and the
friction velocity (u∗0 ). The former is estimated by the terrain type from Table 3.1. The parameter L shows
large uncertainties when finding it quantitatively from atmospheric measurements. Thus, using Table 3.2,
it can be estimated qualitatively from the weather conditions that were described for each stability class.
The velocity u∗0 , after knowing the remaining parameters, is determined solving the respective wind profile
equation with the wind speed measured at a given altitude.

3.2 Kinematics and reference frames

Considering the origin of the reference frame at the launch site, powerful rockets that reach high altitudes
may suffer some errors calculating the trajectory due to the Earth’s rotation. The alternative is to define
the center of the Earth as the origin of the inertial reference frame and set the local frame at the launch
site (point o, see Figure 3.1). The local frame is defined as a topocentric-horizon coordinate frame with

Figure 3.1: Inertial (SXY Z) and topocentric-horizon (oxyz) reference frames [25].

18
the x axis pointing southward and the y eastward. Thus, as this is a right-handed coordinate frame, z
axis points upward collinear with the Earth’s radius direction (or zenith).
The elevation (El) is the angle measured in the vertical plane, at the origin of the local frame, from the
horizon up to the rocket (see Figure 3.2). Likewise, the azimuth (Az) is defined as the angle measured in

Figure 3.2: Elevation-azimuth topocentric coordinate frame [37].

the horizontal plane from north (negative x axis direction) to the rocket. The coordinate transformations
from spherical coordinates (El, Az, r) to Cartesian coordinates (x, y, z) are:

x = −r cos El cos Az , (3.26)

y = r cos El sin Az , (3.27)

z = r sin El . (3.28)

The total acceleration in the inertial reference frame is given by [25]

∂2r dΩ ∂r
a = ao + 2
+ × r + 2Ω × + Ω × (Ω × r) , (3.29)
∂t dt ∂t

where Ω is the Earth’s angular velocity (Ω = 7.2921 × 10−5 rad/s), r is the position of the rocket relative
to the local frame and ao is the linear acceleration of this frame given as Ω × (Ω × Ro ), where Ro is
the position of point o relative to the Earth’s center (see Figure 3.1). The rocket’s velocity seen from
∂2r
an observer at the launch site is v = ∂t2 , where the partial derivative stands for the time derivative of
dΩ
r with reference to the local frame. The term dt × r is the angular acceleration of the moving frame,
∂r
2Ω × ∂t is the Coriolis acceleration and Ω × (Ω × r) is the centripetal acceleration. In order to get the
trajectory seen by an observer at the launch site, we must write these vectors in the local frame (oxyz)
coordinates.

19
3.3 Rocket dynamics

Besides the forces due to the Earth’s rotation, also act on the rocket the drag (D) and the weight (W).
Then, from the momentum conservation, and due to the ejection of propellant, the rocket equation of
motion becomes [38]

dm Ve
ma = Ve − A(Pe − Pa ) +D+W, (3.30)
dt Ve

where m is the rocket’s mass, Ve is the exhaust velocity relative to the rocket, A is the motor nozzle
area, Pe is the exhaust pressure and Pa is the ambient pressure. The first two terms on the right-hand
side of (3.30) correspond to the force responsible for the acceleration due to the loss os mass [39]: the
thrust (T). Thus, using this definition and replacing (3.29) in (3.30), we get

∂2r dΩ ∂r
m = −mao − m × r − 2mΩ × − mΩ × (Ω × r) + T + D + W . (3.31)
∂t2 dt ∂t

Solving (3.31) the position of the rocket seen from the launch site can be obtained.
The thrust is always aligned with the rocket’s longitudinal axis because the rocket is considered as not
steerable. Since the rocket may fly with angle of attack until it stabilizes and the thrust becomes collinear
with the true air speed vector, the effective propulsive force may be lower in this direction. Therefore,
the adopted model to the thrust force is

vt
T = Tp cos α (3.32)
kvt k

where Tp is the propulsion from the motors’ thrust profile.


The aerodynamic force is assumed to be merely the drag since the rocket only flies with angle of
attack during a short interval of time after the launch (or after any perturbation) and it describes a
low amplitude oscillation. The lift that is produced during these moments contributes essentially to
stabilize the rocket, its average is considered to be zero, and its effect can be implemented as described
in Section 3.3.1. Hence, the aerodynamic force is

1
D= ρkvt k SCD vt , (3.33)
2

where S is the reference area (usually chosen as the maximum sectional area of the rocket or, after
the apogee, the reference area of the recovery device) and CD is the drag coefficient (determined in
Section 4.5).
The weight points to the Earth’s center and is found by

µ⊕ m
W=− z, (3.34)
(R0 + z)2

where µ⊕ = 3.986 × 105 km3 s−2 is the Earth’s gravitational parameter and z is the unit vector of the

20
vertical coordinate.
During the launch, when the rocket is constrained by the guiding rods, the drag and weight normal
to the launch direction are counterbalanced by the rods reaction and only the tangential forces to the
rods contribute to the launch. Thus, (3.31) becomes

∂2r dΩ ∂r 1
m = −mao − m × r − 2mΩ × − mΩ × (Ω × r) − ρkvk SCD v + (Tp + W sin El)el , (3.35)
∂t2 dt ∂t 2

where el is the unit vector of the rods’ direction determined by the launch elevation and azimuth as

el = (cos El cos Az, cos El sin Az, sin El) . (3.36)

3.3.1 Dynamic stability

After leaving the platform, and when a disturbance occurs, it is considered that the rocket acquires an
angle of attack that will decrease gradually and force it to oscillate about the true air speed direction
[24], as mentioned in Section 2.3. To generate this damped oscillating movement, we consider that the
rocket is subjected to two moments [40]: a stabilizing (or restoring) moment and a damping moment.
The stabilizing moment comes from the normal aerodynamic force acting on the CP which causes the
rocket to rotate about the CM since these points must distance from each other by the static margin (see
Figure 2.3). The normal force depends on the normal coefficient, CN , which, assuming small angles of
attack, is given by

CN = CNα α , (3.37)

where CNα is the rocket’s normal force coefficient derivative. Therefore, the stabilizing moment is [41]

1 2
Ms = N (xcp − xcm ) = ρV SCNα α(xcp − xcm ) , (3.38)
2

where xcm and xcp are, respectively, the rocket’s CM and CP distances from the reference point (often
considered as the nose tip, which we also consider throughout this work) and V is the flow velocity.
One source of the damping moment results from the aerodynamic resistance of the air while the rocket
is rotating [40]. During the rotation, the angle of attack of each part of the rocket changes due to the
tangential velocity of this motion. Considering a component i with its CP at distance xcpi − xcm from
the rocket’s CM, the variation of the angle of attack for this component, assuming again small angles, is
(see Figure 3.3)

α̇(xcpi − xcm )
∆α = , (3.39)
V

where α̇ is the angular velocity. Like the stabilizing moment, only the normal force of this additional
resistance contributes to dampen the rotation. Thus, using (3.39), and since the aerodynamic damping

21
Figure 3.3: Additional angle of attack during rocket’s rotation [40].

moment (Mda ) is determined by summing all the elemental moments along the rocket [40], we get [42]

n
X n
X
Mda = Mdai = Ni (xcpi − xcm )
i=1 i=1
n
1 2 X
= ρV S CNαi αi (xcpi − xcm )
2 i=1
n
1 X
= ρV S α̇ CNαi (xcpi − xcm )2 , (3.40)
2 i=1

where n is the total number of components of the rocket that contribute to this moment.
The other contribution to the damping moment comes from the Coriolis acceleration due to the change
of the gas flow through the nozzle [40], also called jet damping. Knowing the Coriolis acceleration as

ac = 2ve α̇ , (3.41)

where ve is the exhaust velocity, and taking an element of propulsive mass (dm) with length dx at distance
x from the CM, the elemental jet damping comes as [40]

dMdj = 2ve α̇x dm = 2ve α̇xρp Snozzle dx = 2α̇xṁ dx , (3.42)

where ρp is the density of the propellant, Snozzle is the nozzle section area and ṁ is the mass burning
rate. Integrating (3.42) from the CM up to the nozzle, we get the jet damping as
Z xnozzle
Mdj = 2α̇xṁ dx = α̇ṁ(xnozzle − xcm )2 , (3.43)
xcm

where xnozzle is the nozzle distance from the reference point.


Once the moments acting on the rocket are known, we can find the governing equation for the motion
of the angle of attack assuming the rotation is two dimensional (in the plane formed by the rocket’s
velocity and wind vectors). The angular momentum about the rocket’s CM is [43]

Hcm = I α̇ , (3.44)

where I is the transversal moment of inertia relative to the CM , and the balance of moments about the
same point is given by the angular momentum derivative as

X d
M= Hcm . (3.45)
dt

22
Therefore, from (3.44) and (3.45), and knowing that the stabilizing and damping moments counteract
the rocket’s rotation, we get

−Ms − Mda − Mdj = I˙α̇ + I α̈ , (3.46)

which, applying (3.38), (3.40) and (3.43), leads to


 
n
1 X 1
I α̈ +  ρV S CNαi (xcpi − xcm )2 + ṁ(xnozzle − xcm )2 + I˙ α̇ + ρV 2 SCNα (xcp − xcm )α = 0 . (3.47)
2 i=1
2

The coefficients of α̇ and α in (3.47) are named the damping and stabilizing moment coefficients [28, 42]
– CMd and CMs , respectively.
As expected, (3.47) represents a damped harmonic oscillator, generically formulated as

α̈(t) + 2ξωn α̇(t) + wn2 α(t) = 0 , (3.48)

where ξ is the damping ratio and ωn is the natural frequency. If ξ < 1 the rocket is underdamped
and it will oscillate until the angle of attack decreases to zero; if ξ = 1 the rocket is critically damped
and it returns smoothly to zero angle of attack; if ξ > 1 the rocket is overdamped and, although the
disturbance decreases, it always flies with some angle of attack. The natural frequency tells how fast are
these oscillations. For the case in study, comparing (3.47) and (3.48), the damping ratio and natural
frequency are given by:

CM
ξ= p d , (3.49)
2 ICMs
r
CMs
ωn = . (3.50)
I

As we can deduce from (3.50), the oscillations of the rocket become faster if we move the CM toward the
nose (increasing the CMs ). On the other hand, increasing the inertia of the rocket makes the oscillations
slower, as expected.

23
24
Chapter 4

Rocket design

The proposed structure for a general multi-stage rocket developed in the present thesis is shown in
Figure 4.1. Although the figure represents a three-stage rocket, this structure can be extended to any
number of stages and was developed with the goal of modeling several types of rockets, specially the
most common. For each stage the fundamental components are: a connector (or nose, in the case of
the last stage), a body tube and fins. Internally, this structure only considers the motors since data
from additional components can be joined with the body tube properties. The last stage includes the
nose and in the first stage the rocket can also be supplied with an internal cluster of motors or external
boosters in order to achieve parallel staging. The reference from which the position and properties of the
components are specified was defined as being the top of the respective component (this will be explained
in Section 5.2).

3rd Stage 2nd Stage 1st Stage

Figure 4.1: Multi-stage rocket design.

The rocket’s CM, moment of inertia, CP and drag coefficient may depend and change according to
the mass, shape and position of the several components. There are cases where the rocket’s properties
are impossible to determine directly due to lack of information. Therefore, this chapter presents methods
to find or estimate them for the main elements of the rocket. Both these values and the properties taken
directly from measurements of the rocket (that can still be used), must be defined in the respective list
inside the general structure developed for the model of the rocket (see Section 5.2). It must be emphasized
that, although the following methods are focused in model rocketry, the developed simulator can still be
used to process other types of rockets or models if the input data is coherently employed.

25
4.1 Mass

During the ignition and engine burnout phases, the propellant is expelled through the nozzle to generate
thrust and, also when the delay charge burns, the mass of the motor decreases over time. Thus, the
rocket’s mass, in each instant, is found from

m(t) = ms (t) + mm (t) , (4.1)

where ms (t) is the structural mass (without the motor(s)), changing from stage to stage, and mm (t)
is the mass of the motor(s). Considering one motor, structured as represented in Figure 2.1, mm (t) is
obtained by

mm (t) = mf + mp (t) + md (t) + me (t)


Z tb Z tb +td
= mf + mp0 − ṁp (t) dt + md0 − ṁd dt + me (t) , (4.2)
0 tb

where mf is the motor’s final mass, mp0 and md0 are, respectively, the propellant and delay charge initial
masses, tb and td are the burnout and delay times, ṁp and ṁd are the propellant and delay charge mass
burning rates and me is the ejection charge mass. Note that in (4.2) we are considering the entire motor
since we have to estimate the mass of the delay charge. However, the structural (casing) and variable
masses must be defined in different lists under the rocket structure (see Section 5.2).
The propellant’s mass burning rate is given by [39]

Tp (t)
ṁp (t) = , (4.3)
g0 Isp

where g0 is the standard sea level acceleration of the Earth, Tp (t) is the thrust profile of the motor and
Isp is the specific impulse, which represents the linear momentum produced per unit weight of propellant
consumed (measured in seconds) [25]. Therefore, the specific impulse can be found from

I
Isp = , (4.4)
mp0 g0

where the total impulse, I, is


Z tb
I= Tp (t) dt . (4.5)
0

Again from the motor design in Figure 2.1, the delay charge mass burning rate may be estimated as

md0 mm0 − mp0 − me − mf


ṁd = = , (4.6)
td td

where mm0 is the motor’s initial mass.


