You are on page 1of 241

An Investigation into Static and Dynamic

Analysis of an Airborne Tethered Aerostat

A Thesis Submitted

in Partial Fulfillment of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

by

AJIT KUMAR

to the

Department of Aerospace Engineering


Indian Institute of Technology, Kanpur

May, 2017
iii
Dedicated to

My Wife Gunjan & Son Abhyudaya


vii

ACKNOWLEDGMENT

I express my sincere gratitude to my thesis supervisors Prof. A. K. Ghosh and Dr. S.

C. Sati, for their invaluable guidance and encouragement throughout the course of this

research work. I consider it as a privilege and great opportunity to have association with

them. They have been a constant source of inspiration to learn various aspects of flight

dynamics and research activities. I am highly indebted to them for their active enthusiasm

and untiring support.

I would like to express my sincere thanks to all my colleagues, lab mates and friends

who provided useful help during the course of present research work. I am thankful to all,

who made the research work as a memorable one and interesting.

Finally, I take this opportunity to express my deepest and sincere thanks to my parents

and family for providing the required emotional support and making this journey so simple,

smooth, beautiful and memorable.

Ajit Kumar

IIT Kanpur, 2017


ix

SYNOPSIS

The present work is predominantly motivated by the requirement to investigate into

static and dynamic analysis of an actual experimental aerostat during its design and

development phase. During the past few decades, tethered aerostats have become a proven

aerial platform for a variety of civil and military applications. In addition to providing

long and persistent endurance, the tethered aerostats are cheaper and without noise as

compared to their counterparts such as an unmanned aerial vehicle. Tethered aerostats are

passive lighter-than-air vehicles restrained by a single tether fixed to the ground. These

contain a gas (usually helium) having lower density as compared with ambient air which is

enclosed in an envelope and the difference in their densities gives rise to buoyancy. Hence

aerostat envelope derives the lifting force mainly by the buoyancy effect in contrast to

the conventional aircraft, where relative motion generates lift. Due to unavailability of on

board propulsion and control system, the performance of tethered aerostats is primarily

driven by the effect of natural wind which sometimes causes large displacements and

attitude. The present work addresses the aspects of static and dynamic analysis for the

experimental aerostat including the method of estimation of all the required parameters

to carry out these analysis.

The major components of the present experimental aerostat envelope include main

helium filled hull compartment, helium filled fins for stability, air filled ballonet for tem-

perature & height compensation, patches for creating hard points on the surface, cordage

for securely holding the envelope, pressurization system for maintaining the pressure and

actual payloads for specific purposes. The aerostat envelope is anchored to the ground

through a tether. Before presenting the static and dynamic models of the experimental

aerostat, a set of input parameters have been estimated which are required for this pur-

pose. These parameters uniquely describe the aerostat and include aerostatic properties,

masses and moments of inertia (including apparent mass), aerodynamic coefficients and

derivatives and tether cable properties. The mass and aerostatic properties for the teth-

ered aerostat have been estimated using the solid modeling approach. The solid modeling

is a modern software based tool, which models any system to component level to esti-

mate mass, geometric and inertia properties. The analytical method to estimate these
x

properties is based on the numerical integration approach where a particular property

is evaluated for a small element and then integrated to the whole body. The analytical

approach is suitable for simple geometry - hull in the present aerostat case, but provides

an efficient way to validate the approach of solid modeling. Hence before applying the

solid modeling technique to the present aerostat system, validation has been done with

analytical approach for hull. Certain other mass and inertia properties are estimated using

a combination of solid modeling, analytical approach and direct measurement.

The aerodynamic coefficients and derivatives are required to describe the aerodynamic

forces on the aerostat envelope. To carry out the equilibrium and static analysis in lon-

gitudinal direction, only lift, drag and pitching moment coefficients are required. These

coefficients for the present aerostat has been estimated based on three different approaches

namely semi-empirical method, Computational Fluid Dynamics (CFD) and wind tunnel

testing. To quickly estimate the lift, drag and moment coefficient of the aerostat envelope

in the absence of any reliable aerodynamic data, a semi empirical method is available in

the literature and is based on hull-fin interference factors which are defined in the an-

alytical model and obtained from experimental data as functions of certain geometrical

parameters. This method applies to the low speed regime, when the flow is attached and

no flow separation has occurred over the aerostat hull. The CFD approach to estimate

the aerodynamic coefficients is based on the numerical solution of Navier-Stokes equations

using commercial software. Since the ease of availability of high performance computing

facility, this has become quite feasible. The CFD is also able to present the variation of

pressure and velocity around the aerostat envelope surface as well as to visualize the flow

pattern. The wind tunnel testing has also been carried out for the scaled down model of

the envelope where a Fiber Reinforced Plastic (FRP) model has been fabricated, properly

instrumented and tested in National Wind Tunnel Facility (NWTF), IIT Kanpur. A com-

parative study has also been done for the results using these three different methods which

suggests that the results are in good agreement for low magnitude of angle of attack.

For the dynamic analysis, additional coefficients as well as dynamic derivatives will

also be required. These derivatives have been estimated using approximations available

in the literature. The present aerostat has got three fins, one vertical fin on the upper

surface and two inclined fins in inverted-Y configuration. This is primarily done to save

weight. The aerostat envelope’s two inclined fins are approximated by their components
xi

in horizontal and vertical directions for use in estimating these derivatives.

Under a constant horizontal wind speed, the aerostat attains equilibrium with a trim

angle and tether tension and the tether attains equilibrium with a particular shape from

the confluence point to the anchor point. This also results in the horizontal and vertical

shift of aerostat envelope from the zero wind position. The objective of carrying out

equilibrium analysis is to estimate the tether tension, trim angle of attack and the tether

cable shape for a given horizontal wind speed. The forces under the equilibrium condition

for an aerostat include buoyancy force, gravity force, aerodynamic force and tether-cable

force where the tether-cable force can be expressed in terms of other forces. The buoyancy

force and tether force are additional terms in comparison to an aircraft. The expression

for estimating trim angle of attack has been proposed by balancing the forces and the

pitching moment and the results are plotted for the considered range of wind speed. It is

important to estimate trim angle of attack for the aerostat under the operating wind speed

to examine whether the estimated trim angle lies within the range of specified limits. It

is also important to examine this aerostat envelope for pitch static stability throughout

the equilibrium range to ensure that it has a tendency to return to its original equilibrium

state following unforeseen disturbances. Using the total moment coefficient, a generalized

criterion for the static stability of aerostat has been proposed. The sensitivity studies

are also carried out for the variation of trim angle with respect to temperature, geometric

and aerodynamic parameters. For estimation of cable shape, equilibrium cable parameters

have been expressed using the available literature followed by two proposed expressions for

polygonal approximations. These approximations have been validated with the available

literature for a small size aerostat. As is the case with medium and large size aerostats,

wind speed and density variations with height cannot be neglected. Since the tether can

be divided into finite number of elements from anchor point to confluence point, wind

speed and air density variations with height can be easily accounted for in polygonal

approximations. Once, the tether tension and angle at the confluence point are known,

the method proceeds downward with approximation of these parameters for the next lower

element. An error analysis has also been carried out for the case of uniform wind speed

and density for the small aerostat.

To predict the dynamic stability behavior of the tethered aerostat envelope under a

uniform horizontal wind speed, a generalized six degrees of freedom dynamic modeling
xii

approach for the envelope and a suitable dynamic model for tether is required. Unlike an

aircraft, the model for tethered aerostat also requires to consider the forces and moments

due to buoyancy, apparent mass and tether. The buoyancy and apparent mass can be

easily taken into account in the generalized six degrees of freedom equations whereas tether

modeling requires additional assumptions and also adds further to complexity. Using the

proposed approximations for the evaluation of tether cable shape, the dynamic stability

analysis of this aerostat has been carried out using small perturbation assumptions for

the evaluation of eigen-values in longitudinal and lateral directions. The eigen-values

indicate dynamic stability of the aerostat system. An aerostat envelope may be subjected

to initial disturbances during its operation, particularly during launch and recovery. These

disturbances may be in terms of distance, angle, velocity, angular velocity or a combination

of these. It is also important to predict the motion of aerostat envelope to these initial

disturbances. Once the eigenvalues and corresponding eigenvectors are evaluated, the

response of this aerostat envelope due to initial disturbances has been generated. To

account for the effect of non-linearity, detailed motions and peak tether tensions under the

influence of a time-dependent wind vector, a non linear dynamic analysis of the tethered

aerostat is then carried out where tether is modeled as lumped masses. The theoretical

model incorporates dynamic motion of the aerostat envelope and a dynamic tether. The

proposed non-linear model has been used to study the response of the present aerostat to

time dependent wind vector and also to simulated wind gusts.

To demonstrate the efficacy of the proposed static and dynamic models for the aerostat

and correlation among them, a comparative study has been carried out. The equilibrium

analysis approach has been compared with nonlinear dynamic modeling approach for a

constant horizontal wind speed. The equilibrium analysis and nonlinear analysis results

have also been compared with the limited available experimental flight data. The com-

parisons demonstrate the efficacy of the proposed static and dynamic models.
Table of Contents

List of Figures xix

List of Tables xxiii

List of Symbols xxvii

1 Introduction 1

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Tethered Aerostats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2.1 Subsystems of Tethered Aerostat . . . . . . . . . . . . . . . . . . . . 5

1.2.2 Aerostat Envelope . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.2.3 Advantages and Applications of Tethered Aerostat . . . . . . . . . . 9

1.3 Fundamentals of Tethered Aerostat . . . . . . . . . . . . . . . . . . . . . . . 9

1.3.1 Line of Sight Coverage . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.3.2 Principle of Aerostatics . . . . . . . . . . . . . . . . . . . . . . . . . 11

1.3.2.1 The Atmospheric Air Properties . . . . . . . . . . . . . . . 12

1.3.2.2 Off-standard Atmospheric Air Properties . . . . . . . . . . 13

1.3.2.3 Other factors affecting lift . . . . . . . . . . . . . . . . . . . 14

1.3.3 Shape and Volume Estimation of Aerostat Envelope . . . . . . . . . 14

1.4 Objectives of Present Research . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.5 Aerostats Considered for Analysis . . . . . . . . . . . . . . . . . . . . . . . 17

1.6 Layout of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2 Estimation of Input Parameters for Static and Dynamic Analysis 19

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.2 Aerostatic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

xiii
xiv

2.2.1 Analytical Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.2.2 Solid modeling of Aerostat Envelope . . . . . . . . . . . . . . . . . . 25

2.2.2.1 Hull Modeling . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.2.2.2 Fin Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.2.2.3 Ballonet Modeling . . . . . . . . . . . . . . . . . . . . . . . 27

2.2.2.4 Cordages, Patches, Accessories and Payloads modeling . . 28

2.2.2.5 Comparison of Hull Aerostatic Properties . . . . . . . . . . 28

2.2.3 Total Mass of Aerostat Envelope . . . . . . . . . . . . . . . . . . . . 28

2.2.4 Total Moments of Inertia of Aerostat Envelope . . . . . . . . . . . . 29

2.2.5 Center of Buoyancy of Aerostat Envelope . . . . . . . . . . . . . . . 30

2.2.6 Structural Center of Mass of Aerostat Envelope . . . . . . . . . . . . 31

2.3 Apparent air mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.4 Aerodynamic Coefficients and Derivatives . . . . . . . . . . . . . . . . . . . 33

2.4.1 Semi-Empirical Method . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.4.2 Computational Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . 41

2.4.2.1 Geometrical Description . . . . . . . . . . . . . . . . . . . . 41

2.4.2.2 Grid Generation and Boundary Conditions . . . . . . . . . 43

2.4.2.3 Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . 43

2.4.2.4 Results of CFD Analysis . . . . . . . . . . . . . . . . . . . 44

2.4.3 Wind Tunnel Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

2.4.3.1 Wind Tunnel Model Details . . . . . . . . . . . . . . . . . . 51

2.4.3.2 Test Instrumentation . . . . . . . . . . . . . . . . . . . . . 51

2.4.3.3 Model Mounting Scheme and Results . . . . . . . . . . . . 52

2.4.4 Comparison of Aerodynamic Coefficients for Different Methods . . . 53

2.4.5 Calculation of Dynamic and other Stability Derivatives . . . . . . . 57

2.4.6 Longitudinal Dynamic Derivatives . . . . . . . . . . . . . . . . . . . 58

2.4.7 Lateral Dynamic Derivatives . . . . . . . . . . . . . . . . . . . . . . 58

2.5 Tether cable properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3 Equilibrium and Static Analysis of the Tethered Aerostat 63

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

3.2 Tether Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64


xv

3.3 Equilibrium Analysis of Aerostat Envelope . . . . . . . . . . . . . . . . . . . 70

3.3.1 Moment Coefficient about CG . . . . . . . . . . . . . . . . . . . . . 73

3.3.2 Static Longitudinal Stability . . . . . . . . . . . . . . . . . . . . . . 73

3.3.3 Results for Aerostat Equilibrium and Static Stability Analysis . . . 74

3.3.4 Results for Sensitivity Analysis on Aerostat equilibrium . . . . . . . 78

3.3.4.1 Center of Gravity (CG) Variation . . . . . . . . . . . . . . 78

3.3.4.2 Confluence Point (CP) Variation . . . . . . . . . . . . . . . 80

3.3.4.3 Center of Buoyancy (CB) Variation . . . . . . . . . . . . . 81

3.3.4.4 Aerodynamic Coefficients Variation . . . . . . . . . . . . . 83

3.3.5 Approximate Expressions . . . . . . . . . . . . . . . . . . . . . . . . 85

3.4 Equilibrium Cable Configuration . . . . . . . . . . . . . . . . . . . . . . . . 87

3.4.1 Cable Configuration for Uniform Wind Speed and Density . . . . . . 87

3.4.2 Approximation for Wind Speed and Air Density Variation . . . . . . 90

3.4.2.1 Polygonal Approximation 1 . . . . . . . . . . . . . . . . . . 91

3.4.2.2 Polygonal Approximation 2 . . . . . . . . . . . . . . . . . . 91

3.4.3 Results on Equilibrium Cable Configuration . . . . . . . . . . . . . . 92

3.5 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

4 Dynamic Stability Analysis of the Tethered Aerostat 105

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

4.2 Description of Aerostat System . . . . . . . . . . . . . . . . . . . . . . . . . 106

4.3 Rigid Body Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . 107

4.4 Aerostat Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 109

4.4.1 Apparent Masses and Moments of Inertia . . . . . . . . . . . . . . . 112

4.4.2 Aerodynamic Forces and Moments . . . . . . . . . . . . . . . . . . . 114

4.4.3 Tether-Cable Forces and Moments . . . . . . . . . . . . . . . . . . . 115

4.4.4 Buoyancy Forces and Moments . . . . . . . . . . . . . . . . . . . . . 119

4.4.5 Gravity Forces and Moments . . . . . . . . . . . . . . . . . . . . . . 119

4.4.6 Envelope Equations of Motion . . . . . . . . . . . . . . . . . . . . . 120

4.5 Stability Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

4.6 Envelope Stability Characteristics . . . . . . . . . . . . . . . . . . . . . . . . 123

4.7 Response of Envelope due to Initial Disturbances . . . . . . . . . . . . . . . 126


xvi

4.7.1 Real Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

4.7.2 Complex Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

4.8 Validation of Dynamic Stability Model . . . . . . . . . . . . . . . . . . . . . 128

4.9 Results and Discussions for Test Aerostat . . . . . . . . . . . . . . . . . . . 131

4.9.1 Longitudinal Stability Characteristics . . . . . . . . . . . . . . . . . 132

4.9.2 Lateral Stability Characteristics . . . . . . . . . . . . . . . . . . . . 137

4.9.3 Response to Initial Disturbances . . . . . . . . . . . . . . . . . . . . 141

4.10 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

5 Non Linear Dynamic Analysis of the Tethered Aerostat 153

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

5.2 Aerostat System Description . . . . . . . . . . . . . . . . . . . . . . . . . . 154

5.3 Aerostat Envelope Dynamic Model . . . . . . . . . . . . . . . . . . . . . . . 156

5.3.1 Buoyancy Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

5.3.2 Gravity Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

5.3.3 Aerodynamic Forces and Moments . . . . . . . . . . . . . . . . . . . 157

5.3.4 Apparent Mass Terms . . . . . . . . . . . . . . . . . . . . . . . . . . 158

5.3.5 Terminal Tether Segment . . . . . . . . . . . . . . . . . . . . . . . . 159

5.3.6 Aerostat Envelope Dynamic Equations . . . . . . . . . . . . . . . . . 161

5.4 Tether Dynamic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

5.5 Solution Procedure of Equations . . . . . . . . . . . . . . . . . . . . . . . . 163

5.6 System Input Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

5.7 Simulations Results and Discussions . . . . . . . . . . . . . . . . . . . . . . 165

5.7.1 Simulation 1: Wind Variation in Magnitude . . . . . . . . . . . . . . 165

5.7.2 Simulation 2: Stepped Wind Variation in Magnitude . . . . . . . . . 169

5.7.3 Simulation 3: Stepped Wind Variation both in Magnitude and Di-

rection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

5.7.4 Simulation 4: Wind Gust in Horizontal Direction . . . . . . . . . . . 178

5.7.5 Simulation 5: Wind Gust in Vertical Direction . . . . . . . . . . . . 181

5.7.6 Simulation 6: Wind Gust in Side Direction . . . . . . . . . . . . . . 185

5.8 Comparison of Nonlinear Analysis with Available Experimental Data . . . . 189

5.9 Comparison of Equilibrium Analysis with Available Experimental Data . . 193


xvii

5.10 Comparison of Nonlinear Analysis with Equilibrium Analysis . . . . . . . . 193

5.11 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

6 Concluding Remarks 195

6.1 Innovative Research Work Carried Out . . . . . . . . . . . . . . . . . . . . . 195

6.2 Scope of Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

Bibliography 199
List of Figures

1.1 Schematic of a typical aerostat system . . . . . . . . . . . . . . . . . . . . . 5

1.2 Geometry for radial line of sight coverage [23] . . . . . . . . . . . . . . . . . 10

1.3 Variation of coverage radius with aerostat altitude . . . . . . . . . . . . . . 11

1.4 The mechanism of aerostatic lift [21] . . . . . . . . . . . . . . . . . . . . . . 12

1.5 Flow chart for volume estimation of aerostat envelope . . . . . . . . . . . . 15

2.1 Hull geometry of aerostat envelope . . . . . . . . . . . . . . . . . . . . . . . 20

2.2 The complete solid model of aerostat envelope . . . . . . . . . . . . . . . . . 27

2.3 Variation of apparent mass and inertia coefficients with fineness ratio [34] . 32

2.4 Schematic of steady-state analytical model for semi-empirical method [18] . 34

2.5 Fin planform definition for semi-empirical method [18] . . . . . . . . . . . . 34

2.6 Fins cross flow drag coefficient as function of aspect and taper ratio [52] . . 38

2.7 3D model for aerostat envelope used for CFD analysis . . . . . . . . . . . . 42

2.8 3D external domain for aerostat envelope used for CFD analysis . . . . . . 42

2.9 Unstructured mesh for CFD analysis . . . . . . . . . . . . . . . . . . . . . . 44

2.10 Mesh clustering near wall of hull and fins of aerostat envelope . . . . . . . . 44

2.11 Pressure coefficient variation over hull at 0 deg angle of attack . . . . . . . 46

2.12 Pressure contour at an angle of attack 0 deg . . . . . . . . . . . . . . . . . . 47

2.13 Velocity contour at an angle of attack 0 deg . . . . . . . . . . . . . . . . . . 47

2.14 Pressure contour at 28m from nose and an angle of attack 0 deg . . . . . . 48

2.15 Pressure contour at 10m from nose and an angle of attack 0 deg . . . . . . 48

2.16 Velocity vector plot at an angle of attack 10 deg . . . . . . . . . . . . . . . 48

2.17 Pressure contour at an angle of attack 20 deg . . . . . . . . . . . . . . . . . 49

2.18 Velocity contour at an angle of attack 20 deg . . . . . . . . . . . . . . . . . 49

2.19 Pressure contour at 10m from nose and an angle of attack 20 deg . . . . . . 49

xix
xx

2.20 Pressure contour at 28m from nose and an angle of attack 20 deg . . . . . . 50

2.21 Pressure contour at 10m from nose and an angle of attack -20 deg . . . . . 50

2.22 Pressure contour at 28m from nose and an angle of attack -20 deg . . . . . 50

2.23 Aerostat envelope model mounted on robotic arm . . . . . . . . . . . . . . . 53

2.24 Aerostat envelope model mounted on turn table . . . . . . . . . . . . . . . . 53

2.25 Lift coefficient as a function of angle of attack for different methods . . . . 55

2.26 Drag coefficient as a function of angle of attack for different methods . . . . 56

2.27 Moment coefficient about nose vs. angle of attack for different methods . . 56

2.28 Projection of tails for aerostat envelope for estimating stability derivatives . 57

2.29 Static load testing of tether sample . . . . . . . . . . . . . . . . . . . . . . . 61

3.1 Forces and moments acting on aerostat in equilibrium condition . . . . . . . 64

3.2 Tether tension variation with wind speed and αt for small aerostat . . . . . 66

3.3 Angle with horizontal variation with wind speed and αt for small aerostat . 67

3.4 Tether tension variation with wind speed and αt for test aerostat . . . . . . 69

3.5 Angle with horizontal variation with wind speed and αt for test aerostat . . 69

3.6 Variation of trim angle of attack with wind speed for test aerostat . . . . . 75

3.7 Variation of dCmg /dα with wind speed for test aerostat . . . . . . . . . . . 76

3.8 Total moment coefficient with αt at 5 m/s wind speed for test aerostat . . . 77

3.9 Total moment coefficient with αt at 25 m/s wind speed for test aerostat . . 77

3.10 Trim angle with wind speed for CG sensitivity for test aerostat . . . . . . . 79

3.11 dCmg /dα with wind speed for CG sensitivity for test aerostat . . . . . . . . 79

3.12 Trim angle of attack with wind speed for CP sensitivity for test aerostat . . 80

3.13 dCmg /dα with wind speed for CP sensitivity for test aerostat . . . . . . . . 81

3.14 Trim angle of attack with wind speed for CB sensitivity for test aerostat . . 82

3.15 dCmg /dα with wind speed for CB sensitivity for test aerostat . . . . . . . . 83

3.16 αt with V∞ for aerodynamic coefficients sensitivity for test aerostat . . . . . 84

3.17 dCmg /dα with V∞ for aerodynamic coefficients sensitivity for test aerostat . 84

3.18 Trim angle with V∞ for approximate expression for test aerostat . . . . . . 86

3.19 dCmg /dα with V∞ for approximate expression for test aerostat . . . . . . . 86

3.20 The coordinate system and forces acting on the tether cable [42] . . . . . . 88

3.21 Polygonal approximation of tether cable element . . . . . . . . . . . . . . . 91


xxi

3.22 Non-dimensional cable profile for small aerostat using PA-1 & V∞ = 20 m/s 93

3.23 Non-dimensional cable tension for small aerostat using PA-1 & V∞ = 20 m/s 94

3.24 Non-dimensional cable angle for small aerostat using PA-1 & V∞ = 20 m/s 94

3.25 Non-dimensional cable profile for small aerostat using PA-2 & V∞ = 20 m/s 95

3.26 Non-dimensional cable tension for small aerostat using PA-2 & V∞ = 20 m/s 95

3.27 Non-dimensional cable angle for small aerostat using PA-2 & V∞ = 20 m/s 96

3.28 Percentage error with number of elements for small aerostat for V∞ = 20 m/s 96

3.29 Non-dimensional cable profiles for small aerostat with different V∞ . . . . . 97

3.30 Non-dimensional cable tensions for small aerostat with different V∞ . . . . 98

3.31 Non-dimensional cable angles for small aerostat with different V∞ . . . . . . 98

3.32 Non-dimensional cable profiles for test aerostat with different V∞ . . . . . . 99

3.33 Non-dimensional cable tensions for test aerostat with different V∞ . . . . . 100

3.34 Non-dimensional cable angles for test aerostat with different V∞ . . . . . . 100

3.35 Non-dimensional cable profiles for test aerostat with fixed operational height101

3.36 A typical measured wind speed with height . . . . . . . . . . . . . . . . . . 102

3.37 Non-dimensional cable profile for test aerostat for the measured wind speed 102

3.38 Non-dimensional cable tension for test aerostat for the measured wind speed 103

3.39 Non-dimensional cable angle for test aerostat for the measured wind speed . 103

4.1 Tethered aerostat schematic considered in dynamic stability analysis [42] . . 107

4.2 Orientation of stability axis system [13] . . . . . . . . . . . . . . . . . . . . 110

4.3 Longitudinal root locus presented for the 7.64m long reference balloon in [42]129

4.4 Longitudinal root locus generated for the 7.64m long reference balloon . . . 129

4.5 Lateral root locus presented for the 7.64m long reference balloon in [42] . . 130

4.6 Lateral root locus generated for the 7.64m long reference balloon . . . . . . 130

4.7 Variation of longitudinal damping parameter with velocity for test aerostat 133

4.8 Variation of longitudinal circular frequency with velocity for test aerostat . 133

4.9 Longitudinal root locus plot with velocity as a parameter . . . . . . . . . . 134

4.10 Variation of lateral damping parameter with velocity for test aerostat . . . 138

4.11 Variation of lateral circular frequency with velocity for test aerostat . . . . 138

4.12 Lateral root locus plot with velocity as a parameter . . . . . . . . . . . . . 139

4.13 Longitudinal response due to initial disturbance case LO1 . . . . . . . . . . 143


xxii

4.14 Longitudinal response due to initial disturbance case LO2 . . . . . . . . . . 143

4.15 Longitudinal response due to initial disturbance case LO3 . . . . . . . . . . 144

4.16 Longitudinal response due to initial disturbance case LO4 . . . . . . . . . . 144

4.17 Longitudinal response due to initial disturbance case LO5 . . . . . . . . . . 145

4.18 Longitudinal response due to initial disturbance case LO6 . . . . . . . . . . 145

4.19 Longitudinal response due to initial disturbance case LO7 . . . . . . . . . . 146

4.20 Lateral response due to initial disturbance case LA1 . . . . . . . . . . . . . 146

4.21 Lateral response due to initial disturbance case LA2 . . . . . . . . . . . . . 148

4.22 Lateral response due to initial disturbance case LA3 . . . . . . . . . . . . . 148

4.23 Lateral response due to initial disturbance case LA4 . . . . . . . . . . . . . 149

4.24 Lateral response due to initial disturbance case LA5 . . . . . . . . . . . . . 149

4.25 Lateral response due to initial disturbance case LA6 . . . . . . . . . . . . . 150

4.26 Lateral response due to initial disturbance case LA7 . . . . . . . . . . . . . 150

5.1 Tethered aerostat schematic for nonlinear analysis . . . . . . . . . . . . . . 155

5.2 Wind variation in magnitude for simulation 1 . . . . . . . . . . . . . . . . . 166

5.3 Body frame linear velocity components variation for simulation 1 . . . . . . 167

5.4 Body frame angular velocity components variation for simulation 1 . . . . . 167

5.5 Euler angles variation for simulation 1 . . . . . . . . . . . . . . . . . . . . . 168

5.6 Earth-fixed position of center of mass variation for simulation 1 . . . . . . . 168

5.7 Tether tension at confluence point variation for simulation 1 . . . . . . . . . 169

5.8 Wind variation in magnitude for simulation 2 . . . . . . . . . . . . . . . . . 170

5.9 Body frame linear velocity components variation for simulation 2 . . . . . . 171

5.10 Body frame angular velocity components variation for simulation 2 . . . . . 171

5.11 Euler angles variation for simulation 2 . . . . . . . . . . . . . . . . . . . . . 172

5.12 Earth-fixed position of center of mass variation for simulation 2 . . . . . . . 172

5.13 Tether tension at confluence point variation for simulation 2 . . . . . . . . . 173

5.14 Wind variation in magnitude and direction for simulation 3 . . . . . . . . . 175

5.15 Body frame linear velocity components variation for simulation 3 . . . . . . 175

5.16 Body frame angular velocity components variation for simulation 3 . . . . . 176

5.17 Euler angles variation for simulation 3 . . . . . . . . . . . . . . . . . . . . . 176

5.18 Earth-fixed position of center of mass variation for simulation 3 . . . . . . . 177


xxiii

5.19 Tether tension at confluence point variation for simulation 3 . . . . . . . . . 177

5.20 Horizontal wind gust for simulation 4 . . . . . . . . . . . . . . . . . . . . . . 179

5.21 Aerostat forward and vertical speed variation for simulation 4 . . . . . . . . 179

5.22 Aerostat pitch rate and pitch angle variation for simulation 4 . . . . . . . . 180

5.23 Aerostat horizontal and vertical displacement variation for simulation 4 . . 180

5.24 Tether tension at confluence point variation for simulation 4 . . . . . . . . . 181

5.25 Vertical wind gust for simulation 5 . . . . . . . . . . . . . . . . . . . . . . . 183

5.26 Aerostat forward and vertical speed variation for simulation 5 . . . . . . . . 183

5.27 Aerostat pitch rate and pitch angle variation for simulation 5 . . . . . . . . 184

5.28 Aerostat horizontal and vertical displacement variation for simulation 5 . . 184

5.29 Tether tension at confluence point variation for simulation 5 . . . . . . . . . 185

5.30 Side wind gust for simulation 6 . . . . . . . . . . . . . . . . . . . . . . . . . 186

5.31 Aerostat forward and side speed variation for simulation 6 . . . . . . . . . . 187

5.32 Aerostat pitch rate and yaw rate variation for simulation 6 . . . . . . . . . 187

5.33 Aerostat roll angle and yaw angle variation for simulation 6 . . . . . . . . . 188

5.34 Aerostat side and vertical position variation for simulation 6 . . . . . . . . . 188

5.35 Tether tension at confluence point variation for simulation 6 . . . . . . . . . 189

5.36 Measured wind speed in horizontal direction . . . . . . . . . . . . . . . . . . 191

5.37 Comparison between computed and measured pitch angle . . . . . . . . . . 192

5.38 Comparison between computed and measured tether tension . . . . . . . . . 192


List of Tables

2.1 Hull parameter comparison for analytical and solid modeling approach . . . 28

2.2 Aerostat envelope fin parameters and other assumptions for lift curve slope

estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.3 Estimation of parameters for Semi Empirical Method . . . . . . . . . . . . . 39

2.4 Lift, drag and pitching moment coefficient about envelope nose for the test

aerostat envelope evaluated by semi-empirical method . . . . . . . . . . . . 40

2.5 Lift, drag and pitching moment coefficient about envelope nose for the test

aerostat envelope evaluated by CFD . . . . . . . . . . . . . . . . . . . . . . 45

2.6 Lift, drag and pitching moment coefficient about envelope nose for the test

aerostat envelope evaluated by wind tunnel testing . . . . . . . . . . . . . . 54

3.1 Parameters for evaluating tether tension vector for small aerostat . . . . . . 65

3.2 Parameters for evaluating tether tension vector and carrying out equilib-

rium and static analysis for test aerostat (ISA+15 deg) . . . . . . . . . . . 68

3.3 Tether tension and angle with horizontal at the confluence point for small

and test aerostat for αt = 5 deg . . . . . . . . . . . . . . . . . . . . . . . . . 70

3.4 Tether parameters for the small and test aerostat . . . . . . . . . . . . . . . 92

4.1 Additional parameters required for the test aerostat envelope for dynamic

stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

4.2 Aerodynamic derivatives of test aerostat envelope about center of mass . . 132

4.3 Longitudinal eigenvalues and eigenvectors for wind speed of 5 m/s . . . . . 134

4.4 Longitudinal eigenvalues and eigenvectors for wind speed of 20 m/s . . . . . 135

4.5 Lateral eigenvalues and eigenvectors for wind speed of 5 m/s . . . . . . . . 139

4.6 Lateral eigenvalues and eigenvectors for wind speed of 20 m/s . . . . . . . . 140

4.7 Response to initial disturbance cases for longitudinal motions . . . . . . . . 141

xxv
xxvi

4.8 Response to initial disturbance cases for lateral motions . . . . . . . . . . . 141

5.1 Additional aerostat parameters required for nonlinear analysis . . . . . . . . 164

5.2 The operating conditions and flight configuration for flight data of test

aerostat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
List of Symbols

a Distance along envelope center line from nose to reference point, m

Ac Hull cross-sectional area, m2

AG Characteristic matrix for longitudinal motion

Ah Hull surface area, m2

apht , avt , apvt Lift curve slopes of projected horizontal tail, vertical tail and projected ver-

tical tail respectively based on aerostat envelope reference area

At Aspect ratio of aerostat envelope fin

AT Characteristic matrix for lateral motion

B Buoyancy force or weight of air displaced, N

Bb Net upward lift B − Wb , N

Bd Disposable lift of the gas bag, N

Bg Gross lift of the gas bag Vg (ρa − ρg ) g = B − WG , N

c̄ Aerostat balloon or envelope length, m

Ca , Cn Axial and normal aerodynamic force coefficient on the aerostat envelope

CD Drag coefficient

CDc Tether-cable drag coefficient

(Cdc )f Fins cross-flow drag coefficient, referenced to Sf

(Cdc )h Hull cross-flow drag coefficient, referenced to J1

xxvii
xxviii

(Cdf )0 fins zero-angle axial drag coefficient, referenced to Sf

(Cdh )0 Hull zero-angle axial drag coefficient, referenced to Sh

Cl , Cm , Cn Rolling, pitching and yawing-moment coefficient about body axes

CL Lift coefficient

Cmg Total pitching-moment coefficient about center of mass

Cmn Aerodynamic pitching-moment coefficient about nose

(Cn∗α )f Derivative of the isolated fins normal force coefficient with respect to αt , at αt = 0

and referenced to Sf

(Ct )f Fins leading-edge suction coefficient, referenced to Sf

Cv Viscous damping coefficient of tether cable segment of length l0 , N-s/m

CX , CY , CZ Force coefficients along body axes

D Aerodynamic drag force, N

dc Tether-cable diameter, m

Dh Maximum diameter of aerostat envelope hull, m

Dj Drag vector on the j th tether segment or lumped mass, N

En n × n identity matrix

Fa , Fn Axial and normal aerodynamic force on the aerostat envelope, N

F A , M A Aerostat aerodynamic forces and moments in body-axes, N and N-m

F AM , M AM Apparent mass force and moment in body-axes, N and N-m

FB Aerostat buoyancy force in body-axes, N

FL Free lift expressed as percentage of gross lift

FRe Equivalent fineness ratio of hull

F T , FT Tether terminal segment tension vector and magnitude for nonlinear analysis, N
xxix

FT0 Initial tether tension in terminal segment, N

FW Aerostat weight including inflation gas in body-axes, N

FX , FY , FZ External forces acting on envelope along stability axes, N

g Acceleration due to gravity, m/s2

ha Height of the aerostat envelope from ground, m

hp Pressure height of the envelope operation, m

hvt , hpvt The vertical distances between CG and mean aerodynamic chord of vertical tail

and projected vertical tail respectively

Iat Apparent moment of inertia in transverse direction

I AM , I AI Apparent mass and inertia matrices

IB Aerostat envelope inertia matrix for body-axes

ib , j b , kb Aerostat envelope body axes unit vectors

ie , j e , ke Aerostat envelope earth-fixed axes unit vectors (Figure 5.1)

ii , j i , ki Aerostat envelope inertial axes unit vectors (Figure 4.1)

is , j s , ks Aerostat envelope stability axes unit vectors

Ixb , Iyb , Izb , Ixz b Mass moments of inertia about envelope center of mass in the body axes

system including apparent inertia, kg-m2

Ixs , Iys , Izs , Ixz s Mass moments of inertia about envelope center of mass in the stability

axes system including apparent inertia, kg-m2

k Helium gas purity expressed in fraction of unity

k1 , k2 , k3 Apparent mass coefficients in the body frame

kr Apparent inertia coefficient for rotation about longest diameter

Ks Stiffness of the tether cable segment of length l0 , N/m

Kt Tether length factor as defined in Equation (1.10)


xxx

kxx Tether x-force per unit of x-displacement at confluence point, N/m

kxz Tether x-force per unit of z-displacement, N/m

kxθ Tether x-force per unit of pitch displacement, N/rad

kyy Tether y-force per unit of y-displacement at confluence point, N/m

kyφ Tether y-force per unit of roll displacement, N/rad

kyψ Tether y-force per unit of yaw displacement, N/rad

kzx Tether z-force per unit of x-displacement, N/m

kzz Tether z-force per unit of z-displacement at confluence point, N/m

kzθ Tether z-force per unit of pitch displacement, N/rad

kθx Tether pitching moment per unit of x-displacement, N-m/m

kθz Tether pitching moment per unit of z-displacement, N-m/m

kθθ Total tether pitching moment per unit of pitch displacement in the body-axis sys-

tem for pitch about center of mass, N-m/rad

kθθD Portion of kθθ due to rotation of envelope relative to steady tension vector at

confluence point, N-m/rad

kθθT Portion of kθθ due to displacement of confluence point, N-m/rad

kφy Tether rolling moment per unit of y-displacement, N-m/m

kφφ Tether rolling moment per unit of roll displacement in the body-axis system for

roll about center of mass, N-m/rad

kφψ Tether rolling moment per unit of yaw displacement, N-m/rad

kψy Tether yawing moment per unit of y-displacement, N-m/m

kψφ Tether yawing moment per unit of roll displacement, N-m/rad

kψψ Tether yawing moment per unit of yaw displacement in the body-axis system for

yaw about center of mass, N-m/rad


xxxi

l Tether cable length, m

L Aerodynamic lift force, N

l0 Length of tether cable segment for nonlinear analysis, m

lc Length along the tether cable from anchor point, m

(lf )1 Distance from the hull nose to the fins aerodynamic center

(lf )2 Distance from the hull nose to the fins center of cross-flow force

lh Axial distance from envelope nose to fin starting point, m

li Inertial position of tether lumped masses nodes, m

lpht , lvt , lpvt Horizontal distance between CG and mean aerodynamic chord of projected

horizontal tail, vertical tail and projected vertical tail respectively, m

ls Line of sight radius from the anchor point, m

lts Stretched length of the terminal segment, m

mb Combined mass of envelope structure and inflation gas, ms + mg , kg

mg Mass of inflation gas, kg

Mg Total moment on aerostat envelope about centre of gravity, N-m

Mn Aerodynamic moment about envelope nose, N-m

ms Envelope structural mass including payload, kg

MS1 , MS2 Moments defined by Equations (4.32)

mt Mass of length l0 of tether cable segment or lumped mass, kg

mx , my , mz Masses defined by Equations (4.33)

MX , MY , MZ Rolling, pitching, and yawing moments about stability axes, N-m

mxb , myb , mzb Aerodynamic apparent masses associated with accelerations in body axes,

kg

mxo , myo , mzo Total masses of envelope for accelerations along stability axes, kg
xxxii

mxs , mys , mzs Apparent masses associated with accelerations in stability axes, kg

nd Cable drag per unit length for cable normal to the wind, N/m

N Number of tether segments or lumped masses

p̄ wc /2n

Pb , Qb , Rb Rolling, pitching, and yawing rates about the body axes, rad/sec

ph Pressure for ISA at a pressure height hp , N/m2

ph0 Pressure at ISA sea level, N/m2

PL Payload capacity of aerostat at pressure height hp , kg

Pn The normal drag force per unit length of tether cable, N/m

ps , qs , rs Perturbation rolling, pitching, and yawing rates about stability axes, rad/sec

Ps , Qs , Rs Rolling, pitching, and yawing rates about the stability axes, rad/sec
p
q̄ 1 + p̄2

q0 Steady-state dynamic pressure = ρa V∞ 2 /2, N/m2

cp
r cb
cg , r cg Position vector from the aerostat mass center to the center of buoyancy and

confluence point, m

RE Earth’s mean radius, m

rh Hull radius

S Characteristic or reference area of envelope, Vh 2/3 , m2

Sf Fins reference area, m2

Sfh Fins area which is inside the hull (Figure 2.5), m2

Sh Hull reference area taken same as S, m2

Sl Meridian cross section area of envelope hull, m2

t Time, sec
xxxiii

T Tether -cable tension during linear analysis, N

T0 , T1 Tensions of tether cable at lower and upper ends during dynamic stability analysis,

TB E
E , T B Matrix for transformation from earth-fixed axes to body axes and vice versa

(Figure 5.1)

T SI , T SI Matrix for transformation from inertial axes to stability axes and vice versa (Fig-

ure 4.1)

Tg Wind gust time period, s

Th Temperature for ISA at a pressure height hp , K

Th0 Temperature at ISA sea level, K

uA , vA , wA Aerodynamic velocity components in body-axes, m/sec

Ub , Vb , Wb Velocities components of envelope center of mass along the body axes, m/s

Ug , Um Wind gust speed and maximum wind gust speed, m/s

uj , vj , wj Inertial velocity components of tether lumped masses at nodes, m/s

us , vs , ws Perturbation velocities of envelope center of mass along stability axes, m/sec

Us , Vs , Ws Velocity components of envelope center of mass along the stability axes, m/s

uwind , vwind , wwind Wind velocity components in body-axes, m/sec

V∞ Steady wind velocity, m/s

v ∗ , w∗ y-and z-perturbation velocity of envelope with respect to airstream in stability axes,

m/s

V A , VA Aerodynamic velocity vector and magnitude, m/s

Vg Helium volume of envelope, m3

Vh Hull volume of envelope, m3

vi Inertial velocity of tether lumped masses nodes, m/s


xxxiv

Vn Component of wind velocity normal to cable, m/s

Vpht , Vvt , Vpvt Tail volume ratios of projected horizontal tail, vertical tail and projected

vertical tail respectively

WB Sum of structural weight and gas weight of envelope, N

wc Tether-cable weight per unit length, N/m

WG Gas weight inside the envelope, N

WS Structural weight of the envelope including payload, N

wx , wy , wz Earth-fixed axes wind velocity components, m/s

x, y, z Perturbation displacements in the stability axis system, m

x1 , z1 Coordinates of balloon center of mass with respect to tether-cable anchor point;

x1 is horizontal and positive in downstream direction, z1 is vertical and positive

upwards, m

XB Axial distance of centre of buoyancy from envelope nose, m

xc , yc , zc Earth-fixed positions of envelope confluence point for nonlinear analysis, m

XC Axial distance of confluence point from envelope nose, m

xcs , zcs Coordinates defined by Equations (4.26)

xe , ye , ze Inertial positions of aerostat mass centre for nonlinear analysis, m

XG Axial distance of combined centre of gravity from envelope nose, m

xi , yi , zi Perturbation position components of envelope in the inertial axes system (Fig-

ure 4.1), m

xj , yj , zj Inertial position components of tether lumped masses at nodes, m

Xp Vector of unknowns in nonlinear analysis

XS Axial distance of structural centre of gravity from envelope nose, m

ZB Normal distance of centre of buoyancy from envelope nose, m


xxxv

ZC Normal distance of confluence point from envelope nose, m

ZG Normal distance of combined centre of gravity from envelope nose, m

ZS Normal distance of structural centre of gravity from envelope nose, m

α Angle of attack, rad

αt Envelope steady-state or trim angle of attack, rad

β Angle of sideslip, rad

δ Displacement defined by Equation (4.24)

∆l Finite length of tether element for polygonal approximation, m

δn Angle between nth cable element and horizontal

∆x, ∆y, ∆z Components of terminal difference vector, m

∆T Temperature difference between standard and off-standard condition, K

ε Strain in tether cable expressed as percentage

η Real part of characteristic root of stability equations, damping parameter, 1/sec

ηf Fin-efficiency factor accounting for the effect of the hull on the fins

ηk Hull-efficiency factor accounting for the effect of fins on the hull

γ Tether-cable angle with respect to the horizontal, rad

γ0 , γ1 Angles between the horizontal and tether cable at lower and upper ends, respec-

tively, rad

γn Tether-cable angle with respect to the horizontal at nth joint, rad

λ Characteristic root of stability equations, η ± iω, 1/sec

λ̄ Variable defined by Equation (3.29)

λ̄0 , λ¯1 λ̄ at lower and upper ends, respectively, of tether cable

ω Imaginary part of characteristic root of stability equations, circular frequency,

rad/sec
xxxvi

Ω Angular velocity vector of the envelope-fixed axis system with respect to the inertial

reference frame, rad/sec

ρa Atmospheric air density at operating conditions, kg/m3

ρg Helium gas density at operating conditions, kg/m3

ρg0 Impure helium gas density at ISA sea level, kg/m3

ρh Atmospheric air density for ISA at a pressure height hp , kg/m3

ρh0 Atmospheric air density at ISA sea level, kg/m3

τ Variable defined by Equation (3.26)

τ0 , τ1 τ at lower and upper ends, respectively, of tether cable

θs , φs , ψs Perturbations of Θs , Φs and Ψs respectively, rad

Θs Φs , Ψs Euler angles of pitch, roll, and yaw, respectively for stability axes, rad

Θb , Φb , Ψb Euler angles of pitch, roll, and yaw, respectively for body axes, rad

ABBREVIATIONS

CB Center of buoyancy

CM Center of mass or gravity

CP Confluence point

SCM Structural center of mass or gravity


Chapter 1

Introduction

1.1 Motivation

During the design and development of a medium size aerostat, the requirement was iden-

tified to develop suitable mathematical models for carrying out the static and dynamic

analysis for this aerostat. The present aerostat is helium filled and is tethered to the

ground through a single tether. The additional factors to be considered for this aerostat

as compared to an aircraft include buoyancy effect, tether effects and apparent mass ef-

fects. Also, as there is no propulsion system and active control system for this tethered

aerostat, the performance is primarily driven by the effect of natural wind which some-

times causes large displacements and attitude. The motivation for carrying out the present

research work is the requirement to develop suitable mathematical models for equilibrium

and static analysis, dynamic stability analysis, initial disturbance response as well as non-

linear model for wind response including gusts for this airborne tethered aerostat system.

