Sei sulla pagina 1di 108

SURFACE ACCURACY ANALYSIS AND MATHEMATICAL

MODELING OF DEPLOYABLE LARGE APERTURE


ELASTIC ANTENNA REFLECTORS

By Michael J. Coleman

Bachelor of Arts in Mathematics


Boston University, May 2004

A Dissertation submitted to

The Faculty of
The Columbian College of Arts and Sciences
of The George Washington University
in partial fulfillment of the requirements
for the degree of Doctor of Philosophy

August 31, 2010

Dissertation directed by

Frank E. Baginski
Professor of Mathematics
The Columbian College of Arts and Sciences of The George Washington University certifies that

Michael J. Coleman has passed the Final Examination for the degree of Doctor of Philosophy as of

May 14, 2010. This is the final and approved form of the dissertation.

SURFACE ACCURACY ANALYSIS AND MATHEMATICAL MODELING OF


DEPLOYABLE LARGE APERTURE ELASTIC ANTENNA REFLECTORS

Michael J. Coleman

Dissertation Research Committee:

Frank E. Baginski, Professor of Mathematics

Dissertation Director

Katharine F. Gurski, Assistant Professor of Mathematics, Howard University

Dissertation Committee Member

Xiaofeng Ren, Associate Professor of Mathematics

Dissertation Committee Member

ii
c Copyright 2010 Michael J. Coleman

All Rights Reserved

iii
Acknowledgements
The research presented in this dissertation has been supported by the National Aeronautics and

Space Administration Grants NNX07AR67G and NNX09AH08G. I would like to begin by thanking

our technical contact at NASAs Glenn Research Center, Dr. Robert R. Romanofsky. This work would

not have been possible without his attention and guidance on this project throughout my research

time at GW and NASA/GRC in Cleveland. His colleague at NASA, Dr. Kevin M. Lambert also gave

me very useful guidance on antenna theory.

I am particularly grateful for the undivided attention of my advisor, Frank Baginski throughout

my years of study and research. His constant support and advice have been an integral part in

the development of my mathematical, presentation and writing skills. Frank gave many hours (at

times we even met on weekends) to review the countless pages and hours of presentation material

that I have created throughout my graduate career. He is a certainly a devoted advisor and cares

very much for the work that we do. He also makes the best crab dip!

Many thanks to all others who agreed to serve on my dissertation committee: Robbie Robinson,

Katie Gurski, Xiaofeng Ren and Magda Musielak. I appreciate their help and support for my defense

as well as the support theyve offered me over my years in the graduate program.

Thank you also to my close friends whom Ive known throughout my time at GWU and in the

DC region in general. I would especially like to thank Tyler White, Joe Herning and Dzung Trac

who helped me set up my computer, Apollo. Most of the results obtained in this dissertation were

computed on Apollo and so I am very thankful for their assistance.

Finally, I would not have made it to this stage without the unceasing love and support of my

family: Mom, Dad, Meaghan, Grandma and Grandpa, Grandma and the Kitty (rest in peace),

Jimmy and Lauren, Ray and Michelle, Diane and Kenny, and all my aunts and uncles. Special

thanks to Mom and Dad for your encouragement in both the best and worst of times. Mom, I hope

that this document can well complement that first calculator of mine!

George Washington University Michael J. Coleman

May 14, 2010

iv
Abstract
SURFACE ACCURACY ANALYSIS AND MATHEMATICAL MODELING OF
DEPLOYABLE LARGE APERTURE ELASTIC ANTENNA REFLECTORS

One class of deployable large aperture antenna consists of thin lightweight parabolic reflec-

tors. A reflector of this type is a deployable structure that consists of an inflatable elastic membrane

that is supported about its perimeter by a set of elastic tendons and is subjected to a constant hy-

drostatic pressure. A design may not hold the parabolic shape to within a desired tolerance due

to an elastic deformation of the surface, particularly near the rim. We can compute the equilib-

rium configuration of the reflector system using an optimizationbased solution procedure that

calculates the total system energy and determines a configuration of minimum energy. Analysis

of the equilibrium configuration reveals the behavior of the reflector shape under various loading

conditions. The pressure, film strain energy, tendon strain energy, and gravitational energy are all

considered in this analysis. The surface accuracy of the antenna reflector is measured by an RMS

calculation while the reflector phase error component of the efficiency is determined by computing

the power density at boresight. Our error computation methods are tailored for the faceted surface

of our model and they are more accurate for this particular problem than the commonly applied

Ruze Equation.

Previous analytical work on parabolic antennas focused on axisymmetric geometries and loads.

Symmetric equilibria are not assumed in our analysis. In addition, this dissertation contains two

principle original findings: (1) the typical supporting tendon system tends to flatten a parabolic

reflector near its edge. We find that surface accuracy can be significantly improved by fixing

the edge of the inflated reflector to a rigid structure; (2) for large membranes assembled from

flat sheets of thin material, we demonstrate that the surface accuracy of the resulting inflated

membrane reflector can be improved by altering the cutting pattern of the flat components.

Our findings demonstrate that the proper choice of design parameters can increase the perfor-

mance of inflatable antennas, opening up new antenna applications where higher resolution and

greater sensitivity are desired. These include space applications involving high data rates and high

v
bandwidths, such as lunar surface wireless local networks and orbiting relay satellites. A light

weight inflatable antenna is also an ideal component in aerostat, airship and free balloon systems

that supports communication, surveillance and remote sensing applications.

vi
Contents

Acknowledgements iv

Abstract v

List of Figures viii

List of Tables ix

Chapter 1 Introduction 1

1.1 Parabolic Reflectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.2 Shell and Membrane Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.3 Mathematical Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

Chapter 2 Inflatable Elastic Reflector Model 14

2.1 Mathematical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.2 Flat Panel Constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

Chapter 3 Parabolic Reflector Efficiency 30

3.1 Electromagnetic Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.2 Reflectivity Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.3 Antenna Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.4 Geometric Error Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

Chapter 4 Demonstration Cases 64

4.1 A Test Case: The Mylar Balloon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4.2 Analysis of Molded Reflectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

vii
Chapter 5 Parametric Studies 76

5.1 Parametric Study for RMS versus N g and p0 . . . . . . . . . . . . . . . . . . . . . . . . . 76

5.2 Parametric Study for RMS versus and . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.3 Tendon Supported vs. Fixed Boundary Reflectors . . . . . . . . . . . . . . . . . . . . . 82

Chapter 6 Conclusions and Future Research 85

References 90

Appendix A Surface Metrology Data 94

viii
List of Figures

1.1 Lenticular deployable reflectors and their support structures. . . . . . . . . . . . . . . 2

1.2 Cross sectional diagram of an axisymmetric parabolic reflector. . . . . . . . . . . . . . 6

1.3 Cross sectional diagram of an offaxis parabolic reflector. . . . . . . . . . . . . . . . . 7

2.1 Inflatable parabolic reflector antenna. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.2 Faceted surface for molded antennas. Dots highlight the reflector rim. . . . . . . . . 17

2.3 Deformation of a material body M under a deformation x. . . . . . . . . . . . . . . . . 19

2.4 Stress components 2i at point p in material M . . . . . . . . . . . . . . . . . . . . . . . 21

2.5 Flat gore construction. Generating ribs indicated by dots. . . . . . . . . . . . . . . . . 28

2.6 Four gores of the reference configuration for the flat gore construction. . . . . . . . . 29

2.7 Flat ring construction. Generating rings indicated by dots. . . . . . . . . . . . . . . . . 29

3.1 Charge with velocity v flowing through a differential cross section, d a. . . . . . . . 34

3.2 A contour C attached to two surfaces passing through different current. . . . . . . . 38

3.3 Boundary conditions for Maxwells Equations. . . . . . . . . . . . . . . . . . . . . . . . 41

3.4 Cross sectional diagram of a paraboloid and its geometry. . . . . . . . . . . . . . . . . 46

3.5 Wave front shape in the near and far field of an AUT. . . . . . . . . . . . . . . . . . . . 49

3.6 Geometry of rays extending to a point P in the far field. . . . . . . . . . . . . . . . . . 50

3.7 Far field patterns for an ideal axisymmetric parabolic reflector antenna. . . . . . . . 58

3.8 Deformed triangle T n with centroid (x n , yn , zn ). The vertical displacement, z, is

labeled and the pathlength for this triangle is n = f n + an . . . . . . . . . . . . . . . . 60

4.1 Cross section of the Mylar Balloon with generating curve z(x). . . . . . . . . . . . . 65

4.2 Generating curve of modeled mylar balloon. . . . . . . . . . . . . . . . . . . . . . . . . 66

4.3 Fully inflated mylar balloon (actual and modeled). . . . . . . . . . . . . . . . . . . . . 67

ix
4.4 Reflector Distortion Pattern for High and Low Tendon Tension. . . . . . . . . . . . . 71

4.5 Effect of tendon force distribution for Case 3 of Table ??. . . . . . . . . . . . . . . . . 72

4.6 Effect of tendon force distribution for Case 4 of Table ??. . . . . . . . . . . . . . . . . 73

4.7 Shell separation for small offaxis antenna reflector. . . . . . . . . . . . . . . . . . . . 74

5.1 Aperture radiation phase plane at 40 GHz with N g = 16. . . . . . . . . . . . . . . . . . 78

5.2 Phase distributions for two values of p0 at 40 GHz. . . . . . . . . . . . . . . . . . . . . 79

5.3 Parabolic cross section for a flat ring configuration with = 0.05. . . . . . . . . . . . 80

5.4 Effect of rim modification factor on RMS for various tendon loading. . . . . . . . . . . 81

5.5 Comparison of error for a fixed rim and a tendonsupported reflector. . . . . . . . . 83

A.1 Surface metrology plot of axisymmetric reflector for = 90 . . . . . . . . . . . . . . . 96

A.2 Surface metrology plot of axisymmetric reflector for = 76 . . . . . . . . . . . . . . . 97

A.3 Surface metrology plot of offaxis reflector; exhibits substantial rim deflection. . . 97

x
List of Tables

3.1 Charge and mass of the three fundamental charged particles. . . . . . . . . . . . . . 31

3.2 Common IEEE Radiation bands. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.3 Comparison of predicted versus actual errors for various grid sizes. . . . . . . . . . . 63