The ejection charge, consisting of black powder, is consumed instantaneously at t = tb + td and, for

26
model rocket motors, it ranges from 0.3 g to 0.5 g [19]. In alternative, knowing the gases from the burst
must over pressurize the rocket and assuming that all the mass turns into gas, we can estimate me from
the ideal gas law

me
pv = R̄Tbp , (4.7)
Mbp

where v is the internal rocket’s volume, Mbp = 34.75 g mol−1 is the molecular weight of black powder and
Tbp = 2600 K its combustion temperature [44]. Since p = F/Sint (where F is the required force to open
the rocket and Sint is the interior sectional area) and v = lint Sint (being lint the interior length), (4.7)
yields

Mbp
me = F lint . (4.8)
R̄Tc

4.2 Center of mass

We consider the rocket as being an axisymmetric body, so the CM lies on the longitudinal axis and its
location from the reference point (nose tip) is given, for n pieces, by [43]

n
P
xcmi mi
i=1
xcm = n
P , (4.9)
mi
i=1

where xcmi is the CM of piece i of the rocket (also from the reference) and mi its respective mass.
Considering that the rocket is made from an homogeneous material, the CM of each component
matches its centroid [43]. Since these parts are, usually, solids of revolution or other common shapes,
their centroids are easily known. Thus, we can get xcmi by adding the CM (or centroid) given from the
top of the component with the distance from the nose tip to the same component. Again, we alert that
the following methods are only applied to homogeneous bodies.

Nose

The cone’s CM, from the nose’s apex, is [43]

3
xcmcone = ln , (4.10)
4

where ln is the nose length. For other shapes, the nose’s CM is found from
Z ln
1
xcmnose = ln − πxr2 (x) dx , (4.11)
vnose 0

27
R ln
where vnose = 0
πr2 (x) dx is the nose volume and r(x) its respective profile given, for the remaining
shapes, as (see Appendix A for deduction of the ogive’s profile)
!
d x2
rparabolic (x) = 1− 2 , (4.12)
2 ln
s 2 !
ln2 d ln2 d
rogive (x) = + − x2 − − , (4.13)
d 4 d 4

where d is the nose base diameter which is equal to the body tube diameter in case there is only one
stage.

Body tube

The body tube, since it is cylindrically shaped, has its centroid located at half of its length. Hence, its
CM is given as

lb
xcmb = xb + , (4.14)
2

where xb is the distance from the nose’s apex to the beginning of the tube and lb its length.

Connector

Connectors consist of hollow cones sections. The CM of these parts, relative to the nose tip, is given by
(see Appendix B)
!
lc dcf + 2dca
xcmc = xc + , (4.15)
3 dcf + dca

where xc is the distance of this part from the nose tip, lc is the connector’s length and dcf and dca are,
respectively, the fore and aft diameters of the connector.

Fins

The fins’ CM (assumed as flat plates) is

xcmf in = xf in + x̄f in , (4.16)

where xf in is the distance from the nose tip up to the root of the leading edge and x̄f in is the fin’s
centroid.
Since the fins may exhibit several different shapes, the Barrowman Method [45], described and used
in Section 4.4 to find the CP, simplifies their geometry to trapezoids with the root and tip edges parallel
and the same or slightly less area than the original fin [45] (see Figure 4.2). With this geometry it is
possible to get many shapes (including rectangles and triangles) by selecting different values for the fin’s
dimensions. Therefore, dividing the generalized shape in two triangles (see Figure 4.2), x̄f in can be found

28
b

x
s
cr
41

42 ct

Figure 4.2: Generalized fin geometry and respective division to find its centroid.

from the centroids of these two parts as

A41 x̄41 + A42 x̄42


x̄f in = , (4.17)
A41 + A42

where A41 and A42 are the areas of the triangles and x̄41 and x̄42 their centroids. Considering x1 ,
x2 and x3 as the coordinates of the vertexes of a triangle along the x-axis, its centroid’s x coordinate is
given by

x1 + x2 + x3
x̄4 = . (4.18)
3

Then, from the generalized fin geometry (see Figure 4.2), and placing a local reference frame at the point
where the root chord and the leading edge meet, the centroids of the two triangles are:

s + cr
x̄41 = , (4.19)
3
cr + 2s + ct
x̄42 = . (4.20)
3

Motor

During the burnout and coasting phases, the motor is losing mass. Thus, recalling the motor constituents
from Section 2.2, its CM (including the casing mass) is changing over time as

1 h i
xcmm (t) = xm + xcmf mf + xcmp (t)mp (t) + xcmd (t)md (t) + xcme (t)me (t) , (4.21)
mm (t)

where xm is the distance from the nose tip to the top of the motor, xcmf is the motor’s final CM (in this
case equal to half of its length) and xcmp , xcmd and xcme are the propellant, delay and ejection charge’s
CM from the top of the motor, respectively. These are found from:

xcmp (t) = lg + le (t) + ld (t) + lp (t)/2 , for t ≤ tb , (4.22)

xcmd (t) = lg + le (t) + ld (t)/2 , for t ≤ tb + td , (4.23)

xcme (t) = lg + le (t)/2 , for t ≤ tb + td , (4.24)

29
where lg is a gap considered as the thickness of the clay cap plus the empty space at the top of the motor
and lp , ld and le are the propellant, delay and ejection charge lengths, respectively. Since every motor’s
casing is cylindrically shaped, with internal radius rintm , these lengths are found from

2
m = ρv = ρπrintm
l, (4.25)

which gives:

mp (t)
lp (t) = 2 , (4.26)
ρp πrintm

md (t)
ld (t) = 2 , (4.27)
ρd πrintm

me (t)
le (t) = 2 , (4.28)
ρe πrintm

where ρp is the propellant’s density (equal to 1750 kg m−3 for black powder and 1520 kg m−3 to 1750 kg m−3
for composite propellant [44]) and ρd is the delay charge density. Since the ejection charge consists of
black powder, its density is equal to 1750 kg m−3 for any of the two types of motors. The delay charge
density may be estimated from (4.27) with ld and md evaluated at t = 0, being the first predicted by

ld |t=0 ≈ lm − lnozzle − lp |t=0 − le |t=0 − lg , (4.29)

where lm is the motor’s length and lnozzle is the length of the nozzle (from the rear of the motor to the
propellant).
If the motor is a core burning, its propellant’s length is constant and found from (4.26) with mp
2 2 2
evaluated at t = 0 and rintm
= rext p
− rintp
, where rextp and rintp are the initial external and internal
propellant radius, respectively.

4.3 Moments of inertia

From Section 3.3.1, we only need to find the transversal moment of inertia relative to the CM (I), and not
the longitudinal moment of inertia, since the roll motion is not considered in the attitude of the rocket
but just the pitch motion from the developed model of the angle of attack. The moment of inertia may
be found from

n
X
I= Ii , (4.30)
i=1

where Ii is the moment of inertia of a given component relative to the rocket’s CM found from the parallel
axis theorem [43]

Ii = I¯i + mi (xcmi − xcm )2 , (4.31)

30
where I¯i is the central moment of inertia of the respective component, which is known for simple solids
of revolution or can be found by integration.

Nose

The moment of inertia of the nose cone relative to the nose tip is [43]
 
3 1 2
Icone = mn r + ln2 , (4.32)
5 4 n

where mn is the nose mass and rn its base radius. In order to have its moment of inertia relative to
the rocket CM, we must apply (4.31) two times: firstly to determine its central moment of inertia and,
secondly, taking this value, to find the moment of inertia relative to the rocket CM.
To other nose shapes, the moment of inertia relative to the base diameter axis is determined by [43]
Z ln  
1 2
In = r (x) + x2 ρn πr2 (x) dx , (4.33)
0 4

where ρn is the nose density and r(x) is given by (4.12) or (4.13). Their moments of inertia relative to
the rocket CM is then found by the same method described to the conical nose.

Body tube

The body tube is a hollow cylinder. Thus, its central moment of inertia of the body tube is [43]

1 h i
I¯b = 2
mb 3(rext 2
+ rint ) + lb2 , (4.34)
12

where rext and rint are the external and internal radius of the body, respectively, and mb is the body
tube’s mass.

Connector

The connector’s transversal moment of inertia relative to the smaller diameter is (see Appendix B)
!
1 dcf + 3dca
Ic = mc lc2 , (4.35)
6 dcf + dca

where mc is the connector’s mass.

Fins

Since the fins are considered as flat plates made of an homogeneous material, their moments of inertia
can be found from [43]

If in = ρf tf If0 in , (4.36)

where ρf is the fin’s density, tf is the fin’s thickness and If0 in is the area moment of inertia. Like in
Section 4.2, dividing the fin in two triangular pieces and applying the parallel axis theorem to each one

31
of them, If0 in relative to the top of the fin is determined by

If0 in = I¯4
0
1
+ A41 x̄241 + I¯4
0
1
+ A42 x̄242 (4.37)

where the central area moments of inertia of the two triangular pieces, having in mind Figure 4.2, are
given by [46]:

c3r b − c2r sb + s2 cr b
I¯4
0
= , (4.38)
1
36

c3 b − c2t (ct + s − cr )b + (ct + s − cr )2 ct b


I¯4
0
= t . (4.39)
2
36

Motor

In order to find the motor’s central moment of inertia, we must add the moments of inertia of each
component relative to the motor’s CM applying (4.31), where the principal moments of inertia of the
casing, propellant, delay and ejection charges are given from (4.34).

4.4 Center of pressure and CNα

Since its publication, the Barrowman Method provides a series of reliable equations that have been widely
used to estimate the aerodynamic characteristics of model and high power rockets. This method makes
several assumptions in order to find those equations [45]:

1. The angle of attack is near zero;


2. Viscous forces are negligible;
3. The flow is steady, irrotational and subsonic;
4. The rocket is thin compared to its length;
5. The nose of the rocket comes smoothly to a point;
6. The rocket is an axially symmetric rigid body;
7. The fins are thin flat plates.

Although there are some restrictions, model and high power rockets meet these requirements specially
during and shortly after lift-off which, not considering wind perturbations, is when the CP and CNα
play an important role, according to Section 3.3.1. Also, since model rockets have, usually, a general
geometry, the Barrowman Method divides the rocket in simple geometric parts (like in Section 4.2), finds
its properties and recombine them to find the final results. The third assumption is invalid for sounding
rockets due to the high velocities that they reach. In this case, and as the CP is used to determine the
angle of attack, we consider that the rocket oscillations are already damped before reaching the transonic
regime. At such high speeds the rocket has to fly without angle of attack in order to minimize the drag
forces applied on its structure.
According to the Barrowman Method, for subsonic velocities, CNα depends only on the shape of the

32
rocket and is found by summing all the derivatives of the parts of the rocket as [45]

n
X
CNα = CNα i . (4.40)
i=1

The CP is found from [45]


Pn
i=1 CNα i xcpi
xcp = , (4.41)
Cnα

where xcpi is the CP location of each component of the rocket from the reference point. From these
equations we get a constant CP, which is a satisfactory result since the rocket meets the first and third
assumptions of this method at least until it stops swinging.

Nose

For the most common noses geometries used in model rocketry, the Barrowman Method finds that [45]:

CNαnose = 2 , (4.42)
2ln
xcpcone = , (4.43)
3
ln
xcpparaboloid = . (4.44)
2
xcpogive = 0.466ln , (4.45)

Body tube

Since the cylindrical body has constant cross sectional area, according to the Barrowman Method and
its assumptions, it does not produce lift [45]. Thus, Cnαbody = 0.

Connector

For connectors, from the Barrowman Method, we have [45]


" 2  2 #
dca dcf
CNαc = 2 − , (4.46)
d d

where d is the rocket’s diameter at the base of the nose. If we have a reducer or boat-tail, then CNαc
comes out negative.
The CP location is given by
 
dcf
lc  1− dca

xcpc = xc + 1 +  2  . (4.47)
3 d
1 − dcfca

Fins

The Barrowman Method only allows the rocket to have 3, 4 or 6 fins [45]. Recalling the fin geometry in

33
b

s
cr
lc̄

ct

Figure 4.3: Generalized fin geometry.

Figure 4.3, this method founds that

4nkf ( db )2
CNαf in = q , (4.48)
1 + 1 + ( cr2l+cc̄
t
) 2

where n is the number of fins and kf is the fin interference factor, due to the presence of the body, given
by:

rext
kf = 1 + , for n = 3, 4 , (4.49)
rext + b
5rext
kf = 1 + , for n = 6 , (4.50)
rext + b

where rext is the external rocket radius between the fins.


Fin’s CP is found by [45]
 
s(cr + 2ct ) 1 cr ct
xcpf in = xf in + + cr + ct − . (4.51)
3(cr + ct ) 6 cr + ct

4.5 Drag coefficient

The drag force opposes the true speed direction and appears due to several effects of the flow around the
rocket. As seen in (3.33), the drag coefficient contributes to the aerodynamic force and it may change
throughout the flight. It can be estimated by adding all the drag coefficients of the external components
of the rocket and, for each one of them, the drag may come from different sources: pressure drag, skin
friction drag, base drag and wave drag.
Pressure drag arises due to the distribution of normal forces on the components and the skin friction
drag, on the other hand, is a tangential force generated by the viscosity of the flow that creates a boundary
layer on the surface of the rocket [47]. Therefore, assuming a flight without angle of attack, the nose,
connectors and fins generate these two kinds of drag and the body tubes of each stage only produce skin
friction drag. For conical noses, the pressure drag coefficient variation is represented in Figure 4.4. Since
the drag coefficient for cones describe, approximately, a linear variation with the half vertex angle (ε)

34
Figure 4.4: Pressure drag coefficient of wedges, cones and similar shapes as a function of the half-vertex
angle (ε) [47].

from 0 to 90°, the pressure drag coefficient may be found from

CDp (ε) = 0.233 + 0.011ε . (4.52)

The total drag coefficient for the nose and connector sections is

Sb Sw
CDnose/connect = CDp + CDf ,, (4.53)
S S

where CDf is the skin friction drag coefficient (determined in Section 4.5.1), Sb is the nose base area and
Sw is the wetted area. Note that, since the connector sections describe a cone without the apex, Sb , in
that case, must represent the difference between the aft and fore areas in order to discard the pressure
drag already included in the nose coefficient.
The fin’s drag coefficient, is given, for the two surfaces, as [48]
 
tf Sw
CDf ins = 2CDf 1+2 , (4.54)
c̄ S

where c̄ and tf are the fin’s average chord and thickness, respectively, and the wetted area, Sw , must take
into account the two sides of the fin. Considering a flight with angle of attack, there is an extra source
of drag created by the fins, the induced drag. It can be estimated by [48]

CL2 Sf
CDind = , (4.55)
πΛe S

where Λ is the fin aspect ratio, e ≤ 1 is the Oswald efficiency factor (e = 1 for elliptical wings), Sf is the
fin planform area and CL is the fin’s lift coefficient that, assuming the fins as flat plates at low angles of

35
attack, may be approximated by (3.37). Since the flow around the fuselage interacts with the flow passing
through the fins, it is created an interference drag. This extra drag can be interpreted as coming from
the air flowing on fins that are extended into the fuselage [48]. Then, the interference drag coefficient is
measurable by

cr rext
CDint = 2CDf n, (4.56)
S

where n is the number of fins and cr is the fin’s root chord.