It is also required to generate all the required inputs including aerostatic, geometric, mass,

inertia, aerodynamic and tether cable properties for carrying out static and dynamic anal-

ysis of this tethered aerostat. A comparative study for aerodynamic coefficients using

different methods as well as computed and measured flight data is also desirable.

The equilibrium analysis is useful in predicting the trim angle of attack of the aerostat

envelope, the static stability of the aerostat envelope and the tether cable shape under

the uniform wind speed. The trim angle of attack prediction is useful in deciding the

strength of the tether cable as the tether tension is dependent on it. Also the trim angle

is a very useful information for directional payloads as these payloads have to be designed

1
2

taking into account the possible variation in trim angle. The static stability analysis for

the tethered aerostat is useful in predicting the tendency of aerostat envelope to return

to its equilibrium position. The equilibrium and static analysis also estimates the tether

cable profile from the upper attachment point to the anchor point which will in turn

predict the operational height as well as horizontal sway of the aerostat from the launch

point. The requirement of carrying out simplified linearized dynamic stability analysis is

based on the estimation of characteristic roots as well as different modes of motion with

frequency and damping. The response to initial disturbances can also be estimated in

this model. But unlike aircraft, the simplified linearized dynamic model is not suitable if

we are interested in response studies under different types wind conditions. Hence it is

required to develop nonlinear model for the tethered aerostat which will give the response

to different types of wind conditions. The response to winds is required primarily to

estimate the angular rates, attitude, position and tether tension as function of time. The

angular rates are vital input parameters for designing stabilization system for the payloads

and taking higher angular rates may result in heavier and bulkier stabilization system.

Generating all the input parameters for carrying out the static and dynamic analysis is

also a very important assignment. The input parameters are in terms of the aerostat

envelope geometric, mass, inertia, aerostatic, aerodynamic properties and tether cable

stiffness and damping properties.

A tethered aerostat attains equilibrium under a uniform horizontal wind speed with

a certain value of pitch angle, also called trim angle of attack, a horizontal shift from

launch point and a vertical shift from initial height. Previous work on equilibrium analysis

of tethered aerostat envelope include pitch angle estimation for the range of operating

uniform wind speed [42], pitch angle, static margin estimation and static stability analysis

with simplified assumption of small angle [22, 17]. The approach presented in [22] was

adapted in [39] for carrying out stability analysis of aerostat. But for the practical case of

a tethered aerostat, a generalized approach for equilibrium analysis of tethered aerostat is

required which covers trim angle estimation, static stability analysis and sensitivity studies

for the entire range of operating wind speed including for large angles. The tether tension

vector also needs to be quantified.

Once the tether tension and angle with horizontal are known at the confluence point, it

is required to estimate these parameters at the anchor point and also the cable shape from
3

confluence point to anchor point. The method and relevant expressions are presented for

the case of uniform wind speed and air density in vertical direction by Neumark [36] which

is applicable only for the cases where these variations are negligible e.g. for a small size

aerostat due to low height of operation. As is the case with larger aerostats, wind speed

and density variations with height cannot be neglected. To account for such variations,

Berteaux [6] suggested an approximate step-wise change of velocity with vertical position.

Following this assumption, expressions for polygonal approximations are required to be

developed and compared with results of Neumark [36] for a small aerostat before applying

to the test aerostat configuration.

The linearized stability models of tethered aerostats have been presented by Redd et

al. [42] and DeLaurier [10, 11, 12]. These models provide insights into dynamic stability

and, in the case of the latter, mean response to turbulence; they are limited to small

perturbations and 3 degrees of freedom. The linearized model presented by Redd et

al. [42] is very efficient and forms a benchmark for carrying out stability analysis using

small perturbations assumptions. But the tether model employed in this analysis is based

on Neumark [36] which is applicable only for small aerostats as described in previous

paragraph due to assumption of uniform wind speed and air density in vertical direction.

If the tether model derived from Berteaux [6] approach is used in linearized stability model

of Redd et al. [42], it becomes suitable for application to larger aerostats. But still the

basic assumption of flexible and inextensible cable remains.

The effect of aerodynamic non-linearity, detailed motions and peak tether tensions

under the influence of a time-dependent wind vector cannot be determined in a linearized

model. To overcome this limitation, nonlinear models of the tethered aerostats have been

presented in [1, 15]. A modeling and estimation method for airships was proposed by

Li and Nahon [29]. Tether modeling is complex in [1], where beam elements are used.

Using beam elements allows modeling of bending moments but requires a finite element

nonlinear solver, and so the computational burden is much higher. Computations are

further decreased by treating each link as a body of revolution and assuming that tether

spin is negligible to the dynamics, where each link then only has two degrees of freedom

as done in [15]. The lumped-mass model for tether as proposed in [19, 28] is appealing

because they result in extremely simple, computationally efficient models, even for large

numbers of elements. Hence the aerostat model as proposed in [15] is combined with the
4

tether model as proposed in [19] to develop an efficient nonlinear model for the application

to the present test aerostat.

Generating the aerodynamic coefficients for static and dynamic analysis is also a chal-

lenging task. The semi empirical method for aerodynamic properties estimation is based

on hull-fin interference factors which are defined in the analytical model and obtained

from experimental data as functions of certain geometrical parameters, as proposed by

Jones & DeLaurier [18] and later applied to un-symmetrically finned bodies of revolutions

by Gill et al. [14]. In absence of any data, the semi-empirical method can be applied

to get a quick estimate of aerodynamic coefficients. The more advanced computational

fluid dynamics (CFD) technique has been applied by Kale et al. [20] to generate drag

coefficient of an aerostat envelope. But no literature is available on the determination of

lift, drag and moment coefficient as a function of angle of attack for an aerostat envelope

using CFD. Hence as a proposed method, these coefficients for the aerostat envelope are

estimated using CFD. The experimental method of wind tunnel testing has also been car-

ried out to generate coefficients and comparative studies have been done on results using

the three different methods. The approximation method for dynamic derivatives has been

presented in [13] and the same method has been used for the present aerostat also. The

aerostatic, mass, geometric and inertia properties of the aerostat have been estimated

using a combination of solid modeling approach combined with fundamental definitions.

1.2 Tethered Aerostats

Tethered aerostats are passive lighter-than-air vehicles restrained by a single tether fixed

to the ground. These contain a gas (usually helium) having lower density as compared

with ambient air which is enclosed in an envelope and the difference in their densities

gives rise to buoyancy. Hence aerostat envelope derives the lifting force mainly by the

buoyancy effect in contrast to the conventional aircraft, where relative motion generates

lift. The envelope gas is generally helium because it is inert and provides adequate lifting

capability. Ground based sensors have limited line of sight range due to the limitations

posed by earth’s curvature (horizon effect). Mounting these sensors on elevated platforms

like tower, aircraft and balloon can increase the line of sight range. The limitation of

the height up to which a tower can be built, is obvious. Aircraft have limited endurance
5

(on-station time) of few hours whereas aerostats can remain operational continuously for

days. Aerostats have been proven platforms for these sensors especially in surveillance

and communication role for a variety of civil and military applications.

Aerostat systems provide help in raising the electronic payloads for increasing their line

of sight range so as to overcome the terrain obstructions like trees, buildings, mountains

and similar obstructions. Aerostat system is a mission-oriented vehicle with attributes like

payload platform availability at high altitudes, increased line of sight coverage for payload

and long on-station time. Payloads along with operational conditions are the deciding

factors for the size estimation of the aerostat envelope [23].

1.2.1 Subsystems of Tethered Aerostat

A tethered aerostat is a multidisciplinary system consisting of various subsystems. The

key component is a balloon filled with lighter-than-air gas that enables the system to take

flight and remain aloft. These balloons are usually referred to as envelopes. A typical

schematic of the present tethered aerostat system is shown in Figure 1.1.

Ballonet Fin
Envelope Hull
Helium
Pressurization System Air

Confluence Lines
Payloads
Confluence Point

Tether

Mooring Structure

Figure 1.1: Schematic of a typical aerostat system


6

Apart from the envelope, the other components of a tethered aerostat system may in-

clude a winch-mooring platform, tether, pressurization system, payloads and other ground

systems as briefly described below:

Aerostat Envelope

The aerostat envelope is a pressurized fabric enclosure containing the lifting gas. Air or

gas filled fins are attached at the rear end of the hull and provide stability to the aerostat.

Air filled ballonet provides the space for gas expansion or contraction due to variation in

altitude or temperature. A series of ropes called confluence lines connect the hull to a

single point called confluence point, to which the main tether is attached.

Winch-mooring Structure

When the aerostat is brought down periodically for gas topping up, it is moored on the

mooring structure. The aerostat is also moored during initial launch and for maintenance

purpose. This structure arrests the aerostat envelope and allows the alignment with the

wind direction. When the aerostat is launched, this platform also acts as a winch and

releases the tether in a controlled manner.

Payloads

The payloads may include a variety of surveillance and tactical equipment such as high-

resolution video cameras, electro-optical/ infrared (EO/IR) sensors, communication re-

peaters, acoustic detectors, and radar. These payloads are attached on the surface of

envelope through attachment points or patches. The payloads receive power through

tether. The signals are also communicated through tether.

Pressurization System

The pressurization system consists of blowers and deflation valves. This is attached

through the air compartment of envelope and is important for maintaining the shape

of pressurized envelope. Depending on the temperature and height variation, air is either

sent inside the air compartment through blower or removed using the deflation valve from

the air compartment to maintain constant differential pressure inside the envelope.
7

Tether

The tether holds the aerostat envelope at the confluence point and the lower end of the

tether is connected to the mooring structure. Besides providing strength, the tether also

provides power to the airborne components through electrical cable and allows signal

communication through fiber optic option.

Other Ground Systems

Other ground systems include gas trailers, cranes, fork lifters, power supply system, ground

station, etc. These systems are ground based and provide assistance for aerostat operation,

maintenance and repair.

1.2.2 Aerostat Envelope

Aerostat envelopes are typically made from light weight and durable fabrics having low

gas permeability such as polyurethane coated nylon or polyester. To provide long life and

protection from ultraviolet radiation, some envelope manufacturers offer laminated fabric.

The envelope may also contain a structural feature such as a frame or a keel. For models

without a structural feature, such as the present aerostat envelope, the envelope relies

solely on the lift gas to maintain its shape. The aerostat envelope is the aerodynamically

shaped fabric enclosure filled with lighter than air gas and provides lifting capability due

to buoyancy effect. The important components of an aerostat envelope are presented in

Figure 1.1 and are briefly summarized below:

Hull

Hull is the main helium containing body. The differential pressure inside the hull is selected

slightly more than the dynamic pressure corresponding to the maximum operating wind

speed so that no dimples are formed during aerostat operation. This excess differential

pressure provides rigidity to the envelope due to membrane action and thus the envelope

can be assumed to be a rigid body for carrying out the static and dynamic analysis. For

the case of present test aerostat envelope, the hull volume is about 2023 m3 .
8

Fins

The fins may be air or helium filled depending upon the requirement. Fins provide sta-

bility to the envelope and help align the aerostat envelope opposite to the wind direction

to minimize drag and hence horizontal sway or blow by. If the fins are air filled then

they are interconnected with ballonet compartment and if fins are helium filled they are

interconnected with the hull compartment. For the case of present test aerostat envelope,

the fins are helium filled having a volume of about 84 m3 .

Ballonet

Ballonet is a separate air filled fabric compartment within the hull volume. This air is

pressurized either by means of a blower fan or by increasing temperature or height and

this pressure, in turn, is transmitted to the inner hull to establish its rigidity. As the gas

expands by increasing altitude or temperature, the ballonet is correspondingly compressed

and its air is allowed to go to outside atmosphere through deflation valve. Therefore, no

gas is lost and the hull’s internal pressure stays approximately constant. For the case of

present test aerostat envelope, the ballonet volume is about 17% of the hull volume. For

higher size aerostat for operation at higher altitudes, this percentage goes up.

Lines or cordages

Different lines or cordages on the aerostat envelope include confluence lines, mooring lines,

handling lines, fin lines and nose lines. The confluence lines are the load bearing lines

which hold the aerostat envelope firmly to the confluence point. Due to lack of rigidity

and strength, the aerostat envelope cannot be held at one or two points and multiple

symmetric lines are required for this purpose. The upper portions of these lines are spread

over the surface of the envelope and lower portions converge to a single point, known as

confluence point. The number of lines at the confluence point can be reduced by converging

lines in between, thus efficiently spreading the load and also making termination efficient.

Mooring lines are used when the aerostat is in mooring condition on the ground. Handling

lines are provided to handle the aerostat envelope during testing or repair. Fin lines are

used to provide rigidity to the fins and nose lines are provided to support latching of

envelope to the mooring platform.


9

Patches

Patches on the surface of the aerostat envelope are provided to hold it properly. These are

also the hard points on the surface of the aerostat envelope. But unlike the metal hard

points, specified loads can only be applied in a particular direction, generally tangent to

the surface as applying loads in other directions greatly reduces the load bearing capability.

Patches are provided for holding confluence lines, mooring lines, handling lines, fin lines,

nose lines and payloads.

1.2.3 Advantages and Applications of Tethered Aerostat

Tethered aerostat system offers many advantages as compared to its counterparts, some

of which include long endurance, low operation and maintenance cost, ease in operation,

ease in directional tracking, silent operation and no runway requirement [48]. Tethered

aerostats can be used in a variety of military and civil applications. Military applications

include intelligence, surveillance, and reconnaissance (ISR); providing high-resolution im-

agery to ground installations, and providing communications and data-relay to wide areas

over any terrain. Civil application include natural disaster relief, search and rescue, tacti-

cal communications relay, harbor and coastal monitoring, critical infrastructure security,

personnel location and tracking, traffic monitoring, crowd management and event surveil-

lance [45].

1.3 Fundamentals of Tethered Aerostat

Mounting aerostat to a certain height increases its line of sight coverage and this line

of sight should be known for that height to assess the limiting capacity of the aerostat.

Unlike the conventional aircrafts, the basic source of lift on any lighter than air vehicle

is buoyancy which is governed by Archimedes’ principle. Atmospheric properties for off-

standard atmosphere as well as helium gas properties are required to be evaluated to

estimate this lifting force. This section covers some fundamental aspects or theory of

operation and some important relevant terms of tethered aerostat.


10

1.3.1 Line of Sight Coverage

The coverage area of the aerostat is determined by calculating the radial distance to the

horizon from the aerostat launch point. Figure 1.2 shows an aerostat at a height of ha

from the launch point and Earth is represented approximately as a sphere of radius RE .

A tangent from the aerostat to the Earth’s surface will be the line of sight for any payload

Figure 1.2: Geometry for radial line of sight coverage [23]

mounted on it. The radial distance between the launch point and the point of tangent ls

is known as the radial line of sight or coverage radius. It is noted that this is a simplified

model and location and terrain has not been taken into account. Using simple geometric

relation in Figure 1.2, the expression for ls is written as below [9]:

 
−1 RE
ls = RE cos (1.1)
RE + ha

The radial line of sight or coverage radius using Equation (1.1) for a range of altitudes

from 0 to 25 km is plotted in Figure 1.3. It is observed that the slope of coverage radius

curve is continuously decreasing with increasing the altitude. For the present test aerostat,

the height of operation is 1 km and the corresponding coverage radius is about 113 km.

Presently, the aerostats operate maximum up to about 5 km altitude worldwide and the

corresponding coverage radius is about 250 km. The altitude of operation of stratospheric

airship may be about 20 km from ground and may cover a radius of about 510 km.
11

600

500
coverage Radius (km)

400

300

200

100

0
0 5 10 15 20 25
Altitude (km)

Figure 1.3: Variation of coverage radius with aerostat altitude

1.3.2 Principle of Aerostatics

The basic principle of aerostatics is Archimedes’ law which states that the upward buoyant

force that is exerted on a body immersed in a fluid, whether fully or partially submerged,

is equal to the weight of the fluid that the body displaces and acts in the upward direction

at the center of mass of the displaced fluid. The aerostat envelope comes under fully

submerged body as it is completely surrounded by the atmospheric air. The total or

upward buoyancy force may be written as:

B = Vg ρa g (1.2)

Here Vg is the volume of air displaced or helium volume, ρa is the atmospheric air density

and g is the acceleration due to gravity. If helium gas weight is taken into consideration,

then the gross lift Bg may be written as:

Bg = Vg (ρa − ρg ) g (1.3)
12

The buoyancy force acts on the center of buoyancy which is same as the center of mass

of the displaced fluid. Figure 1.4 presents the basic mechanism of buoyant lift generation

on an aerostat envelope. If the envelope internal pressure at the bottom point is assumed

zero i.e. internal and external pressure same, the air pressure as well as helium pressure

will both decrease with height. But this decrease will be more for air as air density is more

as compared to helium density, giving rise to a net upward force [21].

Figure 1.4: The mechanism of aerostatic lift [21]

1.3.2.1 The Atmospheric Air Properties

As mentioned earlier in this section, the air and helium density along with helium volume

are required for buoyancy calculation. It was also pointed out that presently, the aerostats

operate maximum up to about 5 km height which is within the troposphere limit. The air

and helium density for the real condition have been presented in this section [21].

The International Standard Atmosphere (ISA) sea level conditions are as below:

Temperature (Th0 ) : 288.150 K

Pressure (ph0 ) : 101325 N/m2

Density of air (ρh0 ) : 1.225 kg/m3

For ISA and within the tropospheric region, the temperature, density and pressure at
13

a pressure height hp (m) are given by the following relations:

Th = 288.15 − 0.0065hp (1.4)

 4.3
ρh Th
= (1.5)
ρh0 Th0
 5.3
ph Th
= (1.6)
ph0 Th0

Both pressure and density at any altitude in the ISA can be seen to be dependent on, and

defined from, the temperature alone.

1.3.2.2 Off-standard Atmospheric Air Properties

The real atmosphere will be different from ISA most of the time and this off-standard

atmosphere is defined by a temperature difference ∆T from ISA. The pressure is same for

ISA and off-standard atmosphere and the density is defined be the following equation [21]:

 
Th
ρa = ρh (1.7)
Th + ∆T

Using Equations (1.4), (1.5) and (1.7) we can estimate the air density at the height for

any specified real temperature conditions. In Equation (1.3), the unknown remains the

helium gas density only.

The helium gas density is estimated in the same way as done for the air. However,

one additional factor i.e. helium gas purity is also required to be considered as the helium

purity level will continuously degrade with time right after the initial filling. The following

relation is used to calculate the density of impure helium at ISA sea level [21]:

ρg0 = 0.169k + 1.225 (1 − k) (1.8)

where the density of air at ISA sea level is 1.225 kg/m3 , the density of 100% pure helium

at ISA sea level is 0.169 kg/m3 and k is the helium purity. To calculate the helium density

at the Off-standard operating conditions, it is multiplied by the same density factor as

that for air as below [21]:


ρa
ρg = ρg0 (1.9)
ρh0
14

1.3.2.3 Other factors affecting lift

Apart from the parameters mentioned earlier, other factors which affects buoyancy include

internal pressure, superheat - an effect where the temperature inside the envelope is higher

than the ambient air temperature, rain, snow, icing and humidity. The effect of internal

pressure on buoyancy is generally neglected. Due to superheat, there is an increase in

buoyancy because helium density inside the envelope is decreased. Rain, snow and icing

will cause to increase the envelope weight and reduce the buoyancy and hence these factors

have to be accounted beforehand. Humidity decreases the air density and hence reduces

buoyancy [21].

1.3.3 Shape and Volume Estimation of Aerostat Envelope

Aerostat envelopes come in different shapes, sizes, and designs. The shape of the envelope

is designed to provide lift while keeping the wind resistance as low as possible. The

envelopes for certain aerostat should be shaped aerodynamically. These envelopes should

be shaped such as to have minimum drag and required lifting capability. The shape

optimization requires a comparative study of shapes in terms of the required parameters

such as buoyancy capability, surface area, lift, drag, stability, stresses, blow by and ease

of fabrication. The use of advanced computational tools such as Computational Fluid

Dynamics (CFD) and Finite Element Method (FEM) may be required for this purpose

[51]. Based on the requirements outlined, a particular shape is selected. Helium/air filled

fins are added to the aerostat envelope for stability.

The basis of volume estimation of an aerostat envelope is Archimedes’ principle. The

starting point for the volume estimation is given payload capacity and height of operation.

The sequence of procedure followed for volume estimation of aerostat envelope is presented

in Figure 1.5.

For estimating buoyant lift of aerostat envelope, air density, helium density, and he-

lium volume is required. The air density and helium density may be obtained from Equa-

tions (1.7) and (1.9) respectively. The gas volume is obtained by subtracting ballonet air

volume from the total envelope volume. Now using Equation (1.3), payload capacity of

the aerostat may be written in terms of mass as below:

PL = Vg (ρa − ρg ) (1 − FL ) − ms − Kt mc hp (1.10)
15

START

Assume Volume, V

Calculate air density, helium density


and ballonet volume, hence gross lift

Estimate Aerostat Weight including


hull, fins, ballonet, patches, cordages,
accessories and payload

Estimate payload

Is payload No
same as
required

Yes

END

Figure 1.5: Flow chart for volume estimation of aerostat envelope

Here FL is the free lift expressed as percentage of gross lift. The free lift is required

to always keep some amount of tension in the tether. The value of free lift is usually

12-15% of the gross lift expressed in standard conditions. Kt is the tether length factor,

generally about 1.1, which is to cater to maintain height even after some sway of aerostat

in horizontal direction. mc is the tether cable mass per unit length. The additional factors

which may be required to be considered while applying the above equation are included

in section 1.3.2.3. The weight of aerostat envelope includes hull fabric, fin fabric, ballonet

fabric, joints, adhesive, patches, cordages and accessories. Equation (1.10) is a nonlinear

equation in terms of Vg , because for structural mass is a function of Vg , and can be solved

using Newton iteration method [8]. Thus we get the required volume of the aerostat

envelope.
16

1.4 Objectives of Present Research

The objective of the present research is to investigate the static and dynamic analysis of an

airborne tethered aerostat, which has been carried out during its design and development

phase. At the same time the objective is also to estimate all the input parameters required

for carrying out the static and dynamic analysis of the aerostat. During the present

research work, following are the targeted tasks to meet the research objectives:

1. To use the modern tool of solid modeling of the aerostat envelope to generate geo-

metric, mass and inertia properties required for static and dynamic analysis.

2. To examine the lift, drag and moment coefficients of the aerostat envelope using

different approaches namely semi-empirical method, computational fluid dynamics

and wind tunnel testing. A comparative study is also required to be carried out for

the results obtained using these methods.

3. To carry out the equilibrium studies of the aerostat to estimate tether tension, trim

angle of attack and equilibrium cable shapes under uniform horizontal wind speeds.

The criterion for static stability is also required to be proposed and sensitivity studies

to be carried out for possible effect on trim angle of attack and static stability.

Approximate relations are to be proposed for estimating cable shape to account for

air density and uniform wind speed variation along the vertical direction.

4. To propose dynamic stability analysis approach for tethered aerostat which can also

be applied for the situation where the density and uniform wind speed varies along

the vertical direction using the approximations for equilibrium cable shape. The

initial disturbance response is also required to be examined.

5. To propose nonlinear dynamic analysis model for tethered aerostat using the lumped

mass approach and more practical winds are required to be applied along with wind

gusts to generate aerostat response.

6. The available measured aerostat responses are required to be compared with the

computed response for same input wind conditions. The available measured angles

of attack are also required to be compared with the results of equilibrium studies.
17

1.5 Aerostats Considered for Analysis

All the results of static and dynamic analysis and estimation of input parameters have

been presented for a medium size experimental test aerostat referred as test aerostat from

now onwards in the present thesis. This aerostat is capable of taking a payload of 300 kg

up to a height of 1000m AMSL. The hull volume for this aerostat is about 2023 m3 with

maximum hull diameter of 11.1 m and envelope length of 33.85 m. The characteristics of

this test aerostat are listed in Tables 3.2, 4.1, 4.2 and 5.1.

However, for estimating and comparing tether tension and equilibrium cable configu-

ration in chapter 3, a small aerostat has also been considered whose characteristics are

presented in Tables 3.1 and 3.4. This small aerostat is capable of taking a payload of 50

kg to a height of 100m AMSL. The hull volume for this aerostat is about 250 m3 with

maximum hull diameter of 5.55 m and envelope length of 16.93 m. This small aerostat is

primarily considered to validate the proposed tether approximation models with standard

results because, due to small height of operation, the density as well as steady wind speed

variation along the vertical direction can be neglected.

To validate the dynamic stability model with an available literature, a 7.64 m long

reference aerostat envelope having a volume of 19 m3 [42] has also been considered in

section 4.8 of chapter 4 only. The properties of this reference aerostat is presented in [42].

Hence the following three aerostat configurations have been considered for the analysis in

the present thesis:

• Test aerostat : For generating results everywhere. Characteristics are presented in

Tables 3.2, 4.1, 4.2 and 5.1.

• Small aerostat : For generating and comparing results of tether tension and cable

configuration only in chapter 3. Characteristics are presented in Tables 3.1 and 3.4.

• Reference aerostat : For validating the dynamic stability model with the available

literature only in section 4.8 of chapter 4. The properties of this reference aerostat

is presented in Table I of reference [42].

Also, ISA+15 deg temperature condition, which approximately represents Indian Refer-

ence Atmosphere, have been considered throughout the analysis unless otherwise stated.
18

1.6 Layout of the Thesis

Chapter 1 describes the motivation for carrying out the present research work. The basic

introduction of a tethered aerostat system including some of the important concepts is

also briefly presented. This chapter also introduces the aerostats for which the results of

static and dynamic analysis are to be presented. The objectives of the present research

work have also been discussed in this chapter.

Chapter 2 presents the methods followed for generating all the required inputs for car-

rying out the static and dynamic analysis of the test aerostat. Solid modeling approach

has been described for generating geometric, mass and inertia properties. Apparent mass

coefficient terms are also estimated. Aerodynamic coefficients have been estimated us-

ing semi-empirical method, computational fluid dynamics and wind tunnel testing. The

approximation techniques for estimating dynamic derivatives are also described.

Chapter 3 presents the equilibrium and static analysis approach for the tethered aero-

stat. The tether tension, trim angle of attack, static stability criteria and equilibrium cable

shapes are evaluated. Approximation techniques for evaluating tether cable configuration

are also proposed.

Chapter 4 describes the dynamic stability analysis approach for the tethered aerostat.

The equations of motion for the aerostat are linearized so that the equations become

decoupled in longitudinal and lateral directions. The stability analysis has been carried

out to estimate characteristic roots or eigenvalues. The response of the test aerostat due

to initial disturbances are also generated.

Chapter 5 presents the nonlinear dynamic analysis of the tethered aerostat where the

equations of motion are formulated and solved without making them linear. Six different

types of test winds including gusts are applied to the tethered aerostat to estimate its

response. The comparison of this nonlinear model with available experimental data is

also carried out. A comparison of equilibrium analysis result with experimental result is

also done. To study the correlation among the static and dynamic model, the equilibrium

analysis approach has been compared with nonlinear dynamic modeling approach for a

constant horizontal wind speed.

Chapter 6 presents the concluding remarks including the innovative research work

carried out as well as some of the potential research areas for future work.
Chapter 2

Estimation of Input Parameters


for Static and Dynamic Analysis

2.1 Introduction

To carry out static and dynamic analysis of a tethered aerostat, a set of input parameters

are required to be estimated which uniquely describe the aerostat. These input parameters

can broadly be classified as below:

• Aerostatic properties

• Masses and moments of inertia (including apparent mass)

• Aerodynamic coefficients and derivatives

• Tether cable properties

This chapter presents the analytical and experimental methods followed for the estima-

tion of these input parameters for the test aerostat. This test aerostat has been briefly

described in section 1.5 of chapter 1. This is a medium size aerostat with a payload ca-

pacity of 300 kg having operating height of 1 km. Aerostatic properties are estimated

by using analytical method combined with solid modeling approach. Mass and inertia

properties are estimated using a combination of solid modeling, analytical approach and

direct measurement. Aerodynamic coefficients are estimated by semi-empirical method,

computational fluid dynamics, wind tunnel testing and analytical methods. A compar-

ative study has also been carried out for results obtained using these different methods.

Tether cable properties are obtained using direct measurement, experiment and literature.

19
20

2.2 Aerostatic Properties

The aerostatic properties for the test aerostat envelope are listed as below:

• Location of structural center of mass

• Location of center of buoyancy

• Buoyancy force

• Mass of inflation gas

• Envelope volume

• Structural weight of envelope

2.2.1 Analytical Method

The analytical approach to estimate the geometric, mass and inertia properties is based on

considering infinitesimal element on the aerostat envelope and by applying basic definition

of the properties, these are estimated for this infinitesimal element and then integrating

them on the whole body give the properties for the entire body. Due to complexity in

geometry, analytical method is only applied to hull. The results obtained for the hull is

compared with that of solid modeling results for validation of solid modeling approach. The

envelope hull of the test aerostat envelope is composed of thee basic shapes of revolution,

namely ellipse in the nose portion, circle in the middle section and parabola in the tail

portion as presented in [20]. This geometry of envelope hull is presented in Figure 2.1,

where Dh is the maximum diameter and c̄ is the hull or envelope length. An element on

Ellipse Circle Parabola

Figure 2.1: Hull geometry of aerostat envelope


21

this hull has been taken at a distance x from the origin, which is at the central axis of

maximum diameter section. The radius of the element is y and the thickness dx along the

axial direction. Depending upon the requirement to estimate surface or volume properties,

surface or volume element is considered. The equations of different portions of hull shape

are presented as below (Figure 2.1):

Nose portion: Ellipse from x = -1.25Dh to 0

x2 y2
+ =1 (2.1)
(1.25Dh )2 (0.5Dh )2

Middle portion: Circle from x = 0 to 1.625Dh

x2 + (y + 3.5Dh )2 = 16Dh 2 (2.2)

Tail portion: Parabola from x = 1.625Dh to 1.8Dh

y 2 = 0.1378Dh (1.8Dh − x) (2.3)

Having the hull geometry defined, analytical method can be applied to evaluate geometric,

mass and inertia properties of the aerostat envelope. Equations (2.1-2.3) may be rewritten

as below: s
x2
y1 = 0.5Dh 1− (2.4)
(1.25Dh )2
q
y2 = −3.5Dh + 16Dh 2 − x2 (2.5)

p
y3 = 0.1378Dh (1.8Dh − x) (2.6)

Here y1 , y2 and y3 are distances from the envelope center-line to the envelope surface for

the three hull curves as function of x.

Volume of Envelope Hull

Volume of the envelope hull is required to estimate the total mass of gas inside hull. If

all the three fins are air filled then volume of hull is sufficient to estimate total gas mass

as the gas density is known as a function of temperature. However in the present case of

test aerostat envelope, all the fins are helium filled and hence fins volume is also required
22

to be added to hull volume for calculating gas mass inside aerostat envelope. Volume of

the small volume element at any point can be written as below:

dVh = πy 2 dx (2.7)

Hence the total hull volume may be obtained by integrating Equation (2.7) along the

entire length of the envelope.


Z 1.8Dh
Vh = πy 2 dx (2.8)
−1.25Dh

Expanding the integral over the total envelope length, we have

Z 0 Z 1.625Dh Z 1.8Dh 
2 2 2
Vh = π y1 dx + y2 dx + y3 dx (2.9)
−1.25Dh 0 1.625Dh

The above integral can be evaluated numerically using the method described in [8], thereby

estimating the hull volume. The hull volume thus we get is 1.4792Dh 3 .

Center of buoyancy of envelope hull

Center of buoyancy of envelope hull is the point at which net buoyant force acts if only

hull is filled with helium gas. Alternatively this is the point of volume center of aerostat

envelope hull. The differential moment of volume element about the envelope hull origin

can be written as (Figure 2.1):

dMvh = πy 2 xdx (2.10)

Using Equations (2.4-2.6), Equation (2.10) may be integrated to get total moment as:

Z 0 Z 1.625Dh Z 1.8Dh 
2 2 2
Mvh = π y1 xdx + y2 xdx + y3 xdx (2.11)
−1.25Dh 0 1.625Dh

If xbh is the center of buoyancy distance from the origin, the moment expression in Equa-

tion (2.11) may also be written as:

Mvh = Vh xbh (2.12)

Using Equations (2.9) and (2.11), xbh can be evaluated from Equation (2.12) as 0.1338Dh .

Hence the center of buoyancy distance from envelope nose is 1.3838Dh .


23

Surface area of envelope hull

The surface area of envelope hull is used to estimate envelope hull weight. The envelope

hull weight is simply obtained by multiplying the surface area of hull by the per unit area

weight of the fabric. To find the surface area, we consider the differential area element on

the envelope hull surface (Figure 2.1). The length of small element along the hull surface

may be written as:


q
ds = (dx)2 + (dy)2 (2.13)

Using Equation (2.13), the surface area of differential surface element may be written as:

s 
 2
dy
dAh = 2πyds = 2πy  1+  dx (2.14)
dx

Hence the hull surface area may me written as below:


s 
1.8Dh  2
dy
Z
Ah = 2π y 1+  dx (2.15)
−1.25Dh dx

Equation (2.15) has been integrated numerically to estimate hull surface area as 7.4481Dh 2

Center of surface of envelope hull

The center of surface of envelope hull is important in the center of gravity estimation of the

aerostat envelope. Since the aerostat envelope hull is a thin surface geometry, the center

of surface may be assumed same as center of gravity. The center of surface of envelope

hull may be estimated following the similar procedure as for the estimation of center of

buoyancy. The differential moment of area element about the origin is written as below

(Figure 2.1):
ds
dMah = 2πyxds = 2πyx dx (2.16)
dx

Using Equations (2.4-2.6), Equation (2.16) may be integrated to get total moment as:

s 
1.8Dh  2
dy
Z
Mah = 2π xy  1 +  dx (2.17)
−1.25Dh dx
24

If xah is the center of surface distance from the origin, the moment expression in Equa-

tion (2.17) may also be written as:

Mah = Ah xah (2.18)

Using Equations (2.15) and (2.17), xah can be evaluated from Equation (2.18) as 0.1664Dh .

Hence the distance of center of surface from envelope nose comes out to be 1.4164Dh .

Mass moment of inertia of envelope hull

The mass moment of inertia of envelope hull is required for studying the dynamics of

aerostat envelope. Due to simplicity, only rotational moment of inertia of envelope hull

about the central axis is estimated here. Other moments of inertia will be estimated

using the solid modeling approach. Using Equation (2.14), differential mass of the surface

element with the surface density mha may be written as:


s 
 2
dy
dmh = mha (2πyds) = 2πmha y  1 +  dx (2.19)
dx

Since the surface element is at a distance y from the center-line, the mass moment of

inertia for this element is written as


s 
 2
dy
dIxxh = y 2 dmh = 2πmha y 3  1+  dx (2.20)
dx

The total moment of inertia is obtained by integrating Equation (2.20) and using Equa-

tions (2.4) to (2.6).

s 
1.8Dh  2
dy
Z
Ixxh = 2πmha y3  1+  dx (2.21)
−1.25Dh dx

Mass moment of inertia of hull helium gas

The mass moment of inertia of hull helium gas is required for studying the dynamics

of aerostat envelope as presented in chapter 4 and chapter 5. Due to simplicity, only

rotational moment of inertia of hull helium gas about the central axis is estimated here.

Other moments of inertia will be estimated using the solid modeling approach. Using
25

Equation (2.7), differential mass of the volume element with density ρg may be written

as:

dmg = ρg πy 2 dx (2.22)

Since the volume element is a disc having radius y, the differential mass moment of inertia

for this element is written as:

y2 πρg 4
dIxxg = dmg = y dx (2.23)
2 2

The total moment of inertia is obtained by integrating Equation (2.23) and using Equa-

tions (2.4) to (2.6).


1.8Dh
πρg
Z
Ixxg = y 4 dx (2.24)
2 −1.25Dh

It is again emphasized that the analytical method has been applied to estimate certain

required parameters for envelope hull only. The values of parameters thus obtained from

analytical method serves useful purpose for the fact that these can be compared with the

values obtained from solid modeling approach explained in section 2.2.2, thus validating

solid modeling approach. Also as the hull is the major component of aerostat envelope,

we immediately have the values of these parameters of a major component and the con-

tributions of other components can be added in it to get the values for entire aerostat

envelope.