4.1 Geometric results of mylar balloon test for two p0 values. . . . . . . . . . . . . . . . . 66

4.2 Mechanical properties of Kapton, reflector dimensions and grid size. . . . . . . . . . 67

4.3 Comparison of energy components for various antenna sizes. . . . . . . . . . . . . . 68

4.4 Gravitational effects on large reflectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.5 Summary of RMS values for various physical testing conditions. . . . . . . . . . . . . 72

4.6 Surface accuracy of offaxis reflectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.1 Grid parameters and dimensions of the reflector being modeled. . . . . . . . . . . . . 76

5.2 Comparison of RMS (N g , p0 ) values in millimeters. . . . . . . . . . . . . . . . . . . . . . 77

5.3 Comparison of RMS (, ) values (mm) for various . . . . . . . . . . . . . . . . . . . . 81

5.4 Collection of RMS (p0 , g) values (mm) for a fixed boundary reflector. . . . . . . . . . 82

5.5 Characterization of fixed rim inflatable antennas; p0 = 12.5 Pa. . . . . . . . . . . . . 84

5.6 Characterization (for comparison) at 80 GHz. . . . . . . . . . . . . . . . . . . . . . . . . 84

A.1 Least squares paraboloid parameters for surface metrology data. . . . . . . . . . . . . 95

xi
Chapter 1

Introduction

An antenna is a device that concentrates a radio signal in a particular direction. There are many

varieties of antennas that have been used throughout the history of communications; see [14] and

[33]. One can choose an antenna that will radiate its signal in all directions in order to reach as

many locations as possible. This is called an isotropic radiator and may be the preferred antenna

type of a radio station, for example. Other types of antennas can direct radiation in a certain

direction in order to concentrate the signals power to a particular destination. While the signal is

only received in a particular direction, it will have greater strength than a signal that is radiated

isotropically.

Reflector antennas are particularly effective for the purpose of concentrating a signal over long

distances. Antennas of this type typically use a parabolic reflector dish to direct the signal in a

particular direction. In the interest of long distance radio communication, this antenna is among

the most widely used [33, Pg. 422]. The common notion of a satellite dish is exactly the type

of antenna which falls under this category and will be the basis of the antenna theory that is

contained in this dissertation.

Future space exploration and near earth remote sensing missions will require large data trans-

fer rates on the order of hundreds of megabits per second [27]. There is also interest in missions

at and beyond Earths orbit and so information must be transmitted over long distances. In order

to support such large amounts of data, an antenna that can receive and transmit a large amount

of radiation must be employed. If the distance between the mission and Earth grows to the order

1
(a) Supporting torus [37] (b) LGardes lenticular deployable antenna [36]

Figure 1.1: Lenticular deployable reflectors and their support structures.

of astronomical units, there will be additional burdens on the antenna to radiate enough power to

compensate for a large loss of the signal over such a large distance.

The power that radiates from the antenna increases with the size of the reflector [33, Pg.

433]. Therefore, the need for such large data transfer rates leads to the necessity for aperture

antennas which have large parabolic reflectors. The possibility of 10 meter diameter reflectors has

been mentioned [27]. Unfortunately, the limited volume of a launch vehicle or spacecraft places

restriction on the size of a rigid antenna reflector. To allow larger reflectors to serve these missions,

a deployable antenna is desirable. Such an antenna has a reflector which is made of a lightweight

material, can be collapsed or folded for compact stowing, and is deployable at an appropriate time.

Deployable space structures have a history that extends back to around 1960. The background

on space structures that is presented here is documented by R. Freeland, G. Bilyeu and M. Mikulas

in [14] and [15]. Goodyear developed some of the first versions of inflatable antennas; one was

known as a search radar antenna. This antenna was primarily supported by metallic trusses that

unfolded like an accordion to expand the antenna surface and support it. Goodyear also developed

a Radar Calibration Sphere consisting of a large number of flat hexagonal membrane pieces. The

sphere was inflated to a diameter of 6 meters and was metalized for reflectivity.

The lenticular inflatable parabolic reflector is a design much nearer to the one we will consider.

It consists of an inflatable structure formed by two opposite facing paraboloids seamed together

and supported about the periphery by an inflatable toroidal shaped membrane; see Figure 1.1(a).

2
The diameter of the reflector for this particular antenna is on the order of 10 meters. An antenna

of this type was developed by LGarde, Inc. and selected for NASAs INSTEP Inflatable Antenna

Experiment in 1996 [15]. It includes long supporting beams which attach the reflector to the

satellite; see Figure 1.1(b). The experiment was able to verify that these large inflatable antennas:

could be built at low costs;

have high mechanical packaging efficiency;

have low weight;

have high deployment reliability;

have surface precision to within a few millimeters RMS.

In 2000, astromesh reflectors were successfully deployed for operation at 14 and 30 GHz [34].

There is now interest in operating large deployable antennas at higher radiation frequencies which

will require greater surface accuracy. Distortions to the parabolic reflector will perturb the signal

of the antenna and cause a loss of the signals power to the observer. Any such distortion is

more conspicuous at higher radiation frequencies, as we will show in Chapter 3. An estimate for

the tolerance of the reflector surface accuracy (for the current performance target of inflatable

antennas) is an RMS value of no more than 0.5 mm [27]. Maintenance of the parabolic shape is

therefore critical for successful operation of the antenna. Initial findings suggest that the edges of

the inflated antenna reflectors we are considering are distorted to a flattened ring causing larger

RMS values and an overall reduction of transmitted power.

Investigation of the reflector shape and response of the reflector to internal and external forces

is the central topic of this dissertation. We will model the inflatable antenna reflector so that the

configuration can be calculated based on: physical parameters that quantify the forces acting on

the entire system, mechanical properties of the materials used to construct the reflector, and design

parameters that adjust the lengths and sizes of select apparatus within the system. It is ultimately

our goal to identify those parameters that can be used to improve the efficiency of the antenna

reflector. From this, we can draw some conclusions regarding the support structure, the design

of the current inflatable antenna prototypes and possible construction patterns for the reflector

membrane that may increase performance.

3
The geometric properties of parabolic reflectors will be discussed in Section 1.1. There we focus

on the geometry of rigid paraboloids as reflecting surfaces and neglect the elastic properties which

we include later. Section 1.2 will provide an overview of the literature that addresses membrane

and elastic reflector problems. Background of the mathematical theory required for development

of the model and surface accuracy calculations that we perform in subsequent chapters is given in

Section 1.3.

In Chapter 2, we outline a model for the shape of the deployable antenna system that we

study. This model is based on one that was developed for the analysis of high altitude large

scientific research balloons by F. Baginski and his collaborators; see [2] [4]. The model uses a

piecewise linear surface of triangular facets to approximate the reflector surface. The model also

includes analysis of large scale deformations and wrinkling which was developed by Allen C. Pipkin

in [26]. In Section 2.2, specific reference configurations will be introduced which feature flat

panel constructions that approximate a paraboloid. The benefits and drawbacks of these differing

reference configurations will also be discussed in Chapter 2.

In Chapter 3, methods of measuring the efficiency and accuracy of a given reflector configura-

tion will be discussed. This will entail a discussion of electromagnetic theory which is necessary for

understanding how antennas reflect radiation. Some direct computations for the surface accuracy

based solely on the geometry of the reflector configuration are also considered. These calculations

depend on both the displacement and orientation of the surfaces triangular facets.

Results of our model are presented in Chapters 4 and 5. The first of these chapters includes the

results of the model applied to a mylar balloon. Mylar balloons are commonly found in party shops

or grocery stores and are comprised of two flat circular membranes which are seamed together and

then inflated; more details regarding these balloons are in Chapter 4. This test case will compare

the results of the model with known analytical results for the shape of a mylar balloon. There

are also tests involving molded parabolic reflectors. Reflectors of the molded variety are actually

cast on mandrels so that the unstrained state of the reflector material is parabolic. The results of

these experiments show that high tendon forces cause deflection of the reflector surface near the

rim. Furthermore, we show that nonsymmetric loading at the rim on both symmetric and offaxis

antennas causes high surface inaccuracy.

In Chapter 5, we will present more recent results of the models predictions for flat panel

4
construction methods. We perform several parametric studies involving parameters that adjust the

cutting patterns of the panels. Modifications to the material patterns near the rim are introduced

to combat the rim deflection that is described in Chapter 4. These changes improve the surface

accuracy of the deployed antennas. Changes to the inflation pressure also help improve surface

accuracy but excessive pressure causes surface distortions to appear.

In Appendix A, we present surface metrology data on two inflatable antennas that we observed

at NASAs Glenn Research Center in 2008. The data sets are for a 2.13 meter inflatable astromesh

reflector and a 0.3 meter offaxis reflector. The data sets were taken by a Leica LR200 laser

scanner which allows us to plot the surface points and see the configuration of the reflectors. We

also calculate best fit paraboloids to this data and then determine the reflectors surface error.

1.1 Parabolic Reflectors

A parabolic reflector takes advantage of the geometric properties of a paraboloid. If one places an

electrical transmitter at the focal point of the paraboloid, then the rays of radiation will be reflected

in rays that are parallel to the symmetric axis of the antennas coordinate frame. This is the feature

that causes the signal to be amplified in the direction that the antennas reflector points. In

Figure 1.2 we see the characteristics of a typical antenna with an axisymmetric parabolic reflector.

Some important terminology for parabolic antenna reflectors is listed here.

Feed: an apparatus located at the focal point of the parabolic reflector which transmits or

receives a signal.

Focal plane: a plane orthogonal to the symmetric axis of the paraboloid passing through the

focal point of the antenna.

Focal length: a geometric property of the antenna which is the distance from the vertex of

the paraboloid along the symmetric axis to the focal point.

Aperture: an imaginary projection of the reflector onto the focal plane. For the antennas of

interest here, the aperture is always circular.

F/D value: the ratio of the focal length to the apertures diameter.

5
Since the aperture is always circular for a parabolic reflector, these antennas are a subset of

antennas known as circular aperture antennas. The geometry is entirely determined by the focal

length and the aperture diameter. In a system of cylindrical coordinates (r, , z), an axisymmetric

reflector is obtained by generating the surface of revolution from the curve

4F z = r 2 (1.1)

D
for 0 r 2
where F is the focal length of the antenna and D is the aperture diameter.

Circular
Aperture


n
F Focal Point

Parabolic
Reflector

Figure 1.2: Cross sectional diagram of an axisymmetric parabolic reflector.