Due to the lower pressure at the rear of the rocket, the boundary layer separates from the surface and
increases the drag. This fact originates the base drag which can be found by [47]
 3
0.029 dbt
CDbase = p , (4.57)
CDbody d

where CDbody is the drag coefficient of the fore body (not counting with the fins) and db is the boat-tail
aft diameter. If dbt = d, the rocket does not have a boat-tail and the base drag coefficient is maximum,
as expected.
Model rockets are provided with launch lugs that guide the rocket through the rods. They are an
extra source of drag estimated as [48]

1.2Sll + 0.0045Swll
CDll = , (4.58)
S

where Sll and Swll are, respectively, the sectional and wetted (inner and outer surfaces) areas of the
launch lugs.

4.5.1 Skin friction drag

Skin friction drag changes according the regime of the flow. In turbulent flow, the skin friction drag is
higher than in laminar flow (although turbulence, in some cases, is desirable in order to delay separation
and decrease the pressure drag). As the air hits the rocket’s surface, the flow is laminar and rapidly turns
to turbulent with increasing Reynolds number. Between this two regimes appears a laminar-turbulent
transition when critical Reynolds is reached (Recr ) and the flow becomes fully turbulent above transition
Reynolds (Retr ).
Reynolds number, representing the relative order of magnitude between inertial and viscous forces of
the fluid, is defined as [49]

ρV L
ReL = , (4.59)
µ

where L is the surface’s length which must be selected according the respective part of the rocket and µ
is the dynamic viscosity of the air found from the empirical Sutherland’s law as [49]

T 3/2
µ = 1.458 × 10−6 . (4.60)
T + 110.4

36
Since transition is hard to predict, a way to find the skin friction drag coefficient in a flat plate
(considering zero pressure gradient) is to assume Recr ≈ Retr , yielding [49]

 Retr h   i
CDf = CDturb l
−CDturb xtr − CDlam xtr
ReL
Retr h i
= 0.074Re−0.2
L − 0.074Re −0.2
tr − 1.33Re−0.5
tr , (4.61)
ReL

where Retr ≈ 3 × 105 to 106 . Equation (4.61) determines the turbulent drag coefficient for the entire
plate and replaces the turbulent contribution by the corresponding laminar value up to the point (xtr )
where the transition occurs [49].

4.5.2 Compressibility effects

Mach number is the dimensionless parameter [49]

V
M= , (4.62)
a

where a is the speed of sound given as

p
a= γRT , (4.63)

where γ ≈ 1.4 for air is the ratio of specific heats at constant pressure and volume.
From M > 0.3, the flow no longer can be assumed as incompressible since there are changes in density
and temperature and shock waves begin to increase the drag. The Prandtl-Glauert rule is commonly
used to relate incompressible and compressible coefficients (CD and CDc , respectively) for slender and
planar bodies [50]. This rule is given as

CD
CDc = p , (4.64)
|1 − M2 |

that must be applied between 0.3 < M < 0.7 and 1.2 < M < 5. In transonic and low supersonic regimes
(0.7 < M < 1.2), an interpolated function should be used to estimate the compressible drag coefficient
and avoid the singularity at M = 1. For a conical nose, the wave drag due to shock waves can be estimated
by the empirical equation [51]
   1.69
0.096 ε
CDw = 0.083 + for M > 1 , (4.65)
M2 10

where ε is again the half vertex angle in degrees. Equation (4.65) should match the subsonic drag
coefficient at M = 0.7 which is when the transonic effects begin to appear.

37
There are also corrections for the skin friction coefficient. These relations are [47]:

CDlamc = CDlam , (4.66)

CDturbc = CDturb (1 − 0.12M2 ) , (4.67)

for subsonic (M < 1) and

CDlam
CDlamc = , (4.68)
(1 + 0.045M2 )0.25
CDturb
CDturbc = , (4.69)
(1 + 0.15M2 )0.58

for supersonic (M > 1) regimes.

4.5.3 Recovery device

After the apogee, or when it is desired, the recovery system is deployed and the main source of drag
comes from the parachute or streamer. Since the respective drag coefficient depends on the shape of the
recovery device, Table 4.1 lists several ranges of values for common parachute designs. The tabulated

Shape dproj /dref CD


Flat circular 0.67 – 0.7 0.75 – 0.80
Conical 0.7 0.75 – 0.90
Biconical 0.7 0.75 – 0.92
Triconical 0.7 0.80 – 0.96
Hemispherical 0.66 0.62 – 0.77
Annular 0.94 0.85 – 0.95
Cross 0.66 – 0.72 0.60 – 0.85

Table 4.1: Solid textile parachutes’ projected to reference diameter (dproj /dref ) and respective drag
coefficient [52].

coefficients are associated to the reference area which is given as the total area of the canopy including
the vent or other openings. To find the projected area when the parachute is inflated, are also listed the
values for the ratio between the projected and reference diameters.
If the recovery device is a streamer, its drag coefficient may be found from the Open Rocket docu-
mentation that formulates the empirical function [15]
  
ρs + 25 ls + 1
CDstreamer = 0.034 , (4.70)
105 ls

where ρs is the streamer’s material density (in kg/m3 ) and ls is the streamer’s length (in m).

38
Chapter 5

Rocket trajectory simulator

The developed rocket trajectory simulator aims to:

ˆ determine the rocket trajectory and respective flight data given known/estimated inputs;
ˆ give the user flexibility to change the default trajectory and rocket’s models;
ˆ enable to simulate from micro rockets up to sounding rockets’ models;
ˆ forecast the landing region from a Monte Carlo simulation;
ˆ optimize the rocket characteristics.

This tool is developed in Mathematica ® [53] since it is a powerful language in functional programming
and to manipulate lists. Due to these characteristics, this language gives flexibility to implement the
features we desire: structure of lists for the rocket, listable databases and modular programming. Al-
though Mathematica ® becomes slower as it is an interpreted language, the mentioned advantages surpass
this handicap. Besides the following sections aim to explain the methods applied to develop the simu-
lator, their goal is also to expose some of the advantages Mathematica ® offers while programming and
manipulating data.

5.1 Simulator description

The simulator’s main function is RocketTrajectory (see Figure 5.1) which integrates the trajectory equa-
tions and receives all the required inputs. This function calls the RocketProperties function that computes
every parameter of the rocket over time. These properties are determined by combining the data of the
several stages regarding the developed structure that gathers the data of the rocket and considering the
instants of the ignitions and ejections (see Section 5.2).
The most relevant inputs are the rocket structure, the atmospheric density, the wind and the CD
models that, since they are external parameters, the user is allowed to define or change them according
to his needs. The outputs from the simulation are three functions giving the position (one for each
Cartesian coordinate) plus a function giving the angle of attack. These are computed in three steps: one
for the constrained trajectory due to the guiding rod, another for the climbing phase and the last one for

39
Inputs

RocketT rajectory RocketP roperties

Outputs

Figure 5.1: Simulator’s algorithm.

the recovery trajectory. The angle of attack starts to be computed only in the second phase, since it is
when the rocket suffers the first perturbation from the wind. After this process, the three parts of the
trajectory are merged together and, having these time dependent functions, it is possible to determine
its derivatives and get other flight data such as total velocity, total acceleration, Mach and flight path
angle.
If there are windy conditions, the rocket needs to speed up until it becomes stable. Therefore, a
guiding rod is required to restrain the trajectory during the launch in the desired direction. The rod’s
length is needed as input in order to determine the instant after which the rocket’s trajectory becomes
unrestricted and the model of the angle of attack starts to be computed. The user must be careful in
selecting the rod’s length since he should know the required length of the rod that allows a given thrust to
overcome the wind conditions and avoid an abrupt increase of the angle of attack over time. Nevertheless,
it is also recommended that the rocket’s static margin is previously determined to assure it stays always
positive (stable rocket).
During the simulation the rocket is treated as a particle since there are only three degrees of freedom
(x, y and z) and the model developed for the attitude is only used to estimate the loss of thrust along the
true speed direction. Therefore, since the rocket’s nose is always a steady point in the rocket’s reference
frame (unlike the CM), it is used to give the rocket’s position over time and, consequently, the initial
nose distance to the ground is required as input.
The rockets used in the simulator must always have three or more fins, boosters or cluster motors
because we did not consider the case when the rocket has two components of each type. In this situation,
the rocket acquires two different transversal inertias which is incompatible with the model developed for
the angle of attack that assumes the rocket has only one transversal moment of inertia.
As default, the recovery system is deployed at the apogee. However, there is an optional argument
that orders the simulator to start to compute the recovery trajectory at the ejection instant determined
by the data within the rocket structure.

40
5.2 Rocket assembly

The list concerning the model of the rocket has a specific structure in order to easily access its several
data, to be quickly processed by the RocketProperties function and to allow any number of stages. Every
parameter inside this structure may be defined directly by the user (from measurements or other available
data) and some of them can be estimated from the functions already included in the simulator that apply
the methods presented in Chapter 4.
Figure 5.2 shows the scheme in which the assembling of the rocket’s data is structured. The stages

Rocket
Assembly List
Boosters List
Number
Position
Thrust Data List
Structure Data List
Delay Booster
Stages List
Stage 1 List
Stage 2 List

Stage i List
Motor List
Body List
Fins List
Connector/Nose List
Delays List
Miscellany List
Motor Cluster List
Number Motors
Motor List
Recovery List
Reference Area
Drag Coefficient

Figure 5.2: Rocket assembly list structure.

list contains as many lists as the number of stages. Each one of these lists, having in mind the rocket’s
structure from Chapter 4, holds other lists for the rocket’s components plus a list to indicate the time
intervals between events (explained later).
The motor list is described in Figure 5.3. Its first entry is the distance between the top of the motor
and the top of the body tube. Likewise, in the thrust data list, the first element is the distance between
the top of the combustion chamber and the top of the motor. This list contains time dependent functions
since the propellant is burning over time. The moments of inertia should also be given relative to the top
of the combustion chamber in order to have a steady reference. Otherwise, considering the propellant’s
moment of inertia, its reference would change over time. Although the simulator does not require the
longitudinal moment of inertia, we keep an entry for this property because an improved version may need
it in the future. The body list is defined as depicted in Figure 5.4.

41
Motor List
Position
Thrust Data List
Position
Burnout
Isp
Thrust (t)
Mass (t)
CG (t)
Inertia List
Transv. Inertia (t)
Long. Inertia (t)
Structure Data List
Mass
CG
Inertia List
Transv. Inertia
Long. Inertia
Diameter

Figure 5.3: Motor list structure.

Body List

Mass CG Inertia List CP CNα Length Diameter

Transv. Inertia Long. Inertia

Figure 5.4: Body list structure.

In the fins list, showed in Figure 5.5, there is an additional CM in order to find the distance of the
fin’s CM from the rocket’s CM, which is required for the parallel axis theorem. Also, the CP and the
CNα in this list are assumed to be the values resulting from the ensemble of fins in each stage, as it was
determined in Section 4.4.

Fins List
Number
Position
Fins Data List
Mass
CG List
Axial CG
Radial CG
Inertia List
Transv. Inertia
Long. Inertia
CP
CNα
Root chord
Tip chord
Span
Thickness

Figure 5.5: Fins list structure.

42
The connector (or nose) list is structured as depicted in Figure 5.6. In the delays list, defined in
Figure 5.7, the delay of the body represents the time when the respective stage’s body tube is discarded
after the burnout; the delay of the connector gives the instant when this component is discarded after the
previous delay (if there is no connector, this delay must be zero); the third delay gives the time interval
until the next stage is ignited. When the delays list belongs to the nose, the sum of these three delays
represent the coasting time until the recovery system is deployed if the ejection charge option is chosen
in the RocketTrajectory function.

Connector/Nose List

Mass CG Inertia List CP CNα Length

Transv. Inertia Long. Inertia

Figure 5.6: Connector or nose list structure.

Since the boosters are external components, the structure data list, within the boosters list (see
Figure 5.2), contains additional elements: the lengths of the cylindrical tube and nose structures, the CP
and the CNα . The last two must give the respective property for the complete structure of the booster.
All the boosters must also be identical and, when adding them, the user should pay attention if the
new rocket’s CM does not surpass the CP. The delay of the boosters represents the time between their
burnout and the instant they are discarded, which, at most, should be the instant when the first stage’s
body tube is also discarded. If the rocket does not have boosters, their number must be zero and the
RocketProperties function ignores the other entries within the boosters list.

Delays List

Delay Body Delay Connector/Nose Delay To Next Ignition

Figure 5.7: Delays list structure.

Inside the miscellany list (see Figure 5.2), the cluster list allows to define a cluster of identical motors
around the first stage’s motor and equally spaced between them. The recovery list contains the properties
of the recovery device (reference area and CD ) to be used during the descending phase of the simulation.