2.2.2 Solid modeling of Aerostat Envelope

As evident in the analytical approach for estimating aerostatic properties that only simple

geometry can be handled in the method. Aerostat envelope hull is an example of such a

simple geometry since the equations of curve of revolution are well defined. Hence enve-

lope hull modeling was carried out in section 2.2.1 using analytical approach and many

aerostatic properties were evaluated. But for the complete aerostat envelope, analytical

approach becomes quite complex and hence is not suitable. Solid modeling is a mod-

ern tool which can efficiently handle complex geometries and help provide the required

aerostatic properties. CATIA R (Computer Aided Three Dimensional Interactive Appli-

cation) is a multi-platform software suite for Computer Aided Design (CAD), Computer

Aided Manufacturing (CAM), Computer Aided Engineering (CAE), Product Lifecycle


26

Management (PLM) and 3D [50]. It is an excellent 3D modeling software extensively used

in the industry. The present test aerostat envelope is modeled using CATIA R V5. It

was mentioned in section 1.2.2 that the aerostat envelope consists of hull, fins, ballonet,

cordages, patches and payloads. Although the complete modeling containing each fine

details of these components will be required for fabricating the aerostat envelope, approx-

imate modeling will be sufficient for our purpose to evaluate aerostatic properties. The

approach for carrying out the modeling is summarized as below:

2.2.2.1 Hull Modeling

Aerostat envelope hull is a thin surface of revolution with geometry defined in Equa-

tions (2.1) to (2.3). Generative shape design feature of CATIA R software is used to

model the envelope hull. The coordinates of envelope hull may be generated in an Excel

sheet and the same is imported in CATIA R using the macros feature to create a curve.

It should be noted in this method that we have sufficient number of data points near the

nose and tail of envelope hull to facilitate uniform curvature. The hull curves may also be

directly drawn in the CATIA R where it automatically caters for number of data points.

The later method is followed for generating the hull surface curve in the present modeling

approach. The line curve is now revolved about the central axis to get the complete hull

surface. The maximum diameter Dh of the present test aerostat envelope is 11.1 m and

the length from nose to tail c̄ is 33.85 m. The hull surface we now get is shown in Fig-

ure 2.2(a). The small holes visible on the rear side of hull are provided to fill the helium in

the fins. Consistent coordinate system has been followed for modeling of different envelope

components. If we provide the mass of unit surface area in the model then we get all the

structural mass and inertia properties for envelope hull.

2.2.2.2 Fin Modeling

The envelope fin is also modeled as thin surface similar to hull. Fin coordinates at the tip

and center-line are generated in an Excel file and the same is imported in CATIA using

macros feature. Since the fin is symmetric about the mid plane, only the coordinates of

half curve are required. A curve is generated using these coordinates at the tip and center-

line. These two curves are joined by a surface using multi-section surface feature. Now

using symmetry feature, second half fin surface is created as mirror image of the first half.
27

(a) Hull modeling (b) Fin modeling

(c) Ballonet modeling (d) Complete model of envelope

Figure 2.2: The complete solid model of aerostat envelope

Fin tip is then closed by rotating the half curve at the fin tip. There are three fins and

hence copy of this fin is rotated two times at the desired positions. Now the fins portion

which is inside the hull is removed for all three fins thereby completing the modeling as

presented in Figure 2.2(b). The lines visible in this figure indicate the partition of panels

and the opposite partitions are tightly held together to maintain the fin in desired shape.

2.2.2.3 Ballonet Modeling

The ballonet is the air filled compartment inside the envelope hull. This is also a surface of

revolution but part of it intersects with hull to create joints. Ballonet surface is generated

in a similar manner as hull. The part which is intersecting with the hull is removed using

the trim option. The ballonet modeling is presented in Figure 2.2(c). The solid modeling of

complete aerostat including hull, fins and ballonet is presented in Figure 2.2(d). Although

the ballonet modeling is presented here for the case where the air is completely filled in the

ballonet, it will be presented later on that for carrying out static and dynamic analysis,

fully deflated ballonet has been considered. In such a case, the total weight of the ballonet

is assumed to be distributed uniformly over the common area of hull and ballonet.
28

2.2.2.4 Cordages, Patches, Accessories and Payloads modeling

The cordages are modeled as lines. The patches are modeled as point masses as the sizes

of patches are small. Accessories and payloads are also modeled as point masses for the

same reason.

2.2.2.5 Comparison of Hull Aerostatic Properties

To validate the solid modeling approach, results obtained for hull surface using the solid

modeling approach is compared with that using analytical approach followed in section

2.2.1 of this chapter for the maximum hull diameter Dh = 11.1 m. The comparison has

been presented in Table 2.1.

Table 2.1: Hull parameter comparison for analytical and solid modeling approach
Hull parameter Analytical values Solid modeling values
Volume 2023.049 m3 2023.138 m3
center of buoyancy 15.360 m 15.357 m
Surface area 917.680 m 2 917.795 m2
center of surface 15.722 m 15.742 m
Rotational MI 6976.637 kg-m2 6976.632 kg-m2
Helium gas MI 4018.823 kg-m2 4018.857 kg-m2

The parameters for the comparison are hull volume, center of buoyancy, surface area,

center of surface, hull rotational mass moment of inertia and helium gas mass moment of

inertia. It comes out that there is an excellent agreement between the results of analytical

approach and solid modeling approach and thereby solid modeling approach is validated.

2.2.3 Total Mass of Aerostat Envelope

The total mass of aerostat envelope consists of the different components and these com-

ponents as well as the method followed to estimate them are summarized as below:

Structural mass of envelope

The fabric mass of test aerostat envelope is obtained simply by weighing the envelope in

packed condition. It is to be noted that mass of pack cover, if any has to be subtracted

from the weighed mass. The fabric mass thus obtained will consist of hull, fins, ballonet,

patches and cordages. If the accessories mass and payload mass are added in the fabric

mass, we obtain structural mass of the aerostat envelope ms .


29

Helium gas mass

From the solid modeling of aerostat envelope as presented in Figure 2.2(d), the total

envelope volume is estimated. This volume is the summation of hull and fins volume

since at the height of operation, the ballonet volume is small. Hence all the static and

dynamic analysis have been carried out for the case when the ballonet is empty. The

density of helium at ISA sea level is 0.169 kg/m3 [21] and the helium density at the

analysis condition, which is ISA+15 deg with a height of 1 km, has been estimated using

Equations (1.8) and (1.9) as presented in chapter 1. We now have helium volume Vg and

density ρg and hence the mass of helium gas may be estimated as mg = ρg Vg .

Apparent air mass

When the aerostat accelerates in a particular direction, a certain amount of air also moves

with it which is known as apparent mass. For the present test aerostat envelope, apparent

mass in axial and lateral directions will be different. The apparent mass will be discussed

in detail in section 2.3.

We now have structural mass, helium mass and apparent air mass components. These

masses are added to get the total mass of aerostat envelope.

2.2.4 Total Moments of Inertia of Aerostat Envelope

The mass moments of inertia similar to envelope mass have contributions from structure,

helium gas and apparent terms. These inertia components with the method followed to

estimate them are summarized as below:

Moments of Inertia of Structure

The structural moment of inertia will consist of components from hull, ballonet, fins,

cordages, patches, accessories and payloads. The moments of inertia of hull, ballonet, fins

and cordages were directly obtained from solid modeling. Since the static and dynamic

analysis have been carried out for empty ballonet condition, the ballonet mass have been

assumed distributed at the hull ballonet intersection surface. All the mass of joints were

distributed uniformly over that surface. The mass moment of inertia we thus obtain will be

about center of mass of solid model. These mass moments of inertia have been transferred

to the combined center of mass of structure and helium gas using the relation as described
30

in reference [32]. The patches, accessories and payloads were considered as point masses

and moments of inertia were obtained accordingly and transferred to the combined center

of mass of structure and helium gas using the relation as described in reference [32]. These

moments of inertia were then added to get the total structural mass moments of inertia.

Moments of Inertia of Helium Gas

Helium gas mass moments of inertia were obtained using the solid model in which the gas

of known density was filled in hull and fins with empty ballonet. As described earlier, the

moments of inertia thus obtained will be about center of mass of helium gas i.e. center of

buoyancy and will have to be transferred to the combined center of mass of structure and

helium gas using the relation as described in reference [32].

Moments of Inertia of Apparent Air

As described earlier, when the aerostat accelerates in a particular direction, a certain

amount of air also moves with it which is known as apparent mass. This also contributes

to moment of inertia. The apparent air inertia has been estimated in section 2.3.

We now have structural moments of inertia, helium gas moments of inertia and ap-

parent air moments of inertia about combined center of mass of structure and helium

gas. These moments of inertia are added to get the total moments of inertia of aerostat

envelope.

2.2.5 Center of Buoyancy of Aerostat Envelope

The envelope hull center of buoyancy was estimated in section 2.2.2.5 using the solid

modeling approach which was also validated with analytical approach. The center of

buoyancy is basically the point on which the buoyancy force acts. It is also the center of

volume of the helium gas. The center of buoyancy of aerostat envelope has been estimated

using the solid modeling approach. As the ballonet is empty, the volume of hull and fins

was filled with helium gas whose density is known and the volume center was estimated

using CATIA R . The helium gas mass was obtained in a similar manner in section 2.2.3.

As the aerostat envelope is symmetric about x-z plane the center of buoyancy lies on the

plane of symmetry.
31

2.2.6 Structural Center of Mass of Aerostat Envelope

The structural mass of the aerostat envelope was estimated in section 2.2.3. The various

components included were hull, fins, ballonet, cordages, patches, accessories and payloads.

The center of mass of hull, fins, ballonet and cordages was obtained from the solid modeling

of envelope. The patches, accessories and payloads were treated as point mass whose mass

and location are known. Thus the structural center of mass has been estimated using the

following relation:
P P P
mx my mz
XS = P ; YS = P ; ZS = P (2.25)
m m m

Since the aerostat envelope structure is symmetric about x-z plane, the structural center

of mass lies on the plane of symmetry and hence YS = 0. Since the equations of motion

are written about the combined center of mass of structure and helium gas in chapter 4

and 5, combined center of mass of structure and helium gas XG and ZG is also obtained

using Equation (2.25). All the aerostatic, mass and inertia properties estimated for the

test aerostat envelope have been presented in Tables 3.2, 4.1 and 5.1.

2.3 Apparent air mass

When a body fully immersed in a fluid is accelerating, a certain amount of the surrounding

fluid (air in the present case) moves along with the body. That portion of the fluid set

in motion is referred to as the aerodynamic apparent mass, or simply apparent mass.

Because the mass of the air for the present case of test aerostat is of the same order of

magnitude as the combined masses of the envelope structure and helium, the apparent

mass effect has to be considered [41].

The apparent mass of air associated with an ellipsoid motion parallel to axis of revolu-

tion is different from that for motion at right angles to the axis of revolution [34, 7]. The

apparent air masses for the present test aerostat envelope hull are evaluated assuming it to

be an ellipsoid of revolution. Since the hull geometry is not an exact ellipsoid of revolution

for the test aerostat envelope, it is converted into an equivalent ellipsoid of revolution with

the same area of meridian cross section Sl as the hull and same length c̄ as described in

reference [33]. This leads to an equivalent fineness ratio as below:

πc̄2
FRe = (2.26)
4Sl
32

For ellipsoid of revolution, the additional apparent masses are determined using the pro-

cedure as described in [34], which is a collection of the tables of the factor of apparent

mass, and presented as below:

Motion parallel to axis of revolution:

Additional apparent mass = ρa Vh k1

Motion at right angles to axis of revolution:

Additional apparent mass = ρa Vh k3

Rotation about longest diameter:

Additional apparent moment of inertia = kr . Moment of inertia of displaced fluid

The table for apparent mass and inertia coefficients k1 , k3 , k3 − k1 and kr as a function

of fineness ration is presented in reference [34] and is plotted in Figure 2.3. It is observed

0.8

0.6 k1
k3
k3 − k1
0.4 kr

0.2

0
2 4 6 8 10
Fineness ratio

Figure 2.3: Variation of apparent mass and inertia coefficients with fineness ratio [34]

that for high fineness ratio ellipsoids, apparent mass coefficient for motion parallel to axis

of revolution becomes negligible whereas the apparent mass coefficient for motion at right

angles to axis of revolution becomes close to unity. The additional apparent masses for y

and z direction will be the same with the coefficients k2 = k3 . Using the table presented
33

in [34] and by using the equivalent fineness ratio from Equation (2.26), we have estimated

k1 , k2 = k3 and kr . Linear interpolation is used to estimate values between data points.

Apparent masses and inertia properties for the test aerostat envelope have been presented

in Tables 4.1 and 5.1.

2.4 Aerodynamic Coefficients and Derivatives

The aerodynamic coefficients and derivatives are required to describe the aerodynamic

forces on the aerostat envelope. The representation of the aerodynamic model employed

for this test aerostat envelope for carrying out dynamic analysis has been has been covered

in chapters 4 and 5. To carry out the equilibrium and static analysis in longitudinal

direction, only lift, drag and moment coefficients are required. But for the dynamic

analysis, additional coefficients as well as dynamic derivatives will also be required. The

semi-empirical method [18] provides a quick estimate of the aerodynamic coefficients as it is

semi-empirical in nature and is based on certain geometric properties. The computational

fluid dynamics, also known as CFD is more complex in nature as it involves modeling and

computational efforts. The drag on the aerostat envelope has been estimated by using

computational fluid dynamics approach in [51], but the variation of lift, drag and moment

coefficients as well as flow pattern using CFD is not present in literature. Wind tunnel

testing is a reliable experimental technique used to generate aerodynamic coefficients.

The wind tunnel testing is most difficult as it requires the hardware fabrication as well as

extensive and precise testing in wind tunnel. This section employs these three techniques

to estimate static longitudinal aerodynamic coefficients for the test aerostat envelope.

In all the method a reference area is taken as Vh 2/3 and reference length as envelope

length c̄. Although the results of wind tunnel testing have been used for carrying out

static and dynamic analysis of aerostat envelope, the results of semi empirical method

and computational fluid dynamics have been compared with wind tunnel results. The

comparison may provide a basis for assuming the aerodynamic coefficients in the absence of

practical results. The other coefficients as well as dynamic derivatives have been estimated

using the theoretical approach as described in reference [13].


34

2.4.1 Semi-Empirical Method

The semi empirical method for aerodynamic properties estimation is based on hull-fin

interference factors which are defined in the analytical model and obtained from exper-

imental data as functions of certain geometrical parameters, as proposed by Jones and

DeLaurier [18] and later applied to un-symmetrically finned bodies of revolutions by Gill

et al. [14]. This method applies to the low speed regime, when the flow is attached and

no flow separation has occurred over the aerostat hull. Hence the semi empirical model

appears to give a good aerodynamic representation of finned axisymmetric bodies with

attached flow. Figure 2.4 presents the schematic of this steady-state model for semi-

empirical method, where the flow is at an angle of attack αt and the forces on hull and fins

are also shown. Figure 2.5 defines the fin planform terms for the semi-empirical method.

As presented in [18], the expressions for forces and moments may be written as below:

Figure 2.4: Schematic of steady-state analytical model for semi-empirical method [18]

Figure 2.5: Fin planform definition for semi-empirical method [18]


35

The equation for the normal force is written as

Fn = q0 {(k3 − k1 ) ηk I1 sin (2αt ) cos (αt /2) + (Cdc )h sin αt sin |αt | J1
(2.27)
+ Sf [(Cnα ∗ )f ηf (sin (2αt ) /2) + (Cdc )f sin αt sin |αt |]}

Axial force is written as

Sf ] cos2 αt

Fa = q0 {[(Cdh )0 Sh + Cdf 0
(2.28)
− (k3 − k1 ) ηk I1 sin (2αt ) sin (αt /2) − (Ct )f Sf }

Moment about the nose is written as

Mn = −q0 [(k3 − k1 ) ηk I3 sin (2αt ) cos (αt /2)

+ (Cdc )h J2 sin αt sin |αt | + Sf ηf (lf )1 (Cnα ∗ )f (sin (2αt ) /2) (2.29)

+ Sf (lf )2 (Cdc )f sin αt sin |αt |]

The integrals appearing in these equations are defined as

lh
dAc
Z
I1 = dξ = Ah (2.30)
0 dξ

lh
dAc
Z
I3 = ξ dξ (2.31)
0 dξ
Z lh
J1 = 2rh dξ (2.32)
0
Z lh
J2 = 2rh ξdξ (2.33)
0

These equations are made non-dimensional by the following definitions:

Fa , Fn = (Ca , Cn ) q̄0 S (2.34)

Mn = Cmn q̄0 Sc̄ (2.35)


 
I1 , J1 , Sf , Sh = Iˆ1 , Jˆ1 , Sˆf , Sˆh S (2.36)
 
I3 , J2 = Iˆ3 , Jˆ2 Sc̄ (2.37)
h    i
(lf )1 , (lf )2 = lˆf , lˆf c̄ (2.38)
1 2
36

Also, it was assumed that the αt magnitudes for attached flow are less than 30 deg. Hence

cos αt /2 ≈ 1, and Equations (2.27-2.29) become [18]:

h i
Cn = (k3 − k1 ) ηk Iˆ1 + 0.5 (Cnα ∗ )f ηf Sˆf sin (2αt )
h i (2.39)
+ (Cdc )h Jˆ1 + (Cdc )f Sˆf sin αt sin |αt |

h i
Ca = (Cdh )0 Sˆh + Cdf 0 Sˆf cos2 αt

(2.40)
− (k3 − k1 ) ηk Iˆ1 sin (2αt ) sin (αt /2) − (Ct )f Sˆf
h i
Cmn = − (k3 − k1 ) ηk Iˆ3 + 0.5(ˆlf )1 (Cnα ∗ )f ηf Sˆf sin (2αt )
h i (2.41)
− (Cdc )h Jˆ2 + (Cdc )f (ˆlf )2 Sˆf sin αt sin |αt |

For the application of these equations to the given aerostat envelope configuration, the

unknown parameters in Equations (2.39) to (2.41) are required to be evaluated. The

process followed to evaluate these parameters is similar as in [18] and are explained as

below:

The Integrals (Iˆ1 , Jˆ1 , Iˆ3 , Jˆ2 )

We observe that these integrals depend upon the geometry of aerostat envelope from the

nose to the fin starting point (Figure 2.4). Since the equations of hull are known, these

integrals are evaluated using numerical integration [8].

Apparent Mass Coefficient (k3 − k1 )

Apparent mass terms are covered in section 2.3 of this chapter. The apparent mass

coefficients for motion along and perpendicular to the axis of revolution as a function of

fineness ratio are presented in Figure 2.3 as taken from reference [34]. Equivalent fineness

ratio as evaluated from Equation (2.26) is used for the estimation of (k3 − k1 ) from the

table presented in [34]. A linear interpolation is used to estimate the values between data

points.
37

Fin Lift Curve Slope (Cnα ∗ )f

The fin lift curve slope is estimated using the formula available in [40], which is applicable

for subsonic speed as:

 
∗ 2πAt Sexp
(Cnα )f = r (2.42)
2 2

tan2 Λmax
 Sf
2 + 4 + Atη2βt 1 + βt 2
t
t

The required fin parameters and other assumptions for the present test aerostat to estimate

fin lift curve slope is presented in Table 2.2.

Table 2.2: Aerostat envelope fin parameters and other assumptions for lift curve slope
estimation
Parameter Value
Fin tip chord 3.144 m
Fin root chord 8.768 m
Fin span 20.5 m
Fins reference area, Sf 111.46 m2
Fins area inside the hull, Sfh 38.70 m2
βt 1.0
ηt 0.95

Other terms in Equation (2.42) are evaluated using the same process as available in

reference [40]. The exposed fin area Sexp has been obtained by subtracting the fin area

lying inside hull Sfh from the total fin area Sf . It should be noted that in above formula

lift curve slope is for a fin which has zero dihedral angle. If the fin has a substantial

dihedral, then the lift curve slope will change on two accounts. The projected area of the

fin is reduced and the angle of attack does not remain same. However, if the projected

area is taken into account and considering small angle of attack, Equation (2.42) can be

used with reasonable accuracy [14].

Hull Cross-flow Drag Coefficient, Referenced to J1 (Cdc )h

Hull cross-flow drag coefficient is chosen to be cross-flow drag coefficient of a circular

cylinder at the hull’s maximum cross-flow Reynolds number [18]. The Reynolds number

is estimated for a wind speed of 20 m/s as below [4]:

ρa V∞ Dh
Re =
µ
38

The Reynolds number comes out to be about 152 × 105 which is very much in the turbulent

region. From Hoerner [16], (Cdc )h for this range of Reynolds number is estimated.

Fins Cross-flow Drag Coefficient, Referenced to Sf : (Cdc )f

Fin cross flow drag coefficient is a function of the aspect ratio and the taper ratio of the

fin. It is provided in Wardlaw [52] as a plot for various taper ratios and is reproduced in

Figure 2.6.

7
Λ=0
Λ=0.5
6 Λ=1

4
(Cdc )f

0
0 0.5 1 1.5 2 2.5 3
Aspect Ratio

Figure 2.6: Fins cross flow drag coefficient as function of aspect and taper ratio [52]

From this figure, the value of (Cdc )f has been obtained. For intermediate values of

aspect ratio and taper ratio, a linear interpolation is used.

Hull Zero-angle Axial Drag Coefficient, Referenced to Sh : (Cdh )0

The relation to calculate (Cdh )wet has been provided in Hoerner [16] as below:

3/2 3
(Cdh )wet
 
Dh Dh
= 1 + 1.5 +7 (2.43)
C fh c̄ c̄
39

The coefficient Cfh is available as a function of Reynolds number in Hoerner [16]. The

wet area of the aerostat envelope is the hull surface area. Once (Cdh )wet is known (Cdh )0

can simply be evaluated by changing the reference area.


Fins Zero-angle Axial Drag Coefficient, Referenced to Sf : Cdf 0

The relation to calculate Cdf wet
has been provided in Hoerner [16] as below:

  4
Cdf
 
wet t t
=1+2 + 60 (2.44)
2Cff c̄ c̄

The coefficient Cff is available as a function of Reynolds number in Hoerner [16]. The
 
wet area is the fin surface area. Once Cdf wet is known Cdf 0 can simply be evaluated

by changing the reference area.

Fin and Hull Efficiency Factors (ηk , ηf )

The fin and hull efficiency factor as function of non-dimensional parameters have been

plotted in reference [18]. These plots were obtained by curve fitting the values of some

known airships and the efficiency factors for the present test aerostat envelope were ob-

tained from these plots.

By following the process just explained, we get all the parameters required in Equa-

tions (2.39-2.41) for the estimation of aerodynamic coefficients of the test aerostat enve-

lope. All these parameters are presented in Table 2.3.

Table 2.3: Estimation of parameters for Semi Empirical Method


Parameter Value(SI units) Parameter Value(SI units)
I1 50.9441 (Cdc )h 1.0
I3 -1.2049e+003 k3 − k1 0.7024
J1 237.3575 ηk 1.1
J2 3.2187e+003 (Cnα ∗ )f 2.2062
Iˆ1 0.3185 ηf  0.53
ˆ
I3 -0.2225 ˆlf 0.7990
 1
Jˆ1 1.4839 ˆlf 0.8891
2
Jˆ2 0.5944 (Cd ) h 0 0.0412
Ŝf 0.6968 Ŝh  1.0
(Cdc )h 0.32 Cdf 0 0.0059

If we substitute these parameters in Equations (2.39-2.41), we get the aerodynamic

coefficients of the test aerostat. The normal and axial force coefficients are converted to
40

lift and drag coefficients by using the relation as presented in Anderson [4].

CL = Cn cos α − Ca sin α

CD = Cn sin α + Ca cos α

The lift coefficient, drag coefficient and pitching moment coefficient about envelope nose

obtained using semi empirical method are presented in Table 2.4 for angle of attack from

-20 deg to +20 deg. It is observed that the lift and moment coefficients about envelope

nose are symmetric about the origin and the drag coefficient is symmetric about the y-axis.

This is expected as the method considers the projected horizontal fins in place of inverted-

Y fins, making the system symmetric for these coefficients. It is also observed that the lift

and moment coefficient about nose vanishes for 0 deg angle of attack as expected. The lift

coefficient is having positive slope whereas the pitching moment coefficient about envelope

nose is having negative slope. These coefficients have also been plotted in Figures 2.25

to 2.27 and these plots will be discussed in section 2.4.4.

Table 2.4: Lift, drag and pitching moment coefficient about envelope nose for the test
aerostat envelope evaluated by semi-empirical method
Angle of Lift Drag Moment coefficient
attack (deg) coefficient coefficient about nose
-20 -0.5472 0.1254 0.1935
-18 -0.4866 0.1073 0.1676
-16 -0.4265 0.0919 0.1429
-14 -0.3671 0.0791 0.1195
-12 -0.3089 0.0686 0.0975
-10 -0.2522 0.0604 0.0770
-8 -0.1972 0.0543 0.0580
-6 -0.1442 0.0499 0.0408
-4 -0.0936 0.0472 0.0253
-2 -0.0454 0.0457 0.0117
0 0 0.0453 0
2 0.0454 0.0457 -0.0117
4 0.0936 0.0472 -0.0253
6 0.1442 0.0499 -0.0408
8 0.1972 0.0543 -0.0580
10 0.2522 0.0604 -0.0770
12 0.3089 0.0686 -0.0975
14 0.3671 0.0791 -0.1195
16 0.4265 0.0919 -0.1429
18 0.4866 0.1073 -0.1676
20 0.5472 0.1254 -0.1935
41

2.4.2 Computational Fluid Dynamics

The Computational Fluid Dynamics (CFD) approach to estimate an aerodynamic coef-

ficients is based on the numerical solution of Navier-Stokes equations using commercial

software. Since the ease of availability of high performance computing facility, this has

become quite feasible. Applying the fundamental laws of mechanics to a fluid gives the

governing equations for a fluid and the conservation of mass and momentum equations are

written as [3, 49]:


∂ρa
+ ∇. (ρa V ) = 0 (2.45)
∂t
∂V
ρa + ρa (V .∇) V = −∇p + ρa g + ∇.τij (2.46)
∂t

Here V is the flow velocity at a point, g is the acceleration due to gravity vector, p is

the pressure and τij are the stresses. These equations form a set of coupled, nonlinear

partial differential equations. It is not possible to solve these equations analytically for

most engineering problems. However, it is possible to obtain approximate computer-

based solutions to the governing equations using the CFD approach with a commercial

software such as ANSYS FLUENT R . For all flows, ANSYS FLUENT R solves conservation

equations for mass and momentum. For flows involving heat transfer or compressibility,

an additional equation for energy conservation is solved. Additional transport equations

are also solved when the flow is turbulent [5].

For the case of test aerostat envelope an external flow analysis using CFD has been

carried out. A wind speed of 20 m/s and angle of attack from -20 deg to +20 deg have

been considered with air as working medium at standard atmospheric conditions. As the

aerodynamic contribution of payload frames and structures is very small, it has not been

taken into consideration. Similar assumption has also been made in semi empirical method

and the wind tunnel testing results have also been presented for clean configuration.

2.4.2.1 Geometrical Description

Solid 3D modeling of the aerostat envelope was carried out in CATIA R software as ex-

plained in section 2.2.2 in detail. The cordages and payload are not considered for carrying

out CFD analysis because these are not going to contribute significantly as compared to

the hull and fins and also because of great simplification achieved in the analysis. Fig-

ure 2.7 shows the 3D model of test aerostat envelope without cordages and payloads. Once
42

Figure 2.7: 3D model for aerostat envelope used for CFD analysis

Figure 2.8: 3D external domain for aerostat envelope used for CFD analysis

the necessary geometrical approximations have been made, model is saved in IGES for-

mat file. This file is imported in ICEM-CFD R meshing tool as a pre-processor of ANSYS

FLUENT R . Surfaces and curves which are too small in comparison to hull or not required

for the CFD analysis have been cleaned up. After the cleaning of the geometry of aerostat,

it is required to define the computational domain. A cylindrical computational domain

with radius of 5 times the length of the aerostat envelope is considered. The upstream and

downstream domains are extended by a value equal to 6 and 8 times the length of aerostat

envelope respectively. Figure 2.8 shows the 3D external domain for aerostat envelope used

for CFD analysis.


43

2.4.2.2 Grid Generation and Boundary Conditions

After the cleaning of geometry and defining computational domain, grid generation is per-

formed. During the grid generation in CFD, we are having a choice between a structured

grid for simple geometry and unstructured grid for complex geometry. While the struc-

tured grid has regular connectivity and conserves storage space, the unstructured grid has

nonuniform pattern and required large storage space. Other risks of using structured or

block-structured meshes with complicated geometries include the oversimplification of the

geometry, mesh quality issues, and a less efficient mesh distribution that results in a high

cell count. Keeping these points in view along with complexity of the geometry of aero-

stat envelope, an unstructured tetrahedral grid is generated for full scale (1:1) geometry

of aerostat envelope. The salient features of the generated grid are that it uses an Octree-

based algorithm to fill the volume with tetrahedral elements, clustering is made near the

wall of hull and fins and that the total cell count for full geometry of aerostat envelope is

8.5 million.

Figure 2.9 shows the unstructured mesh generated for aerostat envelope. Mesh is

imported in FLUENT R solver for further definition of boundary conditions. A wind

speed of 20 m/s is taken for CFD simulations. It is a low subsonic flow. For this, all

far field boundary is considered as Pressure- far- field boundary condition to specify flow

directions. Rest of the geometries is considered as “Wall”. Standard air properties at sea

level have been used as working parameters. Air is considered as working fluid. The mesh

clustering near the wall of hull and fins is depicted in Figure 2.10.

2.4.2.3 Turbulence Model

As presented in section 2.4.1, the flow over the aerostat envelope is predominantly tur-

bulent. Hence it is also required to model the turbulence during the numerical solution

of conservation equations. The flow behavior over the aerostat envelope is considered to

be in steady state, viscous, incompressible and turbulent. The choice of turbulence model

depends on considerations such as the physics encompassed in the flow, the established

practice for a specific class of problem, the level of accuracy required, the available com-

putational resources, and the amount of time available for the simulation. For modeling

turbulence in the present analysis, the Spalart-Allmaras model [5, 46] has been used. The

Spalart-Allmaras model is a one-equation model that solves a modeled transport equation


44

Figure 2.9: Unstructured mesh for CFD analysis

Figure 2.10: Mesh clustering near wall of hull and fins of aerostat envelope

for the kinematic eddy (turbulent) viscosity. The Spalart-Allmaras model was designed

specifically for aerospace applications involving wall-bounded flows and has been shown

to give good results for boundary layers subjected to adverse pressure gradients [5].

2.4.2.4 Results of CFD Analysis

A CFD simulation at different angles of attack for given parameters have been carried

out. For low subsonic speed, flow around aerostat separates as well as circulates. Several

iterations are carried out to achieve convergence. A convergence criterion of 1e-06 is

maintained with second order accuracy. Truncation errors have been overcome by using
45

finer meshes. For this, the aerostat boundary mesh has been refined and adapted for grid

independence study to get an accurate solution. Round-off errors have been minimized by

using double precision solver. The lift coefficient, drag coefficient and pitching moment

coefficient about envelope nose obtained from the CFD analysis are presented in Table 2.5.

It is observed that lift and moment coefficients about envelope nose do not vanish for 0

deg angle of attack and these coefficients are not completely symmetric about the origin as

opposed to semi empirical method because the CFD approach considers the real fins and

not the projected horizontal fins. The lift coefficient is having a slightly negative value for

0 deg angle of attack which may be attributed to the effect of providing inverted-Y fins

on the lower portion of the hull. These coefficients have also been plotted in Figures 2.25

to 2.27 and these plots have been discussed in section 2.4.4. The pressure coefficient plot

for 0 deg angle of attack along the length of aerostat has been plotted in Figure 2.11 shows

the fluctuating pattern on the rear portion of aerostat hull because of presence of fins.

Table 2.5: Lift, drag and pitching moment coefficient about envelope nose for the test
aerostat envelope evaluated by CFD
Angle of Lift Drag Moment coefficient
attack (deg) coefficient coefficient about nose
-20 -0.3555 0.1153 0.1276
-18 -0.3262 0.0966 0.1164
-16 -0.3070 0.0787 0.1147
-14 -0.2801 0.0692 0.1078
-12 -0.2407 0.0612 0.0922
-10 -0.2118 0.0535 0.0853
-8 -0.1724 0.0479 0.0704
-6 -0.1321 0.0434 0.0552
-4 -0.0914 0.0400 0.0398
-2 -0.0454 0.0376 0.0204
0 -0.0071 0.0372 0.0074
2 0.0398 0.0372 -0.0123
4 0.0810 0.0391 -0.0273
6 0.1216 0.0422 -0.0417
8 0.1617 0.0465 -0.0558
10 0.1955 0.0519 -0.0651
12 0.2278 0.0583 -0.0733
14 0.2622 0.0668 -0.0840
16 0.2922 0.0778 -0.0922
18 0.2831 0.1051 -0.0777
20 0.2659 0.1443 -0.0604

The computed flow fields obtained after CFD analysis of aerostat for various angles of

attack have been plotted as contour plots. Section planes perpendicular to the central axis
46

Figure 2.11: Pressure coefficient variation over hull at 0 deg angle of attack

and at a distance of 10m and 28m from nose were created for clear visualization of flow

behavior at these places. Figures 2.12 to 2.15 present the results for 0 deg angle of attack.

Figure 2.12 represents pressure contour and the pressure distribution is symmetric except

near the fins. This is expected as fins are not symmetric about a horizontal plane passing

through the central axis. The maximum pressure is observed near the envelope nose where

the flow is retarded. Figure 2.13 presents velocity contour which again demonstrates

symmetry except near the fins. The maximum velocity is observed near the envelope

surface somewhere in the middle were the pressure is lowest. The pressure contour at a

sectional plane 28 m from nose is presented in Figure 2.14 where the highest pressure is

observed near the fin tip and the lowest pressure is observed on the both sides of the fins.

Also as observed in Figure 2.14, the pressure contour is symmetric about a vertical plane

passing through central axis of the aerostat envelope because the fins are also symmetric

about this plane. The pressure contour at a sectional plane 10m from nose is presented

in Figure 2.15 where complete symmetry is observed indicating that the effect of fins are

not present.

The velocity vector plot at an angle of attack 10 deg is presented in Figure 2.16 and

it is observed that maximum velocity is around the upper portion of the lower fins near

the leading edge where the pressure is expected to be the lowest. For this typical case,

maximum velocity is about 32 m/s. Figures 2.17 to 2.20 present the contour plot results

for 20 deg angle of attack. The sectional pressure contour for the aerostat envelope has

been presented in Figure 2.17 where it is observed that due to positive angle of attack the

stagnation point shifts downward whereas the low pressure zone is created in the space
47

Figure 2.12: Pressure contour at an angle of attack 0 deg

Figure 2.13: Velocity contour at an angle of attack 0 deg

between lower fins. The velocity contour is presented in Figure 2.18 and it is observed here

that velocity is having maximum magnitude in the zone where pressure is minimum, as

expected. Figure 2.18 also indicates that flow separation occurs much earlier as compared

to the case of 0 deg angle of attack. The pressure contours at a distance of 10 m and 28

m from the nose have been presented in Figure 2.19 and Figure 2.20 respectively. The

pressure contours at a distance of 10m and 28m from the nose for -20 deg angle of attack

have been presented in Figure 2.21 and Figure 2.22 respectively. It is noted that low

pressure zone for positive angle of attack is on the upper side whereas for negative angle

of attack it is on the lower side. As angle of attack increases the corresponding change in

flow field can be clearly seen as it is visible in contour plots.


48

Figure 2.14: Pressure contour at 28m from nose and an angle of attack 0 deg

Figure 2.15: Pressure contour at 10m from nose and an angle of attack 0 deg

Figure 2.16: Velocity vector plot at an angle of attack 10 deg


49

Figure 2.17: Pressure contour at an angle of attack 20 deg

Figure 2.18: Velocity contour at an angle of attack 20 deg

Figure 2.19: Pressure contour at 10m from nose and an angle of attack 20 deg
50

Figure 2.20: Pressure contour at 28m from nose and an angle of attack 20 deg

Figure 2.21: Pressure contour at 10m from nose and an angle of attack -20 deg

Figure 2.22: Pressure contour at 28m from nose and an angle of attack -20 deg
51

2.4.3 Wind Tunnel Testing

To measure the aerodynamic forces and moments on the aerostat envelope experimentally,

wind tunnel testing has been carried out on scaled down model. The test has been carried

out at National Wind Tunnel Facility (NWTF), IIT Kanpur. NWTF is a low speed closed

circuit wind tunnel with a test section of cross section 3 m × 2.25 m. The tunnel is able

to produce flow with velocity up to 80 m/s at a turbulence level of less than 0.1%. The

wind tunnel is powered by 1000 kW variable speed DC motor through a 4.64 m diameter

12-bladed axial flow fan. The wind tunnel test was conducted for angle of attack range 0

to ±20 deg at wind velocity of 30 m/s [31].

2.4.3.1 Wind Tunnel Model Details

A rigid scaled down FRP model of the test aerostat envelope was fabricated that was

machined properly to obtain the desired shape accurately. The length of model is 1.693

m. The fins were also fabricated with FRP using same length scale. For the measurement

of forces and moments, the model was instrumented with a six component strain-gauge

balance in the test section using sting support system.

2.4.3.2 Test Instrumentation

The force measurement system consists of a windows-based host computer installed with

LabVIEW R -based application software. The Application software links the host computer

through its PCI-8336 board to the PXI-8336 board installed in the PXI-System. The

force signal of six-component strain-gauge balance is acquired using a high-accuracy 18-

bit data acquisition module (PXI-6289) of the PXI System through Universal Strain-Gage

Signal Conditioner (SCXI System). The application software also performs all the required

functions like balance test and calibration, data acquisition, and analysis. The accuracy

of the force measurement system is maintained within ±0.66% for axial force, ±0.06%

for normal force, ±0.16% for side force, and ±0.14% for rolling moment of the balance

calibration range.

Calibration of the balance is done for the accurate measurement of forces acting on

the model during a test. The percentage error over the range in which the balance is

calibrated is within ±0.3% of the balance full-scale range. Calibration procedure uses a

single component calibration rig. For loading, pre-calibrated dead weights are used and all
52

the loadings are performed by using gravity-loading methodology. Levelling of the balance

is performed at each loading point. Total calibration procedure including initialization,

bridge nulling, data acquisition, monitoring of the acquired test data, computing the in-

verse matrix and the final acceptance check of the balance are performed by LabVIEW R

based application software. The calibration process consists of creating a calibration data

file by applying a series of known loads to the balance and acquiring its electrical sig-

nal output. The balance calibration matrix is calculated using this calibration data file.

The inverse of calibration matrix, also known as load matrix, is used for computing the

aerodynamic forces and moments.

2.4.3.3 Model Mounting Scheme and Results

The six component strain gauge based forces and moments measurement system is first

calibrated and cross checked in the desired load ranges, which are pre-determined based on

expected coefficients values. The aerostat envelope model mounted on the robotic arm and

turn table is shown in Figures 2.23 and 2.24 respectively. The model was mounted on the

turn table through an internal strain gauge balance fixed to a sting support mechanism.

The sting was further attached between two turn tables having synchronized motion. For

reducing the moments on the balance due to model weight, a front-end balance adapter is

attached at the appropriate longitudinal position inside the model. This balance adapter

transfers forces and moments to the six-component internal strain gauge balance. The

balance is locked to the balance adapter using axial bolt and roll lock pin. An internal

balance with 50 kg axial force limit, 100 kg side force limit and 200 kg normal force limit

was used to measure forces and moments experienced by the model. The wind tunnel study

on the aerostat model was conducted using Robotic-Arm as well as Beta Mechanism model

mounting systems. While the robotic arm is capable of giving all the required orientations

to the model (Figure 2.23), beta mechanism consisting of electric motor is required at the

junction point between the turn tables to give sideslip orientation (Figure 2.24). The lift

coefficient, drag coefficient and pitching moment coefficient about envelope nose obtained

from wind tunnel testing are presented in Table 2.6 [31]. Here also, a slight negative value

of lift coefficient is observed for 0 deg angle of attack. It comes out from the results of

CFD and wind tunnel testing that the net effect of two inverted-Y fin on the hull is a slight

negative angle of attack of envelope with projected horizontal fins. These coefficients have
53

Figure 2.23: Aerostat envelope model mounted on robotic arm

Figure 2.24: Aerostat envelope model mounted on turn table

also been plotted in Figures 2.25 to 2.27 and these plots will be discussed in section 2.4.4.