Another type of parabolic antenna that we will consider is an offaxis reflector. The optical

properties of an offaxis reflector are identical to those of an axisymmetric reflector. The only

difference lies in the construction of the reflector surface which can be visualized by taking a

cylindrical cut of a paraboloid (not an axisymmetric cut). A key advantage of the offaxis reflector

is that the feed and its supporting apparatus do not block the reflected radiation. The paraboloid

from which the offaxis reflector is taken is called the parent paraboloid and the focal length of the

antenna is the focal length of this paraboloid. See Figure 1.3 for a visualization of this construction

method.

1.2 Shell and Membrane Theory

The elasticity of inflatable antenna reflectors is an important component of the mathematical

model that we present in Chapter 2. Elastic shells of various shapes had been studied long be-

6
D
Circular Aperture

Focal Point
F

Parent Paraboloid
Reflector

Figure 1.3: Cross sectional diagram of an offaxis parabolic reflector.

fore the 1960s when deployable structures were first conceived. An early analysis of a membrane

problem is Dr. I. Henckys 1915 paper ber den Spannungszustand in kreisrunden Platten mit

verschwindender Biegungssteifigkeit (On the Stress State in Circular Plates With Vanishing Bending

Stiffness) [18]. Henckys paper discusses (in cylindrical coordinates) an isotropic circular mem-

brane that is fixed at the rim and exposed to a symmetric lateral (z directed) loading, p. He

includes bending rigidity in his analysis and models the axisymmetric vertical displacement of this

membrane, by

Eh2
2 hr 1 B(, F ) = p (1.2)
12(1 2 )

2 2 F + E r 1 B(, ) = 0 (1.3)

where E, and h are the Youngs Modulus, Poissons Ratio and thickness of the plate, respectively,

is the Laplacian in polar coordinates and B(u, v) = (u r vr ) r . The boundary condition (a) = 0

must be satisfied where a is the radius of the circular membrane. Equations (1.2) and (1.3) are

equivalent to the Von Krmn Equations for a lateral load on a plate [7, Pg. 285]. The antenna

problem that we consider, however, involves a uniform pressure load on the membrane surface

rather than a lateral load and has outward force around the rim for boundary support rather than

a fixed boundary.

W. Fichter recognized the absence of the radial component of the loading in Henckys problem

and developed results for the problem with a true uniform pressure load on the membrane in

7
[13]. Fichter showed that the solutions for lateral load problem of Equations (1.2) and (1.3), and

uniform pressure load are quite similar for small loading but diverge slowly when the load grows

larger. At higher loading, the lateral displacement of the membrane reveals a slightly more spher-

ical shape when under uniform pressure and a more uniform distribution of the stress resultants.

Since the surface accuracy of antenna reflectors is important in application, more recent work

has included results on the causes, magnitudes and distribution of deformations on membranes.

Among these studies are gravitational affects on singlycurved membranes having the shape of a

parabolic cylinder [22]. The existence of wrinklefree solutions to circular membranes problems

has also been explored in [5] and is of importance to deployable antennas; in order to be effective,

the deployed reflector should not have wrinkled areas.

In more recent antenna specific work, Greschik explores the sensitivity of reflector membranes

to certain environmental factors and physical assumptions in [16]. He analyzes the possible errors

of reflector surface accuracy computations when uniformity of the reflector material is assumed,

extreme temperatures are neglected, or regions of wrinkling are ignored. It was found, for ex-

ample, that RMS values could be miscalculated by 1 mm for shallow reflectors (F /D = 4) at a

pressure of 125 Pa when wrinkling is ignored.

In their 1997 paper, Marker and Jenkins reviewed the typical W curve error pattern for a

parabolic antenna membrane which is a result following from Henckys analysis [21]. They found

that modifications of the boundary position could reduce the maximum deviation of a deformed

reflector up to 58%. We also investigate, in this project, the possibility of reducing parabolic

surface distortions using geometric modifications for the specific antenna reflectors of interest.

The articles cited here focus primarily on axisymmetric membranes and axisymmetric loadings.

While the initial configurations for our model are axisymmetric, the model is not constrained to

compute axisymmetric equilibrium configurations. For this reason, we are able to apply non

symmetric loading in our analysis. Membrane wrinkling is also included in the model and is a

critical feature to consider since an antenna reflector should be as free of wrinkling as possible.

Our results add to the current knowledge of reflector deformation. We find that distortion

of the reflector membrane supported about the boundary by elastic tendons does not exhibit the

typical W profile error pattern. Rather, there is deflection of the membrane near the rim and

substantial (order of 3 mm) vertical displacement at the vertex. We combat these distortions by

8
finding construction patterns for the membrane (adjustable by a certain set of geometric parame-

ters) which will deform to a surface that is sufficient to serve as an effective antenna reflector at

high radio frequency.

1.3 Mathematical Preliminaries

In this section, we will introduce some of the background topics that are used throughout this

dissertation. These few notes also serve as a convenient reference for the reader. First, we present

some notational conventions.

1.3.1 Notational Conventions

The applications covered in the following chapters draw on the theories of elasticity, electromag-

netism and antennas. The standard notations used for these subjects differ and occasionally over-

lap. We will adhere to standard notation as often as possible and clarify which conventions we

adopt in order to remain consistent with notation throughout this dissertation. Unless otherwise

noted, we follow these conventions:

Lower case italic roman letters (a, b, x, y): scalar quantities or scalar functions. The context

will be clear.

Lower case bold roman letters (a, b, x, y): vectors in R2 or R3 or vector valued functions

(also referred to as first order tensors).

The set of vectors {i, j, k} is the standard basis of R3 .

Any vector displaying a hat (


n, for example) has unit length. We will suppress the hat

notation for the standard basis vectors of R3 .

Upper case bold roman letters (A, B, X, Y): second order tensors (for elasticity theory), vector

fields or phasors (for electromagnetic theory). Whether a notation of this type represents a

tensor, vector field or phasor, will be clear in context. Tensors are introduced in Section 1.3.3

and phasors in Section 3.1.2.

Greek letters of all cases and types will be defined as they are used.

9
1.3.2 Vector Analysis

The development of Maxwells Equations and a solution are included to give the reader a com-

plete picture of the reflectors role in the antennas operation. The following vector relations are

important for that section.

The triple product rule is convenient and relates the cross product of three vector fields. For

any three vectors a, b and c,

a (b c) = b(a c) c(a b). (1.4)

In addition to the usual definitions of divergence and curl, the following relations are worth con-

sidering especially in this framework. These formulations provide a more intuitive definition of

these two differential operators [23, Sections 15 and 16]. In each definition, we have the vector

field F C 1 (R3 ), the point x R3 and a closed ball V centered at x with radius .

Definition 1.3.1. The divergence of a vector field, F is the limit of its surface integral per unit volume

as the volume enclosed by the surface goes to zero. That is,

1
ZZ
F(x) = lim
F n
dS.
0 V V

Definition 1.3.2. The curl of a vector field F is the limit of the ratio of the integral of its cross product

with the outward drawn normal, over a closed surface, to the volume enclosed by the surface as the

volume goes to zero. That is,

1
ZZ
F(x) = lim
n
F dS.
0 V V

We will take advantage of familiar corollaries of Stokes Theorem as they are often presented

in Vector Calculus. The version of Stokes Theorem specific to the curl of a vector field as well as

the Divergence Theorem are of particular interest here.

Theorem 1.3.1. (Divergence Theorem) If F is a C 1 vector field and B R3 is a volume with boundary

B, then Z ZZ ZZ
F dV =
F n
dS (1.5)
B B

10
is the unit outward normal of B, and d V and dS are Lebesgue volume and surface area
where n

measures, respectively [29, Theorem 10.51].

Theorem 1.3.2. (Stokes Theorem) If F is a C 1 vector field and R3 is a surface then

ZZ I
( F) n
dS = (F t) ds (1.6)

is the unit outward normal of , t is the unit tangent along the boundary , and dS and
where n

ds are Lebesgue surface area and arc length measure, respectively [29, Theorem 10.50].

The following vector calculus identities are useful so we present them for reference. If F and

G are differentiable vector fields and F a differentiable scalar field, then

(F ) = 0 (1.7)

( F) = 0 (1.8)

F = ( F) 2 F (1.9)

(F G) = G ( F) F ( G) (1.10)

1.3.3 Tensor Algebra

In order to fully treat elasticity, we must introduce a little tensor theory. Unlike force, a vector is

not a sufficient device to express the notion of stress in a material. Tensors are also necessary to

formulate expressions for the deformation of a material body. The relationship between the stress

and the strain in a material play a critical role in the model we use to determine the configuration

of the reflector surface. We follow the introduction given by Antman in [1, Section 11.1] since the

notation is consistent with that of our presentation of the model in Chapter 2.

Let {e1 , e2 , e3 } be a basis for R3 . We can express a vector (first order tensor) by

3
X
v = vi ei R3
i=1

where v = (v1 , v2 , v3 ).

Definition 1.3.3. A second order tensor A is a transformation of R3 onto itself.

11
We can represent the second order tensor A as a matrix. The tensor A is evaluated at a v R3

and denoted by A v R3 . A product of two tensors A and B is denoted by A B and satisfies the

relationship

(A B) v = A (B v)

for all v R3 . We will encounter the notion of a dyadic product of two vectors in our model. If

a, b R3 , then the dyadic product a b is a second order tensor that satisfies

(a b) v = (b v)a

for all v R3 . In the case of the first order tensors a = (a1 , a2 , a3 ) and b = (b1 , b2 , b3 ), we can

represent the dyadic product in matrix form by


a 1 b1 a 1 b2 a 1 b3

a b = a 2 b1 a 2 b3 .

a 2 b2


a 3 b1 a 3 b2 a 3 b3

Notice that the matrix representing the dyadic tensor a a would be symmetric. It is also the case

that any symmetric tensor can be decomposed into a dyadic product of vectors.

Definition 1.3.4. The spectral representation of a symmetric tensor A is

3
X
A = i ei ei (1.11)
i=1

where {ei } is an orthonormal basis of eigenvectors that correspond to the eigenvalues of the tensor A

[1, Section 11.1].

If we consider two second order tensors Ci j and Dkl , then the product of these two would be

the fourth order tensor

Ei jkl = Ci j Dkl .