5.3 Atmospheric model

The atmosphere is an important element influencing the trajectory since it is the fluid in which rockets
fly. As the atmospheric properties change throughout the flight, they must be known at every point of
the rocket trajectory. To simulate trajectories of model rockets, as these reach low apogees, it can be
important to take into account some atmospheric variations due to the location of the launch and due
to the weather conditions. As this work includes simulation of sounding rockets trajectories, it is also
required to know the atmospheric properties until a high altitude.
One of the existing atmospheric models is the Committee on Space Research (COSPAR) International

43
Reference Atmosphere 1986 (CIRA-86). This is an empirical model ranging from 0 km to 2000 km height
[54]. Based on this model, after its publication, occurred several developments. One set of the developed
tables consists of the monthly zonal mean values of temperature from 0 km to 120 km height with almost
global coverage (80°N to 80°S) [54]. These tables also show pressure data but it is only changing with
altitude. Due to the altitude range, latitude coverage and monthly variation, this is the implemented
model to approximate the atmospheric properties up to the lower thermosphere (high enough to small
sounding rockets) in every populated places on Earth all year long.
Corrections are applied in this model as CIRA-86 gives wrong temperatures (999.9 K) at the two
lowest altitude layers (0.1 km and 2.2 km) for each month from 70°S to 80°S. Therefore, interpolating all
data above 2.2 km in the corresponding latitudes, these values are replaced by extrapolated temperatures
at the respective altitudes. Also, since pressure and temperature at the surface are not provided, the
corrected model is improved by including these properties found by extrapolation at 0 km height. After we
get this upgraded version of CIRA-86, its data is rearranged in lists structured as depicted in Figure 5.8.
With this structure, pressure and temperature are easily interpolated in altitude, month and latitude and

Cira-86 Upgraded
List
Data 1 List
Data 2 List

Data i List
Variables List
Altitude
Month
Latitude
Pressure
Temperature

Figure 5.8: CIRA-86 upgraded version structure.

the obtained three variable functions are included in the atmospheric database. Using these functions,
the density is found from (3.3) assuming the atmosphere is an ideal gas (as gases often do approach ideal
gas behavior [32]).
This model is stored in a database in order to be easily explored and get access to its properties and
details. Also, this database allows to define other atmospheric models and new functions representing
extra atmospheric parameters. The atmospheric database is a listable function (i.e., applies the function
to each element in a list), named AtmosphereData, thus it has not a specific structure. We can access its
data by providing the name of the model and the required properties. For example,

AtmosphereData [ ‘ ‘ Cira86 ’ ’ , { ‘ ‘ Temperature ’ ’ , ‘ ‘ Density ’ ’ } ] ,

gives a list with the temperature and density functions for the CIRA-86 model. If another model is
stored, we just have to change the name of the model to get these properties. Also, if we want to know
the units of each property (or another information previously saved), we must type

AtmosphereData [ ‘ ‘ Cira86 ’ ’ , { ‘ ‘ Temperature ’ ’ , ‘ ‘ Density ’ ’ } , ‘ ‘ Units ’ ’ ] .

44
At the launch site, the local conditions must be taken into account. Therefore, the simulator contains
functions that receive the atmospheric models and adjust them in order to match the measured pressure
and temperature at the corresponding altitude besides tending exponentially (at a chosen rate) to the
original model at higher altitudes. Taking as example the correction for the temperature, this function
is given by
 
 zl − z
Tc (z) = Tm (z) + Tzl − Tm (zl ) exp for z ≥ zl , (5.1)
k

where Tc is the corrected temperature, Tm is the temperature from the atmospheric model, Tzl is the
measured temperature at the launch altitude (zl ) and k is a constant to define the exponential rate. The
CIRA-86 temperature (at 40°N in June) and the corresponding correction for a local temperature of 300 K
at 100 m above sea level with k = 4 is depicted in Figure 5.9. After getting the adjusted pressure and
temperature functions, we use (3.3) to get the density also dependent on the local weather conditions.

Temperature (K)
300 CIRA-86
CIRA-86 with local conditions
280

260

240

220
Altitude (km)
5 10 15 20 25

Figure 5.9: Temperature profiles for the CIRA-86 model.

5.4 Wind model

The developed wind model is a combination between the functions for the ABL and the wind data provided
in the CIRA-86 model at higher altitudes. Therefore, to model the ABL according to Section 3.1.1,
there is required the local wind speed, its respective measurement altitude, the terrain’s specification
(given by the surface roughness length – z0 ), the launch latitude and the constants κ and c. Since the
RocketTrajectory function requires the wind model to be defined by its three wind speed components,
the wind direction (downwind angle relative to north in the clockwise direction) is also necessary. This
model does not consider vertical wind.
The user must be careful when increasing the roughness length because it will extend the zero wind
region up to a higher height. If this region is greater or equal than the height of the guiding rods, the
simulator gives a zero angle of attack over time since there is no wind at the instant supposed to determine
the initial condition of the angle of attack.
There are optional arguments in the wind model function in order to determine the stability of the
ABL. The default option is a neutral atmosphere. If the atmospheric stability is known, the user must
indicate the respective conditions (stable or unstable) and give an estimated guess for the Obukhov

45
length (L) based on Table 3.2. Otherwise, giving qualitative information about the ABL, the wind model
function deduces the atmospheric stability (as described in the Section 3.1). This is accomplished through
other optional arguments: time of day (day/night), weather (clear/cloudy) and temperature gradient
(positive/negative). In this case, the Obukhov length is also required and its possible values must comply
with the stability found by the wind model.
In Figures 5.10, 5.11 and 5.12 are represented the wind profiles for a neutral, stable and unstable

Wind speed (m/s)


6
5
4
3
2
1
Altitude (m)
200 400 600 800 1000 1200

Figure 5.10: Neutral wind speed profile for the ABL.

Wind speed (m/s)

Altitude (m)
50 100 150 200 250 300

Figure 5.11: Stable wind speed profile for the ABL.

Wind speed (m/s)


3.5
3.0
2.5
2.0
1.5
1.0
0.5
Altitude (m)
200 400 600 800 1000

Figure 5.12: Unstable wind speed profile for the ABL.

ABL, respectively. These profiles were modeled considering a local wind speed of 2 m s−1 at 2 m height,
in a high grass terrain (z0 = 7 cm) at the latitude of 37°N with κ = 0.4 and c = 0.4. Also, L = 125 m for
the stable profile and L = −150 m for the unstable. See Section 6.4.3 in order to observe the influence of
L in the wind profiles.

46
Above the ABL, the wind speed remains constant with height until a certain altitude (defined by the
user) at which the wind starts to approximate the CIRA-86 values at a chosen rate. Since the CIRA-86
model gives westerly and easterly (negative values) winds, the direction of the wind from the ABL also
starts to match the west-east direction. This fit between the two models is accomplished by a vector
sum in which the intensity of the ABL vector begins to decrease above the user-defined altitude and the
CIRA-86 vector’s intensity increases from zero until its original value.

5.4.1 Wind gusts

During the launch the wind may increase suddenly with height, leading the rocket to suffer some wind
gusts. The effect of these perturbations is implemented in the RocketTrajectory function by changing the
solution from the equation of the angle of attack to a new one with its initial angle condition determined
by the rocket and wind gust’s velocities at the perturbation’s instant. Since the angle of attack in the
simulator only affects the thrust force, the effect of the wind gusts can only change the trajectory if the
new solution gives a non zero angle during any burning.
The wind gusts’ input is defined by a list containing as many lists as the number of gusts (see
Figure 5.13). Each one of these lists contains the perturbation’s instant (after the ignition) and the gust

Wind Gusts List

Gust 1 List Gust 2 List Gust i List


{Instant 1, Gust 1 vector} {Instant 2, Gust 2 vector} {Instant i, Gust i vector}

Figure 5.13: Wind gusts list structure.

speed vector. The effect of a wind gust in the angle of attack at 2 s after the ignition is depicted in
Figure 5.14. This perturbation has an intensity of 25 m s−1 along the wind direction.

α (º)

10

t (s)
1 2 3 4 5 6 7

-5

-10

Figure 5.14: Angle of attack until the apogee with a wind gust perturbation.

5.5 Motors data

Some model rocket motors data are stored, by default, in another database. As example, and as described
for the atmospheric database, the motor’s properties may be accessed by

47
MotorData [ ‘ ‘ C6−0 ’ ’ , { ‘ ‘ Thrust ’ ’ , ‘ ‘ Isp ’ ’ , ‘ ‘ Mass ’ ’ , ‘ ‘CG’ ’ , ‘ ‘ I n e r t i a ’ ’ } ] .

Again, other motors can be stored in order to automatically change all the motor’s properties within the
structure of the rocket only by specifying the name of the motor. These new motors may be defined
by any mathematical model the user may require, which gives freedom to change a given property and
observe its influence in the simulation.
For all the model rocket motors already in the database, their properties are determined considering
an end burning combustion, a delay charge (for the ones which possess a delay time) and an ejection
charge. The thrust profiles are obtained by an interpolated function fitting the points given in the data
sheets available on-line at www.thrustcurve.org. As an example, Figure 5.15 represents the obtained

Thrust (N)
30

25

20

15

10

t (s)
0.5 1.0 1.5 2.0

Figure 5.15: Thrust profile for the D12 motor.

thrust curve of the motor D12. Although the main motors’ properties are found from the thrust profile
and other info provided in the data sheets by the methods described in Chapter 4, they depend on other
data which can only be determined by measurements taken from the motors’ dimensions. Therefore, this
data must be inserted in the database in order to automatically compute the properties of the motors.

5.6 Drag coefficient model

The simulator provides a drag coefficient model to be used in the RocketTrajectory function. If de-
sired, another model can be developed which may depend on time, the coordinates variables and their
derivatives. It is also possible to define a constant drag coefficient.
The developed model requires the list containing the structure of the rocket to determine the drag
coefficients for each one of the external components and get the total CD , depending on Mach, Rex and
time, in respect to the rocket’s reference area. The coefficients for each component are computed by
calling previously defined functions that apply the equations from Section 4.5.
Using this drag coefficient model, we are considering that all the components contribute to the skin
friction drag but the pressure drag comes only from the nose, connectors and fins (due to the normal
surface facing the flow). Also, in this model the rocket’s nose is the only wave drag source. To compute
the wave drag coefficient from (4.65), and since it is not defined in the transonic regime bellow M = 1,

48
a fitted function is determined to match the incompressible CD at M = 0.7 and the wave plus the
incompressible CD at M = 1. In order to match smoothly the other functions, this fit is performed by a
third degree polynom. The final drag coefficient function for the conical nose, including the wave drag,
is represented in Figure 5.16.

CD
0.38
0.37
0.36
0.35
0.34
0.33
0.32

Mach
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Figure 5.16: Conical nose drag coefficient with Mach.

5.7 Validation of the simulator

The developed tool requires validation in order to assure the simulator meets the specifications and
correctly computes the expected output. This may be performed by comparing our results with a known
analytic solution from a specific launch.
The altitude over time (until the burnout) from a vertical rocket launch, considering the simplest case
without atmosphere (no drag and no wind) and constant thrust, can be found by [39]
 
ve m0 − ṁt 1
H(t) = (m0 − ṁt) ln + ṁt − g0 t2 , (5.2)
ṁ m0 2

where m0 is the initial mass of the rocket, ṁ is the propellant burning rate and ve = g0 Isp is the
exhaust velocity. Equation (5.2) also ignores the effects of the Earth’s rotation and assumes a constant
acceleration of gravity. Therefore, using a sounding rocket model, supplied with a motor providing 13 kN
of thrust during 32 s (burnout), and specifying the same assumptions in the simulator, we can compare
the analytical and numerical solutions from the relative error shown in Figure 5.17. It can be observed
that the error presents values in the order of magnitude of 10−6 %. Although this error gives reasonable
results under the mentioned conditions, more tests should be performed in order to increase the confidence
in the solutions from the developed simulator.
To test the drag effect, we can make a terminal velocity test for a given rocket in free fall after the
apogee. From an equilibrium between the gravitational force and the drag force as

1 2
mg = ρV SCd , (5.3)
2

49
Relative Error (%)

3. × 10-6

2. × 10-6

1. × 10-6

t (s)
5 10 15 20 25 30

Figure 5.17: Altitude’s relative error between analytical and numerical solution until the burnout.

the terminal velocity can be found by


r
2mg
V = . (5.4)
ρSCd

As an example, from (5.4), we conclude that if the rocket loses mass, it must reach a new velocity.
Therefore, we can also verify if the simulator is computing this quasi-static process. Testing a vertical
launch under the mentioned forces, for a density of 1.225 kg m−3 (equal as the density at the launch
site altitude used in the atmospheric model) and for a model rocket with a drag coefficient of 0.6, a
reference area of 4.9 cm2 (2.5 cm in diameter) and a mass of 59.7 g, we get, from (5.4), 57.0 m s−1 for the
terminal velocity. Decreasing 30 g in the rocket’s mass, in order to verify the simulator’s behavior, the
new terminal velocity comes 40.2 m s−1 .
In Figure 5.18 it is represented the rocket’s free falling velocity from the simulator, 10 s after the lift off,
under the mentioned conditions. As we can see, the rocket tends asymptotically to the terminal velocity
of about −57 m s−1 and, after losing 30 g, it tends to a new terminal velocity of about −40 m s−1 . These
velocities are consistent with the ones deduced in (5.4) and, as expected, the velocity tends asymptotically
to each one of the equilibrium states.

Velocity (m/s)

-10

-20

-30

-40

-50
t (s)
15 20 25 30 35 40

Figure 5.18: Descent velocity from a vertical launch in which the rocket loses mass at 25 s.