2.4.4 Comparison of Aerodynamic Coefficients for Different Methods

The lift coefficient, drag coefficient and pitching moment coefficient about aerostat en-

velope nose obtained from different methods viz. semi empirical method, computational

fluid dynamics and wind tunnel testing are plotted in Figure 2.25, Figure 2.26 and Fig-

ure 2.27 respectively for angle of attack from -20 deg to +20 deg. Following observations

can be made from these plots:

1. It is observed that the trend of variation of the coefficients with angle of attack

remains same in all the three methods i.e. lift coefficients increases with increase in
54

Table 2.6: Lift, drag and pitching moment coefficient about envelope nose for the test
aerostat envelope evaluated by wind tunnel testing
Angle of Lift Drag Moment coefficient
attack (deg) coefficient coefficient about nose
-19.9833 -0.6865 0.2127 0.3544
-17.9937 -0.6745 0.1548 0.3513
-15.9833 -0.5685 0.1234 0.2868
-13.9937 -0.4829 0.0996 0.2382
-11.9937 -0.3908 0.0806 0.1851
-9.9833 -0.3149 0.0662 0.1447
-7.9938 -0.2415 0.0551 0.1070
-5.9833 -0.1724 0.0474 0.0729
-3.9938 -0.1145 0.0427 0.0480
-2.0042 -0.0618 0.0395 0.0274
0.0063 -0.0157 0.0389 0.0125
2.0062 0.0312 0.0402 -0.0029
4.0167 0.0934 0.0443 -0.0301
6.0167 0.1514 0.0506 -0.0542
8.0063 0.2176 0.0589 -0.0851
10.0063 0.2917 0.0707 -0.1227
12.0063 0.3644 0.0866 -0.1594
14.0167 0.4284 0.1047 -0.1914
16.0063 0.5023 0.1281 -0.2300
17.9958 0.5693 0.1548 -0.2645
20.0063 0.5915 0.2056 -0.2697

angle of attack, the drag coefficient first decreases and then increases with increase

in angle of attack thereby forming a drag polar and the moment coefficient decreases

with increase in angle of attack.

2. The lift and moment coefficients can be approximated by a linear curve whereas the

drag coefficient can be approximated by a quadratic curve.

3. The agreement of aerodynamic coefficients is very good within angle of attack range

from -5 deg to +5 deg. Beyond this range, the difference in general increases with

increase in absolute angle of attack.

4. The slope of lift coefficient curve as well as the magnitude of pitching moment

coefficient curve is highest for wind tunnel testing whereas lowest for computational

fluid dynamics.

5. The drag coefficient is lowest for the computational fluid dynamics approach and

highest for wind tunnel testing except for low magnitude of angle of attack, where

the coefficient is highest for semi empirical method.


55

0.8
Semi Empirical Method
CFD
0.6
Wind Tunnel Testing

0.4

0.2
CL

−0.2

−0.4

−0.6

−0.8
−20 −15 −10 −5 0 5 10 15 20
Angle of attack (deg)

Figure 2.25: Lift coefficient as a function of angle of attack for different methods

6. A flow separation is observed in CFD result beyond the angle of attack range of ±16

deg which results in decrease in lift curve slope and increase in drag. Similar flow

separation behavior is observed in wind tunnel testing beyond ±18 deg whereas no

flow separation is observed in semi empirical result. The boundary layer effects have

not been taken into account in semi empirical method and hence it is not expected

to predict flow separation.

The computational fluid dynamics approach is based on purely theoretical approach where

the flow equations are solved using digital computer. The wind tunnel approach on the

other hand is purely experimental in nature. The semi empirical method is a combination

of theoretical and experimental approach as explained in section 2.4.1 and the values of

aerodynamic coefficients obtained using semi-empirical method also lie mostly between

computational fluid dynamics and wind tunnel results. The possible reasons for the dif-

ference between the CFD and wind tunnel results may include the imperfection in model

geometry, limited wind tunnel cross section and the disturbance in the flow in the tunnel

due to presence of vertical column and horizontal robotic arm just downside the model.
56

0.4
Semi Empirical Method
CFD
0.35
Wind Tunnel Testing

0.3

0.25
CD

0.2

0.15

0.1

0.05

0
−20 −15 −10 −5 0 5 10 15 20
Angle of attack (deg)

Figure 2.26: Drag coefficient as a function of angle of attack for different methods

0.4
Semi Empirical Method
CFD
0.3
Wind Tunnel Testing

0.2

0.1
Cmn

−0.1

−0.2

−0.3

−0.4
−20 −15 −10 −5 0 5 10 15 20
Angle of attack (deg)

Figure 2.27: Moment coefficient about nose vs. angle of attack for different methods
57

2.4.5 Calculation of Dynamic and other Stability Derivatives

The dynamic and other derivatives for the test aerostat envelope have been estimated based

on relations presented in chapter 5 of reference [13] and also presented in reference [26] for

aerostat envelope. The test aerostat envelope has three fins in inverted-Y configuration

i.e. one vertical fin and two inclined fins. To approximate the geometry so that the

relations present in the above references can be applied, the inclined fins are projected in

horizontal and vertical directions to get the components as illustrated in Figure 2.28. Thus

we have Projected Horizontal Tails (PHT), Vertical Tail (VT) and Projected Vertical Tails

(PVT) for the envelope. The two projected horizontal tails are treated in combination

because they have same dimensions and produce same effect on longitudinal dynamic

derivatives. The vertical tail and two projected vertical tails are treated separately because

they have different dimensions and may or may not produce similar effect on lateral

dynamic derivatives. Also, the contributions of hull to the derivatives are neglected in

present approximation which is primarily done for simplification [26] although the hull

contributions to the derivatives may be significant. The wind speed for test aerostat

operation is within low subsonic range and wind speed will not have any significant effect

on these derivatives.

Figure 2.28: Projection of tails for aerostat envelope for estimating stability derivatives

As illustrated in reference [40] and also applied in section 2.4.1 of this chapter the

lift curve slope and the distance of mean aerodynamic chord from the center of gravity

i.e. the moment arm are evaluated for projected horizontal tail, vertical tail and projected

vertical tail respectively. The slopes are denoted by apht , avt and apvt based on the aerostat
58

2/3
envelope reference area which is Vh . The horizontal distances between center of mass and

mean aerodynamic chord of tails are denoted by lpht , lvt and lpvt for projected horizontal

tail, vertical tail and projected vertical tail respectively. The corresponding tail volume

ratios as defined in [13] are denoted by Vpht , Vvt and Vpvt . The vertical distances between

center of mass and mean aerodynamic chord of vertical tail and projected vertical tail are

hvt and hpvt respectively.

2.4.6 Longitudinal Dynamic Derivatives

The relevant longitudinal dynamic derivatives are CLq and Cmq . As illustrated in reference

[13], the hull contributions are neglected and the expression for these derivatives may be

written as:
lpht
CLq = 2apht Vpht = 2apht (2.47)

 2
lpht
Cmq = −2apht (2.48)

2.4.7 Lateral Dynamic Derivatives

The lateral dynamic derivatives are obtained for vertical tail and projected vertical tails

separately and then properly added to give the resultant dynamic derivatives for the

complete vertical tail.


 
 ∂σ
CYβ vt = −avt 1 − ≈ −avt (2.49)
∂β
 
 ∂σ hvt hvt
Clβ vt = −avt 1 − ≈ −avt (2.50)
∂β c̄ c̄
 
 ∂σ lvt lvt
Cnβ vt = avt 1 − ≈ avt (2.51)
∂β c̄ c̄
 
 hvt ∂σ hvt
CYp vt = −avt 2 − ≈ −2avt (2.52)
c̄ ∂ p̂ c̄

hvt 2
   
 hvt ∂σ hvt
Clp vt = −avt 2 − ≈ −2avt (2.53)
c̄ ∂ p̂ c̄ c̄
 
 hvt ∂σ lvt lvt hvt
Cnp vt = avt 2 − ≈ 2avt 2 (2.54)
c̄ ∂ p̂ c̄ c̄
 
lvt ∂σ lvt
(CYr )vt = avt 2 + ≈ 2avt (2.55)
c̄ ∂ r̂ c̄
 
lvt ∂σ hvt lvt hvt
(Clr )vt = avt 2 + ≈ 2avt 2 (2.56)
c̄ ∂ r̂ c̄ c̄
59

   2
lvt ∂σ lvt lvt
(Cnr )vt = −avt 2 + ≈ −2avt (2.57)
c̄ ∂ r̂ c̄ c̄

Similarly, the dynamic derivatives for projected vertical tail are approximated as below:

 
 ∂σ
C Yβ pvt
= −apvt 1 − ≈ −apvt (2.58)
∂β

 
 ∂σ hpvt hpvt
Clβ pvt = apvt 1 − ≈ apvt (2.59)
∂β c̄ c̄
 
 ∂σ lpvt lpvt
Cnβ pvt = apvt 1 − ≈ apvt (2.60)
∂β c̄ c̄
 
 hpvt ∂σ hpvt
CYp pvt = apvt 2 − ≈ 2apvt (2.61)
c̄ ∂ p̂ c̄

hpvt 2
   
 hpvt ∂σ hpvt
Clp pvt = −apvt 2 − ≈ −2apvt (2.62)
c̄ ∂ p̂ c̄ c̄
 
 hpvt ∂σ lpvt lpvt hpvt
Cnp pvt = −apvt 2 − ≈ −2apvt (2.63)
c̄ ∂ p̂ c̄ c̄2
 
lpvt ∂σ lpvt
(CYr )pvt = apvt 2 + ≈ 2apvt (2.64)
c̄ ∂ r̂ c̄
 
lpvt ∂σ hpvt lpvt hpvt
(Clr )pvt = −apvt 2 + ≈ −2apvt (2.65)
c̄ ∂ r̂ c̄ c̄2
   2
lpvt ∂σ lpvt lpvt
(Cnr )pvt = −apvt 2 + ≈ −2apvt (2.66)
c̄ ∂ r̂ c̄ c̄

Now the dynamic derivatives for the complete tail is evaluated as the algebraic sum of the

contributions from vertical tail and double the contributions of one projected vertical tail

as below:
 
C Yβ = C Yβ vt
+ 2 C Yβ pvt
(2.67)

 
Cl β = Cl β vt
+ 2 Cl β pvt
(2.68)

 
Cnβ = Cnβ vt
+ 2 Cnβ pvt
(2.69)

 
CYp = CYp vt
+ 2 C Yp pvt
(2.70)

 
Clp = Clp vt
+ 2 Clp pvt
(2.71)

 
Cnp = Cnp vt
+ 2 Cnp pvt
(2.72)

CYr = (CYr )vt + 2 (CYr )pvt (2.73)


60

Clr = (Clr )vt + 2 (Clr )pvt (2.74)

Cnr = (Cnr )vt + 2 (Cnr )pvt (2.75)

All the aerodynamic derivatives estimated for the test aerostat envelope are presented

in Table 4.2. The lift, drag and moment coefficient as well as their α derivatives are

presented from the wind tunnel testing results with linear approximation for lift and

moment coefficient and quadratic approximation for drag coefficient. All other derivatives

are estimated based on the method described in this section.

2.5 Tether cable properties

For carrying out linear dynamic analysis of the tethered aerostat using small perturbation

theory as explained in chapter 4, the tether has been assumed as flexible and inextensible

cable having small mass compared to aerostat. The properties required for linear dynamic

modeling are tether diameter and its weight per unit length. These two properties are very

simple to measure. Other required parameters are all dependent on aerostat configuration.

As will be explained in chapter 5 that for nonlinear dynamic modeling of the tethered

aerostat the tether will be modeled as a spring-mass-damper system and hence required

tether properties are its mass, stiffness and damping properties of a finite segment of

tether. The mass of a tether sample is immediately known as weight per unit length of

tether is known. The stiffness of the tether segment can be extracted from load-strain

curve. The load-strain curve for the test aerostat tether sample has been generated using

Universal Testing Machine (UTM) and the results have been presented in Figure 2.29.

This figure shows the results for two tether samples. The test for first tether sample has

been stopped as the load reached a preset value whereas the test for second tether sample

has been continued till breakage. It is observed that the slope of load-strain curve is low

on low loads but increases with increase in load and becomes almost constant at higher

loads. As will be seen in chapter 5, during normal operation, the tension on the tether may

be around 10000 N. Hence the slope of load-strain curve around this value is of interest.

If the slope of load-strain curve is dT /dε expressed as N/%, then the stiffness of a tether
61

segment having length l0 in N/m can be written as:

dT /dε
Ks = × 100 (2.76)
l0

Figure 2.29: Static load testing of tether sample

The damping property of the tether may be measured experimentally as suggested in

reference [53]. But for the present test aerostat, the damping property has been assumed

same as presented in reference [15] and extrapolated to length l0 using ‘dashpot in se-

ries’ relation. The spring-mass-damper properties of the tether cable estimated for the

test aerostat has been listed in Table 5.1.


Chapter 3

Equilibrium and Static Analysis of


the Tethered Aerostat

3.1 Introduction

This chapter presents the equilibrium and static analysis for the tethered aerostats. As

presented in Figure 1.1, the aerostats are tethered to the anchor point or mooring structure

unlike the normal aircraft. Under a constant horizontal wind speed, the aerostat attains

equilibrium with a trim angle and tether tension and the tether attains equilibrium with

a particular shape from the confluence point to the anchor point. This also results in

the horizontal and vertical shift of aerostat envelope from the zero wind position. The

objective of carrying out equilibrium analysis is to estimate the tether tension, trim angle

of attack and the tether cable shape for a given horizontal wind speed. The wind speed is

assumed constant in magnitude for a particular analysis case and parallel to the horizontal

direction. It is also assumed that the aerostat envelope is aligned along the wind direction

with the wind flowing from envelope nose to tail, thus equilibrium only in longitudinal

direction is considered. The more practical wind conditions are considered in Chapter 5

which have been applied to a nonlinear tethered aerostat model. All the techniques for

generating required inputs for carrying out the static and equilibrium analysis have already

been discussed in chapter 2.

63
64

3.2 Tether Tension

As mentioned earlier, the aerostat configuration establishes its equilibrium under given

constant horizontal wind speed V∞ with a certain value of pitch angle and a corresponding

horizontal and vertical shift from zero wind condition. The aerostat in this condition is

acted upon by its weight, buoyancy force, aerodynamic force and tether tension [22] as

shown in Figure 3.1.

Figure 3.1: Forces and moments acting on aerostat in equilibrium condition

The weight includes the weight of structure of aerostat envelope WS and helium gas weight

WG acting downward at the combined center of mass of envelope structure and helium

gas. The buoyancy force B acts at the center of buoyancy in the upward direction. The

aerodynamic lift and drag force are represented by L and D respectively. The aerodynamic

forces are also associated with a pitching moment which can be referred to any arbitrary

point. In the present case, the pitching moment has been referred about the envelope nose.

These forces and moments are discussed in detail in section 3.3 of this chapter. Balancing

the forces in horizontal and vertical directions, tether tension T1 and angle with horizontal

γ1 at the confluence point of the aerostat can be estimated as below:


65

q
T1 = D2 + (Bg + L − WS )2 (3.1)
 
−1 Bg + L − WS
γ1 = tan (3.2)
D

where, Bg = B − WG = Vg (ρa − ρg ) g is the upward buoyancy less the gas weight. Using

Equations 3.1 and 3.2, the tether tension and the angle of tether tension vector with

horizontal at the aerostat envelope confluence point are evaluated. The results for these

parameters are presented here for both the small aerostat and the test aerostat. The

operating height for small aerostat is 100 m whereas the operating height for the test

aerostat is 1000 m. The other characteristics of the small and test aerostat are presented

in section 1.5 of chapter 1.

The required parameters for the small aerostat is presented in Table 3.1. It is evident

from this table that hull and helium volume are same and hence the fins are air filled.

The air density ρa and gross lift Bg corresponds to a height of 100 m above the mean

sea level for ISA+15 deg temperature condition. The lift and drag coefficients are taken

from wind tunnel testing. The lift coefficient is modeled as linear curve whereas the drag

coefficient is modeled as complete quadratic curve. The aerodynamic coefficients for the

small aerostat is same as the test aerostat as the shape of both the aerostats are same.

Table 3.1: Parameters for evaluating tether tension vector for small aerostat
Parameter Value
Helium volume, Vg 250 m3
Hull volume, Vh 250 m3
Reference area, S Vh 2/3
Air density, ρa 1.1531 kg/m3
Envelope structure weight, WS 1510.74 N
Gross lift, Bg 2386.5 N
Lift coefficient, CL 1.8313αt -0.0245
Drag coefficient, CD 1.4090αt 2 +0.0074αt +0.0294

Figures 3.2 and 3.3 present the variation of tether tension and angle with horizontal

with wind speed and angle of attack for the small aerostat. It is observed in Figure 3.2

that tether tension increases with increase in wind speed except for the case of zero angle

of attack in which case the tether tension decreases slightly with increase in wind speed.

This is due to the fact that for zero angle of attack the lift coefficient is having a slightly

negative value which results in vertically downward force thereby reducing the tether
66

tension slightly with an increase in wind speed. For all other angles, the lift coefficient is

positive and hence the lift force increases in vertically upward direction thereby increasing

the lift. For the case of zero wind speed, the tether tension comes out to be 876 N and

this tether tension does not vary with angle of attack as expected. This tether tension

may also be obtained by subtracting the structural weight of envelope from the gross lift

in Table 3.1. For the case of 25 m/s wind speed, the tether tension for 0 deg angle of

attack is about 673 N which becomes 7605 N for 15 deg angle of attack. Hence flying the

aerostat at high angle of attack and also high wind speed at the same time is undesirable.

8000
αt = 0°
7000 αt = 5°
αt = 10°
6000 αt = 15°

5000
T1 (N)

4000

3000

2000

1000

0
0 5 10 15 20 25
V∞ (m/s)

Figure 3.2: Tether tension variation with wind speed and αt for small aerostat

It is observed in Figure 3.3 that at zero wind speed, the angle with horizontal remains

at 90 deg as expected since there is no horizontal force. It is also observed that with

increase in wind speed, the angle with horizontal decreases for all the angles of attack.

For the 0 deg angle of attack this decrease is extremely high, primarily due to the fact that

the lift force is decreasing with increase in wind speed and the drag force is increasing.

For higher angles of attack, the trend becomes regular. For the case of angle of attack 5
67

90

88

86

84

82
γ1 (deg)

80

78

76
αt = 0°
74 αt = 5°
αt = 10°
72
αt = 15°
70
0 5 10 15 20 25
V∞ (m/s)

Figure 3.3: Angle with horizontal variation with wind speed and αt for small aerostat

deg, the angle with horizontal falls from 90 deg for zero wind speed to 78.3 deg for 25 m/s

wind speed. This goes down further to 76 deg for 15 deg angle of attack.

Table 3.2 presents the parameters required for the calculation of tether tension and

carrying out equilibrium and static analysis of the test aerostat. This is an aerostat which

has got all three fins helium filled as the difference between helium volume Vg and hull

volume Vh is the total volume of three fins. The air density ρa and gross lift Bg corresponds

to a height of 1000 m above the mean sea level for ISA+15 deg temperature condition.

The lift, drag and moment coefficients are taken from wind tunnel testing. The lift and

moment coefficient about envelope nose are modeled as linear curves whereas the drag

coefficient is modeled as complete quadratic curve. The center of buoyancy coordinates

(XB , ZB ) were obtained from the solid modeling approach explained in chapter 2. The

center of mass coordinates (XG , ZG ) are the combined center of mass of aerostat envelope

structure and helium gas.

The results for the tether tension and angle with horizontal for the test aerostat are

presented in Figures 3.4 and 3.5 respectively. Here also we observe similar pattern of tether
68

Table 3.2: Parameters for evaluating tether tension vector and carrying out equilibrium
and static analysis for test aerostat (ISA+15 deg)
Parameter Value
Atmospheric air density, ρa 1.0554 kg/m3
Helium volume, Vg 2107 m3
Hull volume, Vh 2023 m3
Reference area, S Vh 2/3
Envelope length or reference length, c̄ 33.85m
Envelope structure weight, WS 11248 N
Gross lift, Bg 18406.3 N
Axial distance of CP from nose, XC 11.2m
Vertical distance of CP from nose, ZC 10.9m
Axial distance of CM from nose, XG 18.24 m
Vertical distance of CM from nose, ZG 1.85 m
Axial distance of CB from nose, XB 15.93 m
Vertical distance of CB from nose, ZB 0.0325m
Lift coefficient, CL 1.8313αt -0.0245
Drag coefficient, CD 1.4090αt 2 +0.0074αt +0.0294
Pitching moment coefficient about nose, Cmn -0.8782αt +0.0214

tension variation as for small aerostat. It is observed in these figures that for higher angles

of attack the tether tension increase sharply with increase in wind speed. This is due to the

fact that for higher angles of attack, the aerodynamic forces become dominant and varies

as square of the wind speed. For the case of zero wind speed, the tether tension comes out

to be 7158 N and this tether tension does not vary with angle of attack as expected. As

also stated earlier, this tether tension may also be obtained by subtracting the structural

weight of envelope from the gross lift in Table 3.2. For the case of 25 m/s wind speed, the

tether tension for 0 deg angle of attack is about 6067 N which becomes 31880 N for 15 deg

angle of attack, which indicates that operating aerostat in these conditions will require

stronger and hence heavier tether.

It is observed in Figure 3.5 that for test aerostat, the angle with horizontal range for

25 m/s wind speed is between 75.2 deg to 81.4 deg. It is also observed in both the cases

that angle with horizontal drops drastically with wind speed for higher wind speed. This

indicates that the aerostat is likely to shift more in horizontal direction. The results for

tether tension and angle with horizontal in tabular form for an angle of attack of 5 deg

are presented in Table 3.3 for both small aerostat and test aerostat. The results for tether

tension and angle with horizontal in this table will be used for studying the equilibrium

shape of tether cable in section 3.4.


69

4
x 10
3.5
αt = 0°
αt = 5°
3
αt = 10°
αt = 15°
2.5
T1 (N)

1.5

0.5
0 5 10 15 20 25
V∞ (m/s)

Figure 3.4: Tether tension variation with wind speed and αt for test aerostat

90

88

86

84

82
γ1 (deg)

80

78

76
αt = 0°
74 αt = 5°
αt = 10°
72
αt = 15°
70
0 5 10 15 20 25
V∞ (m/s)

Figure 3.5: Angle with horizontal variation with wind speed and αt for test aerostat
70

Table 3.3: Tether tension and angle with horizontal at the confluence point for small and
test aerostat for αt = 5 deg
V∞ (m/s) Small aerostat Test aerostat
T1 (N ) γ1 (deg) T1 (N ) γ1 (deg)
5 953.44 88.60 7444.33 89.34
10 1189.02 85.50 8307.57 87.63
15 1586.30 82.40 9758.88 85.45
20 2146.83 79.99 11807.38 83.30
25 2870.59 78.28 14457.58 81.44

3.3 Equilibrium Analysis of Aerostat Envelope

As explained in section 3.1 that under a constant horizontal wind speed the aerostat attains

equilibrium with a trim angle and tether tension and the tether attains equilibrium with a

particular shape. It is important to estimate trim angle of attack for the aerostat under the

operating wind speed to examine whether the estimated trim angle lies within the range of

specified limit. The static stability analysis is also required to be carried out to ascertain

the tendency of aerostat to return to equilibrium position if perturbed. Sensitivity studies

are required to estimate the effect of variation of relevant parameters on the aerostat trim

angle and static stability. These aspects of the aerostat envelope have been covered in

this section. The results are presented only for the test aerostat envelope. The forces

and moment acting on an aerostat envelope under equilibrium condition are shown in

Figure 3.1. These forces and moment are briefly described as below:

Buoyancy Force (B)

The buoyancy force acts vertically upward at the center of buoyancy of aerostat envelope.

It is equal to the weight of air displaced and is estimated as B = Vg ρa g. If the helium gas

weight is also taken into account, then the gross lift Bg = Vg (ρa − ρg )g can be written as

mentioned in section 3.2.

Gravity Force (WS + WG )

The gravity force on the aerostat envelope acts vertically downward at the center of mass.

This force is obtained by summation of the structural weight of the aerostat envelope (WS )

and the helium gas weight (WG ). This force is also called the total weight of the aerostat

envelope (WB ) acting at the combined center of gravity of aerostat structure and helium.
71

Aerodynamic force

The aerodynamic force on the aerostat envelope is represented by lift, drag and moment

at a reference point. There are various techniques available for estimating aerodynamic

force on the aerostat envelope which has already been explained in detail in chapter 2.

The coefficients for estimating aerodynamic force here has been directly taken from wind

tunnel testing results of scaled down model. The moment reference point here is envelope

leading edge or nose and the reference length is envelope length. The reference area for
2/3
the coefficients is S = Vh . The aerodynamic lift, drag and moment coefficients about

envelope nose are denoted as CL , CD and Cmn then the aerodynamic forces and moments

are expressed as:


1
L = ρa V∞ 2 SCL (3.3)
2
1
D = ρa V∞ 2 SCD (3.4)
2
1
Mn = ρa V∞ 2 Sc̄Cmn (3.5)
2

The forces expressed in Equations 3.3 and 3.4 are perpendicular to the wind direction and

along the wind direction respectively. The forces along axial and normal directions are

expressed as follows [4]:

Fa = D cos αt − L sin αt (3.6)

Fn = D sin αt + L cos αt (3.7)

Tether Tension (T1 )

The tether tension acts along the tether at the confluence point with an angle as shown

in Figure 3.1. It has been shown in section 3.2 that the tether tension and the angle with

horizontal can be expressed as a combination of buoyancy, gravity and aerodynamic force

in equilibrium condition.

In Figure 3.1, GA and GN represent the axial and normal directions where G is the

combined structural and helium gas center of mass of aerostat envelope. The buoyancy

force is B acting at center of buoyancy. The gas weight WG also acts at the center of

buoyancy. Thus the gross lift is Bg = B − WG . The structural center of mass and

center of buoyancy are located at distances XS & ZS and XB & ZB respectively from
72

envelope nose; αt is the equilibrium trim angle of attack, the angle GA-axis makes with

the horizontal wind V∞ . The confluence point distances from the envelope nose are XC

and ZC respectively. The combined structural and gas center of mass or center of mass

or center of gravity distances from the envelope nose are XG and ZG respectively. For the

present purpose, the envelope structure weight WS acting at the structure center of mass

and helium gas weight acting at the center of buoyancy is equivalent to the total weight

WB = WS + WG acting at the combined center of mass. The tether tension T1 acts at an

angle γ1 with the horizontal direction as also shown in Figure 3.1.

Summing the forces along and perpendicular direction of the wind we have,

X
Fh = 0 =⇒ D − T1 cos γ1 = 0 (3.8)

X
Fv = 0 =⇒ Bg + L − WS − T1 sin γ1 = 0 (3.9)

Using Equations (3.8) and (3.9), we have

T1 cos γ1 = D (3.10)

T1 sin γ1 = Bg + L − WS (3.11)

Taking moment about center of gravity G, we have

Mg = Mn + Fa ZG + Fn XG + (XG − XB ) B cos αt − (ZG − ZB ) B sin αt

− T1 sin γ1 [(XG − XC ) cos αt + (ZC − ZG ) sin αt ] (3.12)

+ T1 cos γ1 [(ZC − ZG ) cos αt − (XG − XC ) sin αt ]

Putting expressions from Equations (3.10) and (3.11) in Equation (3.12), we get the ex-

pressions for total pitching moment about center of gravity as below:

Mg = Mn + Fa ZG + Fn XG + (XG − XB ) B cos αt − (ZG − ZB ) B sin αt

− (Bg + L − WS ) [(XG − XC ) cos αt + (ZC − ZG ) sin αt ] (3.13)

+ D [(ZC − ZG ) cos αt − (XG − XC ) sin αt ]

For the equilibrium condition of the aerostat envelope, Mg = 0. Putting the values from

Equations (3.6) and (3.7) in Equation (3.13), the moment balance results in the following
73

equation:

Mn + (D cos αt − L sin αt ) ZG + (D sin αt + L cos αt ) XG

+ (XG − XB ) B cos αt − (ZG − ZB ) B sin αt


(3.14)
− (Bg + L − WS ) [(XG − XC ) cos αt + (ZC − ZG ) sin αt ]

+ D [(ZC − ZG ) cos αt − (XG − XC ) sin αt ] = 0

Equation (3.14) is non-linear in αt and can be solved using Newton Iteration Method [8].

Thus a plot of αt vs V∞ for equilibrium can be generated for a given aerostat envelope

configuration.

3.3.1 Moment Coefficient about CG

Equation (3.13) gives the total moment on aerostat envelope about the center of gravity.

In coefficient form, the moment may be written as:

1
Mg = ρa V∞ 2 Sc̄Cmg (3.15)
2

Using Equations (3.13), (3.14) and (3.15) and using q0 = ρa V∞ 2 /2, the moment coefficient

about center of gravity may be written as below:

ZG XG
Cmg = Cmn + (CD cos αt − CL sin αt ) + (CD sin αt + CL cos αt )
c̄ c̄
B
+ [(XG − XB ) cos αt − (ZG − ZB ) sin αt ]
q0 Sc̄
(3.16)
Bg + L − WS
− [(XG − XC ) cos αt + (ZC − ZG ) sin αt ]
q0 Sc̄
CD
+ [(ZC − ZG ) cos αt − (XG − XC ) sin αt ]

The moment coefficient about center of gravity in Equation (3.16) is plotted with angle of

attack in Figures 3.8 and 3.9 to examine the aerostat envelope equilibrium behaviour.

3.3.2 Static Longitudinal Stability

The moment coefficient about center of gravity was obtained in Equation (3.16) and the

moment about center of gravity was obtained in dimensional form in Equation (3.13). Any

of these two equations may be used to estimate trim angle of attack of aerostat envelope

under a given horizontal wind speed. It is also important to examine this aerostat envelope
74

for pitch static stability throughout the equilibrium range to ensure that it has a tendency

to return to its original equilibrium state following unforeseen disturbances. The criteria

for pitch stable equilibrium [13, 35] are as below:

dCmg
Cmg = 0 & <0 (3.17)

Differentiating Equation (3.16) with respect to α and simplifying after rearrangement, we

get the following expression:

dCmg dCmn ZG
= − (CL cos αt + CD sin αt + CLα sin αt − CDα cos αt )
dα dα c̄
XG
− (CL sin αt − CD cos αt − CLα cos αt − CDα sin αt )
c̄ 
B
− [(XG − XB ) sin αt + (ZG − ZB ) cos αt ]
q0 Sc̄
 
1 Bg − WS
+ CL + [(XG − XC ) sin αt − (ZC − ZG ) cos αt ] (3.18)
c̄ q0 S
1
− [(XG − XC ) cos αt + (ZC − ZG ) sin αt ] CLα

CD
− [(ZC − ZG ) sin αt + (XG − XC ) cos αt ]

1
+ [(ZC − ZG ) cos αt − (XG − XC ) sin αt ] CDα

Equation (3.17) is satisfied if trim angle is finite and the right side of Equation (3.18) turns

out to be less than zero thereby indicating pitch static stability. A plot of dCmg /dα vs.

V∞ is also generated in Figure 3.7 to examine the static stability in the entire operating

wind speed range.

3.3.3 Results for Aerostat Equilibrium and Static Stability Analysis

The method presented in this section has been applied to the test aerostat, which is a

medium size aerostat having hull volume of 2023 m3 with a operating height of 1000 m.

All the three fins of this aerostat are helium filled. This aerostat has been described in

section 1.5 of chapter 1. All the parameters required for carrying out the equilibrium and

static stability analysis for the test aerostat envelope including aerodynamic coefficients are

presented in Table 3.2, which have already been described in section 3.2. The air density

ρa and gross lift Bg correspond to ISA+15 deg temperature condition. The operating

wind speed range considered for this aerostat is 0 to 25 m/s.


75

Figures 3.6 and 3.7 present the results for trim angle of attack αt and dCmg /dα varia-

tion with wind speed for the test aerostat. The trim angle of attack was obtained by solving

Equation (3.14) using Newton Iteration Method [8]. The slope of moment coefficient curve

was obtained from Equation (3.18). Three different atmospheric temperature conditions

are considered for this aerostat as ISA, ISA+15 deg. and ISA+30 deg. It is observed

from Figure 3.6 that as operating temperature increases, angle of attack also increases.

The reason may be attributed to decreased buoyancy force and decreased resulting tether

tension and the combined effect produces less pitch down moment in Equation (3.13). It

is also observed in Figure 3.6 that as wind speed increases, the trim angle of attack also

increases however for higher temperatures the rate of increase is slower. For the considered

operating temperature range, the trim angle of attack for zero wind speed is from -2.3 deg

to 2.5 deg whereas for 25 m/s wind speed the trim angle of attack range is from 3.6 deg to

4.5 deg. Figure 3.7 indicates that dCmg /dα is negative over the entire range of wind speed

indicating static stability of this aerostat. Also it is observed that temperature variation

has minimal effect on the magnitude of dCmg /dα.

2
αt (deg)

−1

ISA
−2
ISA+15° C
ISA+30° C
−3
0 5 10 15 20 25
V∞ (m/s)

Figure 3.6: Variation of trim angle of attack with wind speed for test aerostat
76

−1

−2

−3
dCmg/dα (per radian)

−4

−5

−6

−7

−8
ISA
−9 ISA+15° C
ISA+30° C
−10
0 5 10 15 20 25
V∞ (m/s)

Figure 3.7: Variation of dCmg /dα with wind speed for test aerostat

The total moment coefficient about center of gravity is plotted with respect to angle of

attack in Figure 3.8 and Figure 3.9 for a wind speed of 5 m/s and 25 m/s respectively. It is

observed in these figures that as the angle of attack increases, the total moment coefficient

about center of gravity decreases and also crosses zero moment line. This indicates that

the aerostat envelope can be trimmed at a particular angle of attack and the system is

statically stable. The trim angles in Figure 3.8 and Figure 3.9 for the wind speed are

in agreement with the results in Figure 3.6. It is also observed that with increase in

temperature the moment curve shifts upward for the same angle of attack and the effect

of temperature change has smaller effects at higher wind speed.


77

V∞ = 5 m/s
0.4
ISA
ISA+15° C
0.3
ISA+30° C

0.2

0.1
Cmg

−0.1

−0.2

−0.3

−0.4
−10 −5 0 5 10
αt (deg)

Figure 3.8: Total moment coefficient with αt at 5 m/s wind speed for test aerostat

V∞ = 25 m/s
0.08
ISA
ISA+15° C
0.06 ISA+30° C

0.04
Cmg

0.02

−0.02

−0.04
−10 −5 0 5 10
αt (deg)

Figure 3.9: Total moment coefficient with αt at 25 m/s wind speed for test aerostat
78

3.3.4 Results for Sensitivity Analysis on Aerostat equilibrium

As the operating parameters of an aerostat envelope changes, the performance is liable

to be affected. The performance parameters considered in this section are trim angle

of attack and slope of total moment coefficient about center of mass or CG. Hence a

parametric sensitivity study has been carried out to estimate the effect of variation in

aerostat operating parameters on trim angle of attack and slope of total moment coefficient

about CG. A variation of nominal estimated values ±10% is considered and a uniform

temperature condition of ISA+15 deg is assumed for this purpose.

3.3.4.1 Center of Gravity (CG) Variation

The center of gravity of the aerostat is denoted by two distance components XG and ZG .

Keeping other parameters same as in Table 3.2, only XG and ZG were varied by ±10%

thereby considering a total of 04 cases. The results for trim angle of attack variation with

wind speed are presented in Figure 3.10 and slope of total pitching moment coefficient

variation about CG is presented in Figure 3.11. As observed in Figure 3.10, moving

CG forward towards nose results in large negative trim angle of attack and moving CG

backwards away from nose results in large positive trim angle of attack for low wind speed.

The trend for trim angle variation with wind speed reverses i.e. trim angle decreases with

increasing wind speed for +10% XG values as opposed to normal operating trend observed

in Figure 3.6. The effect of ZG variation on trim angle is observed only for low wind speed.

For the considered center of gravity range, the trim angle of attack for zero wind speed is

from -14.7 deg to 14.7 deg whereas for 25 m/s wind speed the trim angle of attack range

is from 1.4 deg to 6.6 deg. Also as observed in Figure 3.11, slope of total pitching moment

coefficient about CG does not vary much for the above variation in CG and static stability

is indicated for the entire variation range.


79

15

10

5
αt (deg)

−5
0.9XG & 0.9ZG
0.9XG & 1.1ZG
−10
1.1XG & 0.9ZG
1.1XG & 1.1ZG
−15
0 5 10 15 20 25
V∞ (m/s)

Figure 3.10: Trim angle with wind speed for CG sensitivity for test aerostat

−1

−2

−3
dCmg/dα (per radian)

−4

−5

−6

−7
0.9XG & 0.9ZG
−8 0.9XG & 1.1ZG
1.1XG & 0.9ZG
−9
1.1XG & 1.1ZG
−10
0 5 10 15 20 25
V∞ (m/s)

Figure 3.11: dCmg /dα with wind speed for CG sensitivity for test aerostat
80

3.3.4.2 Confluence Point (CP) Variation

The confluence point coordinates with respect to envelope nose are denoted by distance

components XC and ZC . Keeping other parameters same as in Table 3.2, only XC and

ZC were varied by ±10% thereby considering a total of 04 cases here also. The results

for trim angle of attack variation with wind speed are presented in Figure 3.12 and slope

of total pitching moment coefficient variation about CG is presented in Figure 3.13. As

observed in Figure 3.12, increasing both XC and ZC translates the trim angle vs. wind

speed curve upwards but the effect of XC is dominant as compared to ZC . For all the

cases, trim angle increases with increase in wind speed. For the considered confluence

point variation range, the trim angle of attack for zero wind speed is from -4.7 deg to

4.7 deg whereas for 25 m/s wind speed the trim angle of attack range is from 2.8 deg to

5.9 deg. Also as observed in Figure 3.13, slope of total pitching moment coefficient about

CG does not vary much for the above variation in confluence point and static stability is

indicated for the entire variation range.

2
αt (deg)

−2
0.9XC & 0.9ZC
0.9XC & 1.1ZC
−4
1.1XC & 0.9ZC
1.1XC & 1.1ZC
−6
0 5 10 15 20 25
V∞ (m/s)

Figure 3.12: Trim angle of attack with wind speed for CP sensitivity for test aerostat
81

−1

−2

−3
dCmg/dα (per radian)

−4

−5

−6

−7
0.9XC & 0.9ZC
−8 0.9XC & 1.1ZC
1.1XC & 0.9ZC
−9
1.1XC & 1.1ZC
−10
0 5 10 15 20 25
V∞ (m/s)

Figure 3.13: dCmg /dα with wind speed for CP sensitivity for test aerostat

In fact the sensitivity analysis for confluence point variation may be applied as a

method for determination of confluence point. As a thumb rule, the starting point of

confluence point XC and ZC is selected as the maximum diameter of the aerostat envelope

hull. After estimating other parameters, a sensitivity analysis is performed for confluence

point and for the desired variation of trim angle with wind speed; the relevant confluence

point may be selected.

3.3.4.3 Center of Buoyancy (CB) Variation

Similar parametric variation studies were conducted for center of buoyancy. The results

for trim angle of attack variation with wind speed are presented in Figure 3.14 and slope

of total pitching moment coefficient variation about CG is presented in Figure 3.15. It is

observed that moving CB towards nose results in large positive trim angle of attack and

moving CB away from nose results in large negative trim angle of attack for low wind

speed. This trend is opposite to that of CG variation. For 0.9XB , the trend for trim

angle variation with wind speed reverses i.e. trim angle decreases with increasing wind
82

speed. It is also observed that the effect of ZB is negligible on trim angle because center

of buoyancy is very close to center line of aerostat envelope. For the considered center of

buoyancy variation range, the trim angle of attack for zero wind speed is from -15.7 deg

to 15.7 deg whereas for 25 m/s wind speed the trim angle of attack range is from 1.1 deg

to 6.9 deg.As in the previous cases, envelope is statically stable in longitudinal direction

for the entire variation range and this variation has minimal effect on the slope of total

pitching moment curve about CG.

20

15

10

5
αt (deg)

−5

−10 0.9XB & 0.9ZB


0.9XB & 1.1ZB
−15 1.1XB & 0.9ZB
1.1XB & 1.1ZB
−20
0 5 10 15 20 25
V∞ (m/s)

Figure 3.14: Trim angle of attack with wind speed for CB sensitivity for test aerostat
83

−1

−2

−3
dCmg/dα (per radian)

−4

−5

−6

−7
0.9XB & 0.9ZB
−8 0.9XB & 1.1ZB
1.1XB & 0.9ZB
−9
1.1XB & 1.1ZB
−10
0 5 10 15 20 25
V∞ (m/s)

Figure 3.15: dCmg /dα with wind speed for CB sensitivity for test aerostat

3.3.4.4 Aerodynamic Coefficients Variation

The aerodynamic coefficients of concern presently are CL , CD and Cmn . A variation of

±10% of these coefficients is considered for sensitivity analysis. This analysis may also be

required to be carried out to account for possible uncertainty in measurement of data using

wind tunnel. To account for variation, a total of 08 cases are required to be considered.