Writing the entire tensor out is not possible since a four dimensional array would be required to

12
satisfy all the subscripts. For example, fixing i = 1 and l = 2 would leave the portion of the tensor


C11 D12 C12 D12 C13 D12

E1 jk2 = C1 j Dk2 = C11 D22 C13 D22 .

C12 D22


C11 D32 C12 D32 C13 D32

A tensor inner product : is defined in [1, Section 11.1] for dyads. Since the tensors we use

in our model can be expressed as dyads, this definition will be sufficient.

Definition 1.3.5. The tensor inner product : for two tensors of the form a b and c d is given by

(a b) : (c d) = (a c)(b d).

Antman also defines several usual calculus operators which involve tensors [1, Section 11.2].

We present those which we use later.

Definition 1.3.6. The gradient of a vector field F C 1 (R3 ) is a second order tensor of the form

F F F
F = i+ j+ k.
x y z

Definition 1.3.7. The divergence of a second order tensor, T = t1 i + t2 j + t3 k, is

t1 t2 t3
T = + +
x y z

where ti C 1 (R3 ).

13
Chapter 2

Inflatable Elastic Reflector Model

Inflatable antennas are constructed in several different sizes and configurations. Although reflector

diameters on the order of 10 meters or more are conceivable, we have only investigated prototype

antennas with diameters ranging from 0.3 to 2.13 meters. An ideal pressure range for the reflector

assembly is 10 15 Pa, although we also observed the envelope at higher pressures of 25 30 Pa

to investigate the predicted membrane shape [28].

Inflatable antennas are constructed by molding lightweight material into parabolic form to

achieve the desired geometry for the reflector. Two such parabolic membranes are created and

seamed along their rims. The resulting envelope is pressurized and resembles a clam shell. The

entire clam shell is supported by tendons which are attached to the rim on one end and a rigid

frame on the other. Prototypes that have been observed are made of the material CP1 which is

space qualified, and highly transparent. Typically, one of the parabolic sheets is coated with 1200

of aluminum to make the surface a reflective one. The coated half of the clam shell will serve as

the antennas reflector. See Figure 2.1 for a schematic of the inflatable antenna system.

The prototypes investigated at NASAs Glenn Research Center demonstrate some deflection

from parabolic form near the rim of the reflector which subsequently causes a noticeable loss of

power transmitted by the antenna. The distribution of the surface distortions across the reflector

surface is important to understand if we wish to identify what is ultimately responsible for the

reduced antenna performance. These are types of issues that we address in studying the response

of the system to different supporting mechanisms.

14
For large reflectors (diameters closer to the 10 meter range), the molding process may be in-

feasible or too expensive. We therefore consider models for a reflector constructed from flat panels

of material that are cut and seamed in a manner to approximate the parabolic reflector surface.

We will investigate the ability of such a reflector to achieve the desired parabolic shape. For these

constructions, the accuracy of the pressurized reflector will depend on the cutting pattern.

Electrical Feed
Gravity Vector & F Rigid Support Structure
Polar Direction Angle
Transparent Canopy
Reflector Rim
g
k

Aluminum Coated Reflector

Figure 2.1: Inflatable parabolic reflector antenna.

2.1 Mathematical Model

In our mathematical model, we approximate the membrane by a faceted surface, using constant

strain, planestress triangular finite elements. Figure 2.2(a) is an image of the actual faceted

surface that we use for a molded reflector in computations. Although the coated reflector is labeled

in Figure 2.1 (for exposition), we do not actually incorporate the effect of the vaporized aluminum

in our model. The entire membrane is modeled with the material Kapton for which the mechanical

properties are known; see Table 4.2 in Chapter 4. The material parameters will therefore be

identical for both the upper and lower paraboloids in our investigations.

In order to generate the faceted surface, a parabolic mold is created by the parametric equa-

tions

x = a + r cos ; y = a + r sin ; z = r 2 /(4F ), (2.1)

15
for 0 2 and 0 r R f where R f is the radius of the reflector aperture. In Equations (2.1),

a 0 is a translation that is used to generate a mold for an offaxis reflector. For axisymmetric

reflectors, we take a = 0. The paraboloid of Equations (2.1) is then translated and rotated so that

the rim lies in the plane z = 0 and vertex lies on the zaxis. A symmetric paraboloid is formed

over the plane z = 0 to generate the canopy and produce a clam shell mold. The triangles of

the faceted surface are formed by connecting vertices that initially lie on this mold. To simulate

the molding process of the reflector, we take this initial configuration as the unstrained reference

configuration R ; see Figure 2.2(a) for an axisymmetric model. Figure 2.2(b) shows this reference

configuration for offaxis model. Simulating the two flat panel constructions will involve a slightly

different geometry for R to be described in Section 2.2.

The supporting rubber rings are modeled as linearly elastic strings which are hereafter referred

to as supporting tendons. When the envelope is pressurized and the support tendon loads are ap-

plied, the entire inflatable reflector system will seek an equilibrium configuration of minimum

potential energy. In the model, the pressure, tendon stiffness and gravity are treated as parame-

ters which can be modified for parametric studies. Evaluating the shape for various pressure levels

and orientations is of interest as the reflector is likely to operate in different gravity environments.

We follow an optimizationbased solution process which determines the equilibrium position

of the inflated reflector by minimizing the total energy of the discretized system [2]. Consider a

deformed configuration of the inflatable reflector system given by = x(R ) where

x : R R3

is a function that moves each vertex of the unstrained reference configuration to the deformed

configuration. The reference configuration R may be either a subset of R2 or R3 depending on

whether the surface construction is molded or is comprised of developable surfaces which can be

laid out in a plane. Since the faceted surface is ultimately piecewise linear, we require x D where

D is the completion of C 1 (R ; R3 ) with respect to the ||x||1,4 norm.

For a particular configuration of the faceted reflector x(R ) R3 , the total energy of the entire

16
(a) Axisymmetric construction

(b) Offaxis construction

Figure 2.2: Faceted surface for molded antennas. Dots highlight the reflector rim.

inflatable reflector system is calculated by the functional

E T (x) = E f (x) + E p (x) + S f (x) + S t (x) (2.2)

where E f (x) is the gravitational potential energy of the reflector film, E p (x) is the potential energy

of the inflation gas, S f (x) is the relaxed strain energy in the membrane, and S t (x) is the relaxed

strain energy in the supporting tendons.

The following sections include the details of the computations for each of the energy compo-

nents. Following will be the reference configurations for the flat panel designs.

2.1.1 Gravitational and Hydrostatic Pressure Potentials

We include the direction and magnitude of gravity as parameters in our model since the antennas in

question may be operating in a variety of extraterrestrial environments. Given that these antennas

are tested on the ground and used in space applications, one may wish to explore the effect of

zero gravity, a 1g environment and the gravitational environments of other planets. The total

17
gravitational potential energy is calculated by

ZZ
E f (x) = w f (x g) dA (2.3)
R

where w f is the film weight density, g is the directed acceleration due to gravity and dA is the

area measure over the faceted reference configuration R . In our model, we compute this quantity

exactly for each triangular facet and then sum over the set of triangles.

For the deformed configuration , let B be the enclosed interior space. We assume that the

inflation gas is not buoyant and that the differential pressure is the constant, p0 . Therefore, the

hydrostatic pressure potential energy of the system is given by

Z ZZ ZZ
E p (x) = p0 d V =
p0 k n
dS (2.4)
B

where n
is the unit outward normal and dS is the surface area measure over the membrane. The

divergence theorem was applied to obtain the last equality in Equation (2.4). The energy E p can

be calculated exactly for the faceted surface .

2.1.2 Film and Tendon Strain Energy

We calculate the total stain energy stored in the film using a strain energy density function. Be-

fore presenting the details of this energy computation, we recall some terminology and results of

elasticity theory.

Deformation and Strain

Strain is a measure of deformation that a material body undergoes when either an internal or ex-

ternal force is applied. This differs from a rigid body displacement in that said deformation entails

a change in the relative positions of internal material points. Figure 2.3 depicts the deformation

of a body.

Let {e1 , e2 , e3 } be a basis for R3 and {1 , 2 , 3 } be the corresponding independent coordinates

for this basis. A deformation x of a material M has three independent components and can be

18
expressed generally as

x = x 1 (1 , 2 , 3 )e1 + x 2 (1 , 2 , 3 )e2 + x 3 (1 , 2 , 3 )e3 .

Consider two points which are infinitesimally close in the material M , p = (p1 , p2 , p3 ) and p + p

where p = (p1 , p2 , p3 ).

M x(M )
x(p + p)
p + p
e3 p x(p)

x
e2
e1
Figure 2.3: Deformation of a material body M under a deformation x.

A deformation of M will move both points to new locations given by x(p) and x(p + p). Since

p is a small displacement, we assume that x(p + p) x(p) is relatively small. In other words, we

assume that x is a continuous and differentiable material deformation. We may therefore expand

the function components x i in a Taylor series about the component p j , i.e.,

xi 1 2 xi
x i (p j + p j ) = x i (p j ) + p j + (p j )2 + O(p j )3 .
j 2 2j

In the assumption that the difference p j is infinitesimal, we are justified in neglecting the terms

of order (p j )2 and higher. Hence,

xi
x i (p j + p j ) = x i (p j ) + p j . (2.5)
j

Equation (2.5) may expressed in the form x i (p j + p j ) = x i (p j ) + ei j p j + i j p j where

1 1

xi xj xi xj
ei j = + and i j =
2 j i 2 j i

19
are the differential displacements generating strain and rotation, respectively [6, Section 10.2].

Notice that Equation (2.5) represents nine different equations since the indices i and j each run

from 1 to 3. There are nine partial derivatives of the form x i / j for this deformation x which

make up the gradient of the deformation. The gradient x is the deformation gradient and is the

second order tensor explicitly given by

3
3 X
X xi
x = ei e j .
i=1 j=1
j

Since x is a tensor of second order, it may be expressed as the 33 matrix [8, Section 1.8]


xi
x

ij =
j

for 1 i, j 3. If x is a differentiable deformation at a point p, then we can write

x(p + p) x(p) = x(p)p + o ||p|| .




Taking the transpose and multiplying on the left gives,

||x(p + p) x(p)||2 = p T x(p) T x(p)p + o ||p||2 . (2.6)




The CauchyGreen Strain Tensor is the second order tensor produced in Equation (2.6), namely

C = x T x.