In order to test non vertical launches, we may compare them with the solution from a projectile

50
launch. The altitude over time for this type of motion with no drag, no wind and no thrust, is given by

1
z(t) = h0 + vh0 t − g0 t2 , (5.5)
2

where h0 is the initial vertical distance from the ground and vh0 is the initial vertical velocity. Assuming
the projectile is launched from the ground (h0 = 0 m) with an initial total velocity of 30 m s−1 and
70° of elevation, we get the relative error between the analytical and numerical solutions represented in
Figure 5.19. In this test the input concerning the rod’s length is small enough so we can ignore the errors

Relative Error (%)

0.00020

0.00015

0.00010

0.00005

t (s)
1 2 3 4 5

Figure 5.19: Relative error between analytical and numerical solutions until landing from a projectile
launch with 70° of elevation, constant mass, no drag and no wind.

due to the constrained trajectory in the rod. From the values of the relative error, in Figure 5.19, we
conclude the simulator behaves as expected under these conditions. Another verification can be made
from Figure 5.20 which represents the flight path angle during the projectile launch. As expected, the
projectile lifts off and lands with a flight path of 70° from the horizontal reference since its trajectory is
parabolic.

Flight Path (º)

60
40
20
t (s)
1 2 3 4 5
-20
-40
-60

Figure 5.20: Flight path angle from the projectile trajectory with the launch angle (elevation) of 70°.

The results from the tests presented in this section seem to indicate that the simulator behaves as
expected taking into account the respective assumptions for each example. It is foreseen an increase in
those errors when these tests are performed under the full conditions with which the tool was developed.
In that case, although the errors increase due to the variation in the acceleration of gravity with altitude

51
and due the effects of the Earth’s rotation, the numerical solutions should give better approximations to
real-case trajectories.
Other tests that can provide more information about the performance of our tool are real rocket
launches. These were not considered due to the extensive work they demand. Nevertheless, instrumenting
a rocket or taking measurements from the ground to estimate the trajectory through triangulation gives
precious informations that can be compared with the predictions from our simulation tool. Furthermore,
it helps us to deduce how well we modeled other parameters such as the wind and drag coefficient.

52
Chapter 6

Simulation tests and results

Two types of rockets are used to exemplify the results from the developed simulator: a model rocket and
a sounding rocket. The model rocket launch is the most suitable to observe the effect of the wind and the
ABL on the trajectory since these rockets reach lower velocities and apogees. On the other hand, due to
the wide range of rockets handled by the simulator and to extend the limits of the previous launch, we
also present the results from a sounding rocket launch.

6.1 Rockets description

6.1.1 Model rocket specification

This model rocket has two stages wherein the first is based upon a default rocket example provided in
OpenRocket. The rest of the components are defined in order to make the rocket properly modeled. The
second stage is propelled by a C6-0 motor and the first by a D12-0 motor, which is discarded at the
first burnout and the connector is discarded a half second later. Another half second later starts the
second ignition and after the next burnout the rocket coasts until the apogee where the recovery system
is deployed. For each stage, excluding the connector and the nose, Table 6.1 presents their respective
data. The connector has a length of 2.5 cm and the nose is an ogive with 10 cm long. In Table 6.2 are
defined the fin’s dimensions for the two stages.

1st stage (with D12-0 motor) 2nd stage (with C6-0 motor)
Average thrust (N) 10.21 4.74
Burnout (sec) 1.65 1.86
Gross mass (g) 55.62 41.66
Dry mass (g) 18.02 30.86
Length (cm) 8 30
Diameter (cm) 4 2.5
Number of fins 3 3

Table 6.1: Model rocket stages’ data (without connector and nose) [55] [56].

53
1st stage 2nd stage
Number of fins 3 3
Root chord (cm) 8 5
Tip chord (cm) 8 5
Span (cm) 3 3.4
Sweep angle (°) 45 39.8

Table 6.2: Model rocket fins’ data.

The parachute is deployed at the apogee and is taken as hemispherical shaped. Thus, from Table 4.1,
its drag coefficient is chosen to be 0.695 as it is the center of the respective range. To estimate the
parachute’s reference area (the total surface area of the canopy), we choose a terminal velocity of 5 m s−1
since it is below the typical safety landing speed [57]. Therefore, from (5.4), again with a density of
1.225 kg m−3 , we get 406.1 cm2 (or 22.7 cm of diameter), approximately.
Table 6.3 presents the main properties of the model rocket to be used in the first launch.

Model rocket property Value


Gross mass (g) 113.87
Mass after second burnout (g) 44.06
Initial length (cm) 50.5
Length at second ignition (cm) 40
First stage’s structure discharge (after launch) (s) 1.65
Connector’s discharge (after launch) (s) 2.15
Second stage ignition (after launch) (s) 2.65
Parachute’s CD 0.695
Parachute’s canopy diameter (cm) 22.7

Table 6.3: Model rocket’s main characteristics.

6.1.2 Sounding rocket specification

The second launch is based upon a two-stage solid propellant rocket used in the German MiniTEXUS
sounding rocket program. This rocket has a Nike first stage and an Orion second stage which ignites 9 s
after the liftoff [58]. Table 6.4 gives the data about each stage (excluding the connector and nose). The

1st stage (Nike motor) 2nd stage (Orion motor)


Average thrust (kN) 217 13
Burnout (sec) 3.2 32
Gross mass (kg) 599 400
Dry mass (kg) 256 111
Length (m) 3.6 1.8
Diameter (m) 0.42 0.35

Table 6.4: Nike and Orion stages’ data (without connector and nose) [58] [59] [60].

connector is modeled in order to fit each stage’s structure, with the height of 30 cm, and is assumed to
be discarded at half of the delay between the two ignitions. The fins in each stage are trapezoidal shaped

54
and the second stage’s fins are half the size of the ones in the first stage. Their dimensions are given in
Table 6.5. The nose is an ogive with a length of 1.8 m and a mass of 160 kg, which is under the payload
limits for the Nike-Orion rocket [58].

1st stage 2nd stage


Number of fins [58] 3 4
Root chord (cm) 100 50
Tip chord (cm) 40 20
Span (cm) 80 40
Sweep angle (°) 20.56 20.56

Table 6.5: Sounding rocket fins’ data.

This rocket may use a flat circular parachute with 7.3 m of diameter [61]. Hence, again from Table 4.1,
we select a drag coefficient of 0.775, which is the central value of the respective interval. The parachute
is deployed with a maximum dynamic pressure of 12.0 kN m−2 , approximately [61].
After assembling all the components, the essential data of the sounding rocket is given in Table 6.6.

Sounding rocket property Value


Gross mass (kg) 1171.5
Mass after second burnout (kg) 271
Initial length (m) 7.5
Length at second ignition (m) 3.6
First stage’s structure discharge (after launch) (s) 4.65
Connector’s discharge (after launch) (s) 6.1
Second stage ignition (after launch) (s) 9
Parachute’s CD 0.775
Parachute’s canopy diameter (m) 7.3

Table 6.6: Sounding rocket’s main characteristics.

6.2 Rockets characteristics

6.2.1 Model rocket

Taking the model rocket described in Section 6.1.1, we get, from the RocketProperties function, the thrust,
mass, CM, CP, transversal inertia and reference area over time presented in Figure 6.1. Due to the first
discontinuity shown in Figures 6.1b, 6.1c, 6.1d and 6.1e, we can see that the first stage’s structure (not
considering the connector) is discarded at the first burnout. Also in these figures, during the coasting
phase (lasting 1 s), we see the connector’s discharge represented by the second discontinuity. Except from
these sharp changes, the rocket maintains its properties during coasting, leading the mentioned figures
to match with Figure 6.1a when there is no thrust between the burnings.
The reference area is defined as the largest cross section of the rocket (without considering boosters,
for other cases) at each instant. This definition agrees with Figure 6.1f, where it shows the rocket’s
reference area changing nearly 2 s after the launch, which is the instant when the connector is discarded

55
and the rocket loses its largest sectional area.
Comparing Figures 6.1c and 6.1d, we conclude this rocket is always stable since the CM, instead of
the CP, is closer to the nose all the time. Hence, the rocket is well assembled and able to be used in a
trajectory simulation.

Thrust (N) Mass (g)


30
110
25 100
20 90

15 80
70
10
60
5
50
t (s) t (s)
1 2 3 4 5 1 2 3 4 5
(a) Rocket’s thrust. (b) Rocket’s mass.

CM (cm) CP (cm)
40 40

35 35

30 30

25 25
t (s) t (s)
0 1 2 3 4 5 0 1 2 3 4 5
(c) Rocket’s CM relative to the nose tip (reference). (d) Rocket’s CP relative to the nose tip (reference).

Ref. Area (cm2 )


Mom. Inertia (g.m3 )

12

4.0
10

3.5
8

3.0 6
t (s) t (s)
1 2 3 4 5 1 2 3 4 5
(e) Rocket’s transv. mom. of inertia relative to the CM. (f) Rocket’s reference area.

Figure 6.1: Model rocket properties’ functions.

56
6.2.2 Sounding rocket

For the sounding rocket described in Section 6.1.2, its properties are represented in Figure 6.2. During
the burnings, in Figure 6.2b, the rocket decreases its mass linearly since its thrust profile is constant
for each motor (as we only know the motors’ average thrust, see Figure 6.2a) and, consequently, their
burning rates as well. In Figures 6.2b and 6.2e, the discontinuities representing the first stage and the
connector’s discard match with the instants defined in Table 6.6, as well the reference area, in Figure 6.2b,
that changes only when the connector is discarded. From Figures 6.2c and 6.2d, we deduce the CM is
always ahead from the CP and, therefore, this rocket is also stable.

Thrust (kN) Mass (kg)


1200
200
1000
150
800
100
600
50
400

t (s) t (s)
10 20 30 40 10 20 30 40
(a) Rocket’s thrust. (b) Rocket’s mass.

CM (m) CP (m)
5.5 5.5

5.0 5.0
4.5 4.5
4.0 4.0
3.5 3.5
3.0 3.0
2.5 2.5
2.0 2.0
t (s) t (s)
0 10 20 30 40 0 10 20 30 40
(c) Rocket’s CM relative to the nose tip (reference). (d) Rocket’s CP relative to the nose tip (reference).

Mom. Inertia (kg.m3 ) Ref. Area (dm2 )


3500 14
3000
13
2500
2000 12
1500
11
1000
500 10
t (s) t (s)
10 20 30 40 10 20 30 40
(e) Rocket’s transv. mom. of inertia relative to the CM. (f) Rocket’s reference area.

Figure 6.2: Sounding rocket properties’ functions.

57
6.3 Trajectory simulations conditions

To run the simulations we need to set the flight conditions for both launches. Besides the rocket list, the
other trajectory inputs that differ for each launch are the initial nose position and the guiding rod/rail’s
length. Note that the rod’s length is the distance the rocket is traveling constrained in the launch
direction, not its full length from the bottom of the launch platform in a real case scenario.
Taking into account the dimensions of the rockets, the defined initial nose position for the model
rocket launch is 70 cm from the ground and its guiding rod is considered to be 50 cm long. For the
sounding rocket launch, the initial nose position is defined as 9 m from the ground and its guiding rail is
considered as 7 m long.
The launch site, launch direction and atmospheric conditions (density, temperature and wind model)
are the same for both simulations. Since we opt that these hypothetical launches take place around
Lisbon, the defined geodetic coordinates are 38.72°N, −9.14°W and 53 m above the reference ellipsoid
[53]. Both rockets will be launched vertically in order to be easily observed the deviation from the launch
direction. Nevertheless, we also perform another model rocket launch in which the rocket is launched
with 75° of elevation and 90° of azimuth (eastward).
The density, temperature (needed for the drag coefficient model) and wind are defined by the developed
models described in Sections 5.3 and 5.4. The local pressure and temperature inputs correspond to the
extrapolated values given from the CIRA-86 profiles at the launch site and considering the launch takes
place in mid-April (1018.54 mbar and 14.9 ◦C, respectively). To model the wind profile we assume a
neutral atmosphere with a constant wind speed above the ABL, in a high grass terrain, and blowing
northeastward (profile identical to the one represented in Figure 5.10). The required local wind speed is
defined taking into account the semi-empirical Beaufort scale, which ranks the wind speed basing on its
effects over the sea and the land surfaces [62]. This scale assigns to the wind the class gentle breeze when
leaves and small twigs move or light flags extend, limiting its speed between 3.4 m s−1 to 5.2 m s−1 at 6 m
from the ground. For a slower and stronger wind speed, the adjacent classes are, respectively, slight breeze
and moderate breeze. Foreseeing the gentle breeze is the class within the most of the launches may take
place, we define a local wind speed of 4.3 m s−1 (value at the center of the class), blowing northeastward,
at 6 m above the ground. Defining a high grass surface, we get, from Table 3.1, z0 = 7 cm as it is the center
value within the respective range. The previous local conditions are resumed in Table 6.7. Although the
wind direction is considered constant for the vertical launches, a wind direction variation with altitude is

Local condition Value


Temperature (◦C) 14.9
Pressure (mbar) 1018.54
Wind speed at 6 m from the ground (m s−1 ) 4.3
Downwind (from north, clockwise direction) (°) 45 (northeastward)
Surface type – z0 (cm) 7 (high grass)

Table 6.7: Launch site conditions.

58
defined for the non vertical model rocket launch. In this case the downwind is assumed as northeastward
at the ground, changing until the apogee where it takes a northward direction.
For every simulation the drag coefficient is determined by the developed model explained in Section 5.6,
which uses the same density, temperature and wind models from the respective trajectory to determine
the Mach and the Rex .

6.4 Results from the trajectory simulations

6.4.1 Model rocket launch

The obtained vertical model rocket launch trajectory is represented in Figure 6.3. In this trajectory the

Up (m)

500

400

300

200

100

-400 -200 0 200


Ground Track (m)

(b) Trajectory profile.


(a) 3D trajectory.