Figures 3.16 and 3.17 present the result for this analysis. As expected, aerodynamic

coefficients variation has no effect for zero wind speed. For a wind speed of 25m/s, the

trim angle ranges from 3.0 deg. to 6.35 deg. This possible variation in trim angle needs

to be considered to account for possible uncertainty in aerodynamic coefficients.

It is noted during the sensitivity studies that varying the aerostat parameters, including

aerodynamic coefficients, only affects trim angle and has negligible effect on static stability

unlike that for aircraft. In fact putting extreme possible values for the parameters also

does not affect nature of static stability. The static stability appears to be shape dependent

for the test aerostat envelope.


84

14
1.1CL, 1.1CD & 1.1Cmn
12 1.1CL, 1.1CD & 0.9Cmn
1.1CL, 0.9CD & 1.1Cmn
10 1.1CL, 0.9CD & 0.9Cmn
0.9CL, 1.1CD & 1.1Cmn
8 0.9CL, 1.1CD & 0.9Cmn
0.9CL, 0.9CD & 1.1Cmn
αt (deg)

6 0.9CL, 0.9CD & 0.9Cmn

−2
0 5 10 15 20 25
V∞ (m/s)

Figure 3.16: αt with V∞ for aerodynamic coefficients sensitivity for test aerostat

−1

−2

−3
dCmg/dα (per radian)

−4
1.1CL, 1.1CD & 1.1Cmn
−5
1.1CL, 1.1CD & 0.9Cmn
−6 1.1CL, 0.9CD & 1.1Cmn
1.1CL, 0.9CD & 0.9Cmn
−7
0.9CL, 1.1CD & 1.1Cmn
−8 0.9CL, 1.1CD & 0.9Cmn
0.9CL, 0.9CD & 1.1Cmn
−9
0.9CL, 0.9CD & 0.9Cmn
−10
0 5 10 15 20 25
V∞ (m/s)

Figure 3.17: dCmg /dα with V∞ for aerodynamic coefficients sensitivity for test aerostat
85

3.3.5 Approximate Expressions

The expressions for equilibrium analysis have been derived earlier in this section consid-

ering all the terms. However, simpler expressions are possible if we make certain assump-

tions. We observe that ZB and ZG are small as compared to XB and XG and hence can

be neglected. If we neglect ZB and ZG in Equation 3.13, the resulting total moment about

center of gravity can be written as below:

Mg = Mn + Fn XG + (XG − XB ) B cos αt

− (Bg + L − WS ) [(XG − XC ) cos αt + ZC sin αt ] (3.19)

+ D [ZC cos αt − (XG − XC ) sin αt ]

If we further assume that αt is small and hence cos αt ≈ 1 and sin αt ≈ αt and Equa-

tion (3.19), after rearrangement, may be written as below:

Mg = Mn + [XC L − (XB − XC ) B + (XG − XC ) (WS + WG ) + ZC D]


(3.20)
+ [XC D − ZC B + ZC (WS + WG ) − ZC L] αt

Equation (3.20) may be written in coefficient form as

 
XC (XB − XC ) B (XG − XC ) (WS + WG ) ZC
Cmg = Cmn + CL − + + CD
c̄ q0 Sc̄ q0 Sc̄ c̄
  (3.21)
XC ZC B ZC (WS + WG ) ZC
+ CD − + − CL αt
c̄ q0 Sc̄ q0 Sc̄ c̄

Differentiating Equation (3.21) with respect to α and rearranging the terms, we have

dCmg dCmn ZC XC
= − (CL + CLα αt − CDα ) + (CD + CLα + CDα αt )
dα dα c̄ c̄
(3.22)
(B − WS − WG ) ZC

q0 Sc̄

Equations (3.21) and (3.22) provide the approximate expressions for the total moment

coefficient and the slope of total moment coefficient about center of gravity. It is evident

in Equation (3.22) that only the location of confluence point XC and ZC is present in the

expression for the slope of total moment coefficient about the center of gravity.

To verify the accuracy of the approximate expressions, the results for trim angle of

attack αt and dCmg /dα variation with wind speed for the test aerostat as presented in
86

section 3.3.3 and Figures 3.6 and 3.7 are regenerated for ISA+15 deg temperature condi-

tion. Figures 3.18 and 3.19 present the comparison between the complete expression and

approximate expression. It is evident from these figures that approximate expressions can

be applied to carry out the equilibrium analysis of a tethered aerostat with reasonable

accuracy.

2
αt (deg)

−1

−2
Full expression
Approximation
−3
0 5 10 15 20 25
V∞ (m/s)

Figure 3.18: Trim angle with V∞ for approximate expression for test aerostat

−1

−2

−3
dCmg/dα (per radian)

−4

−5

−6

−7

−8

−9 Full expression
Approximation
−10
0 5 10 15 20 25
V∞ (m/s)

Figure 3.19: dCmg /dα with V∞ for approximate expression for test aerostat
87

3.4 Equilibrium Cable Configuration

During its operation, an aerostat encounters wind which causes its tether to take a par-

ticular shape. It is of vital importance to estimate the tether cable shape and tension

variation along its length under the operating wind conditions. Commendable cable mod-

eling for uniform wind speed and air density along the vertical direction to estimate cable

shape and tension has been done in reference [36, 38]. The method presented in [36] has

been applied to a small aerostat to estimate cable parameters in [42]. For the case where

velocity magnitude varies with vertical position, Berteaux [6] suggested an approximate

step-wise change of velocity with vertical position. An optimization technique has also

been proposed [30] for accounting velocity variation. In this section, Neumark’s method

[36] has been used to express equilibrium cable parameters for aerostat assuming uniform

wind speed and density followed by derivation of two types of expressions for polygonal

approximations. As is the case with medium and large size aerostats, wind speed and den-

sity variations with height cannot be neglected. Since the tether can be divided into finite

number of elements from anchor point to confluence point, wind speed and air density

variations with height can be easily accounted for in polygonal approximations. Once, the

tether tension and angle at the confluence point are known, the method proceeds down-

ward with approximation of these parameters for the next lower element. An error analysis

has also been carried out for the case of uniform wind speed and density for the small

aerostat. The polygonal approximations are then applied for the test aerostat considering

the wind speed and density variations [24].

3.4.1 Cable Configuration for Uniform Wind Speed and Density

Once the tether tension and angle with horizontal are known at the confluence point, it is

required to estimate these parameters at the anchor point and also the cable shape from

confluence point to anchor point. The relevant expressions are presented here for the case

of uniform wind speed and air density as adapted from [36, 42]. The coordinate system

and the forces acting on the cable are shown in Figure 3.20. The small tether cable element

as shown in Figure 3.20 is acted upon by the tether tensions (T and T + dT ), gravity force

(wc dl) and drag force which is significant only in the direction perpendicular to the cable

and the drag along the cable may be neglected.


88

Figure 3.20: The coordinate system and forces acting on the tether cable [42]

The wind speed perpendicular to the cable is Vn = V∞ sin γ and hence the cable drag

per unit length Pn may be written as CDc dc q0 sin2 γ where CDc is the cable drag coefficient,

dc is the cable diameter and q0 is the dynamic pressure. For the small element dl, the

drag force normal to the cable is approximated as:

1
Pn dl = nd dl sin2 γ where 2
nd = CDc dc ρa V∞
2

Here nd is the tether cable drag per unit length for a cable normal to the wind. Balancing

the forces in horizontal and vertical direction and neglecting higher order terms gives the

following approximate expressions [42]:

dT = wc dl sin γ (3.23)

T dγ = nd sin2 γ + wc cos γ dl

(3.24)

Using Equations (3.23) and (3.24), the tether tension at the element location may be
89

written as [42]:
T1 τ
T = (3.25)
τ1

where,
 p/ q
q + p − cos γ
τ (γ) = (3.26)
q − p + cos γ
p
where, p = wc /2nd and q = 1 + (p)2 . Substituting Equations (3.25) and (3.26) in

Equation (3.24) and integrating from the lower end to upper end gives [42]:

T1 τ
dl = dγ (3.27)
nd τ1 sin2 γ + 2p cos γ

T1 
l= λ1 − λ0 (3.28)
nd τ1

where,
γ τ (γ)
λ (γ) = ∫ 2 dγ (3.29)
0 sin γ + 2p cos γ

Angle γ0 is obtained by solving Equation (3.28) for λ0 = λ1 − nd τ1 l/T1 and using this

value with Equation (3.29) to obtain

γ0 τ (γ)
λ0 = ∫ 2 dγ (3.30)
0 sin γ + 2p cos γ

The above equation is solved for the unknown limit of integration γ0 by Newton iteration

method [8]. With γ0 known, Equations (3.25) and (3.26) are used to find

τ0
T0 = T1 (3.31)
τ1

where, τ0 = τ (γ0 ). From Figure 3.20, we have dx̃ = dl cos γ. Using Equation (3.27) in this

expression yields
T1
dx̃ = dσ (3.32)
nd τ1

where, dσ = τ cos γdγ/(sin2 γ + 2p cos γ). Equation (3.32) may be integrated numerically

to give [42]:
T1 γ1
x̃1 = ∫ dσ (3.33)
nd τ1 γ0
90

Finally, from Figure 3.20 and Equation (3.23)

dT
dz̃ = dl sin γ = (3.34)
wc

which is integrated to give


T1 − T0
z̃1 = (3.35)
wc

3.4.2 Approximation for Wind Speed and Air Density Variation

The expressions presented in the previous section for tether cable tension, angle with

horizontal and profile are valid only for constant wind speed and density with height

and hence are applicable only for the cases where these variations are negligible e.g. for

a small size aerostat due to low height of operation. Hence it is required to develop

relations which are applicable to variable wind speed and density with height. To account

for such variations, Berteaux [6] suggested an approximate step-wise change of velocity

with vertical position. Following this assumption, expressions for two types of polygonal

approximation have been developed in this section. The basic idea is to represent the

cable through finite number of elements and assume constant wind speed and density over

a particular element. Using the equilibrium of a particular cable element, cable tension

and angle with horizontal for the next cable element is estimated in terms of current

element parameters. If we start from the confluence point of the cable at the top, the

tether tension and angle are known using Equations (3.1) and (3.2) and hence the cable

tension and angle for the second element can be estimated and so on.

Figure 3.21 shows the straight finite tether cable length ∆l, equilibrium of this cable

element and the forces acting on it. This cable element makes the angle δn with the

horizontal direction. As discussed earlier, the forces on this cable element are tension on

the upper and lower ends, its own weight and drag force. Balancing the forces in horizontal

and vertical directions on this tether cable element in Figure 3.21 which represents nth

element, we get

Pn ∆l sin δn + Tn cos γn − Tn+1 cos γn+1 = 0 (3.36)

Tn sin γn − Tn+1 sin γn+1 − Pn ∆l cos δn − wc ∆l = 0 (3.37)

Based on angle δn , two types of approximations are proposed as below:


91

Figure 3.21: Polygonal approximation of tether cable element

3.4.2.1 Polygonal Approximation 1

In this case, it is assumed that δn is the mean of angle γn and γn+1 i.e. δn = (γn +γn+1 )/2.

Then, using Equations (3.36) and (3.37), we have

   
2 sin2 γn +γn+1 ∆l cos γn +γn+1
Tn sin γn − wc ∆l − 0.5CDc dc ρa V∞ 2 2
tan γn+1 =     (3.38)
2 sin2 γn +γn+1 ∆l sin γn +γn+1
Tn cos γn + 0.5CDc dc ρa V∞ 2 2

Since γn+1 is the only unknown parameter here, Equation (3.38) can be solved for γn+1

using Newton iteration method [8]. Then using Equation (3.36), tension in the lower

element can be estimated as below:


   
2 sin2 γn +γn+1 γn +γn+1
Tn cos γn + 0.5CDc dc ρa V∞ 2 ∆l sin 2
Tn+1 = (3.39)
cos γn+1

Using Equations (3.38) and (3.39), tension and angle for the next lower element can be

approximated, thereby evaluating these values for all the lower elements. Summation of

x- and z- components of element length provides x- and z- distance.

3.4.2.2 Polygonal Approximation 2

Further simplification can be achieved and tension and angle for the next element can di-

rectly be written in terms of previous values if δn = γn is assumed. Under this assumption,


92

using Equations (3.36) and (3.37), we have

2 sin2 γ

Tn sin γn − wc ∆l − 0.5CDc dc ρa V∞ n ∆l cos γn
tan γn+1 = 2 sin2 γ
 (3.40)
Tn cos γn + 0.5CDc dc ρa V∞ n ∆l sin γn

Hence, γn+1 is directly obtained as all the parameters in the right side of Equation (3.40)

are known for current element. Further, using Equation (3.36), we get

2 sin2 γ ∆l sin γ
Tn cos γn + 0.5CDc dc ρa V∞ n n
Tn+1 = (3.41)
cos γn+1

Using Equations (3.40) and (3.41), tension and angle for the next lower element can be

directly written in terms of previous values. As for the previous case, summation of x- and

z- components of element length provides x- and z- distance. It is clear that the expressions

for Polygonal Approximation 2 are simpler. As will be presented in the results, for large

number of elements both approximations give similar results.

It is noted that the wind speed V∞ and air density ρa for the elements are required in

steps and hence if the variation of these parameters with height are known beforehand

then that can be accommodated in the expressions for polygonal approximations easily.

3.4.3 Results on Equilibrium Cable Configuration

The results for equilibrium cable configuration are presented for the two aerostats namely

small aerostat and test aerostat as described in section 1.5 of chapter 1. For small aero-

stat, since operating height is small (100 m), a constant air density has been assumed

from ground to operational height. But for the test aerostat, air density variation with

height, as presented in Equations (1.4-1.5) and (1.7), has been taken into account while

using Polygonal Approximation 1 (PA-1) and Polygonal Approximation 2 (PA-2). For

the small and test aerostat parameters, the tether tension and angle with horizontal are

evaluated using Equation (3.1) and Equation (3.2) and are presented in Table 3.3. The

drag coefficient of tether cable element is assumed as 1.17 [16] for the flow considered.

The required tether parameters of the two aerostats are presented in Table 3.4.

Table 3.4: Tether parameters for the small and test aerostat
Parameters Small aerostat Test aerostat
Tether unit mass (kg/m) 0.1 0.3
Tether diameter (mm) 10 17
93

First of all, comparison of Neumark’s method [36] is done with Polygonal Approxi-

mations for varying number of elements. Since a constant density is assumed in Neu-

mark’s method, this comparison is done only for small aerostat with 20 m/s uniform wind

speed. The results are presented in Figure 3.22 to Figure 3.28. Figures 3.22 to 3.24

present the comparison of Neumark’s method with Polygonal Approximation 1 (PA-1)

for non-dimensional position, tether tension and angle with horizontal. Similar results

for Polygonal Approximation 2 (PA-2) have been presented in Figure 3.25 to Figure 3.27.

The percentage errors for these two approximations with increasing number of elements

have been plotted in Figure 3.28. These plots suggest that for large number of elements

both approximations give similar results. But for small number of elements, angle with

horizontal is better estimated using Polygonal Approximation 1 whereas tether tension is

better estimated using Polygonal Approximation 2. Also the percentage error for position

estimation becomes less than 1% if 50 or more elements are used and hence for these many

numbers of elements, both the approximations can be used to generate practical results.

0.8

0.6
z/l

0.4

N=2
0.2 N=5
N=10
N=50
Neumark

0.1 0.2 0.3


x/l

Figure 3.22: Non-dimensional cable profile for small aerostat using PA-1 & V∞ = 20 m/s
94

0.99

0.98
T /T1

0.97

N=2
0.96 N=5
N=10
N=50
Neumark
0.95
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.23: Non-dimensional cable tension for small aerostat using PA-1 & V∞ = 20 m/s

0.95
γ/γ1

0.9

N=2
N=5
N=10
N=50
Neumark
0.85
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.24: Non-dimensional cable angle for small aerostat using PA-1 & V∞ = 20 m/s
95

0.8

0.6
z/l

0.4

N=2
0.2 N=5
N=10
N=50
Neumark

0.1 0.2 0.3


x/l

Figure 3.25: Non-dimensional cable profile for small aerostat using PA-2 & V∞ = 20 m/s

0.99

0.98
T /T1

0.97

N=2
0.96 N=5
N=10
N=50
Neumark
0.95
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.26: Non-dimensional cable tension for small aerostat using PA-2 & V∞ = 20 m/s
96

0.95
γ/γ1

0.9

N=2
N=5
N=10
N=50
Neumark
0.85
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.27: Non-dimensional cable angle for small aerostat using PA-2 & V∞ = 20 m/s

Polygonal approximation 1
15
x
z
10 T
% error

γ
5

0
10 20 30 40 50

Polygonal approximation 2
15
x
z
10 T
% error

γ
5

0
10 20 30 40 50
No. of elements

Figure 3.28: Percentage error with number of elements for small aerostat for V∞ = 20 m/s
97

Figures 3.29 to 3.31 show the equilibrium tether cable non-dimensional position, ten-

sion and angle of small aerostat for uniform wind speed varying from 5 m/s to 25 m/s.

Since the air density variation with height has been neglected in this case, Neumark’s

method [36] has been used. As can be seen from the figures, the operating height de-

creases and blow by increases with the increasing wind speed as expected. For a wind

speed of 25 m/s, the operating height decreases to 96.3 m and blow by increases to 26.8

m. Also with increase in wind speed, tether tension increases and angle with horizontal

decreases as expected.

0.8

0.6
z/l

0.4

5 m/s
0.2 10 m/s
15 m/s
20 m/s
25 m/s

0.1 0.2 0.3 0.4


x/l

Figure 3.29: Non-dimensional cable profiles for small aerostat with different V∞
98

0.95
T /T1

0.9

5 m/s
10 m/s
15 m/s
20 m/s
25 m/s
0.85
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.30: Non-dimensional cable tensions for small aerostat with different V∞

0.95
γ/γ1

0.9

5 m/s
10 m/s
15 m/s
20 m/s
25 m/s
0.85
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.31: Non-dimensional cable angles for small aerostat with different V∞
99

Figures 3.32 to 3.34 show the equilibrium tether cable non-dimensional position, ten-

sion and angle of test aerostat for uniform wind speed varying from 5 m/s to 25 m/s.

Since the air density variation with height has been taken into account, Polygonal Ap-

proximation 2 has been used for calculations with 100 elements. Here also, the operating

height decreases and blow by increases with the increasing wind speed as expected. For

a wind speed of 25 m/s, the operating height decreases to 908.8 m and blow by increases

to 388.5 m. If it is required to maintain height of 1000 m, as is generally the case for full

coverage, the tether non-dimensional position is shown in Figure 3.35. The blow by has

increased here to 466 m for 25 m/s wind speed. In this case it is required to release extra

tether of about 120 m to maintain height and hence extra tether weight has to be taken

into account for payload calculations of aerostat.

0.8

0.6
z/l

0.4

5 m/s
0.2 10 m/s
15 m/s
20 m/s
25 m/s

0.1 0.2 0.3 0.4 0.5


x/l

Figure 3.32: Non-dimensional cable profiles for test aerostat with different V∞
100

0.9

0.8
T /T1

0.7

5 m/s
0.6 10 m/s
15 m/s
20 m/s
25 m/s
0.5
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.33: Non-dimensional cable tensions for test aerostat with different V∞

0.9

0.8
γ/γ1

0.7

5 m/s
0.6 10 m/s
15 m/s
20 m/s
25 m/s
0.5
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.34: Non-dimensional cable angles for test aerostat with different V∞
101

0.8

0.6
z/l

0.4

5 m/s
0.2 10 m/s
15 m/s
20 m/s
25 m/s

0.1 0.2 0.3 0.4 0.5 0.6


x/l

Figure 3.35: Non-dimensional cable profiles for test aerostat with fixed operational height

Finally the case of both wind speed and air density variation with height is considered

where the numerical method is most suited. Figure 3.36 shows a typical measured wind

speed with height in a coastal region. The wind speed varies from about 1 m/s near surface

to about 17 m/s at a height of 1000 m as per pattern shown. Polygonal Approximation 2

has been used with 500 elements to estimate the tether cable position, tension and angle

and the result is presented in Figures 3.37 to 3.39. Here the operating height is reduced

to 968 m with an increase in blow by to 238.2 m.


102

20

16
Wind speed (m/s)

12

200 400 600 800 1000


Height (m)

Figure 3.36: A typical measured wind speed with height

0.8

0.6
z/l

0.4

0.2

0.1 0.2 0.3


x/l

Figure 3.37: Non-dimensional cable profile for test aerostat for the measured wind speed
103

0.9
T /T1

0.8

0.7
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.38: Non-dimensional cable tension for test aerostat for the measured wind speed

0.95

0.9
γ/γ1

0.85

0.8

0.75
0.2 0.4 0.6 0.8 1
lc /l

Figure 3.39: Non-dimensional cable angle for test aerostat for the measured wind speed
104

3.5 Chapter Summary

The equilibrium and static analysis has been carried out in this chapter. The tether

tension and angle with horizontal expression has been derived from the free body diagram

of the aerostat. The results for tether tension have been presented for two aerostat namely

small aerostat and test aerostat. After examining the equilibrium of the aerostat under

various forces, expressions have been developed for estimation of trim angle. From the

total moment coefficient expression, the slope of moment curve with respect to angle of

attack has been developed for analyzing static stability. The total moment coefficient

has also been plotted as function of angle of attack. Equilibrium cable configuration

is then estimated first using the theoretical method and then two approximations have

been proposed. The approximations are validated with the theoretical approach for small

aerostat. For test aerostat the theoretical method is not suitable and hence approximation

developed has been applied. The results indicate that the aerostat envelope is statically

stable in the operating range and that the numerical approximations presented are suitable

for the test aerostat.


Chapter 4

Dynamic Stability Analysis of the


Tethered Aerostat

4.1 Introduction

The analysis carried in chapter 3 primarily focused on evaluation of equilibrium behavior

and static performance of the tethered aerostat envelope and that too only in longitudinal

direction. However, to predict the dynamic stability behavior of the tethered aerostat

balloon, a generalized six degrees of freedom dynamic modeling approach for the envelope

and a suitable dynamic model for tether will be required. Unlike an aircraft, the model

for tethered aerostat also requires to consider the forces and moments due to buoyancy,

apparent mass and tether. The buoyancy and apparent mass can be easily taken into ac-

count in the generalized six degrees of freedom equations whereas tether modeling requires

additional assumptions and also adds further to complexity.

Fortunately, like an aircraft, the equations of motion for the aerostat can be decoupled

into two independent sets of equations which can be dealt independently if small perturba-

tions assumptions are invoked about equilibrium position. This type of dynamic stability

analysis has been carried out by Redd et al. [42] which forms a benchmark for carrying

out dynamic stability analysis using small perturbations assumptions. But the analysis

presented in [42] is applicable only for small aerostats flying at low altitude, primarily due

to limiting assumption of a uniform wind speed and density from the ground to the the fly-

ing altitude. The flying altitude for the present test aerostat is 1 km from the ground and

hence the assumption of uniform wind speed and density from the ground to the altitude

105
106

is far from reality. To account for steady wind speed variation along the vertical direction,

Berteaux [6] proposed stepwise approximation of wind speed which has been applied in

chapter 3 to derive expressions for polygonal approximations for aerostat. If the results

of direct integration are replaced by polygonal approximation in the method of Redd et

al. [42], the method becomes suitable to be applied for wind speed as well as air density

variation with altitude. The present chapter follows this approach for dynamic stability

analysis of tethered aerostat for small perturbations. The linear equations of motion are

derived by taking suitable assumptions into account. The equations are then solved to

carry out dynamic stability analysis of tethered aerostat. As the characteristic roots are

available, the response studies are then carried out for initial disturbances. Although the

results have been presented for uniform horizontal wind speed from the ground to the

altitude, the method can also be used for steady wind which varies with altitude.

4.2 Description of Aerostat System

The schematic of aerostat system for dynamic stability analysis is presented in Figure 4.1.

The inertial axes system is fixed at the center of mass of aerostat envelope for the steady

wind conditions and defined by three right-hand orthogonal unit vectors ii , j i and ki .

Two body fixed coordinate systems are defined, namely the body axes system with unit

vectors ib , j b and kb and the stability axes system with unit vectors is , j s and ks . The

ib unit vector of body axes system is parallel to the central axis of the aerostat envelope

and points towards nose whereas the is unit vector of stability axes system is parallel to

the horizontal for steady wind conditions and points forward. It comes out that for steady

wind conditions the inertial and stability axes systems are coincident.

The forces acting on the aerostat envelope include buoyancy force, gravity force, aero-

dynamic force and tether cable force. The buoyancy force combined with the helium gas

weight Bg acts upwards at the center of buoyancy. The aerostat envelope structure weight

WS acts downward at the structure center of mass. The center of mass of aerostat enve-

lope is the combined center of mass of envelope structure and helium gas. The equations

of motion for the present aerostat envelope have been written about the envelope center

of mass. To have a definite and predictable tether force, it is assumed that the aerostat

envelope and confluence lines form a rigid system and hence rigid body equations can
107

CB
CM

SCM

Figure 4.1: Tethered aerostat schematic considered in dynamic stability analysis [42]

be applied to it. Although the aerostat envelope materials are mostly fabric and hence

flexible but under the pressurized conditions, rigidity is ensured. A constant horizontal

wind speed V∞ has been assumed during the analysis with the small perturbations about

equilibrium position. The tether cable is assumed flexible and inextensible.

4.3 Rigid Body Equations of Motion

As mentioned in section 4.2 that the aerostat envelope and confluence lines form a rigid

system and thus rigid body equations of motion are applicable for it. Thus the equations

developed for an aircraft and modified for an aerostat envelope may be used for the

purpose. The rigid body equations of motion are obtained from Newton’s second law,

which states that the summation of all external forces acting on a body is equal to the time
108

rate of change of the momentum of the body, and the summation of the external moments

acting on the body is equal to the time rate of change of the angular momentum. The time

rates of change of linear and angular momentum are referred to an absolute or inertial

reference frame. The basic linear momentum and angular momentum equations from

Newton’s second law, after applying to a small mass element of an aerostat envelope and

then integrating over the whole body, can be expressed in the following vector equations

[13, 35]:
X d
F = (mV E ) (4.1)
dt
X d
M= (H) (4.2)
dt
P P
where F is the vector summation of the external applied forces, M the vector sum-

mation of the external applied moments about the center of gravity, V E the velocity

vector and H the angular momentum vector about the center of gravity. Non-rotating

coordinates are unwarranted because the inertia tensor is a rapidly changing function of

time in the non-rotating axis system and the measurements are made primarily in the

rotating body-fixed axis system.

The Equations (4.1) and (4.2) are transferred to the rotating aerostat envelope balloon-

fixed axis system. If Ω be the angular velocity vector of the envelope-fixed axis system

with respect to the inertial reference frame, the rules for transforming vector derivatives

into rotating coordinate system give:

X δ
F = (mV E ) + Ω × (mV E ) (4.3)
δt

X δ
M= (H) + Ω × H (4.4)
δt

where all quantities are in the rotating body-axis system. The δ/δt operator denotes the

vector of derivatives of the vector components; in a rotating axis system this is not the

same as the derivative of the vector, because the components in the rotating axis system

change due to this rotation even if the vector is constant in inertial space [37].
109

The angular momentum is given by the following expression:


 
 Ix −Ixy −Ixz 
 
H=
−Ixy Iy Ω
−Iyz  (4.5)
 
−Ixz −Iyz Iz

The matrix in Equation (4.5) is the inertia tensor [13] expressed in the balloon or envelope-

fixed axes system.

4.4 Aerostat Equations of Motion

The aerostat envelope equations of motion in the stability axes system are presented in

the present section. The components of Ω in the stability axes system are Ps , Qs , Rs

by definition. The components of V E in the stability axes system are Us , Vs , Ws by

definition. The position of the center of mass in the inertial axes system is represented by

xi , yi and zi . For dynamic modeling of aerostat envelope, we can neglect time derivatives

of mass and the inertia tensor. With these substitutions, the force Equations (4.3) can be

written in scalar form in the stability axes system as:

 
FX = m U̇s + Qs Ws − Rs Vs (4.6a)
 
FY = m V̇s + Rs Us − Ps Ws (4.6b)
 
FZ = m Ẇs + Ps Vs − Qs Us (4.6c)

and the moment Equations (4.4) can be written in scalar form in the stability axes system

as:

MX = Ixs Ṗs − Ixz s Ṙs + Qs Rs (Izs − Iys ) − Ixz s Ps Qs (4.7a)

MY = Iys Q̇s + Rs Ps (Ixs − Izs ) + Ixz s Ps 2 − Rs 2



(4.7b)

MZ = −Ixz s Ṗs + Izs Ṙs + Ps Qs (Iys − Ixs ) + Ixz s Qs Rs (4.7c)

The components of the force and moment acting on the aerostat envelope are composed

of aerodynamic, gravitational, buoyancy and tether cable contributions. Also, since xz is

the plane of symmetry, we have Ixys = Iyz s = 0.


110

Figure 4.2 presents the orientation of the stability axes system with respect to inertial axes

system. The stability axes system is obtained by three consecutive body fixed rotations

about k, j and i axis by angle Ψs , Θs and Φs respectively, starting from inertial axes.

Figure 4.2: Orientation of stability axis system [13]

The relationship between the components of angular velocity in the stability axes
h iT
system [Ps Qs Rs ]T and the Euler rates Ψ̇s Θ̇s Φ̇s can also be written as below

[35]:     
P
 s   1 0 − sin Θs  Φ̇ s 
    
 Q  = 0 cos Φ cos Θs sin Φs  (4.8)
  Θ̇s 
 
 s   s
    
Rs 0 − sin Φs cos Θs cos Φs Ψ̇s

The expression for the velocity of aerostat envelope center of mass in terms of the Euler

angles and velocity components in the stability axes system has also been presented in

[35] and is as below (Figure 4.2):

    
 ẋi  CΘs CΨs SΦs SΘs CΨs − CΦs SΨs CΦs SΘs CΨs + SΦs SΨs   Us 
    
 ẏ  =  C S SΦs SΘs SΨs + CΦs CΨs CΦs SΘs SΨs − SΦs CΨs  (4.9)
  Vs 
 
 i   Θs Ψs
    
żi −SΘs SΦs CΘs CΦs CΘs Ws
111

In Equation (4.9), CΘs = cos Θs , SΘs = sin Θs and so on. The transformation matrix

in Equation (4.9) is represented by TSI which is for transformation from stability axes

system to inertial axes system. The transformation matrix from the inertial axes system

to stability axes system TIS can be obtained by inverting the matrix TSI and is written as

below:  
 CΘs CΨs CΘs SΨs −SΘs 
TIS = 
 
SΦs SΘs CΨs − CΦs SΨs SΦs SΘs SΨs + CΦs CΨs SΦs CΘs 
 (4.10)
 
CΦs SΘs CΨs + SΦs SΨs CΦs SΘs SΨs − SΦs CΨs CΦs CΘs

The notations for stability axes linear velocity components, angular velocity components

and Euler angles are denoted by uppercase letters for the complete motion of aerostat

envelope. The corresponding values for small perturbations will be denoted by lowercase

letters. Equations (4.6) and (4.7) represent the six degrees of freedom dynamic equations

of motion and Equations (4.8) and (4.9) represent kinematic relations for an aerostat

envelope in stability axes system. It is mentioned in section 4.2 that the inertial axes

system is chosen such that the envelope center of mass is the origin under the steady

horizontal wind condition and hence having no initial linear or angular velocity. This is in

contrast with the conventional airplane which moves with a constant initial velocity. Also

the stability axes system is coincident with the inertial axes system initially under steady

wind condition. Hence the total motion of center of mass can simply be written in terms

of small perturbations as below [13]:

Us = us ; V s = vs ; W s = ws (4.11a)

Θs = θs ; Ψs = ψs ; Φs = φs (4.11b)

Ps = ps ; Qs = qs ; Rs = rs (4.11c)

Putting the simplified expressions obtained in Equation (4.11) in the kinematic relations

for the aerostat envelope presented in Equations (4.8) and (4.9) and retaining only first-

order perturbation terms gives:

ps = φ̇s ; qs = θ̇s ; rs = ψ̇s (4.12)

ẋi = us ; ẏi = vs ; żi = ws (4.13)


112

It is now required to simplify the force and moment equations. Substituting Equa-

tions (4.11-4.12) and (4.13) into Equations (4.6) and (4.7), and neglecting higher order

perturbation terms after retaining only first-order terms, results in the following linear

translational equations of motion:

FX = mẍi (4.14a)

FY = mÿi (4.14b)

FZ = mz̈i (4.14c)

and, the rotational equations of motion become

MX = Ixs φ̈s − Ixz s ψ̈s (4.15a)

MY = Iys θ̈s (4.15b)

MZ = −Ixz s φ̈s + Izs ψ̈s (4.15c)

As mentioned earlier, the force and moment terms in Equations (4.14) and (4.15) are

composed of gravity, buoyancy, aerodynamic and tether cable forces.

4.4.1 Apparent Masses and Moments of Inertia

When the aerostat envelope accelerates, a certain amount of surrounding air also accel-

erates with it which has also to be taken into account into the equations of motion. The

method to calculate apparent mass and inertia for the body axes system has been pre-

sented in section 2.3 of chapter 2. The apparent masses for the body axes system are

represented by mxb , myb and mzb . The apparent masses for the stability axes system are

obtained by the following relation [42]:

mxs = mxb cos2 αt + mzb sin2 αt (4.16a)

mys = myb (4.16b)

mzs = mxb sin2 αt + mzb cos2 αt (4.16c)

The total mass of the aerostat envelope consists of the envelope mass including helium gas

mass inside the envelope mb and the apparent masses defined in Equations (4.16). Thus
113

we have total masses mxo , myo and mzo for motions along stability axes as defined below:

mxo = mb + mxs (4.17a)

myo = mb + mys (4.17b)

mzo = mb + mzs (4.17c)

The mass m in Equations (4.14) has to be replaced by the mass terms defined in Equa-

tions (4.17) for motion along the stability axes. The mass moments of inertia were obtained

in section 2.2.4 of chapter 2 included the contributions from structural mass, helium gas

and apparent mass terms and are expressed in the body axes system as Ixb , Iyb , Izb and

Ixz b . These moments of inertia have to be transformed to stability axes system represented

by Ixs , Iys , Izs and Ixz s for use in Equations (4.15) using the transformation relation pre-

sented in [13] as below:

 
 I xs 0 −Ixz s 
 
 0 Iys 0 
 
 
−Ixz s 0 Izs
    (4.18)
 cos α t 0 sin α t   Ixb 0 −Ixz b  cos αt 0 − sin αt 
   
= 0 1 0   0
 Iyb 0   0
 1 0  
   
− sin αt 0 cos αt −Ixz b 0 I zb sin αt 0 cos αt

Thus the force Equations (4.14) after substituting the total masses defined in Equa-

tions (4.17) and considering the terms from aerodynamic, tether cable, buoyancy and

gravity become:

FXA + FXC + FXB + FXG = mxo ẍi (4.19a)

FYA + FYC + FYB + FYG = myo ÿi (4.19b)

FZA + FZC + FZB + FZG = mzo z̈i (4.19c)

and the moment Equations (4.15) take the form

MXA + MXC + MXB + MXG = Ixs φ̈s − Ixz s ψ̈s (4.20a)


114

MYA + MYC + MYB + MYG = Iys θ̈s (4.20b)

MZA + MZC + MZB + MZG = −Ixz s φ̈s + Izs ψ̈s (4.20c)

The above force and moment Equations (4.19) and (4.20) have been written in the stability

axes system. The subscripts A, C, B, and G correspond to aerodynamic, tether-cable,

buoyancy, and gravity terms, respectively.

4.4.2 Aerodynamic Forces and Moments

The aerodynamic forces and moments consist of contributions from steady state trim con-

dition and small perturbations about trim condition. These forces and moments about

the aerostat envelope center of mass are derived in [42] with certain assumptions. Some

important assumptions include the independence of longitudinal and lateral forces and

moments and zero lateral forces and moments for equilibrium trim condition. The pertur-

bation velocity of aerostat envelope with respect to airstream can be written as below:
         
u ∗ u
 s   s   −V∞   us + V∞   V∞ 
 v ∗  =  v  − TIS  0
         
= v −ψ V  ≈  ẏ − ψ V  (4.21)
 s   s     s s ∞   i s ∞ 
         
ws∗ ws 0 ws + θ s V ∞ żi + θs V∞

The perturbation angle of attack α and sideslip angle β can be approximated as [13]:

w∗ żi v∗ ẏi
α= = + θs and β = = − ψs
V∞ V∞ V∞ V∞

Following the similar procedure as explained in [13, 42], the aerodynamic forces and mo-

ments in non-dimensional form are written as below [42]:

   
ρa V∞ S ρa V∞ S
FXA =− (2CD + CDu ) ẋi − (CDα − CL ) żi
2 2
ρa V∞ 2 S ρa V∞ 2 S
 
− (CDα − CL ) θs − CD (4.22a)
2 2
     
ρa Sc̄ ρa V∞ S ρa V∞ Sc̄
FYA = CYβ̇ ÿi + CYβ ẏi + CYp φ̇s
4 2 4
ρa V∞ 2 S
   
ρa V∞ Sc̄ 
− CYβ̇ − CYr ψ̇s − CYβ ψs (4.22b)
4 2
   
ρa V∞ S ρa Sc̄
FZA =− (2CL + CLu ) ẋi − CLα̇ z̈i
2 4
115

   
ρa V∞ S ρa V∞ Sc̄ 
− (CLα + CD ) żi − CLα̇ + CLq θ̇s
2 4
ρa V∞ 2 S ρa V∞ 2 S
 
− (CLα + CD ) θs − CL (4.22c)
2 2
" # " #
ρa S (c̄)2 ρa V∞ S (c̄)2
 
ρa V∞ Sc̄
MXA = Clβ̇ ÿi + Clβ ẏi + Clp φ̇s
4 4 4
" #
ρa V∞ S (c̄)2  ρa V∞ 2 Sc̄
  
− Clβ̇ − Clr ψ̇s − Clβ ψs (4.22d)
4 2
" #
ρa S (c̄)2
 
ρa V∞ Sc̄
M YA = (2Cm + Cmu ) ẋi + Cmα̇ z̈i
2 4
" #
ρa V∞ S (c̄)2
 
ρa V∞ Sc̄ 
+ Cmα żi + Cmα̇ + Cmq θ̇s
2 4
ρa V∞ 2 Sc̄ ρa V∞ 2 Sc̄
 
+ Cm α θs + Cm (4.22e)
2 2
" # " #
ρa S (c̄)2 ρa V∞ S (c̄)2
 
ρa V∞ Sc̄
MZA = Cnβ̇ ÿi + Cnβ ẏi + Cnp φ̇s
4 2 4
" #
ρa V∞ S (c̄)2  ρa V∞ 2 Sc̄
  
− Cnβ̇ − Cnr ψ̇s − Cnβ ψs (4.22f)
4 2

4.4.3 Tether-Cable Forces and Moments

The tether in the present analysis has been modeled as flexible and inextensible cable.

Under a steady horizontal wind velocity V∞ , the tether assumes a particular shape and

there is a variation of tether tension and angle with horizontal along the tether length

(Figure 3.20) from confluence point to anchor point. Hence before developing relations for

tether forces and moments, equilibrium cable configuration along with tether tension and

angle with horizontal variation has to be established.

Equilibrium Cable Shape and Derivatives

The expressions for estimating equilibrium cable shape along with tether tension and

angle with horizontal variation for the case of uniform air density and wind velocity from

the ground to the flying altitude have been presented in [36, 42]. Such type of analysis

is suitable only for small aerostats where flying altitude is less and the variation of air

density and wind velocity from anchor point to the flying altitude is neglected. For the

case of present test aerostat, the flying altitude is 1 km and hence these variations can not

be ignored and thus the method presented in [36, 42] is not suitable for the present test

aerostat.
116

To account for air density and steady wind velocity variation along the vertical direction

from the anchor point to the flying altitude, approximation methods have been proposed

in section 3.4 of chapter 3, where the tether is divided into small elements and the air

density and steady wind velocity is assumed constant on the element. Balancing the forces

on the element gives the tension and angle with horizontal for the next element. For the

present purpose, the expressions derived for polygonal approximation 2 as presented in

Equations (3.40) and (3.41) have been used with 100 elements. The tether tension and

angle with horizontal at the confluence point T1 and γ1 are known using Equations (3.1)

and (3.2) and the corresponding parameters at the anchor point T0 and γ0 are evaluated.

By summing the horizontal and vertical components of cable element, the coordinates of

the confluence point (x̃1 , z̃1 ) as referred in Figure 3.20 are also evaluated. The coordinates

of the center of mass of aerostat envelope with respect to anchor point (x1 , z1 ) are then

evaluated using Equation (5.19) assuming small perturbations.