Note that C is a symmetric tensor by construction. This tensor is also positive definite at all points

p where x is differentiable and it can used to compute the length of a curve deformed by x [8,

Section 1.8].

Not all deformations involve strain. For example, a deformation which consists simply of

translations and rotations of the material will not impose strain on the material. In fact, C = I if

and only if the deformation is a rigid one; see [8, Theorem 1.81]. Since we might expect that

the strain tensor should be zero for rigid deformations, we are lead to consider the material strain

20
tensor [1, Section 12.2]
1 1
G = 2
(C I) = 2
( x T x I ).

The tensor G is also referred to as the CauchySt. Venant strain tensor [8, Section 1.8] or the

engineering strain tensor. Note that G is a symmetric tensor by its construction and by the symmetry

of C. Also, G = 0 if and only if the deformation x is rigid. The matrix equivalent form of this tensor

has the linearized components of the strain of the material


1 12 13

G = 21

2 23


31 32 3

where we adopt the convention i = ii and observe that i j = ji by the symmetry G.

Stress and Constitutive Relations

At any given point, p in a material, the stress is a ratio of the force acting on an infinitesimal planar

slice of material at p to the area of that planar slice. Namely, we can define a stress component by

f ej
i j (p) = lim
A0 A

where A is the area of a planar slice of the material at point p which has normal vector ei and f is

the force acting on that planar slice. See Figure 2.4 for a diagram.

P
A Net stress on A at p:
~ 2 (p) = 2i ei
23
p
22 = 2 M
21
e3
A = e2
n

e2
e1
Figure 2.4: Stress components 2i at point p in material M .

21
Since there are three independent planes at any point, there are nine scalars to fully describe

the stress in a material at that point. These stress components are elements of the Cauchy stress

tensor
1 12 13
~
1
T = 21 2 23 =

~
2

31 32 3
~3

where
~ i are the vector stress components as shown in Figure 2.4. It can be shown that T is a

symmetric tensor using equations of equilibrium for the stress and force components. If F is a

force intensity throughout a static body M that occupies volume B, then the stresses in that body

must satisfy the force equilibrium condition,

Z ZZ ZZ
F dV + dS = 0
T n
B B

where dS is Lebesgue surface area measure [6, Section 9.3]. Applying the Divergence Theorem to

the second integral, we obtain

Z ZZ Z ZZ
F dV + T d V = 0.
B B

Since this equation must hold for any volume, B, we obtain the equilibrium condition

F + T = 0 (2.7)

where the divergence of the second order tensor T can be obtained by the identity presented in

Definition 1.3.7. We must also have a balance of the moments about the origin at all points in the

static body. We have Z ZZ ZZ


r F dV + dS = 0
r T n (2.8)

B B

where T n
is the vector component of the stress tensor in the direction of the normal to the surface

B and r is a position vector in the volume B. Applying the Divergence Theorem to the second

22
integral in Equation (2.8) gives

ZZ Z Z ZZ
r F dV + r T d V = 0.

B B

Again, since this equation must hold for an arbitrary volume, we have

rF + rT = 0. (2.9)


If we consider the i, j and k components of Equation (2.9), we obtain

r2 F3 r3 F2 = (r2
~ 3 r3
~ 2 ) = r2 ( ~ 2 ) + 32 23 ;
~ 3 ) r3 (

r3 F1 r1 F3 = (r3
~ 1 r1
~ 3 ) = r3 ( ~ 3 ) + 13 31 ;
~ 1 ) r1 (

r1 F2 r2 F1 = (r1
~ 2 r2
~ 1 ) = r1 ( ~ 1 ) + 21 12 .
~ 2 ) r2 (

Applying the corresponding components of Equation (2.7), we obtain i j ji = 0 for i 6= j. Thus,

T is symmetric.

The Cauchy stress tensor contains the stress components depending on the coordinate frame

of the deformed configuration. In our model, we calculate the strain energy over the reference

configuration. We will use the second PiolaKirchoff stress tensor since the stress components in

that tensor are in terms of coordinates of the reference configuration. The second PiolaKirchoff

stress tensor is

S = J(x)1 T(x)T

where J = detx [8, Section 2.6]. One can show that S is a symmetric tensor by using the

symmetry of T and noting that

S T = J (x)1 T(x)T = J(x)1 T T (x)T .


T

The strain that an elastic material undergoes is directly related to the stress in the material. The

stress and strain tensors that were defined are related by the generalized Hookes Law. In particu-

lar,

i j = ki1 1 j + ki2 2 j + ki3 3 j + ki4 4 j + ki5 5 j + ki6 6 j . (2.10)

23
This linear system may be inverted and solved for the strain components in terms of the stresses.

For a homogeneous material, small elastic deformations are governed by the more general tensor

relation

Si j = Ki jkl Gkl

where Ki jkl is the fourth order tensor of elastic moduli. This tensor will depend on the nature of the

material and its mechanical properties. Since we know that G and S are both symmetric tensors,

we have the conditions K jikl = Ki jkl = Ki jl k . Thus, there are only 36 independent elements among

the 81 elements in K. These independent elements are the constants ki j in Equation (2.10).

Film Strain Energy

In this section, we derive the total energy contained in the strained film, S f (x). Since the reflector

material has thickness on the order of microns (106 meters), we assume a state of plane stress

in the membrane and a uniform material thickness, h, throughout. A derivation of the stress and

strain tensors gives similar results to those seen above for the three dimensional case; see [6,

Chapters 1 and 2]. In particular, symmetry of the strain and stress tensors still holds and they take

the form
1 0 1 0
G = and S =
0 2 0 2

when expressed in terms of the principal directions.

The strain energy in the film is calculated using the relaxed strain energy density W f by the

equation ZZ
S f (x) = W f (x) dA
R

where dA is the area measure of R . The strain energy density function of a membrane is

1
Wf = 2
G :S (2.11)

where : is the tensor inner product as defined in Definition 1.3.5. Equation (2.11) is due to

Koiter and is derived by Ciarlet in [9, Pg. 548]. The full shell equations of Koiter include a

bending energy component which is O(h3 ). We neglect the bending energy for our problem since

24
the reflector membrane has thickness 106 meters. The strain energy of Equation (2.11) is O(h).

We assume that the material is linearly elastic and isotropic. An elastic body is one which

regains its original configuration when the forces acting on it are removed. A material is said to be

isotropic if its mechanical properties related to stress at a point, are identical in all directions [6,

Section 3.1]. In this case, the strain and stress tensors are related by

hE
G + cof(G) T , (2.12)

S =
1 2

where cof(G) is the 22 cofactor matrix of G, E is the Youngs Modulus of the film and is the

Poissons Ratio of the film. Note that G and S will have the same principal axes (eigenvectors)

since they are related by a linear stressstrain constitutive relation for an isotropic film [2]. Also,

G and S are symmetric tensors so the Spectral Representation Theorem implies that

G = 1 n1 n1 + 2 n2 n2

S = 1 n1 n1 + 2 n2 n2

where n1 and n2 are orthonormal vectors which we take to be the principal directions.

Since each triangle in the faceted surface is assumed to have constant stress within, each tri-

angle will have its own principal directions, principal strains and stress resultants. For a triangular

facet T R , the principal strains are 1 and 2 and the principal stresses are 1 and 2 . Applying

the spectral decompositions of G and S along with the Equation (2.12) to Equation (2.11) gives

12 + 22 + 21 2
W f (T ) = hE .
2(1 2 )

A thin compliant membrane such as those used for these inflatable antennas, will not resist

compressive stresses but will wrinkle instead. We model this wrinkling with the relaxed strain

energy density, W f . The method applied here is due to Pipkin and involves partitioning the faceted

surface into three sets: the set of slack facets, the set of wrinkled facets, and the set of taut facets

[26]. For any given triangular facet, T R , the state is determined by the principal strains and

stresses. The conditions of Equation (2.13) describe the possible states: slack (a), wrinkled (b,c)

25
or taut (d). The relaxed strain energy density for triangle T is then calculated by


0 if 1 < 0 and 2 < 0; (a)






22



if 1 0 and 2 0; (b)

hE
2





W f (T ) = (2.13)
12



hE if 2 0 and 1 0; (c)


2








2 + 22 + 21 2



hE 1 if 1 > 0 and 2 > 0. (d)

2(1 2 )

See [2] for additional details of the relaxed strain energy for an isotropic film.

Tendon Strain Energy

Let Nt be the total number of tendons. The tendons are modeled as linearly elastic strings. There-

fore, the relaxed strain energy density in the mth supporting tendon is



0 if m < 0;



Wt,m = (2.14)
2

m


if m 0,

K t,m
2

where K t,m and m are the tendon stiffness constant and strain of the tendon, respectively [2]. The

strain in the tendon is calculated by

m m,R
m = (2.15)
m,R

where m and m,R are the lengths of the mth tendon in the deformed and reference configurations,

respectively. The initial tension in the supporting tendons can be adjusted by foreshortening all

the lengths, m,R by a percentage . Support loss in a region can be modeled by setting K t,m = 0

for desired values of m. We can also simulate nonsymmetric loading by modifying the stiffness

26
constants K t,m as necessary. The total strain energy in the tendons can then be found by calculating

Nt Z m,R
!
X
S t (x) =
Wt,m (s) ds .
m=1 0

For computational simplicity, the rigid support structure (to which one end of each tendon is at-

tached) is always in the plane z = 0.

2.2 Flat Panel Constructions

As previously mentioned, inflatable antennas are constructed by molding sheets of material into

parabolic form. These antennas are most easily modeled by taking a reference configuration R

which is comprised of vertices that are set initially on a paraboloid. Figure 2.2(a) shows the initial

configuration for an axisymmetric reflector. While this geometry is ideal to approximate the design

of the antennas, it is not clear what impact the molding process might have on the mechanical

properties of the material. Also, since casting material to a certain shape on a precision mandrel is

an expensive process, constructing an antenna of diameter near to 10 meters may be infeasible. It

is reasonable, therefore, to consider other geometric constructions that involve only flat sheets of

material that are seamed together.

2.2.1 Flat Gore Construction

The reference configuration for the flat gore construction is generated by initially setting vertices

along parabolic arcs extending from the vertex to the rim. The arcs fit the desired parabolic shape

of the ideal reflector. There is one arc of vertices on the boundary of each pair of adjacent gores.