Figure 6.3: Model rocket trajectory with a vertical launch and a neutral ABL; Parachute ejection at the
apogee.

weathercock effect is observable. Since the wind blows northeastward, the rocket climbs in the opposite
direction (southwestward) to follow the true air speed. Although this effect increases smoothly with
altitude due to the ABL, we can see a slightly discontinuity in the upward flight path around 200 m of
altitude. This happens because the weathercock is enhanced during the coasting time, between the first
burnout and the second ignition, as the rocket decreases its velocity.
During the recovery phase, the rocket slowly descends toward the downwind direction. Hence, it
surpasses the launch site and lands 380 m north and 382 m east away from that place (coordinates taken
from the simulator). The influence of the ABL is also observable during the recovery phase in which the
trajectory describes a smooth curvature since the wind speed is decreasing as the rocket falls. If the wind
was constant with altitude, the recovery trajectory would describe a linear path.
In Figure 6.4 it is represented the flight data from this launch. It is deduced, from Figure 6.4a, that
the rocket reaches the apogee at about 560 m of altitude and lands 120 s after the lift off. Observing
when the vertical velocity turns negative, in Figure 6.4b, it is concluded the apogee is reached at 9.5 s,

59
Altitude (m) Velocity (m/s)
120
500
100
400 80
300 60

200 40

100 20

t (s) t (s)
20 40 60 80 100 120 2 4 6 8 10 12 14

(a) Profile of altitude vs time. (b) Profile of vertical velocity vs time.

Mach
Accel.(/g)
0.4
20
0.3
10

0.2
t (s)
2 4 6 8 10 12

-10 0.1

-20 t (s)
2 4 6 8 10 12 14
(c) Longitudinal acceleration vs time.
(d) Mach vs time.

α (º) Flight Path (º)


15 100

10 50

5
t (s)
20 40 60 80 100 120
t (s)
0.5 1.0 1.5 2.0 2.5 3.0
-50
-5

(e) Angle of attack vs time. (f) Flight path angle vs time until the landing.

CD
0.9

0.8

0.7

0.6

0.5

t (s)
0 2 4 6 8
(g) Drag coefficient vs time until the apogee.

Figure 6.4: Vertical model rocket launch flight data.

60
approximately. As already mentioned, we also see that the rocket decreases velocity due to the coasting
between stages. This is also observable from the Mach number, in Figure 6.4d, which remains subsonic
throughout the entire flight.
During the burnings, in Figure 6.4c, the acceleration describes a profile similar to the thrust (see
Figure 6.1a). However, it is decreasing over time as the rocket gains velocity, which increases the drag
opposing the thrust force. In contrast, when the rocket is not burning, the acceleration is negative and
increases over time (decreases in absolute value) because the velocity is decreasing. The peak acceleration
at 9.5 s results from the parachute ejection.
The angle of attack, in Figure 6.4e, shows the rocket rapidly damps its oscillations and follows the
true air speed during the majority of the ascending flight. During the first instants the angle of attack is
not computed because its model receives the impulse to start the oscillations only when the rocket leaves
the guiding rod.
From Figure 6.4f, we can see the flight path angle as 90° at the ignition since the launch is vertical.
This angle decreases until the apogee due to the weathercock effect that causes a curvature in the
trajectory. Reaching the apogee, the flight path angle becomes negative because the rocket starts to
descend. There is a peak value of about −90° since the rocket reverses its direction and stabilizes in
a 30° slope, approximately, toward the ground due to the drag force on the parachute. As the rocket
descends, the flight path decreases, which means the velocity is becoming steeper. This fact agrees with
the curvature described by the trajectory due to the ABL, observed in Figure 6.3.
The drag coefficient, in Figure 6.4g, is dependent on the Reynolds and the Mach numbers. Therefore,
when the velocity is increasing, the drag coefficient decreases its value due to the skin friction drag. In-
versely, as the velocity decreases, the drag coefficient starts to increase. When the first stage is discarded,
the CD decreases instantaneously because the rocket loses a great source of drag. On the other hand,

(a) Eastward perspective.

(b) Northward perspective.

Figure 6.5: Model rocket trajectory in a 75° launch eastward with the downwind changing from north-
eastward at the ground to northward at the apogee.

61
when the connector is discarded, although it is lost another source of drag, the CD increases significantly.
This happens because this is the instant when the reference area becomes smaller and the CD has to
change in order to be given in respect to another section.
Launching the model rocket eastward, with 75° of inclination from the ground and changing the down-
wind from northeastward at the surface to northward at the apogee, we get the trajectory in Figure 6.5.
As it is depicted, the upward phase of the trajectory is deviated southward since the weathercock effect
toward southeast (as in the vertical launch) is counteracted by the initial velocity pointing eastward.
Right after the apogee, during the recovery, the rocket falls northward but, gradually, as it approaches
the ground, it glides northeastward due to the 45° variation in the wind direction.

6.4.2 Sounding rocket launch

The trajectory from the vertical sounding rocket launch (again with constant wind direction) is repre-
sented in Figure 6.6. Like the other trajectories, we can see this one is also deviated upwind (southwest-

Up (km)

80

60

40

20

10 20 30 40 50 60
Ground Track (km)
(a) 3D trajectory. (b) Trajectory profile.

Figure 6.6: Sounding rocket trajectory with a vertical launch and wind blowing northeastward; Parachute
ejection at about 28 km of altitude.

ward). However, after the apogee, the rocket continues to veer from the launch site because the parachute
is ejected only when the rocket returns to 28 km of altitude, approximately. The dynamic pressure of
12 kN m−2 (mentioned in Section 6.1.2) is reached at this altitude, enabling the rocket to decrease its
descent rate until it reaches the safety terminal velocity. Also, with this dynamic pressure, the drag
force acting on the parachute is high enough to carry the rocket back northeastward, as observed by the
trajectory curvature at the end of the recovery phase. If the parachute was ejected at the apogee, the
rocket would describe a similar trajectory due to the low density at these altitudes that decrease the drag
force from the parachute (no dynamic pressure, see Figure 6.7).

62
Dyn. Pressure (kN/m2 )

200

150

100

50

t (s)
50 100 150 200 250

Figure 6.7: Sounding rocket’s dynamic pressure until the parachute ejection.

In Figure 6.8 it is represented the data obtained from this launch. As seen from Figure 6.8a, the rocket
reaches an apogee of 80 km at about 150 s after the lift off and the flight lasts 23 min, approximately.
Like the model rocket launch, the velocity, in Figure 6.8b, decreases during the delays between stages and
until the apogee, as expected. After this instant, the velocity increases (in absolute value) as the rocket
continues its trajectory back to the ground without ejecting the parachute. The ejection happens at
250 s, approximately, and the velocity decreases from about 1 km s−1 to nearly the terminal velocity. The
Mach number, in Figure 6.8d, behaves accordingly the velocity, as expected, and reaches the supersonic
regime during the climbing and the descending trajectories. Although the recovery deployment happens
at supersonic speeds, the dynamic pressure suffered by the parachute (see Figure 6.7) lies in the limit
range mentioned in the literature [61] as the atmosphere is still rarefied.
From Figure 6.8c, we see the acceleration increases during the burnings and decreases in absolute
value during the coasting time due to the drag slowing down the rocket. After the burnouts, the absolute
acceleration is around 1g since the dynamic pressure is nearly zero. Owed to the parachute ejection, the
rocket suffers a 120g deceleration that is not represented in Figure 6.8c due to the graphic scale.
The amplitudes of the angle of attack, in Figure 6.8e, are smaller comparing to the ones from the
model rocket launch because the effect of the wind is softened by the higher velocity of the rocket leaving
the guiding rail.
The flight path angle, in Figure 6.8f, starts at 90° since the rocket is launched vertically and, again,
due to the weathercock effect, this angle decreases as the rocket follows the air speed direction. After the
apogee, the flight path increases (in absolute value) because the rocket continues its downward trajectory
without the recovery device. When the parachute is deployed, the flight path decreases instantaneously
to −90° as the rocket reverts its direction. Then, the flight path increases until nearly to the end of the
recovery phase as the descent rate is slowly decreasing due to the higher density. This fact intensifies the
effect of the wind dragging the rocket, as observed in Figure 6.6.
The drag coefficient, in Figure 6.8g, behaves as explained for the model rocket with the exception
during the first burning, where there is a CD increase. This is explained due to the transonic and low
supersonic effects (shock waves) implemented in the drag coefficient model. In Figure 6.8h, the CD begins
to increase several times its value until the apogee. This increase is due to the small Reynolds number
(from which the CD is dependent on), as the rocket is losing velocity and the atmosphere is becoming
rarefied.

63
Altitude (km) Velocity (km/s)
80
1.0

60
0.5

40
t (s)
50 100 150 200 250 300 350
20 -0.5

t (s) -1.0
200 400 600 800 1000 1200 1400
(a) Profile of altitude vs time. (b) Profile of vertical velocity vs time.

Mach
Accel. (/g)
25 3.5

20 3.0
2.5
15
2.0
10
1.5
5 1.0
0.5
t (s)
50 100 150 200 250
t (s)
50 100 150 200 250 300 350
(c) Longitudinal acceleration vs time. (d) Mach vs time.

α (º) Flight Path (º)


6 100

4 50

2
t (s)
200 400 600 800 1000 1200 1400
t (s)
1 2 3 4 5 -50

-2

(e) Angle of attack vs time. (f) Flight path angle vs time until the landing.

CD CD

2.5
0.65
2.0
0.60

0.55 1.5

0.50
1.0
0.45
0.5
t (s) t (s)
5 10 15 20 50 100 150 200 250

(g) Drag coefficient vs time until 20 s after the launch. (h) Drag coefficient vs time from 20 s until the recovery.

Figure 6.8: Vertical sounding rocket launch flight data.

64
6.4.3 The influence of atmospheric stability

In this section we compare the flight profiles from the model rocket launch with the same local wind speed
used in the previous sections but changing the atmospheric stability. For sounding rockets the following
results are irrelevant since they leave rapidly the ABL and it covers a small part ot the trajectory.
Figure 6.9 presents three wind profiles, under three different stable atmospheres, for a measured wind
speed of 4.3 m s−1 at 6 m of altitude. We can see that as the atmosphere becomes more stable, the wind

Velocity (m/s)

15

Very Stable (L = 30 m)
10
Stable (L = 125 m)
Near Stable (L = 350 m)
5

Altitude (m)
100 200 300 400 500 600

Figure 6.9: Wind profiles in a stable atmosphere for a measured wind speed of 4.3 m s−1 at 6 m of altitude.

speed gets stronger at higher altitudes and the ABL thickness decreases. Comparing to Figure 6.10, where
it shows the respective trajectory for each wind profile, we observe that the ascent phase of the trajectory
is negligibly affected as the atmosphere changes from near to very stable conditions. On the other hand,
the recovery phase from each trajectory is strongly influenced by these changes. We can see that the
rocket diverges more from the landing site under stable conditions as it corresponds to the strongest wind
speed profile. Inversely, the rocket lands closer to the launch site in a near stable atmosphere as the wind

Altitude (m)
500
400
300
200
100
Ground Track (m)
-1000 -500 0

Very Stable (L = 30 m) Stable (L = 125 m) Near Stable (L = 350 m)

Figure 6.10: Trajectory profiles of the model rocket launch under three different wind profiles from a
stable atmosphere.

is less intense. Comparing to Figure 6.3b, corresponding to the trajectory in a neutral atmosphere, we
also see that, in this case, the rocket is even less deviated from the launch site and the apogee becomes a
little higher. This is expected since in neutral conditions the wind speed profile intensifies more smoothly
with altitude than in a near stable atmosphere.
Figure 6.11 presents three wind profiles, under three different unstable atmospheres, for the same
previously measured conditions. As it shows, the wind speed gets less intense after a certain altitude

65
Velocity (m/s)

4 Very Unstable (L = -75 m)


Unstable (L = -150 m)
2
Near Unstable (L = -350 m)
Altitude (m)
200 400 600 800 1000 1200 1400
-2

Figure 6.11: Wind profiles in an unstable atmosphere for a measured wind speed of 4.3 m s−1 at 6 m of
altitude.

and may reverse its direction if the atmosphere becomes unstable enough. This is due to the convective
effects since the warmer air at the surface rises and the ABL becomes turbulent. Linking these profiles to
the respective trajectories in Figure 6.12, we observe the rocket lands further from the launch site as the
atmosphere turns less unstable, as expected due to the increase of the wind speed. On the contrary, the
upward trajectory only changes a bit near the apogee but it reaches higher altitudes than in a stable or
neutral atmosphere. The recovery trajectory under a very unstable atmosphere (see Figure 6.12) stands
out owned to its different shape. The curvature presented by this descendant trajectory is explained from
the absence of wind at the apogee (see Figure 6.11) that allows the rocket to fall near vertically. As it
approaches the ground the wind speed increases and the rocket is progressively carried downwind.

Altitude (m)

600

500

Very Unstable (L = -75 m)


400
Unstable (L = -150 m)
300
Near Unstable (L = -350 m)

200

100

Ground Track (m)


-200 -100 0 100 200

Figure 6.12: Trajectory profiles of the model rocket launch under three different wind profiles from an
unstable atmosphere.

To summarize this section, we conclude that the ascending phase of the trajectory is more influ-
enced when the atmospheric stability changes from stable to unstable (or neutral) than when it presents
variations under the same stability class. This is also true for the recovery phase but, in this case,
the downward trajectory already presents large differences when the stability has small deviations but
remains under the same class.

66
Chapter 7

Stochastic simulations

Many of the parameters required to determine the trajectory cannot be known (or well defined, at least)
and the measured data may vary randomly, presenting some uncertainties. Therefore, in order to analyze
the outcome of several uncertain scenarios, it lead us to perform simulations that rely on repeated random
sampling and statistical analysis to compute the results, known as Monte Carlo simulations [63].
To get the outcome due to the uncertainties, we must identify a suitable probability distribution for
each input parameter from which a set of data is randomly generated. Then, all the possible combinations
between the created data correspond to the inputs given in each trajectory simulation. Therefore, the
number of trajectories to be determined, for a given Monte Carlo, is given by

n
Y
Ntraj = Nptsi , (7.1)
i=1

where n is the number of parameters under study during the simulation and Nptsi is the number of points
generated to the parameter i. From Figure 7.1, we deduce that a sample of 1000 points can approximately
represent a normal distribution for a given parameter. However, from (7.1), we find that, even for a few

Prob. Density

0.4

0.3

0.2

0.1

0.0
-2 -1 0 1 2 3

Figure 7.1: Normal distribution N (0, 1) and respective probability density histogram from 1000 samples.

parameters, the resultant Monte Carlo simulation time turns this hypothetical computation impractical in
®
an interpreted language such as Mathematica . Since we do not own such great computational resources
as to minimize the simulation time, a better approach is to perform what if simulations in which it is

67
analyzed the outcome from the uncertainties of only two or three parameters with smaller samples. In
the latter case, the number of points generated to each parameter should be less than the former case in
order to get a similar simulation time.