The contributions of the cable forces come from two sources; first due to static tether

tension T1 acting at the confluence point as shown in Figure 4.1 and second due to change

in position of confluence point with respect to inertial frame. The second source can be

expressed in the form of cable derivatives whose expressions have been presented in [42]

as below:

1
T1 cos γ1 (sin γ1 − sin γ0 ) + nd (z1 − l sin γ0 ) sin3 γ1

kxx = (4.23a)
δ
1
T1 cos γ1 (cos γ0 − cos γ1 ) + nd (l cos γ0 − x̃1 ) sin3 γ1

kxz = (4.23b)
δ
1
T1 sin γ1 (sin γ1 − sin γ0 ) − wc + nd sin2 γ1 cos γ1 (z̃1 − l sin γ0 )
 
kzx = (4.23c)
δ
1
T1 sin γ1 (cos γ0 − cos γ1 ) − wc + nd sin2 γ1 cos γ1 (l cos γ0 − x̃1 )
 
kzz = (4.23d)
δ q
nd τ1 sin2 γ1 + 2p̄ cos γ1

kyy = R γ1 q (4.23e)
τ (γ)
γ0 sin2 γ+2p̄ cos γ

where

δ = x1 (sin γ1 − sin γ0 ) + z1 (cos γ0 − cos γ1 ) − l sin (γ1 − γ0 ) (4.24)

Here nd is the cable drag per unit length for cable normal to the wind and defined in

section 3.4, τ is evaluated from Equation (3.26), p̄ = wc /2nd and wc is the tether-cable

weight per unit length. All the parameters on the right hand side in Equations (4.23)
117

and (4.24) required for the calculations of cable derivatives are thus known. The lateral

derivative kyy in Equation (4.23e) is obtained by numerical integration [8]. These deriva-

tives in addition to tether tension at the confluence point T1 and angle with horizontal γ1

will be used to calculate tether forces and moments.

Cable Forces and Moments

Using the tether tension T1 , angle with horizontal γ1 and cable derivatives, the expressions

for the tether cable forces can be written in the inertial axes system. The components of

force in the inertial axes can be transformed into components in the stability axes using

the transfer relation of Equation (4.10) and also invoking small perturbations assumptions

where higher order perturbation terms are neglected. These forces act at the confluence

point and hence for the calculation of moments, the coordinates of confluence point with

respect to the center of mass of aerostat envelope in the stability axes system are required.

Referring to Figure 3.1, the coordinates in the body axes system are written as
 
X
 G − XC 
 

 0 

 
ZC − ZG

Since the stability axes are obtained by rotating the body axes by an angle −αt about the

j b axes, using the transformation relation presented in [13], the coordinates of confluence

point with respect to the center of mass of aerostat envelope in the stability axes system

may be written as below:


    
 x cs   cos αt 0 sin αt   XG − XC 
    
 0 = 0 1 0   0  (4.25)
    
    
z cs − sin αt 0 cos αt ZC − ZG

which may be expanded to give

xcs = (XG − XC ) cos αt + (ZC − ZG ) sin αt (4.26a)

zcs = (ZC − ZG ) cos αt − (XG − XC ) sin αt (4.26b)


118

The components of Equation (4.26) are used to write the moment expressions in the sta-

bility axes system as the forces in the stability axes system are known. The position of the

confluence point can also be written in terms of position of center of mass and the rota-

tions using similar expression as in Equation (5.19) with small perturbation assumption.

Finally the expressions for forces and moments may be written as below [42]:

FXC = −kxx xi − kxz zi − (kxθ + T1 sin γ1 ) θs + T1 cos γ1 (4.27a)

FYC = −kyy yi + (T1 sin γ1 − kyφ ) φs − (T1 cos γ1 + kyψ ) ψs (4.27b)

FZC = −kzx xi − kzz zi − (T1 cos γ1 − kzθ ) θs + T1 sin γ1 (4.27c)

MXC = −kφy yi − (zcs T1 sin γ1 + kφφ ) φs + (zcs T1 cos γ1 − kφψ ) ψs (4.27d)

MYC = −kθx xi − kθz zi − kθθ θs − xcs T1 sin γ1 + zcs T1 cos γ1 (4.27e)

MZC = −kψy yi + (xcs T1 sin γ1 − kψφ ) φs − (xcs T1 cos γ1 + kψψ ) ψs (4.27f)

where,

kxθ = zcs kxx − xcs kxz (4.28a)

kzθ = zcs kzx − xcs kzz (4.28b)

kθx = zcs kxx − xcs kzx (4.28c)

kθz = zcs kxz − xcs kzz (4.28d)

kθθ = kθθD + kθθT (4.28e)

kθθD = zc2s kxx − zcs xcs (kxz + kzx ) + x2cs kzz (4.28f)

kθθT = zcs (T1 sin γ1 ) + xcs (T1 cos γ1 ) (4.28g)

kyφ = −zcs kyy (4.28h)

kyψ = xcs kyy (4.28i)

kφy = kyφ (4.28j)

kφφ = zc2s kyy (4.28k)

kφψ = −xcs zcs kyy (4.28l)

kψy = kyψ (4.28m)

kψφ = kφψ (4.28n)

kψψ = x2cs kyy (4.28o)


119

4.4.4 Buoyancy Forces and Moments

The buoyancy force acting on the aerostat envelope B has been treated in combination

with the weight of helium gas inside the envelope WG , which acts downward at the center

of buoyancy itself, in the present analysis. This net force is denoted by Bg and is same

as B − WG . The expressions for the buoyancy forces and moments about the center of

mass in the stability axes system are found by multiplying the transformation matrix TIS

with the components in inertial frame and invoking small perturbation assumptions by

neglecting higher order terms. The buoyancy force in inertial frame is [0 0 − Bg ]T .

Also the coordinates of center of buoyancy with respect to center of mass in the stability

axes system are found in a similar manner as in Equation (4.25). The forces and moments

about the center of mass in the stability axes are found as below (Figures 4.1 and 3.1):

FXB = Bg θs (4.29a)

FYB = −Bg φs (4.29b)

FZB = −Bg (4.29c)

MXB = −Bg [(ZG − ZB ) cos αt + (XG − XB ) sin αt ] φs (4.29d)

MYB = Bg [(XG − XB ) cos αt − (ZG − ZB ) sin αt ]

− Bg [(ZG − ZB ) cos αt + (XG − XB ) sin αt ] θs (4.29e)

MZB = −Bg [(XG − XB ) cos αt − (ZG − ZB ) sin αt ] φs (4.29f)

4.4.5 Gravity Forces and Moments

The expressions for the gravity forces and moments about the center of mass in the stability

axis system are found by multiplying the transformation matrix TIS with the components

in inertial frame assuming small perturbations. The gravity force in inertial frame can

be expressed as [0 0 WS ]T . Also the coordinates of structural center of mass with

respect to center of mass in the stability axes system are found in a similar manner as

in Equation (4.25). As the weight of the helium gas has already been taken into account

in the buoyancy terms, only the aerostat envelope structural weight has been considered.

The forces and moments due to gravity are determined as below (Figures 4.1 and 3.1):

FXG = −WS θs (4.30a)


120

FYG = WS φs (4.30b)

FZG = WS (4.30c)

MXG = −WS [(ZS − ZG ) cos αt + (XS − XG ) sin αt ] φs (4.30d)

MYG = WS [(XS − XG ) cos αt − (ZS − ZG ) sin αt ]

− WS [(ZS − ZG ) cos αt + (XS − XG ) sin αt ] θs (4.30e)

MZG = −WS [(XS − XG ) cos αt − (ZS − ZG ) sin αt ] φs (4.30f)

4.4.6 Envelope Equations of Motion

The expressions for aerodynamic, tether cable, buoyancy and gravity forces and moments

were derived in Equations (4.22), (4.27), (4.29 and (4.30) respectively. The simplified

assumption of small perturbations has been used in deriving these forces and moments,

where higher order terms are neglected. Putting these expressions into force and moment

Equations (4.19) and (4.20) results in the equations of motion about the aerostat envelope

center of mass as follows [42]:

x-force:
   
ρa V∞ S ρa V∞ S
(mx ) ẍi + (2CD + CDu ) ẋi + (kxx ) xi + (CDα − CL ) żi
2 2
 2 
ρa V∞ S
+ (kxz ) zi + kxθ + (CDα − CL ) − (Bg − WS ) + T1 sin γ1 θs
2
ρa V∞ 2 S
+ CD − T1 cos γ1 = 0 (4.31a)
2
y-force:
   
ρa V∞ S ρa V∞ Sc̄
(my ) ÿi − CYβ
ẏi + (kyy ) yi − CYp φ̇s
2 4
 
ρa V∞ Sc̄ 
+ [kyφ − T1 sin γ1 + (Bg − WS )] φs + CYβ̇ − CYr ψ̇s
4
ρa V∞ 2 S
 
+ kyψ + T1 cos γ1 + CYβ ψs = 0 (4.31b)
2

z-force:
   
ρa V∞ S ρa V∞ S
(2CL + CLu ) ẋi + (kzx ) xi + (mz ) z̈i + (CLα + CD ) żi
2 2
ρa V∞ 2 S
   
ρa V∞ Sc̄ 
+ (kzz ) zi + CLα̇ + CLq θ̇s + (CLα + CD ) − T1 cos γ1 + kzθ θs
4 2
2
ρa V∞ S
+ CL + Bg − WS − T1 sin γ1 = 0 (4.31c)
2
121

Rolling moment:
" # " #
ρa S (c̄)2 ρa V∞ S (c̄)2
 
ρa V∞ Sc̄
− Clβ̇ ÿi − Clβ ẏi + (kφy ) yi + Ixs φ̈s − Clp φ̇s
4 2 4
" #
ρa V∞ S (c̄)2  
+ [zcs T1 sin γ1 + kφφ + MS1 ] φs − Ixz s ψ̈s + Clβ̇ − Clr ψ̇s
4
ρa V∞ 2 Sc̄
 
+ Clβ − zcs T1 cos γ1 + kφψ ψs = 0 (4.31d)
2

Pitching moment:
" #
ρa S (c̄)2
   
ρa V∞ Sc̄ ρa V∞ Sc̄
− (2Cm + Cmu ) ẋi + (kθx ) xi − Cmα̇ z̈i − Cmα żi
2 4 2
" #
ρa V∞ S (c̄)2 ρa V∞ 2 Sc̄
 

+ (kθz ) zi + (Iys ) θ̈s − Cmα̇ + Cmq θ̇s + kθθ + MS1 − Cmα θs
4 2
ρa V∞ 2 Sc̄
− Cm + xcs T1 sin γ1 − zcs T1 cos γ1 − MS2 = 0 (4.31e)
2
Yawing moment:
" # " #
ρa S (c̄)2 ρa V∞ S (c̄)2
 
ρa V∞ Sc̄
− Cnβ̇ ÿi − Cnβ ẏi + (kψy ) yi − (Ixz s ) φ̈s − Cnp φ̇s
4 2 4
" #
ρa V∞ S (c̄)2  
+ (MS2 + kψφ − xcs T1 sin γ1 ) φs + (Izs ) ψ̈s + Cnβ̇ − Cnr ψ̇s
4
ρa V∞ 2 Sc̄
 
+ Cnβ + xcs T1 cos γ1 + kψψ ψs = 0 (4.31f)
2

where

MS1 = [(XG − XB ) Bg + (XS − XG ) WS ] sin αt

+ [(ZG − ZB ) Bg + (ZS − ZG ) WS ] cos αt (4.32a)

MS2 = [(XG − XB ) Bg + (XS − XG ) WS ] cos αt

− [(ZG − ZB ) Bg + (ZS − ZG ) WS ] sin αt (4.32b)

and

mx = mxo (4.33a)
ρa Sc̄
my = myo − CYβ̇ (4.33b)
4
ρa Sc̄
mz = mzo + CLα̇ (4.33c)
4
122

4.5 Stability Equations

Equations (4.31) contain the steady-state equilibrium terms as well as perturbation terms.

The steady-state trimmed conditions are obtained from longitudinal equations by setting

the perturbation quantities equal to zero. It is observed that solving these steady-state

equations gives the same trim angles as in section 3.3 of chapter 3, which were obtained

by simple balancing of forces and moments, thus validating. The stability equations are

obtained by setting the equilibrium trim portions of the Equations (4.31) equal to zero.

These stability equations are as below:

Longitudinal equations:

x-force:
ρa V∞ S kxx ρa V∞ S kxz
ẍi + (2CD + CDu ) ẋi + xi + (CDα − CL ) żi + zi
2mx mx 2mx mx
ρa V∞ 2 SCDα
 
kxθ
+ + θs = 0 (4.34a)
mx 2mx

z-force:
ρa V∞ S kzx ρa V∞ S kzz
(2CL + CLu ) ẋi + xi + z̈i + (CLα + CD ) żi + zi
2mz mz 2mz mz
ρa V∞ 2 SCLα
 
ρa V∞ Sc̄  kzθ
+ CLα̇ + CLq θ̇s + + θs = 0 (4.34b)
4mz mz 2mz

Pitching moment:
ρa V∞ Sc̄ kθx ρa Sc̄2 ρa V∞ Sc̄ kθz
− (2Cm + Cmu ) ẋi + xi − Cmα̇ z̈i − Cmα żi + zi + θ̈s
2Iys Iys 4Iys 2Iys Iys
ρa V∞ S (c̄)2 ρa V∞ 2 Sc̄
 
 MS1 kθθ
− Cmα̇ + Cmq θ̇s + − Cmα + θs = 0 (4.34c)
4Iys Iys 2Iys Iys

Lateral equations:

y-force:
kyφ ρa V∞ 2 SCL
 
ρa V∞ S kyy ρa V∞ Sc̄
ÿi − CYβ ẏi + yi − CYp φ̇s + − φs
2my my 4my my 2my
" #
ρa V∞ 2 S CYβ + CD

ρa V∞ Sc̄   kyψ
+ CYβ̇ − CYr ψ̇s + + ψs = 0 (4.35a)
4my 2my my

Rolling moment:
ρa S (c̄)2 ρa V∞ Sc̄ kφy ρa V∞ S (c̄)2
− Clβ̇ ÿi − Clβ ẏi + yi + φ̈s − Clp φ̇s
4Ixs 2Ixs I xs 4Ixs
123

kφφ + zcs T1 sin γ1 + MS1 Ixz s ρa V∞ S (c̄)2  


+ φs − ψ̈s + Clβ̇ − Clr ψ̇s
Ixs Ixs 4Ixs
" #
ρa V∞ 2 S c̄Clβ − zcs CD

kφψ
+ + ψs = 0 (4.35b)
Ixs 2Ixs

Yawing moment:
ρa S (c̄)2 ρa V∞ Sc̄ kψy Ixz s ρa V∞ S (c̄)2
− Cnβ̇ ÿi − Cnβ ẏi + yi − φ̈s − Cnp φ̇s
4Izs 2Izs Izs Izs 4Izs
kψφ + MS2 − xcs T1 sin γ1 ρa V∞ S (c̄)2  
+ φs + ψ̈s + Cnβ̇ − Cnr ψ̇s
Izs 4Izs
" #
2
kψψ ρa V∞ S c̄Cnβ + xcs CD
+ + ψs = 0 (4.35c)
I zs 2Izs

Thus we obtain three longitudinal and three lateral stability equations. It is also observed

here that longitudinal and lateral equations of motion are decoupled and hence the equa-

tions can be treated separately for solution and interpretation. Although the stability

equations are written about the aerostat envelope center of mass, the aerodynamic forces

and moments may be referenced to an arbitrary point which is different from center of

mass. In such cases it is necessary to transfer the aerodynamic terms to the envelope

center of mass through the use of appropriate transfer relations.

4.6 Envelope Stability Characteristics

The longitudinal and lateral stability characteristics of the envelope are determined in-

dependently from Equations (4.34) and (4.35), respectively, in a manner similar to that

used in conventional airplane stability analysis [13]. Since the stability equations are ordi-

nary linear differential equations with constants coefficients, they have transient solutions

which are always exponential in form. For example, a typical variable such as θ is of the

form θ = θ̄eλt , where θ̄ is a constant. If these exponential forms are substituted into the

longitudinal and lateral stability equations and the determinants of the coefficients are set

equal to zero, solutions for the characteristic roots λ can be obtained [43]. An alternate

and more efficient way to get the characteristic roots is to write the stability equations in

the standard state space form [35] as given below:

Ẋ = AX + Bµ (4.36)
124

where X is the state vector, µ is the control vector and the matrices A and B contain the

aerostat’s dimensional stability derivatives. For carrying out the stability analysis, input

control vector is not considered; hence matrix A gives the characteristics of the aerostat

system. The longitudinal equations of motion are written in matrix form as below:
     
A
 11 A 12 A ẍ
13   s  B11 B12 B13   ẋs 
     
 21 A22 A23   z̈s  + B21
A    B22 B23   żs 
 
     
A31 A32 A33 θ̈s B31 B32 B33 θ̇s
     (4.37)
C11 C12 C13   xs   0 
    
+C21 C22 C23 

 zs = 0 
 
    
C31 C32 C33 θs 0

Using the definition of Equation (4.36), Equation (4.37) can always be written in the

following form:

   

 s   ẋs 
   
 z̈   żs 
 s   
   
 θ̈s  θ̇s 
   

  = AG   (4.38)
   
 ẋs   xs 
   
   
 ż   zs 
 s   
   
θ̇s θs

The first three equations of Equations (4.38) are taken from Equations (4.34) expressed

in such a manner that each second derivative of the variable is expressed as a function of

all the variables and their first derivative. The last three equations are identity. Using the

definition of Equation (4.13), Equation (4.38) can be written as:

   
 u̇s   us 
   

 ẇs 
 
 ws 

   
q̇s  qs 
   
 
  = AG   (4.39)
   

 ẋs 


 xs 

   

 żs 


 zs 

   
θ̇s θs
125

Similarly, the lateral stability equations are written as below:


   
 v̇s   vs 
   

 ṗs 


 ps 

   
ṙs  rs 
   
 
  = AT   (4.40)
   

 ẏs 


 ys 

   

 φ̇s 


 φs 

   
ψ̇s ψs

The eigenvalues of the matrices AG and AT give the characteristic roots or poles for

longitudinal and lateral equations of motion. Since matrices AG and AT are both of size

6×6, they give six characteristic roots for each of the longitudinal and lateral cases. These

roots appear as complex conjugate pairs (i.e., λ = η ± iω) for oscillatory modes of motion

or as real numbers (i.e., λ = η) for aperiodic modes. Thus, in each of the longitudinal and

lateral cases, it is possible for the envelope to exhibit from three to six modes of motion,

depending on whether the roots are complex conjugate pairs or real. There are three

categories under which all linear control systems fall in terms of stability [27, 44]:

1. When all the roots of the characteristic equation are having negative real part, the

system response due to the initial conditions will decrease to zero as t → ∞. Such

linear systems are said to be stable.

2. If one or more pairs of simple roots are having zero real part, but there are no roots

having positive real part, the response due to initial conditions will be undamped

sinusoidal oscillations. In linear system theory, such a system is said to be unstable.

3. If one or more roots are found to have positive real part, the response will increase in

magnitude as t → ∞. In linear system theory, such a system is said to be unstable.

Computer programs based on Equations (4.39) and Equations (4.40) were developed for

calculating the stability characteristics and plotting the results. The various inputs for the

test aerostat including mass, geometric, inertia and aerodynamic parameters are presented

in Tables 3.2, 4.1 and 4.2. The majority of these parameters were evaluated in Chapter 2

earlier. Since the longitudinal and lateral equations of motion are uncoupled, the results

are presented separately.


126

4.7 Response of Envelope due to Initial Disturbances

An aerostat envelope may be subjected to initial disturbances during its operation, par-

ticularly during launch and recovery. These disturbances may be in terms of distance,

angle, velocity, angular velocity or a combination of these. It is also important to predict

the motion of aerostat envelope to these initial disturbances. Once the eigenvalues and

corresponding eigenvectors for the characteristic matrices are evaluated as described in

section 4.6, the response of this aerostat envelope due to initial disturbances can be cal-

culated following the procedure as described in [25, 43]. As discussed in section 4.6, the

characteristic roots may be real or complex conjugate pair or a combination of these two.

Hence the process to evaluate initial disturbance response is explained here for these two

different cases i.e. real and complex eigenvalues.

4.7.1 Real Eigenvalues

For real eigenvalues, the imaginary component equals zero and indicate non oscillatory

modes of motion. A typical eigenvalue is λ1 and the corresponding eigenvector is repre-

sented as  
a
 1 
 
 a 
 2 
 
 a3 
 
V1=


 (4.41)
 a4 
 
 
 a 
 5 
 
a6

The solution corresponding to this eigenvalue and eigenvector is written as below [43]:

 
a
 1 
 
 a 
 2 
 
a
 
λ1 t  3 
 
X(t) = c1 e   (4.42)
 a4 
 
 
 a 
 5 
 
a6
127

Similarly, the contributions of other eigenvalues and corresponding eigenvectors are eval-

uated. Adding the contributions of all the eigenvalues and corresponding eigenvectors, we

get the complete solution. The constants c1 to c6 are evaluated for a given set of initial

conditions and thus we get the complete solution.

4.7.2 Complex Eigenvalues

For complex eigenvalues, the imaginary component is not equal to zero and indicate oscil-

latory modes of motion. Since the complex eigenvalues always appear as conjugate pairs,

a typical eigenvalue pair is λ1 ± µ1 i and the corresponding eigenvector is written as

 
 a1 ± b1 i 
 

 a2 ± b2 i 

 
a3 ± b3 i 
 

V1=


 (4.43)

 a4 ± b4 i 

 

 a5 ± b5 i 

 
a6 ± b6 i

The solution corresponding to this pair of eigenvalues and eigenvectors is written as below

[43]:  
 a1 ± b1 i 
 

 a2 ± b2 i 

 
a3 ± b3 i 
 
(λ1 ±µ1 i)t 

X(t) = e 

 (4.44)

 a4 ± b4 i 

 

 a5 ± b5 i 

 
a6 ± b6 i

Expanding the above expression and separating real and imaginary parts, the solution

corresponding to this set of eigenvalues and eigenvectors is written as below:


128

   
a
 1 cos µ 1 t − b1 sin µ t
1   a1 sin µ1 t + b1 cos µ1 t 
   
 a cos µ t − b sin µ t   a2 sin µ1 t + b2 cos µ1 t 
 2 1 2 1   
   
a cos µ t − b sin µ t a3 sin µ1 t + b3 cos µ1 t 
   
λ1 t  3 1 3 1 λ1 t 
  
X(t) = c1 e  
 + c2 e 

 (4.45)
 a4 cos µ1 t − b4 sin µ1 t   a4 sin µ1 t + b4 cos µ1 t 
   
   
 a cos µ t − b sin µ t   a5 sin µ1 t + b5 cos µ1 t 
 5 1 5 1   
   
a6 cos µ1 t − b6 sin µ1 t a6 sin µ1 t + b6 cos µ1 t

Similarly, the contributions of other eigenvalues and corresponding eigenvectors are eval-

uated. Adding the contributions of all the eigenvalues and corresponding eigenvectors, we

get the complete solution. The constants c1 to c6 are evaluated for a given set of initial

conditions as done for the previous case and thus we get complete solution.

4.8 Validation of Dynamic Stability Model

Before carrying out the dynamic stability analysis of the test aerostat, the model was

validated with the results presented in reference [42] which is for a 7.64m long balloon

having a volume of 19 m3 . The other details including aerostatic, aerodynamic, mass

and inertia properties of this balloon are presented in Table I in reference [42]. Since the

flying altitude for this balloon is low, no variation in the air density and wind velocity

along the vertical direction is assumed and the tether tension and angle with horizontal

at the anchor point have been estimated using direct integration approach as presented

in reference [42] and also in section 3.4 of chapter 3 of this report. A typical result

for variation of longitudinal characteristic roots in root-locus form with velocity as a

parameter, as presented in reference [42] is shown in Figure 4.3 for all the three modes

of motion. The same result for the same aerostat balloon has been generated using the

approach described earlier in this chapter and is presented in Figure 4.4. Similar results for

lateral modes are presented in Figures 4.5 and Figure 4.6. A comparison of the Figures 4.3

and 4.4, and Figures 4.5 and 4.6 indicate that both the results are same and hence the

approach for dynamic stability analysis is validated. The polygonal approximations used

in place of direct integration for tether modeling have already been validated in section 3.4

of chapter 3 of this report.


129

Figure 4.3: Longitudinal root locus presented for the 7.64m long reference balloon in [42]

10

6
ω (rad/s)

0
−6 −4 −2 0 2 4
η (per sec)

Figure 4.4: Longitudinal root locus generated for the 7.64m long reference balloon
130

Figure 4.5: Lateral root locus presented for the 7.64m long reference balloon in [42]

10

6
ω (rad/s)

0
−6 −4 −2 0 2 4
η (per sec)

Figure 4.6: Lateral root locus generated for the 7.64m long reference balloon
131

4.9 Results and Discussions for Test Aerostat

Since the longitudinal and lateral equations of motion are uncoupled, the results are pre-

sented separately. In each case the results for the test aerostat envelope, whose parameters

are listed in Tables 3.2, 4.1 and 4.2, are presented. Table 3.2 lists the parameters for the

test aerostat which were required for tether tension estimation as well as for carrying

out equilibrium and static analysis in chapter 3. These parameters included aerostatic

properties, geometric properties, mass properties and longitudinal static aerodynamic co-

efficients. But for carrying out dynamic stability analysis additional properties such as

structural center of mass, mass moments of inertia, apparent mass terms, tether cable

static properties as well as dynamic and other aerodynamic derivatives are required, which

are listed in Tables 4.1 and 4.2. The lift and moment coefficient about envelope nose are

modeled as linear curves whereas the drag coefficient is modeled as complete quadratic

curve. The pitching moment coefficient in Table 4.2 has been presented about envelope

nose and this has to be transferred to the center of mass using appropriate transfer relation

[4]. All the input parameters correspond to a height of 1000 m above the mean sea level

for ISA+15 deg temperature condition. The wind speed range considered for the analysis

was from 0 to 25 m/s which is the operating wind speed range for the test aerostat. The

methods which were followed to estimate all the input parameters have been described in

chapter 2.

Table 4.1: Additional parameters required for the test aerostat envelope for dynamic
stability analysis
Aerostat parameter Value
Axial distance of SCM from nose, XS , m 18.94
Vertical distance of SCM from nose, ZS , m 2.4
Apparent air mass along ib , mxb , kg 234.86
Apparent air mass along j b , myb , kg 1750.74
Apparent air mass along kb , mzb , kg 1750.74
Mass of envelope structure and inflation gas, mb , kg 1494
Rolling moment of inertia, Ixb , kg-m 2 31200.23
Pitching moment of inertia, Iyb , kg-m2 186369.73
Yawing moment of inertia, Izb , kg-m2 187165.72
Product of inertia, Ixz b , kg-m 2 2083.86
Tether-cable diameter, dc , m 0.017
Tether-cable length, l, m 1000
Tether-cable weight per unit length, wc , N/m 2.943
Tether-cable normal drag coefficient, CDc 1.17
132

Table 4.2: Aerodynamic derivatives of test aerostat envelope about center of mass
Parameter Value Parameter Value
CL 1.8313αt − 0.0245 C Yβ -1.4729
CLα 1.8313 CYβ̇ 0
CLα̇ 0 C Yp -0.2063
CLq 0.4478 C Yr 0.8386
CLu 0 Cl β -0.1031
CD 2
1.4090αt + 0.0074αt + 0.0294 Clβ̇ 0
CDα 2.818αt + 0.0074 Clp -0.0800
CDu 0 Cl r 0.0587
Cmn −0.8782αt + 0.0214 (about nose) Cnβ 0.4193
Cmnα -0.8782 (about nose) Cnβ̇ 0
Cmα̇ 0 Cnp 0.0587
Cm q -0.1275 Cnr -0.2387
Cm u 0

4.9.1 Longitudinal Stability Characteristics

The calculated values of the longitudinal damping parameter η and circular frequency

ω for the test aerostat envelope in the reference configuration are plotted as a function

of wind velocity in Figures 4.7 and 4.8 respectively. These same values of η and ω are

plotted in root-locus form with velocity as the varying parameter in Figure 4.9. These

figures indicate that the reference configuration has three oscillatory modes of motion for

wind speed less than 14.4 m/sec. For wind speed greater than this value, mode 1 splits

into two real non-oscillatory modes. Since there are no positive real parts, the figures

also indicate that the calculated modes are stable for the operating range of wind speed.

The eigenvalues and corresponding eigenvectors with relative magnitude and direction of

longitudinal parameters for the characteristic modes for two cases namely for a typical low

wind speed (5 m/s) and for a typical high wind speed (20 m/s) are presented in Tables 4.3

and 4.4 respectively.


133

3
0

2
1a

−0.5 1
η (per sec)

−1

1b

−1.5

−2
0 5 10 15 20 25
Wind speed (m/s)

Figure 4.7: Variation of longitudinal damping parameter with velocity for test aerostat

3
ω (rad/sec)

1
1

0 5 10 15 20 25
Wind speed (m/s)

Figure 4.8: Variation of longitudinal circular frequency with velocity for test aerostat
134

3
ω (rad/s)

2 2

1
1
3

−2 −1.5 −1 −0.5 0
η (per sec)

Figure 4.9: Longitudinal root locus plot with velocity as a parameter

Table 4.3: Longitudinal eigenvalues and eigenvectors for wind speed of 5 m/s
Eigenvalue Parameters Eigenvector Magnitude Angle (deg)
-0.1160 + 0.5860i us (m/s) 0.0067 + 0.0369i 0.0375 79.6680
(Mode 1) ws (m/s) -0.0983 + 0.4969i 0.5065 101.1955
qs (rad/s) -0.0153 + 0.0693i 0.0710 102.4679
xs (m) 0.0584 - 0.0230i 0.0628 -21.5274
zs (m) 0.8479 0.8479 0
θs (rad) 0.1188 + 0.0026i 0.1188 1.2725
-0.0876 + 2.7361i us (m/s) -0.0704 - 0.0012i 0.0704 -179.0630
(Mode 2) ws (m/s) -0.9289 0.9289 180.0000
qs (rad/s) 0.1201 + 0.0058i 0.1202 2.7761
xs (m) 0.0004 + 0.0257i 0.0257 89.1028
zs (m) 0.0109 + 0.3392i 0.3393 88.1658
θs (rad) 0.0007 - 0.0439i 0.0439 -89.0582
-0.0074 + 0.0551i us (m/s) 0.0074 - 0.0549i 0.0554 -82.3393
(Mode 3) ws (m/s) -0.0002 + 0.0024i 0.0025 95.5845
qs (rad/s) 0.0000 + 0.0000i 0.0000 70.6046
xs (m) -0.9975 0.9975 180.0000
zs (m) 0.0441 - 0.0016i 0.0442 -2.0763
θs (rad) 0.0004 - 0.0002i 0.0005 -27.0562
135

Table 4.4: Longitudinal eigenvalues and eigenvectors for wind speed of 20 m/s
Eigenvalue Parameters Eigenvector Magnitude Angle (deg)
-0.1990 us (m/s) -0.1478 0.1478 180.0000
(Mode 1a) ws (m/s) 0.1275 0.1275 0
qs (rad/s) 0.0005 0.0005 0
xs (m) 0.7426 0.7426 0
zs (m) -0.6406 0.6406 180.0000
θs (rad) -0.0026 0.0026 180.0000
-1.3709 us (m/s) -0.0959 0.0959 180.0000
(Mode 1b) ws (m/s) -0.8013 0.8013 180.0000
qs (rad/s) 0.0387 0.0387 0
xs (m) 0.0699 0.0699 0
zs (m) 0.5845 0.5845 0
θs (rad) -0.0282 0.0282 180.0000
-0.0251 + 0.7442i us (m/s) 0.0678 + 0.1527i 0.1671 66.0661
(Mode 2) ws (m/s) -0.0193 + 0.5725i 0.5728 91.9288
qs (rad/s) 0.0229 + 0.0111i 0.0255 25.8431
xs (m) 0.2019 - 0.0979i 0.2244 -25.8627
zs (m) 0.7693 0.7693 0
θs (rad) 0.0139 - 0.0313i 0.0342 -66.0857
-0.0424 + 0.0887i us (m/s) -0.0401 + 0.0839i 0.0930 115.5638
(Mode 3) ws (m/s) 0.0119 - 0.0279i 0.0303 -66.8839
qs (rad/s) -0.0000 - 0.0001i 0.0001 -110.2384
xs (m) 0.9463 0.9463 0
zs (m) -0.3079 + 0.0132i 0.3081 177.5523
θs (rad) -0.0008 + 0.0008i 0.0012 134.1978

The different longitudinal modes of motion which are observed for this aerostat and

their characteristics are summarized as below:

Mode 1

This mode has low damping at low wind speed but as the wind speed increases, damping

also increases. For wind speed of less than 14.4 m/s, this mode is oscillatory in nature

whereas for wind speed greater than 14.4 m/s, this oscillatory mode splits into two non

oscillatory modes 1a and 1b as observed in Figure 4.7. Further increase in wind speed

beyond 14.4 m/s decreases damping for mode 1a and at the same time increases damping

for mode 1b. This mode is seen to be a motion in which forward speed us and forward

position xs are very small for low wind speed (Table 4.4). For high wind speed, this mode

splits into two non oscillatory modes. It is to be noted in Tables 4.3 and 4.4 that the

parameters are dimensional and that pitch rate and pitch angle are expressed in radian/s

and radian respectively but are converted to degree/s and degree for comparison.
136

Mode 2

This mode has low damping but high circular frequency for low wind speed as observed

in Figures 4.7 and 4.8. As the wind speed increases, the damping first increases slightly

up to 7 m/s but the damping again decreases with further increase in wind speed. The

circular frequency decreases with increase in wind speed up to 10 m/s and assumes a

nearly constant value after that. Similar to mode 1, the forward speed us and forward

position xs are very small for low wind speed. For high wind speed, all the parameters

are having significant magnitude as observed in Table 4.4.

Mode 3

This mode has very low damping and very low circular frequency resulting in high time

period irrespective of wind speed as observed from Figures 4.7 and 4.8. It is also observed

in Figure 4.9 that all the characteristic roots lie near the origin for all the wind speeds. For

this mode forward speed us , forward position xs and vertical position zs have significant

magnitude for the entire range of wind speed as observed in Tables 4.3 and 4.4. Unlike

mode 1 and mode 2, the behaviour of this mode is almost independent of wind speed.

The vertical speed ws , pitch rate qs and pitch angle θs have negligible magnitude for this

mode.

In conventional aircraft, there are two complex conjugate pairs of eigenvalues for lon-

gitudinal motion indicating two modes of motion, one of long period and lightly damped

named as phugoid mode and the other of short period and heavily damped named as short-

period mode. The phugoid is seen to be a motion in which the pitch rate and the angle of

attack change are very small, but forward speed change and pitch angle change are present

with significant magnitude. The short-period mode, by contrast, is one in which there is

negligible speed variation, while the angle of attack oscillates with an amplitude and phase

not much different from that of change in pitch angle [13]. In contrast, three oscillatory

modes of motion are observed for the test aerostat up to wind speed of 14.4 m/s with the

characteristics indicated above. Beyond this wind speed, mode 1 splits into two non oscil-

latory modes. Mode 3 indicates translation along the horizontal direction whereas modes

1 and 2 indicate a combination of pitching and vertical motion, which is obvious due to the

fact that any change in vertical position will also correspond to a change in tether tension

and hence the pitch angle. In conventional aircraft, the eigenvalues are evaluated at a
137

constant horizontal speed, whereas for a tethered aerostat, entire wind speed range has to

considered for eigenvalues evaluation. The primary purpose of evaluation of eigenvalues

for a tethered aerostat is to examine the dynamic stability and since there are no positive

real part observed in the longitudinal eigenvalues, the test aerostat is dynamically stable

in longitudinal directions.

4.9.2 Lateral Stability Characteristics

The calculated values of the lateral damping parameter η and circular frequency ω for

the test aerostat envelope in the reference configuration are plotted as a function of wind

velocity in Figures 4.10 and 4.11 respectively. These same values of η and ω are plotted

in root-locus form with velocity as the varying parameter in Figure 4.12. These figures

indicate that the reference configuration has three oscillatory modes of motion for the

considered wind speed range. Since there are no positive real parts, the figures also

indicate that the calculated modes are stable for the operating range of wind speed. The

eigenvalues and corresponding eigenvectors with relative magnitude and direction of lateral

parameters for the characteristic modes for two cases namely for a typical low wind speed

(5 m/s) and for a typical high wind speed (20 m/s) are presented in Tables 4.5 and 4.6

respectively.

The different lateral modes of motion which are observed for this aerostat and their

characteristics are summarized as below:

Mode 1

This mode has low damping at low wind speed but as the wind speed increases, damping

also increases as observed in Figure 4.10. Also the circular frequency decreases with

increase in wind speed as depicted in Figure 4.11. This mode is seen to be a motion in

which yaw rate rs and yaw angle ψs are very small for low wind speed. For high wind

speed, other lateral parameters also become significant.


138

0
3

−0.5 2

−1
η (per sec)

−1.5 1

−2

−2.5
0 5 10 15 20 25
Wind speed (m/s)

Figure 4.10: Variation of lateral damping parameter with velocity for test aerostat

1
1.5
ω (rad/sec)

2
0.5

3
0
0 5 10 15 20 25
Wind speed (m/s)

Figure 4.11: Variation of lateral circular frequency with velocity for test aerostat
139

1.5

1
ω (rad/s)

1 2

0.5

3
0
−2.5 −2 −1.5 −1 −0.5 0
η (per sec)

Figure 4.12: Lateral root locus plot with velocity as a parameter

Table 4.5: Lateral eigenvalues and eigenvectors for wind speed of 5 m/s
Eigenvalue Parameters Eigenvector Magnitude Angle (deg)
-0.3311 + 1.7958i vs (m/s) -0.0220 + 0.2671i 0.2680 94.7054
(Mode 1) ps (rad/s) 0.8331 0.8331 0
rs (rad/s) -0.0418 - 0.0407i 0.0583 -135.7384
ys (m) 0.1460 - 0.0147i 0.1468 -5.7412
φs (rad) -0.0827 - 0.4487i 0.4562 -100.4466
ψs (rad) -0.0178 + 0.0265i 0.0319 123.8150
-0.2444 + 0.3152i vs (m/s) 0.2240 - 0.2889i 0.3655 -52.2129
(Mode 2) ps (rad/s) -0.0050 - 0.0016i 0.0053 -162.5545
rs (rad/s) -0.0347 + 0.0492i 0.0602 125.2052
ys (m) -0.9164 0.9164 180.0000
φs (rad) 0.0046 + 0.0124i 0.0132 69.6584
ψs (rad) 0.1508 - 0.0068i 0.1509 -2.5818
-0.0042 + 0.0541i vs (m/s) 0.0042 - 0.0540i 0.0542 -85.5519
(Mode 3) ps (rad/s) 0.0000 - 0.0000i 0.0000 -70.6594
rs (rad/s) 0.0006 + 0.0001i 0.0006 9.8517
ys (m) -0.9985 0.9985 180.0000
φs (rad) -0.0004 - 0.0001i 0.0004 -165.1076
ψs (rad) 0.0010 - 0.0109i 0.0110 -84.5965
140

Table 4.6: Lateral eigenvalues and eigenvectors for wind speed of 20 m/s
Eigenvalue Parameters Eigenvector Magnitude Angle (deg)
-1.6845 + 1.2139i vs (m/s) 0.8143 0.8143 0
(Mode 1) ps (rad/s) 0.2872 - 0.2208i 0.3623 -37.5555
rs (rad/s) -0.1299 + 0.0234i 0.1320 169.8006
ys (m) -0.3182 - 0.2293i 0.3922 -144.2237
φs (rad) -0.1744 + 0.0054i 0.1745 178.2208
ψs (rad) 0.0573 + 0.0274i 0.0636 25.5769
-0.6221 + 1.3672i vs (m/s) 0.7732 0.7732 0
(Mode 2) ps (rad/s) -0.2879 - 0.0520i 0.2926 -169.7683
rs (rad/s) -0.0963 - 0.0146i 0.0975 -171.3675
ys (m) -0.2132 - 0.4685i 0.5147 -114.4655
φs (rad) 0.0479 + 0.1888i 0.1948 75.7661
ψs (rad) 0.0177 + 0.0624i 0.0649 74.1670
-0.0134 + 0.0609i vs (m/s) -0.0134 + 0.0608i 0.0623 102.3917
(Mode 3) ps (rad/s) -0.0000 + 0.0000i 0.0000 126.8014
rs (rad/s) -0.0002 - 0.0001i 0.0002 -154.9125
ys (m) 0.9981 0.9981 0
φs (rad) 0.0004 + 0.0002i 0.0005 24.4097
ψs (rad) -0.0007 + 0.0030i 0.0031 102.6958

Mode 2

This mode has low damping and low circular frequency for low wind speed as observed in

Figures 4.10 and 4.11. As the wind speed increases, the damping also increases till about

14 m/s and remains almost constant after that. The circular frequency for this mode

increases continuously with increasing wind speed. This mode is seen to be a motion in

which roll rate ps and roll angle φs are very small for low wind speed.