The vertices of the ribs used to generate the entire configuration are highlighted in Figure 2.5 as

dots over the corresponding vertices.

The remaining vertices for the reference configuration are positioned on straight lines between

adjacent pairs of these arcs so that each pie slice is a developable surface (a surface that can be

flattened into a plane without any strain or distortion). We then generate the triangular mesh

of this construction with these vertices. The result is a reference configuration that models a

27
Figure 2.5: Flat gore construction. Generating ribs indicated by dots.

surface having no curvature within the panels in the circumferential direction. Notice that the

distribution and pattern of the triangular mesh is similar to the distribution for the molded antenna

of Figure 2.2(a). The configurations differ in that a gore of Figure 2.5 is initially a developable

surface while a typical gore of Figures 2.2(a) and 2.2(b) is initially doubly curved.

The description in the previous paragraph describes the construction of an unstrained spacial

configuration for this reflector. The actual reference configuration, however, would involve cutting

the gores apart and laying them out in the plane as seen in Figure 2.6. Note that the gores are

paired and that each pair is joined by a common section, Si at the vertex to form a super gore.

For example, gores 1 and 2 are both attached to the unshaded region S1 . We construct the

gores in this fashion in order to avoid a clustering of narrow triangles around the paraboloids

vertex. The arrangement of the triangles in the Si region still allows the vertices to lie on a perfect

paraboloid (as shown in Figure 2.5) such that the triangles are unstrained relative to the reference

configuration of Figure 2.6. If N g is the total number of gores, then the reference configuration R

is
Ng N g /2

S j R2 .
[ [
R = i

i=1 j=1

2.2.2 Flat Band Construction

The other flat construction pattern consists of concentric bands of material which is a surface whose

panels have no curvature in the radial direction but are curved in the circumferential direction.

The different panels are modeled in the reference configuration by first fixing rings of vertices to

28
v

1 2 3 4

Separation of two
gores begins here
S1 S2
Vertex u

Figure 2.6: Four gores of the reference configuration for the flat gore construction.

the desired parabolic shape. These are shown by the rings of dots in Figure 2.7. Rows of triangles

are then generated between the rings to complete the reference configuration. These bands are

sections of cones and are therefore constructible from flat panels of material.

Figure 2.7: Flat ring construction. Generating rings indicated by dots.

In addition to lower construction costs, there are computational advantages to using these flat

panel constructions in our model. We will develop in Chapter 5, a parameter that adjusts the

geometry of the flat band construction. The parameter can then be used to reduce the surface

error. A geometric modification to these constructions translates only to a different cutting pattern

of the panels and therefore very little change in the actual construction procedure. Modifications

are not reasonable to consider in the molded construction case since there is not necessarily a

precision mandrel available to generate a geometry deemed appropriate.

29
Chapter 3

Parabolic Reflector Efficiency

In this chapter, we will establish some methods of quantifying the performance of a particular

antenna reflector shape. We consider two different classes of measurement. The first technique in-

volves characterizing the antenna. Characterization of an antenna entails analyzing how the reflec-

tor in question transmits radiation in certain directions. The electromagnetic fields that comprise

the signal sent must be computed so that both the magnitude and frequency of the transmitted

signal can be determined. These measurements require an understanding of antenna theory as

well as electromagnetic theory.

The other method of measurement relates purely to the geometric distortions in the sense that

only the displacement of the antenna reflector from a parabolic shape is considered. These errors

are referred to as RMS error measurements. We will present two different types of RMS error,

radiometric and Euclidean.

3.1 Electromagnetic Theory

In order to evaluate the effectiveness of a certain antenna geometry, we must understand how it

reflects radiation. The media by which a signal travels is electromagnetic radiation and thus, some

background in Electromagnetic Theory is necessary. We present some material from this theory

and Maxwells system of differential equations that govern the behavior of radiation. Note that
p
the notation for the complex number i = 1 is used throughout. The engineering notation in

electromagnetic theory for the imaginary unit is j.

30
3.1.1 Radiation, Electric and Magnetic Fields

The fundamental electromagnetic particles are the electron, the proton and the neutron. Each is

attributed with a certain charge and mass that is listed in Table 3.1 [17, Section 266]. Charge is

a quantity measured in coulombs and is analogous to the notion of mass in mechanics.

Particle Symbol Charge (coul) Mass (kg)


Electron e 1.602 1019 9.10910 1031
Proton p 0 1.67482 1027
Neutron n +1.602 1019 1.67252 1027

Table 3.1: Charge and mass of the three fundamental charged particles.

Coulombs Law for force, F, acting on electric charges (analogous to Newtons law of gravitation

for masses) is
q1 q2 r 2 r1
F =
40 | r2 r1 |3

where q1 and q2 are point charges positioned at r1 and r2 , respectively. It is important to note

that the force repels the two charges when they have the same sign and attracts them when the

charge signs are opposite. The constant 0 is the permittivity of free space and has an experimental

value of 0 = 8.854 1012 coul2 / N m2 [17, Section 264]. Note also that the coefficient in

Coloumbs Law is
1
= 8.9874 109 N m2 / coul2 .
40

Electric Fields and Gauss Law

Definition 3.1.1. Suppose there are a collection of charges in space which together generate an electric

force field. The electric field is the limit of the ratio of the force on a test charge q to the charge value

as q 0, or
Fq (r)
E(r) = lim
q0 q

where r is the spatial position of the test charge. The electric field has units of volts per meter; see [17,

Section 272].

While Definition 3.1.1 is an intuitive definition, electric fields are rarely measured in this man-

ner due to experimental difficulties. Electric fields are typically computed from the scalar electric

31
potential, V . The electric potential has units of volts. One can find the electrostatic potential

energy at any point r by integrating over the charge density (r ) throughout space [23, Section

24]. We have,
1 1
ZZ Z
V (r) = (r )dr .
40 R3
| r r |

The electric potential energy is related to the electric field by E = V . Note also that an electric

field generated by n point source charges can be found by the vector sum

n
X qi r ri
E(r) = . (3.1)
i=1
40 | r ri |3

Theorem 3.1.1. Gauss Law states that the total electric flux through a closed surface, , is propor-

tional to the total charge enclosed by the surface. In particular,

ZZ n
X qi

E n
dS = . (3.2)
i=1
0

Furthermore, any charge located outside gives no contribution to the value of the integral on the left

hand side of Equation (3.2); see [23, Section 26].

Proof. This law can be shown by considering an arbitrary closed surface enclosing a collection of

n charges. Recalling Equation (3.1), we have

n
1
ZZ
r ri
ZZ
X

E n
dS = qi
n
dS . (3.3)

40 i=1 | r r i |3

To handle the integral on the right hand side, we recall Vi (r) = | r ri |1 which is the potential

generated by the charge qi . Notice that

r ri
Vi = and 2 Vi = 0
| r r i |3

for r 6= ri . These two relations can be seen by applying the Laplacian and gradient operators for

spherical coordinates. Recall that [35, Pg. 188]

2V 2 V csc V csc2 2 V
 
2
V = + + sin + .
r2 r r r2 r2 2

32
It remains to show that

4 if qi is inside ;
ZZ

Vi n
dS = (3.4)

0 if qi is outside .

If qi is located outside of the surface , then 2 Vi = 0 everywhere inside . By Greens Theorem,

the integral in Equation (3.4) vanishes. This establishes that no charge outside contributes to

the flux of E through .

For qi inside , we take the usual approach to handling singularities in the study of Laplaces

Equation [32, Section 4.2]. Take a sphere, with radius which is centered at the singularity ri .

Notice that | r ri | = on and that the normal of is directed inward along the vector r ri .

Hence,
r ri (r ri )
ZZ ZZ

Vi n
dS =
dS.
| r ri |3 | r ri |

We reduced this integral to one over the small sphere since the potential function Vi is harmonic

elsewhere within the closed surface S. Changing to spherical coordinates gives,

Z 2
1
ZZ Z
dS = lim
Vi n 2 sin d d = 4.

0
0 2

Applying the Divergence Theorem Equation (1.5) to Gauss Law, we obtain




ZZ Z ZZ Z

E dV = dV (3.5)
B B
0

where B = . Since Equation (3.5) must hold for any simply connected region B, we have


E = . (3.6)
0

Current and the Continuity Equation

In the previous section, the electric fields were computed assuming that the charges generating

them were fixed for all time. We now introduce the notion of traveling charge which is given the

33
term current.

Definition 3.1.2. Current is the rate of flow of charge in a conducting medium past a given point. If

q(t) is the net charge transported past a particular point at time t, then current is defined by

dq
I =
dt

at said point. The unit of current is the ampere equal to 1 coulomb / second; see [23, Section 72].

Assume that a conducting medium is carrying charged particles qi at corresponding velocities

vi . Then the differential current through an element of area d a is given by

!
n
X
dI = qi Ni vi n
da
i=1

P
where the quantity J = i qi Ni vi is the current density; see Figure 3.1. Current is usually the

result of one class of charge carrier namely, electrons. Hence, we can simplify the expression of

current to

d I = qe N v n
da

where qe = 1.602 1019 coul, the charge of one electron. Here, N is the number of electrons in

the element d a and v is assumed to be a uniform velocity of the electrons.

da
v
e e
e

n
e
e e

e e

Figure 3.1: Charge with velocity v flowing through a differential cross section, d a.

34
It follows that the current passing through any surface S is obtained immediately by

ZZ
I = Jn
dS.
S

For an arbitrary closed surface B, we can relate the current passing through the surface to the

charge density in the enclosed region B. Since current is also the time derivative of charge, we

form the equation ZZZ ZZ


d dq
(r, t) d V = =
J n
dS,
dt B
dt B

where the negative sign indicates an increase in charge density for an inward flow of current.

Using the Dominated Convergence Theorem on the left hand side and the Divergence Theorem on

the right, we see that Z ZZ Z ZZ



dV = J dV
B
t B

for an arbitrary region B. We therefore obtain the Continuity Equation


J+ = 0. (3.7)
t

Magnetic Fields and BiotSavart Law

The magnetic field is as important to electromagnetic theory as the electric field. We begin with

the magnetic induction field which is defined using the notion of a test charge.