7.1 Simulation rocket and conditions

The simulations presented in the following sections use the single stage rocket described in Tables 7.1
and 7.2, which was based on a high power rocket available for sale [64]. With this choice we can simulate

Rocket property Value


Average thrust (N) 1550
Burnout (sec) 3.68
Gross mass (kg) 9.59
Dry mass (kg) 5.47
Body tube length (cm) 80
Nose length (cm) 38.1
Diameter (cm) 7.5
Number of fins 3
Parachute’s CD 0.695
Parachute’s canopy diameter (m) 2.54

Table 7.1: High power rocket data.

a kind of rocket that produces an amount of thrust that lies between the power used in the two types of
rockets from Chapter 6 and, at the same time, achieves low supersonic speeds. The motor M1550-0 is
selected in order to get a ratio between this rocket and the model rocket’s apogees with the same order
of magnitude from the ratio between the sounding rocket and the high power rocket’s apogees.
From Table 4.1, considering this rocket is provided with a hemispherical parachute, we assume its CD
is known and equal to 0.695. Like in Section 6.1.1, the parachute’s reference diameter is found in order to
reach the safety terminal velocity of 5 m s−1 . Therefore, from (5.4) with a density of 1.216 kg m−3 found
from the conditions at the launch site (mentioned later), we get a diameter of 2.54 m.

Fin’s property Value


Nº fins 3
Root chord (cm) 12
Tip chord (cm) 6
Span (cm) 12
Sweep angle (°) 63.4

Table 7.2: High power rocket fins’ data.

The hypothetical launch site for the simulations is the same as the one described in Section 6.3. The
guiding rod is defined as placed vertically, measuring 2 m, and the initial nose distance to the ground is
3 m.

68
In order to study the influence of the exponential rates at which the local weather conditions approx-
imate the data from the developed atmospheric model along the altitude, these conditions must differ
from the defined values in the previous simulations. Since the old conditions correspond to the data
presented by the atmospheric model at the respective altitude and latitude, the constants that define
the rates do not affect the pressure and temperature. Therefore, for the following simulations, the new
local atmospheric conditions are defined as the maximum averages in April 2015 in Lisbon: 1027 mbar
and 21 ◦C [65]. For the wind speed and direction, their conditions are defined to be the same as the ones
in Table 6.7, respectively.

7.2 Landing site uncertainties

As already mentioned, we cannot afford the necessary resources to make a full Monte Carlo simulation.
Therefore, changing one parameter at a time, we can study its impact in the output in order to choose
the most critical to be used in our final simulation. We must have in mind that even if changing one
parameter does not present a significant influence in the output, its effect may be enhanced by changing
simultaneously another parameter from which the former is dependent on. Hence, the following tests are
merely a general approximation to predict the critical inputs that present a great impact in the output
when changing inside a restricted interval.
The uncertainties in the trajectory inputs may arise from sudden fluctuations in the atmospheric
conditions or due to the errors when guessing the value of a given property. In Table 7.3 are presented all
the parameters that contribute to these uncertainties in the trajectory (indirectly through the atmospheric
and drag coefficient models) and their respective influence in the landing site coordinates. The uniform

Landing site
Parameter Distribution
µx (m) µy (m) σx (m) σy (m)
2
Local pressure (mbar) N (1027, 1 ) −7089 7096 0.065 0.065
Pressure rate (m−1 ) U(0.1, 3) −7090 7096 0.427 0.425
Local temperature (◦C) N (21, 22 ) −7089 7096 0.537 0.537
Temperature rate (m−1 ) U(0.1, 4) −7089 7096 0.835 0.840
Local windspeed (m s−1 ) N (4.3, 12 ) −7500 7507 2078 2077
Surface roughness length – z0 (cm) U(4, 10) −7095 7102 286 286
Downwind (from north, clockwise) (◦ ) N (45, 72 ) −7153 6903 956 989

Table 7.3: Mean (µ) and standard deviation (σ) of the landing site in the two directions (x – Southing;
y – Easting) changing each input within a distribution of 20 samples.

distribution is used to pick up values for the parameters that are unknown within an specific interval. On
the other hand, the normal distribution is used to generate values for the parameters that are measured
but may present some disturbances. In this case, the mean represents the expected value under the
current conditions and the standard deviation specifies how much the measured value may deviate from
the mean. The minimum and maximum thresholds in the uniform distributions from the pressure and
temperature rates are defined, respectively, in a way that the atmospheric profiles match the CIRA-86

69
model at about 500 m of altitude and at the apogee. The standard deviations for the local pressure and
temperature are defined as the greatest variation occurred in a 30 min interval during April 15, 2015 in
Lisbon [66]. In the case of the wind speed’s normal distribution, its standard deviation is defined as the
necessary amplitude that changes the local wind to another class according the Beaufort scale [62] (see
fourth paragraph in Section 6.3). Finally, the standard deviation for the wind direction is defined as the
variation occurring at the measured wind speed found from a dataset taken during daytime in a flat grass
surface [67].
Comparing the mean landing coordinates with the standard deviations, in Table 7.3, we conclude the
uncertainties from the pressure and temperature measured at the launch site, and their respective rates to
match the atmospheric model along the altitude, present a negligible impact under the defined conditions.
On the other hand, the impact from the remaining parameters’ uncertainties is much more considerable,
which leads us to neglect the first four parameters from Table 7.3. With only three parameters left,
we still have to compute a billion iterations if we use samples of 1000 values for each parameter, as
illustrated in Figure 7.1. Therefore, for the remaining normal distributions (wind speed and direction),
we take samples of 200 values since, after testing, we concluded that a larger sample would not improve
the landing site dispersion but only saturate the graphic. In respect to the surface roughness length,
since we do not know any characteristic that helps to define it other than the terrain type (defined as
high grass), it can take any value within the range of 4 cm to 10 cm (see Table 3.1). Hence, in order to
reduce the simulation time, we decided this parameter should only take the worst case scenarios, i.e.,
4 cm and 10 cm. Nevertheless, it is still required to replace the CD model by a constant value since it
slows down each iteration and it is not expected to impact the results much. Therefore, the CD used in
the simulation is the mean value of 0.70 calculated from
Z tapogee
CD (t)
CD = , (7.2)
0 tapogee

where CD (t) is the drag coefficient function found from the developed model after simulating a trajectory
under the same conditions used in this chapter.
Figure 7.2 shows the outcome of the Monte Carlo simulation resulting from the input conditions
mentioned above. Since the rocket follows the wind speed direction, it is expected that the uncertainties
from the local wind velocity and the surface roughness length (both only influencing the wind intensity)
change the landing site along this direction, describing a straight line. Analogously, the changes in the
wind direction, due to its uncertainties, make the landing points to describe an arc of a circumference
that gets larger as the distance to the launch site increases.
The dispersion of the landing points, in Figure 7.2, clearly shows the influence of the normal distribu-
tions used to generate the local wind direction and speed applied in each iteration. Since a mean of 45◦
was defined for the wind direction’s normal distribution, the orange dots are heavily concentrated along
the northeast direction. In the same way, the rocket lands less often apart from the northeast direction
as there are fewer inputs that deviate more from the mean. Also due to this principle, the wind speed
uncertainties contribute to the lower density of points near the closest and furthest distances from the

70
Figure 7.2: Landing positions and respective confidence ellipses from the Monte Carlo simulation.

launch site. This pattern can also be seen in Figure 7.3 where the probability density function of the
landing sites is represented. Furthermore, this figure enables us to distinguish better the most proba-
ble landing area since Figure 7.2 gets saturated in this region due to the great amount of trajectories
simulated.
The confidence ellipses in Figure 7.2 are rather eccentric. Their major axes (aligned northeastward)
measure 10.4 km, for a confidence of 95%, and 9.1 km, for 90% of confidence. As the furthest landing site,
collinear with the major axes, is 14.6 km away from the launch location, the major axis from the lowest
confidence level represents almost two thirds of the furthest distance. Thus, we conclude the measured
wind speed, allied with the surface roughness length range, brings a strong uncertainty to the landing site
estimation under the specified conditions. Regarding the minor axes compared to their distances from
the launch site, we see that the wind direction also induces a considerable uncertainty.

0.06

0.04

0.02

Figure 7.3: Probability density function of the landing coordinates resulting from the Monte Carlo
simulation.

71
7.3 Rocket optimization

Taking the previous simulation from a different approach, we get another tool that allows us to perform
optimization. Therefore, after defining an optimization criteria (maximization or minimization of an
output) given an universe of input variables, we can deduce the best alternative that matches our goal.
Changing the rocket parameters in each Monte Carlo iteration (as opposed to Section 7.2, where the
rocket properties are kept constant), we can optimize the rocket characteristics. For example, we are
aiming to determine the best body tube’s length and diameter combination so that the rocket reaches the
highest apogee keeping constant the other properties. To this example a windless atmosphere is defined
(in order to have simpler trajectories) and three options for each tube’s dimension, which are selected
randomly from an uniform distribution. The minimum and maximum length limits (constraints) are 1 m
and 2.5 m, respectively, and for the diameter are 7.5 cm and 13 cm, respectively.
In Figure 7.4 the altitude over time is depicted (until the highest apogee’s instant) for these nine
configurations. As shown, the configuration with the smallest dimensions (1.215 m and 8.4 cm) reaches

Altitude (km) l=1.686m d=10.1cm


l=1.686m d=8.4cm
5
l=1.686m d=12.2cm
4 l=1.215m d=10.1cm

3 l=1.215m d=8.4cm
l=1.215m d=12.2cm
2
l=2.266m d=10.1cm
1 l=2.266m d=8.4cm
t (s) l=2.266m d=12.2cm
5 10 15 20 25

Figure 7.4: Altitude over time from nine launches with different rocket’s length and diameter configura-
tions.

the highest apogee, as opposed to the rocket with the largest dimensions that reach the lowest apogee.
These facts are expected due to the major influence of the rocket’s dimensions on the drag coefficient
and the reference area. The longer and larger the rocket is, the greater is the drag that decreases the
velocity and forces the rocket to reach sooner a lower apogee. From this result, we may say this method
works properly but a more extensive optimization should be performed in the future since the geometric
parameters of the rocket also influence the mass of the components.
Even though this is a simple and expected example, it gives a glimpse of the many possibilities at
our disposal after developing a Monte Carlo simulation to our tool. From this starting point, and having
enough computational resources, more complex optimizations may be processed. Besides, other efficient
optimization methods (not considered in this work) may be developed and applied to the presented tool
so that the expected result can be fast and directly found, saving resources finding the best alternative
from the outputs.

72
Chapter 8

Conclusions

The present work focused on the development of a trajectory simulator tool for fin-stabilized model and
sounding rockets addressing the effects of an ABL profile. This tool was developed under Mathematica ®
and supports from model to sounding rockets with any number of stages. This work presented the tool’s
flexibility to customize the default models and to compute the outcome from the inputs’ uncertainties
(or perform an optimization) from a Monte Carlo simulation.
Methods to estimate the rockets properties (including constant values and time dependent functions)
were presented in this work. Nonetheless, any rocket input can be defined according the user requirements.
The developed atmospheric model takes into account the local conditions at the launch site which tend to
the CIRA-86 profile at a specified rate. Likewise, above the ABL, the wind model approximates the values
given in the CIRA-86 tables. Within the ABL, as the wind plays an important role in the model rockets
trajectory, three profiles were proposed and implemented considering the atmospheric stability, which
can be deduced from the Obukhov length. However, we concluded this parameter is hard to determine
with low errors using common measuring instruments. Therefore, a qualitative alternative was presented
to figure out the atmospheric stability from the weather conditions. Besides the local atmospheric data,
the ABL profiles also take into account the type of the surface. In order to improve the simulations
considering the rocket’s attitude, as our tool possesses three degrees of freedom, we developed a model
to compute the angle of attack over time that can be influenced by wind gusts appearing along the wind
direction. A drag coefficient model was also implemented depending on the Mach, the Reynolds and the
configuration of the rocket.
Two types of simulations were performed to demonstrate the potentiality of the tool: a model rocket
and a sounding rocket launch. Both of them showed that the developed trajectory simulator provides
reliable data considering its restrictions are followed. These results also revealed a large impact of the
wind in the trajectories (due to the weathercock effect), which leads us to reassure the importance to
forecast the atmospheric stability and to model a suitable wind profile. Comparing the same launch
under different atmospheric stabilities with the same local wind, it was concluded the recovery phase
is more sensitive to variations of the Obukhov length (keeping the same stability) than the ascendant
trajectory, which is barely affected. However, different stabilities already induce a considerable impact in

73
the climbing trajectory near the apogee. From the results we also conclude the wind profiles must always
concern the local conditions at the launch since, from the Monte Carlo simulation, and for the defined
uncertainties, they showed a strong influence in estimating the landing site.

8.1 Future work

Although the developed trajectory simulator presented results expected theoretically, it lacks validation
from real case scenarios. Therefore, as an extension to this work, the presented model rocketry highlights
must be deepened and used to perform rocket launches taking into account the methods in this thesis.
Prior to these launches, this tool may be used to estimate the ejection delay after the apogee that enables
the rocket to land near the launch site, thus preventing recovery problems.
Since the presented trajectory simulator is intended to be a flexible tool, we shall propose future
developments in order to expand and improve the simulator’s utilities:

ˆ As this work was focused on fin-stabilized rockets, an active control method may be implemented
in the future. In that case, the wind remains as an important factor to determine the input controls
but it does not influence the trajectory as the rocket follows the predefined path.
ˆ When we changed the wind direction over the altitude, the trajectory presented a strong dependence
on this parameter. It is therefore advised to develop a wind direction profile.
ˆ The way we implemented the wind gusts requires the user to estimate the instant they occur and
only affect the angle of attack. Hence, an approach that joins the gusts within the wind profile can
bring better results and reveal itself as a more practical method.
ˆ Since this tool is primarily aimed to model rocketry, the developed drag coefficient model does not
take into account the free molecular flow in a rarefied atmosphere. If further studies reveal it plays
a significant role when finding the CD of sounding rockets, its effects should be added into the
existing model.