Mode 3

This mode has very low damping and very low circular frequency resulting in high time

period irrespective of wind speed as observed from Figures 4.10 and 4.11. It is also observed

in Figure 4.11 that all the characteristic roots lie near the origin for all wind speeds. For

this mode side velocity vs and side position ys have significant magnitude for the entire

range of wind speed as observed in Tables 4.5 and 4.6. Unlike mode 1 and mode 2, the

behavior of this mode is almost independent of wind speed. The roll rate ps , yaw rate rs ,

roll angle φs and yaw angle ψs have negligible magnitude for this mode.

In conventional aircraft, there are two real eigenvalues and one complex conjugate

pair of eigenvalues for lateral motion indicating three modes of motion, one of very rapid

convergence (roll mode), one very slow convergence (spiral mode) and one is a lightly
141

damped oscillation with a period similar to that of the longitudinal short-period (Dutch

roll mode) [13]. In contrast, three oscillatory modes of motion are observed for the test

aerostat with the characteristics indicated above. Mode 3 indicates translation along the

side direction whereas modes 1 and 2 indicate a combination of all lateral motions. Since

there are no positive real part observed in the longitudinal eigenvalues, the test aerostat

is dynamically stable in lateral directions.

4.9.3 Response to Initial Disturbances

The solution method explained in section 4.7 has been applied to longitudinal as well as

lateral equations of motion for the test aerostat envelope. The response studies are done

for initial conditions listed in Table 4.7 for longitudinal motion and in Table 4.8 for lateral

motion. The unit initial conditions are applied first in isolation and then in combination.

Thus there are seven initial conditions for longitudinal motion and seven initial conditions

for lateral motion. A wind speed of 5 m/s is considered for the analysis. As mentioned

in section 4.7, an aerostat may be subjected to initial disturbances during launch and

recovery. The test aerostat is generally launched and recovered at low wind speed. Hence

the wind speed of 5 m/s is considered for carrying out initial disturbance response studies

in this section.

Table 4.7: Response to initial disturbance cases for longitudinal motions


Case us (m/s) ws (m/s) qs (deg/s) xs (m) zs (m) θs (deg)
LO1 1 0 0 0 0 0
LO2 0 1 0 0 0 0
LO3 0 0 1 0 0 0
LO4 0 0 0 1 0 0
LO5 0 0 0 0 1 0
LO6 0 0 0 0 0 1
LO7 1 1 1 1 1 1

Table 4.8: Response to initial disturbance cases for lateral motions


Case vs (m/s) ps (deg/s) rs (deg/s) ys (m) φs (deg) ψs (deg)
LA1 1 0 0 0 0 0
LA2 0 1 0 0 0 0
LA3 0 0 1 0 0 0
LA4 0 0 0 1 0 0
LA5 0 0 0 0 1 0
LA6 0 0 0 0 0 1
LA7 1 1 1 1 1 1
142

The results for initial disturbances are presented in the form of graphs as shown in

Figures 4.13 through 4.26. For longitudinal motion case the plots are forward velocity,

vertical velocity, pitch rate, forward position, vertical position and pitch angle vs. time in

seconds. For lateral case the plots are side velocity, roll rate, yaw rate, side position, roll

angle and yaw angle vs. time in seconds. The time interval considered is 100 seconds for

both longitudinal and lateral motion cases.

Figure 4.13 presents the response with time for unit initial forward velocity for the

test aerostat envelope. It is observed that initial forward velocity has little effect on

vertical velocity and vertical position and has low damping for forward velocity, forward

position and pitch angle. The pitch rate stabilizes in about 50 seconds. It is observed in

Figures 4.14 and 4.17 that pitch angle and pitch rate are strong function of initial vertical

velocity and vertical position. This is a typical characteristic of aerostats due to presence

of tether where any change in vertical position is coupled with change in tether tension

and hence change in pitch angle. The forward velocity and forward position variation

is negligible for initial pitch rate and pitch angle variation as observed in Figures 4.15

and 4.18. This indicates that these motions are decoupled as also observed in Table 4.3 in

terms of eigenvectors. It is also observed in Figures 4.15 and 4.18 that the vertical velocity

and vertical position variation is small for for initial pitch rate and pitch angle variation and

that the disturbances die out fast indicating moderate damping. Figure 4.16 indicates that

the forward velocity, vertical velocity and vertical position variation is negligible for initial

forward position variation and that the damping for forward position is small whereas the

damping for pitch rate and pitch angle is moderate. Figure 4.19 presents the response for

unit initial disturbances applied to all the longitudinal parameters, which indicates low

damping for forward velocity, position and vertical position and moderate damping for

vertical velocity, pitch rate and angle and the latter parameters stabilize in less than 50 s.
143

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


15 15 1
10 10
0.5
5 5
0 0 0
−5 −5
−0.5
−10 −10
−15 −15 −1
0 50 100 0 50 100 0 50 100

Figure 4.13: Longitudinal response due to initial disturbance case LO1

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


1 1 10

0.5 0.5 5

0 0 0

−0.5 −0.5 −5

−1 −1 −10
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


1 1 10

0.5 0.5
5
0 0
0
−0.5 −0.5
−5
−1 −1

−1.5 −1.5 −10


0 50 100 0 50 100 0 50 100

Figure 4.14: Longitudinal response due to initial disturbance case LO2


144

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.15: Longitudinal response due to initial disturbance case LO3

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.16: Longitudinal response due to initial disturbance case LO4


145

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


1.5 1.5 15
1 1 10
0.5 0.5 5
0 0 0
−0.5 −0.5 −5
−1 −1 −10
−1.5 −1.5 −15
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


1 1 15
10
0.5 0.5
5
0 0 0
−5
−0.5 −0.5
−10
−1 −1 −15
0 50 100 0 50 100 0 50 100

Figure 4.17: Longitudinal response due to initial disturbance case LO5

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


1 1 1.5
1
0.5 0.5
0.5
0 0 0
−0.5
−0.5 −0.5
−1
−1 −1 −1.5
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.18: Longitudinal response due to initial disturbance case LO6


146

Forward Velocity (m/s) Vertical Velocity (m/s) Pitch Rate (deg/s)


2 2 15
10
1 1
5
0 0 0
−5
−1 −1
−10
−2 −2 −15
0 50 100 0 50 100 0 50 100

Forward Position (m) Vertical Position (m) Pitch Angle (deg)


15 15 15
10 10 10
5 5 5
0 0 0
−5 −5 −5
−10 −10 −10
−15 −15 −15
0 50 100 0 50 100 0 50 100

Figure 4.19: Longitudinal response due to initial disturbance case LO7

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 2 3
2
0.5 1
1
0 0 0
−1
−0.5 −1
−2
−1 −2 −3
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


20 2 20

10 1 10

0 0 0

−10 −1 −10

−20 −2 −20
0 50 100 0 50 100 0 50 100

Figure 4.20: Lateral response due to initial disturbance case LA1


147

Figure 4.20 presents the response with time for unit initial side velocity for the test

aerostat envelope. All the lateral parameters are significant for this initial condition but

the side position and yaw angle are strongly dependent as expected whereas the roll rate

dies out quickly in less than 20 seconds indicating strong damping. The side velocity, side

position, yaw rate and yaw angle variation are negligible for initial roll rate and roll angle

as observed in Figures 4.21 and 4.24 indicating decoupling of motion. For unit initial yaw

rate, only the yaw angle and the side position are significant whereas the damping for roll

rate and roll angle is strong as observed in Figure 4.22. It is noted in Figure 4.23 that for

unit side position, only the yaw angle variation is significant and the variation of other

lateral parameters are negligible. Figure 4.25 represents the response with time for unit

initial yaw angle, where it is noted that side velocity, roll rate and roll angle variation is

negligible, yaw rate is strongly damped and side position is lightly damped. Figure 4.26

presents the response for unit initial disturbances applied to all the lateral parameters,

which indicates strong damping for roll rate and light damping for side velocity, yaw rate,

side position, roll angle and yaw angle. It is also observed that lateral motion variables

die out relatively faster as compared to longitudinal motion variables indicating relatively

stronger damping in lateral directions. Since there are no control surfaces on the aerostat

balloon, the analysis results for initial disturbances may also be very useful in selecting

the proper aerostat envelope parameters during design for generating required response

characteristics.
148

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.21: Lateral response due to initial disturbance case LA2

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


2 1 2

1 0.5 1

0 0 0

−1 −0.5 −1

−2 −1 −2
0 50 100 0 50 100 0 50 100

Figure 4.22: Lateral response due to initial disturbance case LA3


149

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.23: Lateral response due to initial disturbance case LA4

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 2 1

0.5 1 0.5

0 0 0

−0.5 −1 −0.5

−1 −2 −1
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.24: Lateral response due to initial disturbance case LA5


150

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


1 1 1

0.5 0.5 0.5

0 0 0

−0.5 −0.5 −0.5

−1 −1 −1
0 50 100 0 50 100 0 50 100

Figure 4.25: Lateral response due to initial disturbance case LA6

Side Velocity (m/s) Roll Rate (deg/s) Yaw Rate (deg/s)


1 3 3
2 2
0.5
1 1
0 0 0
−1 −1
−0.5
−2 −2
−1 −3 −3
0 50 100 0 50 100 0 50 100

Side Position (m) Roll Angle (deg) Yaw Angle (deg)


20 2 20

10 1 10

0 0 0

−10 −1 −10

−20 −2 −20
0 50 100 0 50 100 0 50 100

Figure 4.26: Lateral response due to initial disturbance case LA7


151

4.10 Chapter Summary

The dynamic stability analysis of the tethered aerostat has been carried out in this chapter.

The equations of motion were developed and then linearized with suitable assumptions as

applicable to a tethered aerostat. The tether was modeled as a flexible and inextensible

cable using finite number of elements and polygonal approximation technique, in which

the air density variation along vertical direction has been taken into account. The assump-

tions in this chapter led to decoupling of the longitudinal and lateral equations of motion

and are thus handled separately. Before applying the equations to the test aerostat, the

method has been validated with literature. The equations are written in the state space

form and eigenvalues are evaluated for different operating wind speed and thus dynamic

stability analysis was carried out. The calculated eigenvalues indicated dynamic stability

of this test aerostat. Eigenvectors were also evaluated for extreme operating wind con-

ditions and thereby different modes of motion for this tethered aerostat was described.

Finally response studies were carried out to initial disturbances in longitudinal and lateral

directions separately which reveal useful information.


Chapter 5

Non Linear Dynamic Analysis of


the Tethered Aerostat

5.1 Introduction

During the dynamic stability analysis of the tethered aerostat as carried out in chapter 4,

several simplifying assumptions were made including small perturbations about equilib-

rium trim condition. The purpose of these assumptions was to make the longitudinal

and lateral equations of motion linear as well as decoupled from each other. Although

this type of analysis is suitable for small perturbations, it is useful for carrying out the

dynamic stability analysis of tethered aerostat system. But the fact remains that the

dynamic stability model is limited to small perturbations and 3-degrees of freedom. The

effect of non-linearity, detailed motions and peak tether tensions under the influence of

a time-dependent wind vector cannot be determined. The present chapter describes a

6-degrees-of-freedom, nonlinear dynamic analysis and its application to the test aerostat.

Although the aerodynamics is assumed linear, the equations of motion are not linearized.

The theoretical model incorporates dynamic motion of the aerostat envelope [15] and a

dynamic tether [19]. The ballonet motion inside the envelope has not been considered for

the present analysis. The analysis has been used to study the response of the test aerostat

to time dependent wind vector and also to simulated wind gusts. A comparative study

has also been carried out for equilibrium analysis and nonlinear analysis. The results of

nonlinear analysis have also been compared with available experimental flight data.

153
154

5.2 Aerostat System Description

The aerostat system schematic for carrying out nonlinear dynamic analysis is presented

in Figure 5.1. The aerostat envelope is modeled as a rigid body and the tether is mod-

eled as spring-mass-damper system. The tether is divided into N elastic segments with

spring-mass-damper representation of the segments. The mass of the segment is assumed

concentrated on the lower portion of the segment thereby creating N lumped masses. The

lowest point of the tether is attached to the anchor point and hence the lumped mass of

the bottom-most segment is also attached to the ground. The earth-fixed axes system

is fixed to the ground and defined by three orthogonal unit vectors ie , j e and ke . This

earth-fixed axes system is different from that used in dynamic stability analysis because

during the dynamic stability analysis, the motions of the aerostat envelope was assumed to

be small perturbations about the equilibrium trim condition and hence the center of mass

in equilibrium trim condition was selected as the origin of the inertial axes system. The

time dependent response will be better interpreted in the present earth-fixed axes system.

The equations of motion of lumped or point masses are written in earth-fixed reference

frame. An aerostat body axes with unit vectors ib , j b and kb is located at the aerostat

envelope center of mass, with ib aligned with the aerostat’s longitudinal axis and kb in the

vertical plane of symmetry, same as for dynamic stability model. It is to be noted that the

equations of motion of the aerostat envelope in chapter 4 was written in stability axes sys-

tem, whereas the equations of motion in the present chapter will be written in body axes

system. The aerostat envelope is connected to the tether at the confluence point through

the upper portion of the top most tether segment. The body axes system is defined by

a sequence of three body-fixed rotations. Starting from the earth-fixed axes system, the

body axes system is obtained by rotations about the k, j, and i axes by angles Ψb , Θb , and

Φb , respectively similar to Figure (4.2). The expression for the absolute velocity in terms

of the Euler angles and velocity components in the body axes as presented in reference

[35] and also in Equation (4.9) is rewritten as below:

    
 e  CΘb CΨb
ẋ SΦb SΘb CΨb − CΦb SΨb CΦb SΘb CΨb + SΦb SΨb   Ub 
    
 ẏ  =  C S SΦb SΘb SΨb + CΦb CΨb CΦb SΘb SΨb − SΦb CΨb  (5.1)
  Vb 
 
 e   Θb Ψb
    
że −SΘb SΦb CΘb CΦb CΘb Wb
155

Figure 5.1: Tethered aerostat schematic for nonlinear analysis

where CΘb = cos Θb and SΘb = sin Θb and so on. The transformation matrix in Equa-

tion (5.1) is represented by T E


B which is for transformation from body axes system to

earth-fixed system. The transformation relation as in Equation (5.1) for the lumped or

point masses is not required because of the absence of rigid body rotation and the equa-

tions of motion are directly written in earth-fixed axes system for them assuming point

masses in translational directions.

The transformation matrix from earth-fixed axes system to the body axes system may

be obtained by inverting the transformation matrix of Equation (5.1) as also presented in


156

Equation (4.10):

 
 CΘb CΨb CΘb SΨb −SΘb 
TB
 
E =
SΦb SΘb CΨb − CΦb SΨb SΦb SΘb SΨb + CΦb CΨb SΦb CΘb 
 (5.2)
 
CΦb SΘb CΨb + SΦb SΨb CΦb SΘb SΨb − SΦb CΨb CΦb CΘb

The relationship between the Euler rates (Φ̇b , Θ̇b and Ψ̇b ) and the angular velocities in

the body axes system (Pb , Qb and Rb ) can also be determined from Equation (4.8) by

inverting the matrix as below:


 
  
 Φ̇b  1 sin Φb tan Θb cos Φb tan Θb   Pb 
    
 Θ̇  = 0 cos Φb − sin Φb  (5.3)
  Qb 
 
 b  
    
Ψ̇b 0 sin Φb sec Θb cos Φb sec Θb Rb

The notations for body axes linear velocity components, angular velocity components and

Euler angles are denoted by uppercase letters with subscript b indicating the components in

body axes. These components with subscript s as well as equations of motion in chapter 4

were in the stability axes system whereas the equations of motion are in the body axes

system in the present chapter. It is also noted that earth-fixed axes system in the present

chapter for nonlinear dynamic analysis is different from the inertial axes used in chapter 4

for dynamic stability analysis.

5.3 Aerostat Envelope Dynamic Model

For modeling the aerostat envelope, rigid body assumption with six degrees-of-freedom

is again followed similar to the model for dynamic stability analysis. However, small

perturbations are not assumed in the present model. The earth-fixed positions of the

aerostat envelope center of mass are [xe ye ze ]T and the Euler angles are [Ψb Θb Φb ]T .

The earth-fixed angular velocity of the aerostat envelope Ω has components in the body

axes system as [Pb Qb Rb ]T and the earth-fixed linear velocity of the aerostat envelope

center of mass has components in body axes system as [Ub Vb Wb ]T . As discussed earlier

in chapter 4, the forces and moments on the aerostat envelope have contributions from

buoyancy, gravity, aerodynamic, tether and apparent mass terms.


157

5.3.1 Buoyancy Force

The buoyancy force acting on the aerostat envelope is denoted by B = Vg ρa g, the weight of

atmospheric air displaced. The buoyancy force in earth-fixed axes system can be expressed

as [0 0 − B]T . The expressions for the buoyancy force in the body axes system is found

by multiplying the transformation matrix TEB with the components in earth-fixed axes

system as below:  
 0 
= TB
 
FB E 0 
  (5.4)
 
−B

5.3.2 Gravity Force

Similar to buoyancy force, the gravity force vector in the body axes system is written as

below:  
 0 
TB
 
FW = E

 0 
 (5.5)
 
WB

where WB is the summation of structural weight of the aerostat envelope and inflation gas

weight inside the aerostat envelope.

5.3.3 Aerodynamic Forces and Moments

The model for aerodynamic forces and moments has been adapted from reference [15] and

is represented as below:  
 CX 
 
FA = q0 S 
 CY

 (5.6)
 
CZ
 
C
 l 
 
MA = q0 Sc̄ 
 Cm 
 (5.7)
 
Cn

where CX and CZ are written as below [13]:

CX = CL sin α − CD cos α (5.8a)

CZ = − (CL cos α + CD sin α) (5.8b)


158

Other coefficients are defined as below:

C Y = C Yβ β (5.9a)

Cl = Clβ β + (c̄/2VA ) Clp Pb (5.9b)

Cm = Cm0 + Cmα α + (c̄/2VA ) Cmq Qb (5.9c)

Cn = Cnβ β + (c̄/2VA ) Cnr Rb (5.9d)

where q0 is the dynamic pressure, the characteristic length c̄ is the aerostat envelope

length, and VA is the aerodynamic velocity magnitude. The aerodynamic velocity V A =

[uA vA wA ]T is the difference of the aerostat envelope body axes velocity and the wind

velocity, also expressed in the body axes as below:


         
 uA   Ub   uwind   Ub   wx 
 =  V  − TB
         
 v = V − v E  wy 
  (5.10)
 A   b   wind   b 
         
wA Wb wwind Wb wz

The angle of attack and sideslip angle are then defined in the following manner:

 
−1 wA
α = tan (5.11)
uA

and
 
vA
β = sin−1 (5.12)
VA

The presence of VA in Equations (5.9) adds to non-linearity.

5.3.4 Apparent Mass Terms

The definition of apparent mass terms has been presented in section 2.3 of chapter 2.

When a body immersed in a fluid is accelerating, a certain amount of the surrounding

fluid moves with the body. That portion of the fluid set in motion is referred to as the

aerodynamic apparent mass, or simply apparent mass. For formulating the apparent mass

terms for equations of motion similar results from [15] have been used. The forces and

moments from apparent mass are found by relating the fluid’s kinetic energy to resultant
159

forces and moments, resulting in


   
 U̇b   ẇx 
B
   
F AM = −I AM 
 V̇b  − T E  ẇy 
  
   
Ẇb ẇz
    (5.13)
w
 x  U
 b 
B B
 B  B  
+  wy  − S (Ω) I AM  Vb 
S (Ω) I AM − I AM S (Ω) T E    
   
wz Wb

and    
 Ṗb   Pb 
B
   
M AM = −I AI 
 b −
Q̇  S (Ω) I AI  Q
 b 
 (5.14)
   
Ṙb Rb

where the common convention is used that a cross-product of any vector r with components

rx , ry and rz expressed in a frame C is written

 
 0 −rz ry 
C
 
r× = S (r) = 
 rx 0 −rx 
 (5.15)
 
−ry rx 0

Here the apparent mass and inertia matrices are written as below:
   
mxb 0 0  0 0 0
   
I AM =
 0 myb 0   and I AI =
0 Iat 0 

   
0 0 m zb 0 0 Iat

where mxb , myb , mzb are the aerodynamic apparent masses associated with accelerations

in body axes and Iat is the apparent mass moment of inertia in transverse direction which

are dependent on the aerostat envelope geometry and are already evaluated in chapter 2.

5.3.5 Terminal Tether Segment

The uppermost or terminal tether segment connects the aerostat and tether through a

spring with static stiffness Ks in parallel with a damper with damping coefficient Cv . The

line force or tether tension force is written in terms of components ∆x, ∆y and ∆z of the
160

difference vector formed by subtracting the earth-fixed position of the confluence point

and the uppermost lumped mass. The stretched length of the segment becomes lts =

∆x2 + ∆y 2 + ∆z 2 and the stretch rate is l˙ts . Using the difference vector components,
p

the line force vector is written as


 
 ∆x 
(FT + FT0 )  
FT =  ∆y  (5.16)
lts 



∆z

where FT0 is the initial tether tension in the terminal segment. The tether tension in the

terminal segment may be written using the equations of spring damper system [47] as

below:

FT = Ks (lts − l0 ) + Cv l˙ts (5.17)

The expression for l˙ts may be obtained by differentiating the expression for lts presented

above as
1
l˙ts = (∆x∆ẋ + ∆y∆ẏ + ∆z∆ż) (5.18)
lts

where ∆ẋ is obtained as difference of x-component of earth-fixed velocity of confluence

point and uppermost lumped mass. The earth-fixed position and velocity of confluence

point may be written in terms of earth-fixed position of aerostat envelope center of mass

position and body frame velocity components respectively as below:


   
x x
 c   e 
 y  =  y  + TE cp
   
 c   e  B r cg (5.19)
   
zc ze

   

 c  U
 b 
 ẏ  = T B  V  + T E
E B cp
   
 c   b  B S (Ω) r cg (5.20)
   
żc Wb

where r cp
cg is the position vector from the aerostat center of mass to the confluence point.
161

5.3.6 Aerostat Envelope Dynamic Equations

The rigid body equations of motion are obtained from Newton’s second law, which states

that the summation of all external forces acting on a body is equal to the time rate

of change of the momentum of the body, and the summation of the external moments

acting on the body is equal to the time rate of change of the angular momentum. Using

Equations (5.4-5.6), (5.13), and (5.16) with the linear momentum derivative results in the

following translation dynamic equations [15]:

 
 U̇b 
B
 
((ms + mg ) E 3 + I AM ) 
 V̇b  = F A + F W + F B − T E F T

 
Ẇb
   
U
 b  ẇ
 x 
B B
   
− S (Ω) ((ms + mg ) E 3 + I AM )  Vb  + I AM T E  ẇy 
   (5.21)
   
Wb ẇz
 
 wx 
+ B S (Ω) I AM − I AM B S (Ω) T B
  
E  wy 
 
 
wz

where E 3 is a 3 × 3 identity matrix. Similarly, the rotational dynamics become [15]

 

 b   
B
r cb B cp
   B
(I B + I AI ) 
 Q̇b 
 = M A + S cg F B − S r cg T E F T
 
Ṙb
  (5.22)
 Pb 
− B S (Ω) (I B + I AI ) 
 
Q
 b 

 
Rb

Where and I B is the aerostat envelope inertia matrix in the body axes represented by

Ixb , Iyb , Izb and Ixz b . The vector r cb


cg is the position vector from the aerostat center of

mass to the center of buoyancy. In terms of distances with respect to nose as presented in
162

cp
chapter 3 and Figure 3.1, the vectors r cb
cg and r cg may be written as below:

 
 XG − XB 
r cb
 
cg =
 0 
 (5.23)
 
ZB − ZG

 
 XG − XC 
r cp
 
cg =
 0 
 (5.24)
 
ZC − ZG

5.4 Tether Dynamic Model

The tether configuration, depicted in Figure 5.1, consists of N straight elastic segments

connected by universal joints or nodes at which the mass is assumed to be concentrated.

Since the tether mass is assumed lumped at nodes, the tether tension is constant along the

length of a segment. The location and velocity of j th node in the earth-fixed axes system

is given by the vector lj and v j with components as below:

   
 xj   uj 
   
lj = 
 y j
;
 vj = 
 v j

 (5.25)
   
zj wj

The vector l1 is the fixed location of the ground anchor point i.e. [0 0 0]T and the

vector lN is the location of uppermost lumped mass along the tether with coordinates

[xN yN zN ]T . The N th lumped mass is connected to the confluence point of the aerostat

through the N th segment. The tether tension vector in the j th segment is represented by

F Tj so that F TN = F T , the tether tension at the confluence point. Also the lumped mass

of each segment is constant as equal length segment has been considered and this mass is

obtained by dividing the total tether mass by N and represented by mt . The drag force

on the j th segment in earth-fixed axes is D j and is assumed to act on the j th lumped

mass [19]. Since we have considered the lumped masses, only translational equations are

required. Balancing the forces on the j th node (Figure 5.1), the equations of motion are
163

written as:
     

 j  u̇
 j   0 
     
 ÿj  = mt  v̇j  = −F Tj−1 + F Tj +  0
mt       + Dj
 (5.26)
     
z̈j ẇj mt g

The method to calculate F Tj has already been presented in Equation (5.16) where the

method has been applied to calculate F TN = F T . This method has to be applied for

the relevant segments for evaluation of F Tj for use in Equations (5.26) to get the tension

vectors in earth-fixed frame. For the computation of aerodynamic drag, D j , the segment

is considered to be a cylinder on which the normal drag force is proportional to the

square of the normal components of relative velocity [16], which is taken to be that of

the segment. The relative velocity is [uj − wx v j − wy wj − wz ]T and segment vector is

[xj+1 − xj yj+1 − yj zj+1 − zj ]T . The component of relative velocity which is normal

to the segment vector is computed and is taken for drag computation which is assumed to

act on j th lumped mass. The kinematic relations for the j th lumped mass are written as:
   
 ẋj   uj 
   
 ẏ  =  v  (5.27)
 j   j 
   
żj wj

5.5 Solution Procedure of Equations

Equations (5.1), (5.3), (5.21) and (5.22) are the 12 governing kinematic and dynamic

differential equations of motion for the aerostat envelope. The 12 unknown parameters

for the aerostat envelope are as below:

[Ub Vb Wb Pb Qb Rb Ψb Θb Φb xe ye ze ]T

The equations are dealt in matrix form itself as the same is allowed by the solver. Equa-

tions (5.26) and (5.27) are the 6 governing dynamic and kinematic equations of motion

for the j th lumped mass. Thus we have 6N differential equations of motion for the entire

tether. Also for the entire tether, we have 6N unknowns. The 6 unknown parameters for

j th lumped mass are as below:


164

[uj vj wj xj yj zj ]T

Thus we have a total of 6N + 12 differential equations and an equal number of unknowns

thereby making the system complete. The aerostat envelope is coupled with tether through

N th tether segment. These differential equations are of the form Ẋ p = f (t, X p ), where

X p is the unknown vector of dimension 6N + 12. These equations are solved using ode45

code of MATLAB R . The initial conditions are required for the solution of these equations

which are obtained from inspection.

5.6 System Input Parameters

To demonstrate application of the proposed nonlinear dynamic model and its utility in

analyzing the aerostat and tether dynamics along with tether loads, simulations of the test

aerostat were carried out with a 1000 m length of tether. The test aerostat is having a hull

volume of 2023 m3 and a flying altitude of 1000 m. The required characteristics of the test

aerostat and the aerodynamic properties are presented in Tables 3.2, 4.1 and 4.2. Table 3.2

presents the parameters for the test aerostat which were required for static analysis in

chapter 3 whereas additional parameters required for carrying dynamic stability analysis

in chapter 4 are presented in Tables 4.1 and 4.2. As outlined earlier in the present chapter,

additional parameters required for the test aerostat for carrying out nonlinear dynamic

analysis include the tether spring-mass-damper parameter, total buoyancy force and a

separate apparent moment of inertia in transverse direction. These additional parameters

required for the nonlinear analysis are presented in Table 5.1. Here the tether is divided

into 10 segments or lumped masses. The temperature condition for the analysis is taken as

ISA+15 deg, similar to all other analysis. The method to estimate all the input parameters

has been presented in chapter 2.

Table 5.1: Additional aerostat parameters required for nonlinear analysis


Parameter Estimated Value
Total buoyancy, B 21814.8 N
Lumped mass of cable segment, mt 29.4 kg
Length of cable segment, l0 100 m
Stiffness of cable segment, Ks 14058.5 N/m
Viscous damping coefficient of cable segment, Cv 79.54 N-s/m
Apparent moment of inertia in transverse direction, Iat 71174.1 kg-m2
165

5.7 Simulations Results and Discussions

The analysis has been carried out for three different types of simulated wind which varies

in magnitude as well as in direction as will be explained in the following subsections. Three

wind gusts have also been superimposed to a steady wind to get the gust response of the

test aerostat. The maximum wind speed considered in all these six simulations is 20 m/s,

which is generally the upper limit of operation of such class of aerostat. Simulations 1

and 2 are intended to predict aerostat envelope response to the winds which are varying

in magnitude from zero to 10 m/s and 20 m/s in a single step and two steps respectively

and also for comparison with with equilibrium analysis. In simulation 3, the wind speed

as well as direction changes to simulate the more practical wind. Simulations 4, 5 and 6

are intended to simulate the wind gust along the horizontal, vertical and side direction

along with the constant horizontal wind. In all the simulations, the aerostat is initially

directly above the anchor point with no wind and hence the tether vertical. Orientation

of the aerostat is initially facing in the opposite direction of ie , with Ψb being π. The

initial position of confluence point is at 1000 m height from the ground anchor point. The

initial earth-fixed position of center of mass of aerostat envelope can be estimated using

Equation (5.19) which comes out to be (7 m, 0 m, -1009.1 m). The simulation has been

carried out for 350 s time interval for all the cases since the outputs are stabilized during

the interval. The initial tether tension is calculated as FT0 = B − mb g.

5.7.1 Simulation 1: Wind Variation in Magnitude

Wind condition for this simulation are such that the wind is increased from 0 to 10 m/s

over a 10 s interval, with a direction of 0 deg. The wind speed remains at the constant value

of 10 m/s after that with same direction. This wind condition is depicted in Figure 5.2.

The variation of body frame linear velocity components, body frame angular velocity

components, Euler angles, earth-fixed position of aerostat envelope center of mass and

tether tension at the confluence point for this simulation are presented in Figures 5.3

to 5.7 respectively.

The forward velocity component decreases from 0 to – 4.1 m/s in the initial time

interval of 26 s, again reaches to 0 at 64 s and then with small overshoot, stabilizes at 130

s to its initial value as observed in Figure 5.3. The vertical velocity varies in the initial
166

time interval of 100s and the maximum vertical velocity is about 0.32 m/s. It is seen in

Figures 5.3, 5.4, 5.5 and 5.6 that the lateral components of linear velocity, angular velocity,

Euler angle as well as earth-fixed position of aerostat envelope mass center remain zero in

the time interval as expected because wind has been applied only in longitudinal direction.

The pitch rate varies in the initial time interval of 76 s and stabilizes after that as depicted

in Figure 5.4. The maximum pitch rate was also observed to be 0.36 deg/s at a time of

16 s. The pitch angle varies in the initial time interval of 125 s and stabilizes after that

at a value of 2.1 deg as seen in Figure 5.5. The mass center of aerostat envelope moves to

a maximum distance of 150.7 m at 63.3 s in the horizontal direction and around the same

time the mass center of aerostat envelope comes down to a height of 996.6 m as observed

in Figure 5.6. It is also observed in Figure 5.6 that the mass center stabilizes at 133m

distance in horizontal direction and at 999 m vertical height.

The tether tension force variation in the terminal tether segment i.e. at the confluence

point is shown in Figure 5.7. Initially, the tension force oscillates in the first 150 s, with

a peak value of approximately 7716 N. As the magnitude of wind remains constant, the

tension approaches a steady-state value of 7521 N.

25

20

15
Horizontal Wind (m/s)

10

−5

−10
0 50 100 150 200 250 300 350
time(s)

Figure 5.2: Wind variation in magnitude for simulation 1


167

5
Ub(m/s)

−5
0 50 100 150 200 250 300 350
t(s)
5
Vb(m/s)

−5
0 50 100 150 200 250 300 350
t(s)
0.5
Wb(m/s)

−0.5
0 50 100 150 200 250 300 350
t(s)

Figure 5.3: Body frame linear velocity components variation for simulation 1

5
Pb(deg/s)

−5
0 50 100 150 200 250 300 350
t(s)
0.5
Qb(deg/s)

−0.5
0 50 100 150 200 250 300 350
t(s)
5
Rb(deg/s)

−5
0 50 100 150 200 250 300 350
t(s)

Figure 5.4: Body frame angular velocity components variation for simulation 1
168

Φb(deg)
0

−5
0 50 100 150 200 250 300 350
t(s)
5
Θb(deg)

−5
0 50 100 150 200 250 300 350
t(s)
185
Ψb(deg)

180

175
0 50 100 150 200 250 300 350
t(s)

Figure 5.5: Euler angles variation for simulation 1

200
xe(m)

100

0
0 50 100 150 200 250 300 350
t(s)
200
ye(m)

−200
0 50 100 150 200 250 300 350
t(s)
−990
ze(m)

−1000

−1010
0 50 100 150 200 250 300 350
t(s)

Figure 5.6: Earth-fixed position of center of mass variation for simulation 1


169

7800

7700

7600

7500
FT(N)

7400

7300

7200

7100

7000
0 50 100 150 200 250 300 350
t(s)

Figure 5.7: Tether tension at confluence point variation for simulation 1

5.7.2 Simulation 2: Stepped Wind Variation in Magnitude

Wind condition for this simulation are such that the wind is increased from 0 to 10 m/s

over a 10 s interval, with a direction of 0 deg. The wind speed remains at the constant

value of 10 m/s till 170 s in the same direction. During the period between 170 s and

180 s, the wind speed increases from 10 m/s to 20 m/s in the same direction. This wind

condition is depicted in Figure 5.8. The variation of wind speed from 10 m/s to 20 m/s in

the time interval 170 s to 180 s is considered because by this time all the parameters of the

test aerostat gets stabilized and the effect of variation of stepped wind can be captured

effectively. The variation of body frame linear velocity components, body frame angular

velocity components, Euler angles, earth-fixed position of aerostat envelope center of mass

and tether tension at the confluence point for this simulation are presented in Figures 5.9

to 5.13 respectively. It is observed that till 170 s time the variation in parameters of test

aerostat with time is exactly same as simulation 1 since the wind is taken same till that

time interval hence the variation after 170 s time has been discussed.

The forward velocity component dips to a minimum value of - 6.7 m/s at the time 192
170

s and then stabilizes to its initial value i.e. 0 after 276 s as observed in Figure 5.9. The

vertical velocity varies till about 250 s and the maximum vertical velocity is about 1.4 m/s

at about 183 s. It is again observed in Figures 5.9, 5.10, 5.11 and 5.12 that the lateral

components of linear velocity, angular velocity, Euler angle as well as earth-fixed position

of aerostat envelope center of mass remain zero in the time interval as expected because

wind has been applied only in longitudinal direction. The pitch rate varies in the time

interval between 170 and 230 s and stabilizes after that as depicted in Figure 5.10. The

magnitude of maximum pitch rate was observed to be 0.62 deg/s at a time of 175 s. The

pitch angle dips to -1.5 deg at 183 s and stabilizes after 283 s at a value of 3.7 deg as seen

in Figure 5.11. The center of mass of aerostat envelope moves gradually to a maximum

distance of 316 m in the horizontal direction and the center of mass of aerostat envelope

comes down gradually to a height of 944 m as observed in Figure 5.12.

25

20

15
Horizontal Wind (m/s)

10

−5

−10
0 50 100 150 200 250 300 350
time(s)

Figure 5.8: Wind variation in magnitude for simulation 2


171

Ub(m/s) 10

−10
0 50 100 150 200 250 300 350
t(s)
5
Vb(m/s)

−5
0 50 100 150 200 250 300 350
t(s)
2
Wb(m/s)

−2
0 50 100 150 200 250 300 350
t(s)

Figure 5.9: Body frame linear velocity components variation for simulation 2

5
Pb(deg/s)

−5
0 50 100 150 200 250 300 350
t(s)
1
Qb(deg/s)

−1
0 50 100 150 200 250 300 350
t(s)
5
Rb(deg/s)

−5
0 50 100 150 200 250 300 350
t(s)

Figure 5.10: Body frame angular velocity components variation for simulation 2
172

Φb(deg)
0

−5
0 50 100 150 200 250 300 350
t(s)
5
Θb(deg)

−5
0 50 100 150 200 250 300 350
t(s)
180
Ψb(deg)

180

180
0 50 100 150 200 250 300 350
t(s)

Figure 5.11: Euler angles variation for simulation 2

400
xe(m)

200

0
0 50 100 150 200 250 300 350
t(s)
500
ye(m)

−500
0 50 100 150 200 250 300 350
t(s)
−900
ze(m)

−1000

−1100
0 50 100 150 200 250 300 350
t(s)

Figure 5.12: Earth-fixed position of center of mass variation for simulation 2


173

11000

10500

10000

9500
FT(N)

9000

8500

8000

7500

7000
0 50 100 150 200 250 300 350
t(s)

Figure 5.13: Tether tension at confluence point variation for simulation 2

The tether tension force variation at the confluence point is shown in Figure 5.13. The

tether tension force oscillates between 170 s and 264 s with a peak value of approximately

10830 N. As the magnitude of wind remains constant, the tension approaches a steady-

state value of 10400 N.

5.7.3 Simulation 3: Stepped Wind Variation both in Magnitude and


Direction

The wind variation for this simulation is similar to as presented in reference [15]. Wind

conditions are such that the wind is increased from 0 to 10 m/s over a 10 s interval, with

a direction of 15 deg. The wind speed remains at the constant value of 10 m/s till 170 s

in the same direction. During the period between 170 s and 180 s, the wind rotates to -15

deg while also increasing to 20 m/s. The wind after 180 s remains same in magnitude and

direction. This wind condition is depicted in Figure 5.14. The variation of body frame

linear velocity components, body frame angular velocity components, Euler angles, earth-

fixed position of aerostat envelope center of mass and tether tension at the confluence
174

point for this simulation are presented in Figures 5.15 to 5.19 respectively.

The forward, side and vertical velocity are shown in Figure 5.15. The forward velocity

component dips to a minimum value of - 7 m/s at the time 192 s and then stabilizes to its

initial value i.e. 0. The maximum side velocity observed in is 4 m/s at 200s. As opposed

to simulation 1 and simulation 2, side velocity component is observed in simulation 3

because of presence of inclined wind in horizontal plane. The maximum vertical velocity

is about 1.5 m/s at about 183 s. All the velocity components approaches zero as the

wind conditions remain at constant values. The roll, pitch and yaw rates are presented in

Figure 5.16. The magnitude of maximum roll rate, pitch rate and yaw rate are 1.7 deg/s,

0.84 deg/s and 3.7 deg/s at time 181 s, 175 s and 175 s respectively.

Aerostat roll, pitch and yaw angles are shown in Figure 5.17. The roll angle changes

only in the interval where the wind direction changes. The magnitude of maximum roll

angle is about 2.6 deg at 183 s. Similarly, pitching dynamics of the aerostat are most

significant after wind changes. In response to increasing winds and increasing tether

drag, the steady-state pitch angle increases to maintain equilibrium. As earlier noted, the

aerostat has an initial yaw angle of 180 deg. It varies during the first 100 s and appears

to be reaching a steady-state value of 195 deg until the change in the wind conditions at

170 s. A variation in yaw angle response is noted again before a steady-state value of 165

deg for the yaw angle is attained.

The displacement response of the aerostat due to the specified wind parameters is

shown in Figure 5.18. The altitude of the aerostat decreases from an initial height of

1009.1 m to approximately 999 m over the first 170 s of the simulation, while the aerostat

moves approximately 128 m in the horizontal direction. Altitude eventually decreases to

944 m, and the horizontal displacement reaches a final value of 325 m as the wind increases.

The tether tension force variation at the confluence point is shown in Figure 5.19.

Initially, the tension force oscillates in the first 100 s with a peak value of approximately

7719 N. As the magnitude of wind remains constant, the tension approaches a steady-state

value of 7522 N. Once the wind changes direction, an oscillatory response is seen between

170 and 250 s, and the tension force peaks at a value of 10930 N. As the simulation

progresses and the wind conditions again remain constant, a steady-state value of 10400

N at the confluence point is eventually reached.