Definition 3.1.3. The magnetic induction B is the vector field which satisfies the relation

F = qv B

for all velocities v. The unit of magnetic induction is the tesla (T) equal to 1 Weber / m2 .

A Weber is related to the fundamental SI units by 1 Weber = 1 Joule / Ampere. Let us also

simultaneously introduce the magnetic intensity field H,

B = 0 H

where 0 = 4 107 N / Amp2 is the permeability of free space [33, Pg. 6]. The field H has

35
units of Ampere / meter.

The relation in Definition 3.1.3 is limited to the computation of a magnetic force on a charge.

In order to calculate the magnetic induction at a point, we first derive a new expression that relates

B to current and then apply a law due to Biot and Savart.

Consider a line of current, C, having electrons as the charge carrier. First note that the force

on the element of conducting material d~ from Definition 3.1.3 can be expressed as

dF = N A|d~|qe v B

where N is the number of electrons per unit volume, A is the cross section of the conducting

material and d~ is the differential direction vector of the current. Since d~ and v are parallel

vectors, we can rewrite this expression as

dF = N A|v|qd~ B.

Recalling the relations developed for the computation of current, we find that the force on a test

charge as a result of the moving current is given by

I
F = I d~ B. (3.8)
C

The BiotSavart Law for the force of one loop of current on another is given by

d~2 d~1 (r2 r1 )


I I 
0
F =
4 C1 C2 | r2 r1 | 3

where C1 and C2 are the two paths of the two current loops [23, Section 83].

From the BiotSavart Law and Equation (3.8), we obtain the magnetic field generated by the

first loop of current, C1 as


I1 d~ (r r1 )
I
0
B(r) = .
4 C1 | r r1 | 3

Rather than integrating along C1 , we can (equivalently) integrate over a volume B that encloses

36
the contour C1 and define the current density J that has supp(J) = C1 . Then,

r r1
ZZZ
0
B(r) = J(r1 ) d V.
4 B | r r1 | 3

Take the divergence of both sides to obtain

r r1
Z ZZ
0
 
B(r) = J(r1 ) d V. (3.9)
4 B | r r1 | 3

We refer to the vector identity in Equation (1.10) while noting that J(r1 ) is constant. Thus, Equa-

tion (3.9) becomes

r r1
Z ZZ
0
 
B(r) = J(r1 ) d V. (3.10)
4 B | r r1 | 3

The term within the parenthesis of Equation (3.10) is a curl of a gradient. Therefore, by the vector

identity of Equation (1.8), we obtain

B = 0. (3.11)

This is a significant result since the Divergence Theorem implies that the magnetic flux through

any closed surface S must be zero [23, Pg. 154]. Hence, no isolated magnetic poles exist.

3.1.2 Maxwells Equations

Maxwells Equations consist of Equations (3.6) and (3.11) of the previous section as well as Equa-

tions (3.14) and (3.15) of this section. The last two equations can be derived from widely accepted

observations known as Amperes Law and Faradys Law.

The Laws of Ampere and Faraday

Amperes Law states that the magnetic field due to a distribution of current satisfies

I ZZ
H d~ = Jn
dS
C

where C is the contour boundary of some surface [23, Section 151]. This law, however, fails

for a class of examples involving capacitors as shown in Figure 3.2.

37
S2
S1 Capacitor Plates

I 1
n
C
2
n

Figure 3.2: A contour C attached to two surfaces passing through different current.

Example 3.1.1. From Figure 3.2, we can apply Amperes law to the surface S1 where C is the contour

boundary. We can also use Amperes Law with S2 where C is again the contour boundary. Note that

we obtain ZZ I ZZ
I = Jn
1 dS = H d~ = 2 dS = 0
Jn (3.12)
S1 C S2

which is a contradiction.

This contradiction arises because

ZZ ZZ
Jn
2 dS 1 dS 6= 0.
Jn (3.13)
S2 S1

We notice that the integral difference in Equation (3.13) is the closed surface integral on S1 S2 .

Notice that n
1 is in the direction of the current in Figure 3.2 and so
n1 is needed for the outward

normal. In order to eliminate the contradiction we found in Equation (3.12), we need a vector

field J which has the property that

ZZ ZZ ZZ

J n
dS = Jn
2 dS 1 dS = 0.
Jn
S1 S2 S2 S1

Since S1 S2 is a closed surface, the Divergence Theorem implies that this condition is equivalent

to finding a vector field with J = 0. The Continuity Condition in Equation (3.7) and Gauss

Law in Equation (3.6) suggest that we should select

E
J = J + 0 .
t

38
Taking the divergence, we indeed obtain

E (/0 )
 

J = J + 0 = 0 = 0.
t t t

This major contribution is due Maxwell and yields the modified Amperes Law

I ZZ 
E

H d~ = J + 0 n
dS
C
t

which holds for an arbitrary surface . An application of Stokes Theorem Equation (1.6) on the


right hand side allows us to find


E
H 0 = J. (3.14)
t

Farady hypothesized that the electromotive force in a circuit is related to the magnetic flux by

I ZZ
d
E d~ + dS = 0
Bn

dt

where is the closed contour boundary of the surface . Using Stokes theorem on the left hand

side and using the Dominated Convergence Theorem on the right hand side, we see that

ZZ 
B

E+ dS = 0
n

t

for any simple C 1 surface [23, Pg. 171]. This implies that the integrand must be the zero vector

which yields the equation


B
E+ = 0. (3.15)
t

Phasors and Maxwells Equations

Before proceeding, we will change the mathematical nature of the vector fields that have been

presented in this chapter. We assume that the radiation waves in the scope of this dissertation

will be monochromatic, i.e., characterized by a single frequency, . The time dependence of the

fields is introduced with a factor of e i(t+) to incorporate the oscillation of the radiation waves of

frequency . The angle is the phase of the radiation. From this point forward, the fields E and

39
B are therefore understood to represent

E(r)e i e it and B(r)e i e it ,


 
E(r, t) = B(r, t) =

respectively. These vector fields with the exponential factor amended are known as phasors and

encompass (in addition to E and B) the fields H and J. We can therefore express the Equa-

tions (3.6), (3.11), (3.14) and (3.15) as

E = /0 (3.16)

H = 0 (3.17)

H i0 E = J (3.18)

E + i0 H = 0 (3.19)

p
where is the frequency of the radiation wave and i = 1. Equations (3.16) (3.18) are

Maxwells Equations for monochromatic electromagnetic waves.

We assume that the region where these waves live is composed of a single media and that the

region ends where a new media is encountered. This can be physically represented by a wall that

interferes with radiation waves, a boundary between oil and water, or any other obstruction that

impedes electromagnetic waves. A boundary is essentially any place where refraction or reflection

occurs. The main boundary of interest for our problem, is the antenna reflector surface. The

boundary conditions that we will use are

(H2 H1 ) = Js
n and (E2 E1 ) n
= Ms

where n
is the unit normal of the boundary surface pointing into medium 2. The fields Js and Ms

are the electric and magnetic surface currents at the boundary, respectively. See Figure 3.3 for an

illustration.

40
Ms
Medium 2
n Js

E2 H2 Medium 1
E1 H1

Figure 3.3: Boundary conditions for Maxwells Equations.

Solution to Maxwells Equations

A solution to Maxwells Equations can be found by taking advantage of the two decoupled equa-

tions among Equations (3.16) (3.19). From Equation (3.17), we see that the magnetic field H

has no divergence. By the vector identity in Equation (1.8) there exists a field A such that

H = A. (3.20)

The field A is referred to as the magnetic vector potential. Substituting Equation (3.20) into Equa-

tion (3.19) gives

(E + i0 A) = 0.

By Equation (1.7), there exists a scalar field such that

E + i0 A = . (3.21)

We refer to the scalar field as the electric scalar potential. From Equation (3.18) and the definition

of A, we have

A i0 E = J. (3.22)

Using Equations (1.9) and (3.21), Equation (3.22) becomes

( A) 2 A i0 (iA ) = J.

Linearity of the gradient then gives

2 A + 2 0 0 A (i0 + A) = J.

41
We chose the divergence of A such that the third term on the left hand side is zero [33, Pg. 10].

This leaves the differential equation of A,

2 A + 2 0 0 A = J. (3.23)

Solving this differential equation is equivalent to solving the system of scalar Helmholtz equations

2 + 2 A = J (3.24)


for = x, y, z and 2 = 2 0 0 .

Proposition 3.1.1. Suppose A satisfies the outgoing radiation condition given by

A 1
 
+ iA = O ,
r r2

which models the decay of the radiation as it travels from the current density source. Then the solution

to Equation (3.24) is
exp i| r r |
Z ZZ 

A (r) = J (r ) dr .
R3
4| r r |

See [25, Pg. 460].

Proof. Define the operator L(u) = 2 u + 2 u. A fundamental solution of the homogeneous form

of Equation (3.24) is
exp i| r r |


(r, r ) = (3.25)
| r r |

for r 6= r [12, Pg. 314]. Since there is a singularity in at r, we define the domain

n o
B = r R 3 | r r | R

and apply Greens Identity to the functions and A on the domain B . We obtain,

Z ZZ ZZ
A
 

L A A L dr = A (3.26)
 
dS
B B
n
n

where B has two components: a small sphere of radius centered at r, and a large outer sphere

42
of radius R also centered at r. For the sphere | r r | = , the normal to B is directed toward the

center point, r. Hence,

!
ei| rr | A ei| rr | i| r r | + 1
ZZ
I = A dS
| rr |=
| r r | n
| r r | 2

Z 2 i
ei i + 1
Z 
e A
= A 2 sin d d
0 0
n
2

Z Z 2 
A
  
i i
= e iA A e sin d d
0 0
n

Since the factors and ei are constant over the sphere | r r | = , they can be moved outside

the integral. Since the sphere | r r | = is centered at r, we obtain

I = 4A (r)ei + O(). (3.27)

On the outer component of the boundary, | r r | = R, we have

ZZ ZZ
A A
   
IR = A dS = A dS
| rr |=R
n
n
| rr |=R
r r

rr
where the second equality was obtained since n
= | rr |
for the sphere | r r | = R. Using the

outgoing radiation condition for both and A , we obtain

ZZ   
  
IR = iA + O r 2 A i + O r 2 dS.
| rr |=R

Making appropriate cancellations leaves

ZZ
O r 2 A R2 sin d d.
 