74
Bibliography

[1] Degeratu, M., Georgescu, A. M., Bandoc, G., Alboiu, N. I., Cosoiu, C. I., and Golumbeanu, M.
Atmospheric Boundary Layer Modelling as Mean Velocity Profile Used for Wind Tunnel Tests on
Contaminant Dispersion in the Atmosphere. Journal of Environmental Protection and Ecology,
14(1):22–28, 2013.

[2] Thuillier, R. H. and Lappe, U. Wind and Temperature Profile Characteristics from Observations on
a 1400 ft Tower. Journal of Applied Meteorology, 3:299–306, June 1964.

[3] Seibert, G. The History of Sounding Rockets and Their Contribution to European Space Research.
ESA Publications Division, November 2006.

[4] Shearer, D. A. and Vogt, G. L. Rockets: A Teacher’s Guide with Activities in Science, Mathematics,
and Technology. NASA, Office of Human Resources and Education, Washington, DC.

[5] Page, B. The Rocket Experiments of Robert H. Goddard, 1911 to 1930. The Physics Teacher,
29(8):490–496, November 1991.

[6] Riper, A. B. V. Rockets and Missiles – The Life Story of a Technology. Greenwood Press, 2004.

[7] National Association of Rocketry,. Model Rocket Info. http://www.nar.org/model-rocket-info.


Accessed September 3, 2015.

[8] Colburn, W. Where Did Model Rocketry Really Start? Peak of Flight Newsletter, (314), June 2012.

[9] NASA,. Model Rockets. https://www.grc.nasa.gov/www/k-12/rocket/rktparts.html. Accessed


September 3, 2015.

[10] National Association of Rocketry,. High Power Rocketry. http://www.nar.org/


high-power-rocketry-info. Accessed September 3, 2015.

[11] National Association of Rocketry,. Organizational Statement of the NAR. Available online at
http://www.nar.org/pdf/Organizational%20Statement%20of%20the%20NAR.pdf.

[12] National Association of Rocketry,. Contest Events. http://www.nar.org/contest-flying/


contest-events. Accessed September 3, 2015.

[13] National Association of Rocketry,. Team America. http://www.nar.org/team-america. Accessed


September 3, 2015.

75
[14] NASA,. About Student Launch. http://www.nasa.gov/audience/forstudents/studentlaunch/
about/index.html#.U0IDilfVokg. Accessed September 3, 2015.

[15] Niskanen, S. Development of an Open source model rocket simulation software. Master’s thesis,
Helsinki University Of Technology, 2009.

[16] Fossey, P. RockSim Program Guide. Apogee Components, 2003.

[17] Arya, S. P. Introduction to Micrometeorology. Academic Press, Inc., 1988.

[18] Milligan, T. Phases of a Model Rocket’s Flight. Peak of Flight Newsletter, (117), December 2003.

[19] Stine, G. H. Forty Years of Model Rocketry – A Safety Report. National Association of Rocketry,
1997.

[20] Livermore Unit National Association of Rocketry,. The LUNAR Handbook. LUNAR, January 2008.
Available online at http://www.lunar.org.

[21] Apogee Components,. How Rocket Engines Work. http://www.apogeerockets.com/Tech/How_


Rocket_Engines_Work. Accessed May 28, 2014.

[22] Harris, D., Todd, M., Humble, M., and Locklin, M. Variability in the Performance of
Composite-Propellant Rocket Motors. http://tuhsphysics.ttsd.k12.or.us/Research/IB11/
HumbLockToddHarr/index.htm. Accessed May 28, 2014.

[23] Milligan, T. Rocket Engine Classification System Explained. Peak of Flight Newsletter, (131),
September 2004.

[24] Barrowman, J. Stability of a Model Rocket in Flight. Technical Information Report 30, Centuri
Engineering Company, 1970.

[25] Tewari, A. Atmospheric and Space Flight Dynamics. Birkhauser, 2007.

[26] Milligan, T. How To Multi-Stage Rockets Work – Part 1. Peak of Flight Newsletter, (98), February
2003.

[27] Estes, V. Multi-Staging, 1963. Estes Industries Technical Report TR-2.

[28] Milligan, T. How To Multi-Stage Rockets – Part 2. Peak of Flight Newsletter, (99), February 2003.

[29] Estes Industries,. Cluster Techniques, 1967. Technical Report TR-6.

[30] Stull, R. An Introduction to Boundary Layer Meteorology. Kluwer Academic Publishers, 1988.

[31] Tennekes, H. The Logarithmic Wind Profile. Journal of the Atmospheric Sciences, 30(2):234–238,
March 1973.

[32] Moran, M., Shapiro, H., Boettner, D., and Bailey, M. Fundamentals of Engineering Thermodynamics.
John Wiley and Sons, Inc., 2011.

76
[33] Gryning, S. E., Batchvarova, E., Brümmer, B., Jørgensen, H., and Larsen, S. On the extension
of the wind profile over homogeneous terrain beyond the surface boundary layer. Boundary-Layer
Meteorology, 124(2):251–268, August 2007.

[34] Byun, D. W. Determination of similarity functions of the resistance laws for the planetary boundary
layer using surface-layer similarity functions. Boundary-Layer Meteorology, 57(1–2):17–48, October
1991.

[35] Garrat, J. R. The Atmospheric Boundary Layer. Cambridge University Press, 1992.

[36] Costa, A. Erros e Algarismos Significativos. Gazeta de Fı́sica, 26(4):4–10, Outubro 2003.

[37] Chobotov, V. A. Orbital Mechanics. AIAA Education Series, 3rd edition, 2002.

[38] Wiesel, W. E. Spaceflight Dynamics. McGraw-Hill, 2nd edition, 1997.

[39] Curtis, H. Orbital Mechanics for Engineering Students. Elsevier, 2nd edition, 2010.

[40] Feodosiev, V. I. and Siniarev, G. B. Introduction to Rocket Technology. Academic Press Inc., 1959.

[41] Milligan, T. Basics of Dynamic Flight Analysis (part 2) - The Corrective Moment Coefficient. Peak
of Flight Newsletter, (193), September 2007.

[42] Milligan, T. Basics of Dynamic Flight Analysis (Part 3) - The Damping Moment Coefficient. Peak
of Flight Newsletter, (195), October 2007.

[43] Beer, F., Johnston, E., Mazurek, D., Cornwell, P., and Eisenberg, E. Vector Mechanics for Engineers:
Statics and Dynamics. McGraw-Hill, 9th edition, 2010.

[44] TUDelft,. Solid rocket propellants and their properties. http://www.lr.tudelft.


nl/en/organisation/departments/space-engineering/space-systems-engineering/
expertise-areas/space-propulsion/design-of-elements/rocket-propellants/solids.
Accessed July 7, 2014.

[45] Barrowman, J. Calculating the Center of Pressure of a Model Rocket. Technical Information
Report 33, Centuri Engineering Company.

[46] eFunda Engineering Fundamentals,. Moment of inertia of a triangle. http://www.efunda.com/


math/areas/triangle.cfm. Accessed October 28, 2014.

[47] Hoerner, S. F. Fluid-Dynamic Drag. Published by the author, 1965.

[48] Gregorek, G. M. Aerodynamic Drag of Model Rockets. Model Rocket Technical Report 11, Estes
Industries, Inc., 1970.

[49] Brederode, V. d. Fundamentos de Aerodinâmica Incompressı́vel. Edição do Autor, 1997.

[50] Rathakrishnan, E. Theoretical Aerodynamics. Wiley, 2013.

77
[51] Bonney, E. A. Engineering Supersonic Aerodynamics. McGraw-Hill, 1st edition, 1950.

[52] Knacke, T. W. Parachute Recovery Systems – Design Manual. Para Publishing, 1992.

[53] Wolfram Research, Inc., Mathematica, Version 10.2, Champaign, IL (2015).

[54] ANSI/AIAA. Guide to Reference and Standard Atmosphere Models, February 2009.

[55] NAR,. ESTES C6 Motor Data Sheet. http://nar.org/SandT/pdf/Estes/C6.pdf. Accessed May


28, 2015.

[56] NAR,. ESTES D12 Motor Data Sheet. http://nar.org/SandT/pdf/Estes/D12.pdf. Accessed


May 28, 2015.

[57] Milligan, T. Frequently Asked Questions About Dual Deployment Rockets. Peak of Flight Newsletter,
(324), October 2012.

[58] Encyclopedia Astronautica,. Nike Orion Rocket. http://www.astronautix.com/lvs/nikorion.


htm. Accessed May 7, 2015.

[59] Encyclopedia Astronautica,. Nike Booster. http://www.astronautix.com/engines/m5e1.htm. Ac-


cessed May 7, 2015.

[60] Encyclopedia Astronautica,. Orion Motor. http://www.astronautix.com/engines/orion.htm.


Accessed May 7, 2015.

[61] Suborbital Projects And Operations Directorate,. Sounding Rocket Program Handbook. NASA,
Sounding Rockets Program Office, Goddard Space Flight Center, Wallops Island, Virginia, June
1999.

[62] Lazaridis, M. First Principles of Meteorology and Air Pollution, volume 19. Springer, 2011.

[63] Raychaudhuri, S. Introduction to Monte Carlo simulation. In Proceedings of the 2008 Winter
Simulation Conference, pages 91–100, 2008.

[64] Giant Leap Rocketry,. Almost Ready to Fly Rocket Kits – Vertical Assault 3.0. http://www.
giantleaprocketry.com/products/kits_almost.aspx. Accessed August 14, 2015.

[65] Weather Underground,. Tempo Histórico para LPPT – Month of Abril, 2015. http:
//www.wunderground.com/history/airport/LPPT/2015/4/14/MonthlyHistory.html?&reqdb.
zip=&reqdb.magic=&reqdb.wmo=&MR=1. Accessed September 21, 2015.

[66] Weather Underground,. Tempo Histórico para LPPT – Quarta-feira, Abril 15, 2015.
http://www.wunderground.com/history/airport/LPPT/2015/4/15/DailyHistory.html?
req_city=&req_state=&req_statename=&reqdb.zip=&reqdb.magic=&reqdb.wmo=. Accessed
September 21, 2015.

[67] Mahrt, L. Surface Wind Direction Variability. Journal of Applied Meteorology and Climatology,
50(1):144–152, January 2011.

78
Appendix A

Tangent ogive profile

Figure A.1 represents a circumference arc from which is found the ogive profile y(x), where L is the ogive
length and d is the respective diameter at the base.

d/2
r
L

R
k

Figure A.1: Ogive profile.

Considering the circumference equation

R 2 = x2 + y 2 , (A.1)

we get for x = 0 (where y = k + d/2)

R0 = k + d/2 , (A.2)

and for x = L (where y = k)

p
RL = L2 + k 2 . (A.3)

79
Since R0 = RL , we can join (A.2) and (A.3) yielding

L2 d
k= − . (A.4)
d 4

Substituting k in (A.2), and since R = R0 , we get

L2 d
R= + . (A.5)
d 4

The ogive profile is given by r = y − k, thus

p
r= R 2 − x2 − k , (A.6)

where k and R can be replaced by (A.4) and (A.5), respectively, resulting


s 2 !
L2 d L2 d
r(x) = + − x2 − − . (A.7)
d 4 d 4

80
Appendix B

Connector’s CM and inertia

In Figure B.1 the thick lines represent a connector’s half section where L is the surface length, h is the
height and r1 and r2 are, respectively, the fore and aft radius. In order to find the CM and inertia by
integration, it is used the referential xoy.

r1

o
x
y L

h θ

r2

Figure B.1: Connector’s half section (thick lines).

Since the connector is a conical surface with thickness t, its elemental volume is

dV = t dA , (B.1)

where the elemental area, dA, is

dA = 2π(r1 + x sin θ) dx , (B.2)

and sin θ is given by

r2 − r1
sin θ = . (B.3)
L

Therefore, integrating (B.2) from 0 to L, and using (B.3), the total area is

A = πL(r1 + r2 ) . (B.4)

81
The centroid is found from
Z L
1
ȳ = y dV , (B.5)
V 0

where y is

h
y = x cos θ = x . (B.6)
L

Thus, replacing (B.1), (B.2), (B.3), (B.4) and (B.6) in (B.5), yields
Z L  
1 h x
y= x t2π r1 + (r2 − r1 ) dx =
tπL(r1 + r2 ) 0 L L
Z L  
2h x
= 2 x r1 + (r2 − r1 ) dx =
L (r1 + r2 ) 0 L
 
h r1 + 2r2
= . (B.7)
3 r1 + r2

The transversal inertia relative to the radius r1 , is found from


Z L
I= y 2 dm . (B.8)
0

where the elemental mass is given by

dm = ρ dV , (B.9)

and ρ is the material’s density. Hence, using (B.1), (B.2), (B.3), (B.6) and (B.9) in (B.8), it becomes

Z L  2  
h x
I= x ρt2π r1 + (r2 − r1 ) dx =
0 L L
1 2
= h Lπtρ(r1 + 3r2 ) . (B.10)
6

Since, from B.4, the connector’s mass is

m = ρtπL(r1 + r2 ) , (B.11)

r1 +r2
then, multiplying (B.10) by r1 +r2 , and applying (B.11), the inertia can be found by
 
1 r1 + 3r2
I = mh2 . (B.12)
6 r1 + r2

82

Você também pode gostar