175

Horizontal Wind (m/s)


20

10

−10
0 50 100 150 200 250 300 350
time(s)

20
Wind Direction (deg)

10

−10

−20
0 50 100 150 200 250 300 350
time(s)

Figure 5.14: Wind variation in magnitude and direction for simulation 3

10
Ub(m/s)

−10
0 50 100 150 200 250 300 350
t(s)
5
Vb(m/s)

−5
0 50 100 150 200 250 300 350
t(s)
2
Wb(m/s)

−2
0 50 100 150 200 250 300 350
t(s)

Figure 5.15: Body frame linear velocity components variation for simulation 3
176

Pb(deg/s)
0

−5
0 50 100 150 200 250 300 350
t(s)
1
Qb(deg/s)

−1
0 50 100 150 200 250 300 350
t(s)
5
Rb(deg/s)

−5
0 50 100 150 200 250 300 350
t(s)

Figure 5.16: Body frame angular velocity components variation for simulation 3

5
Φb(deg)

−5
0 50 100 150 200 250 300 350
t(s)
5
Θb(deg)

−5
0 50 100 150 200 250 300 350
t(s)
200
Ψb(deg)

180

160
0 50 100 150 200 250 300 350
t(s)

Figure 5.17: Euler angles variation for simulation 3


177

xe(m) 400

200

0
0 50 100 150 200 250 300 350
t(s)
500
ye(m)

−500
0 50 100 150 200 250 300 350
t(s)
−900
ze(m)

−1000

−1100
0 50 100 150 200 250 300 350
t(s)

Figure 5.18: Earth-fixed position of center of mass variation for simulation 3

11000

10500

10000

9500
FT(N)

9000

8500

8000

7500

7000
0 50 100 150 200 250 300 350
t(s)

Figure 5.19: Tether tension at confluence point variation for simulation 3


178

5.7.4 Simulation 4: Wind Gust in Horizontal Direction

An investigation of the effects from wind gusts on the aerostat is shown using multiple

simulations of the nonlinear model under different conditions. Wind condition for this

simulation are such that the wind is increased from 0 to 10 m/s over a 10 s interval, with a

direction of 0 deg. The wind speed remains at the constant value of 10 m/s after that with

same direction till 170 s. During 170 s to 220 s, horizontal wind gusts were superimposed

to the horizontal wind. Wind gusts were simulated by a Tg = 50 s ‘1-cosine’ gust in addition

to the 10 m/s constant wind as below [2, 15]:

   
Um 2π
Ug = 1 − cos t (5.28)
2 Tg

Four cases were simulated with gusts of maximum magnitude Um = 2.5, 5.0, 7.5, and 10

m/s, respectively, after which the wind speed remained constant at its previous value of

10 m/s, as shown in Figure 5.20.

The forward and vertical velocities of aerostat are shown in Figure 5.21. It is observed

that with increasing gust magnitude, the magnitude of both forward and vertical speed

increases. Figure 5.22 shows the pitch rate and pitch angle. The steady state pitch angle

before the gust was 2.1 deg with a maximum pitch achieved for each just after the gust

ends. The maximum value of 5.7 deg is noted in the 10 m/s case at t=220 s. The 2.5,

5.0, and 7.5 m/s gust simulations produce maximum pitch angles of 2.9, 3.8, and 4.8 deg

at around 220 s. The maximum pitch rate was 0.4 deg/s and was also observed for the 10

m/s gust.

The altitude and horizontal displacement of the aerostat are shown in Figure 5.23. For

each simulation, the horizontal displacement reaches a maximum value at t=216 s, with

the maximum displacement of approximately 273 m occurring as a result of the 10 m/s

gust. The altitude decreases as the wind gusts increase, and the minimum value recorded

for the 10 m/s gust is about 970 m. Both the horizontal displacement and altitude converge

on a steady-state value as the wind speed drops back to a constant value of 10 m/s.
179

25

20

15
Horizontal Wind (m/s)

10

0
Um=2.5 m/s
Um=5.0 m/s
−5 Um=7.5 m/s
Um=10.0 m/s
−10
0 50 100 150 200 250 300 350
time(s)

Figure 5.20: Horizontal wind gust for simulation 4

0
Ub(m/s)

Um=2.5 m/s
−5 Um=5.0 m/s
Um=7.5 m/s
Um=10.0 m/s

−10
0 50 100 150 200 250 300 350
t(s)

1.5

1
Wb(m/s)

0.5

−0.5
0 50 100 150 200 250 300 350
t(s)

Figure 5.21: Aerostat forward and vertical speed variation for simulation 4
180

0.6
Um=2.5 m/s
Um=5.0 m/s
0.4 Um=7.5 m/s
Qb(deg/s)
Um=10.0 m/s
0.2

−0.2

−0.4
0 50 100 150 200 250 300 350
t(s)

4
Θb(deg)

−2
0 50 100 150 200 250 300 350
t(s)

Figure 5.22: Aerostat pitch rate and pitch angle variation for simulation 4

300
Um=2.5 m/s
Um=5.0 m/s
Um=7.5 m/s

200 Um=10.0 m/s


xe(m)

100

0
0 50 100 150 200 250 300 350
t(s)

−960

−970

−980
ze(m)

−990

−1000

−1010
0 50 100 150 200 250 300 350
t(s)

Figure 5.23: Aerostat horizontal and vertical displacement variation for simulation 4
181

9500
Um=2.5 m/s
Um=5.0 m/s
Um=7.5 m/s
9000 Um=10.0 m/s

8500
FT(N)

8000

7500

7000
0 50 100 150 200 250 300 350
t(s)

Figure 5.24: Tether tension at confluence point variation for simulation 4

The tension at the confluence point for each gust is shown in Figure 5.24. In all cases

the maximum tension occurs as the gust reaches around its largest value, after which the

tension decrease dramatically as the aerostat surges forward after the gust is over. Tether

tension at the aerostat is highly dependent on the gust magnitude, with the maximum

tension of 9103 N appearing in the case of the 10 m/s gust at a time of 197 s, which is

around the peak of the wind gust. All four cases eventually converge on a tension of 7523

N as the winds once again become constant after their gusts.

5.7.5 Simulation 5: Wind Gust in Vertical Direction

Wind condition for this simulation are such that the wind is increased from 0 to 10 m/s

over a 10 s interval, with a direction of 0 deg. The wind speed remains at the constant

value of 10 m/s after that with same direction till 170 s. During 170 s to 220 s, vertically

upward wind gusts were superimposed to the horizontal wind with Tg =50 s and Um =2.5,

5.0, 7.5, and 10 m/s, respectively in Equation (5.28), after which the wind speed remained

constant at its previous value of 10 m/s, as shown in Figure 5.25.


182

The forward and vertical velocities of aerostat are shown in Figure 5.26. It is observed

that with increasing gust magnitude, the magnitude of both forward and vertical speed

increases but the forward velocity is strongly dependent. For the 10 m/s gust the maximum

forward speed is observed to be 14.2 m/s at 198 s. Figure 5.27 shows the pitch rate and

pitch angle. The minimum pitch angle is observed to be around -16.7 deg at 203 s. The

aerostat pitches down so much because vertically upward gust increases the tether tension

very much and the aerostat is held at the confluence point resulting in pitching down

action. The 2.5, 5.0, and 7.5 m/s gust simulations produce minimum pitch angles of -11.8,

-7.2 and -2.6 deg at around 200 s. The maximum pitch rate was 1.7 deg/s at 213 s and

was observed for the 10 m/s gust.

The altitude and horizontal displacement of the aerostat are shown in Figure 5.28. It

is observed that due to vertical gust of 7.5 m/s and 10 m/s, the aerostat comes ahead

of the launch point. The maximum distance it comes ahead of launch point is for 10

m/s gust is about 147 m at 214 s. The altitude variation due to the gust also becomes

typical as for all other gust except 2.5 m/s, the aerostat envelope crosses its initial height.

The maximum height of the aerostat is obtained as 1020 m for 10 m/s gust. Both the

horizontal displacement and altitude converge on a steady-state value as the wind speed

drops back to a constant value of 10 m/s.

The tension at the confluence point for each gust is shown in Figure 5.29. In all cases

the maximum tension occurs as the gust reaches around its largest value, after which the

tension decrease dramatically. Tether tension at the aerostat is strongly dependent on the

gust magnitude, with the maximum tension of 21790 N appearing in the case of the 10

m/s gust at a time of 195 s, which is around the peak of the wind gust. All four cases

eventually converge on a tension of 7523 N as the winds once again become constant after

their gusts.
183

15
Horizontal Wind(m/s)
10

−5

−10
0 50 100 150 200 250 300 350
time(s)

15
Um=2.5 m/s
Vertical Wind(m/s)

10 Um=5.0 m/s
Um=7.5 m/s
5 Um=10.0 m/s

−5

−10
0 50 100 150 200 250 300 350
time(s)

Figure 5.25: Vertical wind gust for simulation 5

15
Um=2.5 m/s

10 Um=5.0 m/s
Um=7.5 m/s
Ub(m/s)

Um=10.0 m/s
5

−5

−10
0 50 100 150 200 250 300 350
t(s)

0
Wb(m/s)

−1

−2

−3

−4
0 50 100 150 200 250 300 350
t(s)

Figure 5.26: Aerostat forward and vertical speed variation for simulation 5
184

Qb(deg/s) 1

−1

−2
0 50 100 150 200 250 300 350
t(s)

0
Θb(deg)

−5

−10 Um=2.5 m/s


Um=5.0 m/s

−15 Um=7.5 m/s


Um=10.0 m/s

−20
0 50 100 150 200 250 300 350
t(s)

Figure 5.27: Aerostat pitch rate and pitch angle variation for simulation 5

200

100
xe(m)

0
Um=2.5 m/s
Um=5.0 m/s
−100 Um=7.5 m/s
Um=10.0 m/s

−200
0 50 100 150 200 250 300 350
t(s)

−980

−990
ze(m)

−1000

−1010

−1020
0 50 100 150 200 250 300 350
t(s)

Figure 5.28: Aerostat horizontal and vertical displacement variation for simulation 5
185

4
x 10
2.2
Um=2.5 m/s
Um=5.0 m/s
2
Um=7.5 m/s
Um=10.0 m/s
1.8

1.6
FT(N)

1.4

1.2

0.8

0.6
0 50 100 150 200 250 300 350
t(s)

Figure 5.29: Tether tension at confluence point variation for simulation 5

5.7.6 Simulation 6: Wind Gust in Side Direction

Wind conditions for this simulation are such that the wind is increased from 0 to 10 m/s

over a 10 s interval, with a direction of 0 deg. The wind speed remains at the constant

value of 10 m/s after that with same direction till 170 s. During 170 s to 220 s, side wind

gusts were superimposed to the horizontal wind with Tg =50 s and Um =2.5, 5.0, 7.5, and

10 m/s, respectively in Equation (5.28), after which the wind speed remained constant at

its previous value of 10 m/s, as shown in Figure 5.30.

The forward and side velocities of aerostat are shown in Figure 5.31. It is observed

that with increasing gust magnitude, the magnitude of both forward and vertical speed

increases. For the 10 m/s gust the minimum side speed is observed to be -5.3 m/s at 195

s. Figure 5.32 shows the pitch rate and yaw rate. Both these rates increase with increase

in wind gust. The variation of roll angle and yaw angle is plotted in Figure 5.33. The

minimum roll angle is observed to be -1.9 deg at 216 s and the maximum yaw angle is

observed to be 208 deg at 232 s.

The altitude and side displacement of the aerostat are shown in Figure 5.34. It is
186

observed that side position increases due to increase in gust magnitude. The maximum

side position for all the cases occur at around t=211 s. The maximum magnitude of side

position for 2.5, 5.0, 7.5 and 10 m/s gusts are 32 m, 64 m, 96 m and 129 m respectively.

The altitude decreases as the wind gusts increase, and the minimum value recorded for

the 10 m/s gust is about 990 m. Both the side displacement and altitude converge on a

steady-state value as the wind speed drops back to a constant value of 10 m/s.

The tension at the confluence point for each gust is shown in Figure 5.35. In all cases

the maximum tension occurs as the gust reaches around its largest value, after which the

tension decreases dramatically. Tether tension at the aerostat is highly dependent on the

gust magnitude, with the maximum tension of 8079 N appearing in the case of the 10

m/s gust at a time of 199 s, which is around the peak of the wind gust. All four cases

eventually converge on a tension of 7523 N as the winds once again become constant after

their gusts.

15
Horizontal Wind(m/s)

10

−5

−10
0 50 100 150 200 250 300 350
time(s)

15
Um=2.5 m/s
10
Side Wind(m/s)

Um=5.0 m/s
Um=7.5 m/s
5 Um=10.0 m/s

−5

−10
0 50 100 150 200 250 300 350
time(s)

Figure 5.30: Side wind gust for simulation 6


187

2
Ub(m/s)

−2

−4

−6
0 50 100 150 200 250 300 350
t(s)

0
Vb(m/s)

Um=2.5 m/s
−5 Um=5.0 m/s
Um=7.5 m/s
Um=10.0 m/s

−10
0 50 100 150 200 250 300 350
t(s)

Figure 5.31: Aerostat forward and side speed variation for simulation 6

0.6

0.4
Qb(deg/s)

0.2

−0.2

−0.4
0 50 100 150 200 250 300 350
t(s)

1
Rb(deg/s)

0
Um=2.5 m/s
Um=5.0 m/s
−1 Um=7.5 m/s
Um=10.0 m/s

−2
0 50 100 150 200 250 300 350
t(s)

Figure 5.32: Aerostat pitch rate and yaw rate variation for simulation 6
188

Φb(deg) 1

−1

−2
0 50 100 150 200 250 300 350
t(s)

210
Um=2.5 m/s

200 Um=5.0 m/s


Um=7.5 m/s
Ψb(deg)

Um=10.0 m/s
190

180

170

160
0 50 100 150 200 250 300 350
t(s)

Figure 5.33: Aerostat roll angle and yaw angle variation for simulation 6

150
Um=2.5 m/s
Um=5.0 m/s
100 Um=7.5 m/s
Um=10.0 m/s
ye(m)

50

−50
0 50 100 150 200 250 300 350
t(s)

−990

−995
ze(m)

−1000

−1005

−1010
0 50 100 150 200 250 300 350
t(s)

Figure 5.34: Aerostat side and vertical position variation for simulation 6
189

8200
Um=2.5 m/s
Um=5.0 m/s
Um=7.5 m/s
8000
Um=10.0 m/s

7800
FT(N)

7600

7400

7200

7000
0 50 100 150 200 250 300 350
t(s)

Figure 5.35: Tether tension at confluence point variation for simulation 6

5.8 Comparison of Nonlinear Analysis with Available Ex-

perimental Data

The nonlinear model of tethered aerostat developed in this chapter has been compared

with the available limited flight trial data. The experimental flight data was available for

the test aerostat with a different operating and flight configuration than that mentioned in

Tables 3.2 and 5.1. The flight data was available for the temperature condition ISA+12 deg

and hence air density and total buoyancy force will be affected because the values presented

in these tables correspond to ISA+15 deg. The total mass as well as center of mass location

and confluence point location for the test aerostat envelope for the experimental flight are

presented in Table 5.2 along with air density and buoyancy.

The flight data is available in terms of wind speed, pitch angle and tether tension.

The wind speed was recorded using wind speed sensor mounted above the top fin to get

undisturbed wind. The pitch angle was recorded using an attitude sensor mounted on the

belly of the aerostat. Zero error correction is required for pitch angle due to its initial

inclination with respect to the body X-axis of the aerostat envelope. The tether tension
190

Table 5.2: The operating conditions and flight configuration for flight data of test aerostat
Parameter Value
Air density, ρa 1.0662 kg/m3
Total buoyancy, B 22038 N
Combined envelope mass, mb 1372.6 kg
Axial distance of CP from nose, XC 10.2 m
Axial distance of CM from nose, XG 18.39 m
Vertical distance of CM from nose, ZG 1.79 m

was recorded using a load cell properly mounted at the confluence point. The accuracy

of load cell measurement also depends upon the accuracy of mounting load cell at the

confluence point. The flight data was collected and recorded continuously in real time at

the rate of one sample every second for 150 seconds.

The available data for wind speed variation in x-direction for the time 150 s has been

plotted in Figure 5.36. The wind speed recorded at time t=0 is 3.4 m/s. The wind speed

varies with time as shown, sometimes going up and at other times coming down. The

highest increase in wind speed is observed between t=120 s and t=121 s where the wind

speed changes from 3.4 m/s to 4.6 m/s during 1 s interval. Similarly the highest decrease

in wind speed is observed between t=129 s and t=130 s where the wind speed changes

from 4.6 m/s to 3.4 m/s during 1 s interval. The wind speed at the time t=150 s is 4.6

m/s.

The initial conditions for body fixed linear and angular velocity components are taken

to be zero. Initial yaw angle is taken to be π as previously and initial pitch angle is taken

same as the recorded pitch angle at t=0. The initial roll angle is taken to be zero. The

wind speed in horizontal direction as shown in Figure 5.36 is applied to the nonlinear

dynamic model of the tethered aerostat and the simulation was carried out for 150 s. The

wind speed is available only at 150 data points and the intermediate values are derived

using linear interpolation between data points.

The variation of computed and measured pitch angle with time has been presented in

Figure 5.37. It is observed that the computed and measured pitch angles are in agreement

to some extent, at least regarding the overall range and for many small time intervals

along the 150 s duration. The RMS pitch angle error during this 150 s interval comes out

to be 3.60%. It is also observed that increasing the wind speed increases the pitch angle

of the aerostat envelope for both the computed and measured values. This is particularly

observed at t=67 s, t=120 s and t=138 s. The same trend was also predicted during
191

the equilibrium analysis of this aerostat in chapter 3. Another important phenomenon is

observed at t=120 s where the wind speed change was highest as discussed previously. Due

to sudden change in wind speed, a substantial peak is observed in the available data, which

is also captured by the nonlinear dynamic model as indicated in Figure 5.37, indicating

the efficacy of the model.

The variation of computed and measured tether tension with time has been presented

in Figure 5.38. It is observed that the computed and measured tether tensions are in

agreement only for some small time intervals along the 150 s duration. The RMS tether

tension error during this 150 s interval comes out to be 3.56%. The trend of some tether

tension peaks such as that at t=128 s are also observed to be similar. But it is also

observed that despite some agreements, the large variation in measured tether tension is

not accurately predicted by the computed model. Few possible reasons for this may be

limitation of lumped mass model for tether tension prediction, presence of noise in the

recorded tether tension data and improper positioning of load cell at the confluence point.

5
Wind speed(m/s)

0
0 50 100 150
t(s)

Figure 5.36: Measured wind speed in horizontal direction


192

−4
Computed
Measured
−5

−6
Pitch angle(deg)

−7

−8

−9

−10
0 50 100 150
t(s)

Figure 5.37: Comparison between computed and measured pitch angle

10000
Computed
Measured
9500
Tether Tension at CP(N)

9000

8500

8000

7500

7000
0 50 100 150
t(s)

Figure 5.38: Comparison between computed and measured tether tension


193

5.9 Comparison of Equilibrium Analysis with Available Ex-

perimental Data

The equilibrium analysis gives a quick estimation of trim angle for a given wind speed

and other aerostat parameters as discussed in detail in chapter 3. To compare the pitch

angle estimated by equilibrium analysis with the actual flight pitch angle, it is required

to generate flight data for long duration at constant wind speed so that any transient

behavior can be ruled out. The available wind data for 150 s duration does not correspond

to constant wind speed for long duration as the wind speed changes every now and then

as seen in Figure 5.36. So this data is not very suitable for comparison with equilibrium

analysis. However to get a first hand picture, this wind is taken for comparison with

equilibrium analysis results. To just see the closeness of equilibrium pitch angle, the

theoretical pitch angle is evaluated corresponding to the minimum and maximum wind

speed in the 150 s duration. The minimum wind speed is 2.5 m/s whereas the maximum

wind speed is 4.6 m/s. Corresponding to these range of wind speed, the theoretical

equilibrium pitch angle range is -8.5 deg to -7.3 deg whereas the measured pitch angle

range in the same duration is -8.0 deg to -6.1 deg. This demonstrates that equilibrium

analysis also predicts close values for pitch angle even for variable wind conditions if wind

speeds are low.

5.10 Comparison of Nonlinear Analysis with Equilibrium

Analysis

The equilibrium analysis carried out in chapter 3 considered a uniform horizontal wind

acting on the aerostat envelope continuously. Under this condition, trim angle of attack

was estimated for the range of wind speed. In the present chapter a nonlinear aerostat

model has been developed which can generate response corresponding to a time dependent

wind. Hence a comparison between these two methods can only be done for the response of

angle of attack corresponding to a horizontal wind, which after initial variation becomes

constant with time for long duration. Such winds have been presented in Figures 5.2

and 5.8 for simulation 1 and 2 respectively as described in section 5.7 for the test aerostat.

After initial variation, the horizontal wind stabilizes at 10 m/s for simulation 1 and at 20
194

m/s for simulation 2. The pitch angle response for simulations 1 and 2 have been presented

in Figures 5.5 and 5.11 respectively where the pitch angle of the test aerostat is about 0

deg for zero wind speed. It can be observed in these figures that the stabilized pitch angles

corresponding to stabilized 10 m/s and 20 m/s horizontal wind are 2.094 deg and 3.705

deg whereas the trim angle corresponding 10 m/s and 20 m/s uniform horizontal wind

speed comes out to be 2.095 deg and 3.708 deg in Figure 3.6 using equilibrium analysis

approach. The exact matching of pitch angles estimated using nonlinear analysis and

equilibrium analysis for stabilized winds demonstrate that these two approaches are in

agreement.

5.11 Chapter Summary

A nonlinear dynamic analysis model has been developed for tethered aerostat in this

chapter. The equations of motion for tethered aerostat contain the buoyancy, tether

tension and apparent mass and inertia terms in addition to gravity and aerodynamic terms

as for conventional aircraft. The force and moment equations are kept in matrix form itself

for direct application for solving these equations. The equations of motion for the aerostat

envelope are written in body components similar to a conventional aircraft. The tether

has been modeled using lumped mass approach where the tether is divided into segments

and the mass of the segment is assumed lumped at the lower node. The aerostat envelope

is coupled with the tether through the terminal segment. The analysis model has been

applied to the test aerostat for six cases of simulated winds including gusts. The results

are presented for body frame linear velocity components, angular velocity components

and Euler angles, earth-fixed position of aerostat envelope, and tether tension at the

confluence point. Finally the nonlinear analysis results and equilibrium analysis results are

compared with the available experimental flight data. A comparison of nonlinear analysis

and equilibrium analysis has also been done for stabilized wind speed. The comparison

results demonstrate that the nonlinear model and equilibrium model are in agreement and

that the nonlinear model can be applied to an aerostat to generate practical results.
Chapter 6

Concluding Remarks

6.1 Innovative Research Work Carried Out

The objective of the present research was to investigate the static and dynamic analysis

of an airborne tethered aerostat, which was carried out during its design and development

phase. At the same time the objective was also to estimate all the input parameters

required for carrying out the static and dynamic analysis of the aerostat. During the

research, following innovative tasks were carried out:

1. The modern tool of solid modeling of the aerostat envelope has been used to gener-

ate geometric, mass and inertia properties required for static and dynamic analysis.

Since the majority of ‘hull only’ properties can be estimated analytically, a compar-

ative study has also been carried out between analytical method and solid modeling

method for ‘hull only’ properties to gain confidence before applying the solid model-

ing tool for the entire aerostat envelope. The comparison indicates that the results

for both the methods are same.

2. The lift, drag and moment coefficients of the aerostat envelope have been extensively

examined using three different approaches namely semi-empirical method, compu-

tational fluid dynamics and wind tunnel testing. The agreement of aerodynamic

coefficients for these three methods is very good within angle of attack range from -5

deg to +5 deg. Beyond this range, the difference, in general, increases with increase

in absolute angle of attack. A flow separation is observed in both computational

fluid dynamics and wind tunnel testing approach beyond a certain angle of attack

whereas no flow separation is observed in semi empirical method. The semi empirical

195
196

method is based on the relations where boundary layer effects have not been taken

into account and hence it is not expected also to predict flow separation. It is also

observed that the trend of variation of the coefficients with angle of attack remains

same in all the three methods i.e. lift coefficients increases with increase in angle of

attack, the drag coefficient first decreases and then increases with increase in angle of

attack thereby forming a drag polar and the pitching moment coefficient about nose

decreases with increase in angle of attack. Also the lift and moment coefficients can

be approximated by a linear curve whereas the drag coefficient can be approximated

by a quadratic curve.

3. Equilibrium studies of the aerostat have been proposed to estimate tether tension,

trim angle of attack and equilibrium cable shapes under uniform horizontal wind

speeds. The criterion for static stability is also proposed and sensitivity studies

have been carried out for possible effect on trim angle of attack and static stability.

Approximations have been proposed for estimating cable shape to account for air

density and uniform wind speed variation along the vertical direction. It is observed

that in general the tether tension increases with increase in wind speed. The trim

angle of attack increases with both horizontal wind speed and temperature. By

applying the static stability criteria, it is observed that the test aerostat is statically

stable in the entire operating wind speed and the slope of the total pitching moment

coefficient about center of gravity increases with increase in wind speed and assumes

asymptotic value close to zero at higher wind speeds. The sensitivity analysis results

suggest that the trim angle of attack is sensitive to the changes in aerostat parameters

but the the slope of total moment coefficient about center of gravity does not appear

sensitive. The approximations developed for equilibrium cable shape analysis suggest

that practical results can be obtained by using 50 or more elements.

4. Dynamic stability analysis approach for tethered aerostat has been proposed which

can also be applied for the situation where the density and uniform wind speed varies

along the vertical direction using the approximations for equilibrium cable shape.

The results for the test aerostat indicate that test aerostat is dynamically stable for

the range of wind speed. The initial disturbance response has also been examined for

the test aerostat where it is observed that lateral motion variables die out relatively
197

faster as compared to longitudinal motion variables indicating relatively stronger

damping in lateral directions. The analysis results for initial disturbances may also

be very useful in selecting the proper aerostat envelope parameters during design for

generating required response characteristics.

5. Nonlinear dynamic analysis model for tethered aerostat has been proposed using the

lumped mass approach. The linear model is useful for carrying out the dynamic

stability analysis of tethered aerostat system but the fact remains that the dynamic

stability model is limited to small perturbations and 3-degrees of freedom. The effect

of aerodynamic non-linearity, detailed motions and peak tether tensions under the

influence of a time-dependent wind vector cannot be determined. The winds, which

are changing in magnitude as well as direction including gusts, have been applied to

the nonlinear model to generate aerostat response.

6. The available experimentally measured aerostat responses have been compared with

the computed response for same input wind conditions and the comparison shows

agreement to some extent between the results. However the available tether tension

peaks from the experiment are not predicted by the current nonlinear model. The

available measured angles of attack are also compared with the results of equilibrium

studies and the comparison shows agreement.

6.2 Scope of Future Work

The present work provides an efficient tool for investigating the static and dynamic analysis

of an airborne tethered aerostat along with generating all the required inputs for carrying

out the analysis. The present work has a lot of scope for future research in different

directions, some of which are presented as below:

1. The dynamic derivatives for the present analysis have been estimated using simpli-

fied assumptions neglecting the hull contributions. The estimation of these dynamic

derivatives more accurately using either advanced computational method or exper-

imental method using a test rig in a wind tunnel may be taken up as a potential

future task.

2. The full ballonet volume for the test aerostat envelope is about 17% of hull volume
198

which may even be higher for a larger size aerostat. When the aerostat operates at

the designed height, the height component of the ballonet is empty and the ballonet

remains partially filled which may have oscillatory motion. The effects of these

oscillatory motions have been assumed negligible in the present dynamic analysis

and the same model. The effect of these oscillatory motions may be examined as a

future possible task.

3. The tether has been modeled as flexible and inextensible cable in linear dynamic

model for stability analysis and as lumped masses in non linear analysis. However, a

more realistic model may be beam elements with large deformation. The same may

be taken up as an important possible future scope.

4. The aerostat envelope has been assumed rigid in the present analysis and hence rigid

body equations of motion have been applied for modeling the same. The aerostat

and particularly fins have got tendency to become flexible with large deformations

as observed during high winds. The present analyses have got limitations here as

aero-elastic effects have not been taken into account. Hence considering aero-elastic

effects in dynamic modeling of aerostat will be a quite challenging future scope.

5. During the nonlinear modeling of aerostat, the tether is modeled as spring-mass-

damper system and experiment may be planned as a future scope to accurately

measure tether damping property.

6. The present research work may be extended to include the effect of releasing and

retrieving of the tether coupled with the wind. This becomes particularly impor-

tant for predicting the performance during the launch and recovery of the tethered

aerostats.

7. The aerostat envelope maintains its rigidity completely due to internal pressure and

the value of minimum internal differential pressure at which the aerostat shape tends

to becomes unstable may also be examined as a potential future work considering

the aero-elastic effects.


Bibliography

[1] Aglietti, G., “Dynamic Response of a High-Altitude Tethered Balloon System, ”Jour-

nal of Aircraft, Vol. 46, No. 6, 2009, pp. 2032-2040.

[2] Airship Design Criteria. US Department of Transportation, Federal Aviation Admin-

istration, No. FAA-P-8110-2, 1995, pp. 15.

[3] Anderson, J. D. Jr., Computational Fluid Dynamics: The Basics with Applications,

McGraw Hill Education, 1995.

[4] Anderson, J. D. Jr., Introduction to Flight, Fifth Edition, Tata McGraw-Hill Pub-

lishing Company Limited, New Delhi, 2008.

[5] ANSYS Fluent Theory Guide, Release 15.0, November 2013, ANSYS, Inc.

[6] Berteaux, H.O., Buoy Engineering, Wiley, New York, Chap 4, 1976.

[7] Carichner, G. E. and Nicolai, L. M., Fundamentals of Aircraft and Airship Design,

Volume 2 - Airship Design and Case Studies, AIAA Education Series, Reston, Vir-

ginia, 2013.

[8] Chapra, S.C., Applied Numerical Methods with MATLAB for Engineers and Scien-

tists, Third Edition, McGraw-Hill Higher Education, 2011.

[9] Colozza, A. and Dolce, J. L., “High-Altitude, Long-Endurance Airships for Coastal

Surveillance,”Glenn Research Center, Report No. NASA/TM-2005-213427, pp. 5.

[10] DeLaurier, J. D., “A Stability Analysis for Tethered Aerodynamically Shaped Bal-

loons, ”Journal of Aircraft, Vol. 9, Sept. 1972, pp. 646-651.

199
200

[11] Delaurier, J. D., “A stability analysis of cable-body systems totally immersed in a

fluid stream,”Report No. CR-2021, National Aeronautics and Space Administration,

Washington, D. C., 1972.

[12] DeLaurier, J. D., “Predictions of Tethered-Aerostat Response to Atmospheric Tur-

bulence, ”Journal of Aircraft, Vol. 14, April 1977, pp. 407-409.

[13] Etkin, B., and Reid, L. D., Dynamics of Flight: Stability and Control, Third Edition,

Wiley India Pvt. Ltd., New Delhi, 2011.

[14] Gill, P., Malik, S., and Pant, R., “Estimation of Aerodynamic Characteristics of

Un-Symmetrically Finned Bodies of Revolution,”28th National Conference on Fluid

Mechanics and Fluid Power, National Society of Fluid Mechanics and Fluid Power,

Chandigarh, India, 2001.

[15] Hembree B, and Slegers, N., “Tethered Aerostat Modeling using an Efficient Recursive

Rigid-Body Dynamics Approach, Journal of Aircraft, Vol. 48, No. 2, March-April

2011, pp. 623-632.

[16] Hoerner, S.F., “Fluid-Dynamic Drag-Theoretical, experimental and statistical infor-

mation,”Published by the author, 1965.

[17] Huh, L., Park, Y. M., Chang, B. H. and Lee, Y. G., “Aerodynamic design of the

KARI Mid-sized Aerostat,”KSAS International Journal,Vol. 7, No. 1, May 2006.

[18] Jones, S. P. and DeLaurier, J. D., “Aerodynamic Estimation Techniques for Aerostats

and Airships,”Journal of Aircraft, Vol. 20, No. 2, February 1983, pp. 120-126.

[19] Jones, S.P. and Krausman, J.A., “Nonlinear Dynamic Simulation of a Tethered Aero-

stat,”Journal of Aircraft, Vol. 19, No. 8, August 1982, pp. 679-686.

[20] Kale, S., Joshi, P., and Pant, R., “A generic methodology for determination of drag

coefficient of an aerostat envelope using CFD,”AIAA 5th Aviation, Technology, In-

tegration, and Operations Conferences (ATIO), 26-28 September 2005, Arlington,

Virginia.

[21] Khoury, G. A. and Gillet, J. D., Airship Technology, Cambridge University Press,

Cambridge, England, U.K., 2004.


201

[22] Krishnamurthy, M., and Panda, G.K., “Equilibrium analysis of a tethered Aero-

stat,”Project Document FE 9802, Flight Experiment Division, NAL, Bangalore, In-

dia, 1998.

[23] Kumar, A., Sati, S. C., and Ghosh, A.K., “Design, Testing and Realization of a

Medium Size Aerostat Envelope,”Defence Science Journal, Vol. 66, No. 2, March,

2016, pp. 93-99.

[24] Kumar, A., Sati, S. C., and Ghosh, A.K., “Equilibrium Cable Configuration Esti-

mation for Aerostat using Approximate Method, ”Fifth International Conference on

Soft Computing for Problem Solving (SocPros-2015) at IIT Roorkee, Published by

Springer, 18-20 Dec 2015.

[25] Kumar, A., Sati, S. C., and Ghosh, A.K., “Response Studies of a Tethered Aero-

stat Balloon Subjected to Initial Disturbances,”Journal of Aerospace Sciences and

Technologies, AeSI, India, Vol. 68, No. 1, Feb, 2016, pp. 26-36.

[26] Kumar, R., Srivastava, S., Ghosh, A.K., Gupta, B. and Kumar A., “Parametric

Trend Study During Stability Analysis Of A Tethered Aerostat,”Journal of Aerospace

Sciences and Technologies, AeSI, India, Vol. 63, No. 2, May, 2011.

[27] Kuo, B. C., Automatic Control Systems, Seventh Edition, Wiley, 1995.

[28] Lambert, C. and Nahon, M., “Stability Analysis of a Tethered Aerostat, ”Journal of

Aircraft, Vol. 40, No. 4, 2003, pp. 705-715. 2003.

[29] Li, Y. and Nahon, M., “Modeling and Simulation of Airship Dynamics,”Journal of

Guidance, Control and Dynamics, Vol. 30, No. 6, 2007, pp. 1691-1699.

[30] Maixner, M.R. and McDaniel, D.R., “Preliminary calculations for a flexible cable in

steady, non-uniform flow,”Aerospace Science and Technology, Volume 18, Issue 1, pp.

1-7, 2012.

[31] Singh, C., Mittal, S., Kumar, A., and Gupta, P., “Wind Tunnel Study on Scale

down Aerostat Model for Aerodynamic Forces and Moments Measurement,”National

Conference on Aerial Delivery & Airborne Surveillance Systems for National Security,

ADASS - 2017, ADRDE Agra, India, 10 Oct 2017.


202

[32] Meriam, J. L., and Kraige, L. G., Engineering Mechanics: Dynamics, Sixth Edition,

Wiley India, 2011.

[33] Munk, M.M., Aerodynamics of Airships: Aerodynamic Theory, Vol. 6, Durand, W.F.,

ed., Julius Springer, Berlin, 1936, pp. 32-48.

[34] Munk, M. M., “Some tables of the factor of apparent additional mass,”NACA Report

No. 197, 1924.

[35] Nelson, R. C., Flight Stability and Automatic Control, Second Edition, Tata McGraw

Hill Education Private Limited, New Delhi, 2011.

[36] Neumark, S., “Equilibrium configurations of flying cables of captive balloons, and

cable derivatives for stability calculations”R & M. No. 3333, Brit. A.R.C., 1963.

[37] Peyada, N. K., “Parameter Estimation from Flight Using Feed Forward Neural Net-

works, ”PhD Thesis, Aerospace Engineering Dept., IIT Kanpur, November 2008.

[38] Pode, L., “Tables for computing the equilibrium configuration of a flexible cable in a

uniform current,”DTMB Report No. 687, March 1951.

[39] Rajani, A., Pant, R.S. and Sudhakar, K., “Dynamic Stability Analysis of a Tethered

Aerostat,”Journal of Aircraft, Vol. 47, No. 5, Sep-Oct 2010, pp. 1531-1538.

[40] Raymer, D. P., Aircraft Design: A Conceptual Approach, AIAA Education Series,

1989, pp. 266.

[41] Redd L. T., Benett, R. M. and Bland, S. R., “Experimental and Analytical determi-

nation of stability parameters for a balloon tethered in a wind,”NASA TN D-7222,

1973.

[42] Redd L. T., Bland, S. R. and Benett, R. M., “Stability analysis and trend study of a

balloon tethered in a wind, with experimental comparisons,”Report No. NASA TN

D-7272, Washington, D. C., Oct. 1973.

[43] Simmons, G. F., Differential Equations with Applications and Historical Notes, Sec-

ond Edition, Tata McGraw Hill Education Private Limited, New Delhi, 2010.
203

[44] Srivastava, S., “Stability Analysis and Parameter Trend Study of Single Tether

Aerostats,”M.Tech Thesis, Department of Aerospace Engineering, Indian Institute

of Technology, Kanpur, April, 2009.

[45] “Tethered Aerostat Systems Application Note,”Prepared by Space and Naval Warfare

Systems Center Atlantic, US department of homeland security, September 2013.

[46] Spalart, P., and Allmaras, S., “A one-equation turbulence model for aerodynamic

flows,”AIAA Paper 92-0439, 30th Aerospace Sciences Meeting and Exhibit, Reno,

NV, January 1992.

[47] Thomson, W. T., and Dahleh, M. D., Theory of Vibration with Applications, Fifth

Edition, Pearson, 1997.

[48] URL: https://crowdmanagement.wordpress.com/2014/07/31/ the-main-advantages-

of-aerostats-over-uavs.

[49] URL:http://dragonfly.tam.cornell.edu/teaching/mae5230-cfd-intro-notes.pdf

[50] URL: https://en.wikipedia.org/wiki/CATIA (All Accessed on March 13 2017).

[51] Vijayram, C. and Pant, R., “Multidisciplinary Shape Optimization of Aerostat En-

velopes,”Journal of Aircraft, Vol. 47, No. 3, 2010, pp. 1073-1076.

[52] Wardlaw, A.B., “High-Angle-of-Attack Missile Aerodynamics,”AGARD Lecture Se-

ries 98, Feb. 1979, pp. (5-1), (5-53).

[53] Weber, F. and Distl, H., “Damping Estimation from Free Decay Responses of Cables

with MR Dampers, ”The Scientific World Journal, Vol. 2015 (2015), No. 861954, pp.

1-15. 2015.
205

PUBLICATIONS

Journal Publication

1. A Kumar, S C Sati and A K Ghosh, “Design, Testing and Realization of a Medium


Size Aerostat Envelope,”Defence Science Journal, Vol. 66, No. 2, March 2016, pp.
93-99.

2. A Kumar, A K Ghosh and S C Sati, “Response Studies of a Tethered Aerostat


Balloon Subjected to Initial Disturbances,”Journal of Aerospace Sciences and Tech-
nologies, AeSI, India, Vol. 68, No. 1, Feb, 2016.

3. Kumar, R., Srivastava, S., Gupta, B., Kumar, A., and Ghosh, A.K., “Paramet-
ric Trend Study during the Stability Analysis of a Tethered Aerostat,”Journal of
Aerospace Sciences and Technologies, AeSI, India, Vol. 63, No. 2, May, 2011.

Conference/Symposium

1. A Kumar, S C Sati and A K Ghosh, “Equilibrium Cable Configuration Estima-


tion for Aerostat using Approximate Method,”Fifth International Conference on
Soft Computing for Problem Solving (SocPros-2015) at IIT Roorkee, Published by
Springer, 18-20 Dec 2015.

2. R Bhatia, G Singh, P Gupta, A Kumar, Amitabha Pal, S C Sati , A K Ghosh,


“Dynamic Modeling of Unmanned small size Airship,”Symposium on Applied Aero-
dynamics and Design of Aerospace Vehicle (SAROD 2013), Nov 21-23, 2013, Hyder-
abad, India.

3. A Sharma, A Kumar, P Gupta, Amitabha Pal, S C Sati , A K Ghosh, “Dynamic


Modeling of Medium size Aerostat with Tether,”Symposium on Applied Aerodynam-
ics and Design of Aerospace Vehicle (SAROD 2013), Nov 21-23, 2013, Hyderabad,
India.

4. V Patildar, SVS Chauhan, A Sharma, R Bhatia, A Kumar, P Gupta, A Pal, S C


Sati, “Conceptual Design of Unmanned small size Airship,”Symposium on Applied
Aerodynamics and Design of Aerospace Vehicle (SAROD 2013),”Nov 21-23, 2013,
Hyderabad, India.

5. A Kumar, P Gupta, A Sharma and S Gupta, “Design and Testing of a Medium Size
Aerostat Balloon,”Symposium on Applied aerodynamics and Design of Aerospace
Vehicle (SAROD 2011), Nov 16-18, 2011, Bangalore, India.

You might also like