IR =
| rr |=R

Since and A have the same asymptotic behavior and since we know from Equation (3.25),

43
we can argue that | A | = O r 1 . Hence,


ZZ
O r 3 R2 sin d d = 4O R1 . (3.28)
 
IR =
| rr |=R

For the left hand side of Equation (3.26), recall that L() = 0 in B and L(A ) = J . Then,

ZZ Z Z ZZ

L A A L dr = J dr .
 
B B

Replacing this result along with I of Equation (3.27) and IR of Equation (3.28) into Equa-

tion (3.26) gives

ZZ Z
(r, r ) J (r ) dr = 4A (r) + O() + 4O R1 .

B

Taking the limit 0 and then R gives the desired result.

From Proposition 3.1.1, we combine the solutions for = x, y, z to get the vector solution

exp i| r r |
ZZ Z 

A = J(r ) dr . (3.29)
R3
4| r r |

For the antenna reflector problem, supp(J) is the reflector surface. Therefore, we reduce the

solution of Equation (3.29) to

exp i| r r |
ZZ 

A = J(r ) dS (3.30)
R
4| r r |

where R is the reflector surface. Equation (3.30) is an accepted computation for the field A in

antenna theory; see [33, Chapter 8] and [30].

With this solution for the magnetic vector potential, we can use Equation (3.20) to obtain

H = A.

We then use Equation (3.18) to recover the electric field

1 1
E = ( H J) = ( A J).
i0 i0

44
Using the vector identity from Equation (1.9), we can transform this equation to

1
( A) 2 A J .

E =
i0

Recalling the differential equation (3.23), we obtain

1
( A) + 2 A .

E =
i0

Using the definition of , we get the solution

( A)
E = i0 A. (3.31)
i0

3.2 Reflectivity Properties

In order to transmit large quantities of data, higher radiation frequencies are desired. As men-

tioned in Chapter 1, inflatable antennas are already in use and have been successful for L and S

band radiation frequencies [34]. In Table 3.2 we present the common radiation frequencies as

established by the IEEE [19].

Band L S C X Ku K Ka
(GHz) 12 24 48 8 12 12 18 18 27 27 40
(mm) 300 150 150 75 75 37.5 37.5 25 25 16.7 16.7 11.1 11.1 7.5

Table 3.2: Common IEEE Radiation bands.

The frequency of the radiation, plays an important role in the efficiency of the reflector being

used for an antenna. To understand this, let us first establish the exact geometry of a paraboloid.

Given a focus f = (0, 0, F ) and a directrix plane z = F , a paraboloid is the set of all points p R3

that satisfy

= d p, p
 
d f, p

where d(x, y) is standard Euclidean distance between x R3 and y R3 , p = (x, y, z) is an arbi-

trary point on the paraboloid and p is the point on the directrix plane closest to p. See Figure 3.4

for an illustration. The parabolic curve shown in Figure 3.4 can be revolved about the symmetric

45
axis (zaxis) to generate a paraboloid. In our application, F is the focal distance of the parabolic

reflector as described in Chapter 1.

z Aperture Plane: z = F + C
a

C
f Paraboloid: 4F z = r 2

p
r

D/2 D/2

p Directrix Plane: z = F

Figure 3.4: Cross sectional diagram of a paraboloid and its geometry.

The path that a radiation wave takes from the focus f to aperture point a, is by way of reflection

on the paraboloid at point p. There are two key properties of this path that are determined by the

paraboloids geometry. Before listing these properties, we perform some calculations with the

vectors a, f and p.

x2 + y2 x2 + y2

fp = x, y, F ap = 0, 0, F + C (3.32)
4F 4F

At point p on the paraboloid, a normal vector is

x2 + y2
x y
n = z = , ,1 . (3.33)
4F 2F 2F

46
The two properties of the radiation path are:

1. The radiation from f is reflected at p in a path that is parallel to the symmetric axis. This can

be shown by demonstrating that the angle between n and f p is the same as the angle

between n and a p. This is clear from the relation x y = | x | | y | cos and the equality

(a p) n (f p) n 1 (a p) n (f p) n
 
= = 0. (3.34)
|a p| |n| |f p| |n| |n| |a p| |f p|

The last equality in Equation (3.34) can be established by noting that

(a p) n = | a p | and (f p) n = | f p |

which are derived from Equations (3.32) and (3.33). It must also be shown that the vectors

(a p), (f p) and n are coplanar. This can be checked by using the relations in Equa-

tions (3.32) and (3.33) and then calculating

(a p) (f p) n = 0.

2. The total distance of the path that any line of radiation takes from f to the aperture is the

constant 2F + C. It is important that the aperture have a height value that exceeds the z

components of all points on the paraboloid. Often, the paraboloid lies entirely below the

focal point, and so we may take C = 0. Using Equation (3.32), it is straightforward to show

that

| f p | + | a p | = 2F + C. (3.35)

Since all reflected waves of radiation travel parallel to the zaxis, an observer looking at the

paraboloid from the +z axis (this is referred to as the boresight direction) will see all the reflected

paths of radiation coming directly to him. Equation (3.35) guarantees that said radiation will be

completely in phase, regardless of the where it was reflected on the paraboloid. This holds because

the phase is = ()/c where c is the speed of light and is the constant path length traveled.

The total sum of the radiation (signal seen by the observer) is therefore amplified by the constant

phase and direction of travel of the reflected radiation.

47
If the paraboloid is distorted, then some local areas of the reflector may reflect radiation in

a manner that the traveling to the distance is changed by . The phase error suffered by this

pathlength error is

= . (3.36)
c

If there is a discrepancy in the phase, then some cancellation will occur in the total sum of the

radiation and cause a reduction in the total signal. We see that the phase error is directly propor-

tional to the frequency of the radiation. Hence, phase error increases as the radiation frequency

increases. For this reason, antenna reflectors have lower distortion tolerance limits when operating

at or above KaBand versus below KaBand.

3.3 Antenna Efficiency

Antenna performance can be quantified by the amount of power that is lost between the feed

(electronic source of a signal) and the receiving probe. A collection of losses, each pertaining to a

different property of the antenna or accompanying apparatus, is calculated in order to ultimately

compute the amalgamated power loss and the antennas overall efficiency. For aperture antennas,

there are several factors that contribute to power loss during operation; we list some here.

Spillover loss is radiation from the electrical feed that is lost beyond the edge of the reflector.

In other words, the reflector simply does not intercept this outer edge radiation.

Taper loss is caused by underillumination of the reflector near the edge in order to reduce

spillover loss.

Random surface error loss is caused by Gaussian distributed path length error and can be

calculated by the Ruze Equation [33, Pg. 434].

Cross polarization loss refers to energy lost due to misalignment of the feed with the re-

ceiving probes.

Aperture blockage loss is caused when part of the supporting apparatus obstructs radiation

between the reflector and aperture window.

Feed phase error is nonuniformity of the radiation wave phase caused by the feed output.

48
Reflector phase error is nonuniformity of the radiation wave phase errors caused by re-

flector surface errors.

The scope of this project concerns only the reflector phase error component and so we will

neglect all other efficiencies. One should note that these other sources of power loss are not neces-

sarily subdued even if they are understood and controlled. For example, the maximum efficiency

that can be accomplished just by the trade off between the taper and spillover losses is 82% for

any parabolic antenna system [33, Pg. 434]. The remainder of Section 3.3 focuses on methods to

compute the phase error for a distorted reflector.

3.3.1 Far Field

The shape of a wave leaving a source, such as an AUT (Antenna Under Test), is spherical. When

the generating current of the AUT is sufficiently far away, the radiation waves appear to be planar

as depicted in Figure 3.5.

Spherical Waves Nearly Planar Waves

AUT Probe

Figure 3.5: Wave front shape in the near and far field of an AUT.

The far field region is where propagation lines of radiation arriving at the point P are suffi-

ciently parallel that planar wave structure can be assumed. Equivalently, the observer can make

the judgement that the entire AUT is at a particular distance away and hence contributions of

power at all positions on the aperture are from the same direction.

In order to evaluate the discrepancy between the distances R and r in Figure 3.6, we will need

the Binomial Theorem


n(n 1)
(1 + x)n = 1 + nx + x2 + . . . (3.37)
2!

for |x| < 1 [31, Section 456]. Figure 3.6 shows the parallel approximation that we make for paths

traveling a long distance to the far field. An observer at point P will actually intercept both paths

49
P

r
Symmetric Axis

r R = r cos

Figure 3.6: Geometry of rays extending to a point P in the far field.

and so the vectors R and r both pass through point P. We can write the exact relationship between

R, r and r as

R2 = (r r ) (r r ).

Expanding this product and using the angle between the two vectors r and r gives

2
r r

R2 = r 2 1 2 cos + .
r r

Taking the square root of both sides and applying the Binomial Theorem Equation (3.37) with


1
n= 2
and x = 1 2r r 1 cos , we find

(r )2
R = r r cos + + O(r 2 ). (3.38)
2r

Equation (3.38) contains the far field approximation

R = r r r (3.39)

which is exactly the geometric relation between the distances R and r as they are illustrated in

Figure 3.6. The discrepancy between the far field approximation of R in Equation (3.39) and the

exact computation of R in Equation (3.38) can be bounded by placing a restriction on the third

50
term of Equation (3.38). We require that the third and largest neglected term of Equation (3.38)

be no more than /16. This corresponds to onesixteenth of a wave length or 22.5 phase error,

which is accepted in antenna theory [33, Pg. 23]. We therefore require

(r )2 (D/2)2
=
2r 2r F 16

where D is the diameter of the reflector and r F is the distance to the far field. Solving for r F

gives a definition of the far field distance in terms of the reflectors diameter and the radar band

wavelength.

Definition 3.3.1. The far field is said to begin at a distance where the assumption that parallel lines

are converging yields a path length error of not more than /16 (corresponding to 22.5 phase error).

This distance is given by


2D2
rF =

away from the antenna, where D is the diameter of the aperture and is the wavelength of radiation.

In the far field, the reflected E and H fields are orthogonal to one another as well as the

direction of propagation, r. They are related by the Plane Wave Equation

r E
H = (3.40)

p
where = 0 /0 is the impedance of free space [33, Pg. 26].

The E field in the far field of the AUT is the second term in Equation (3.31) without the radial

component

E = i0 A + (i0 A r)r. (3.41)

3.3.2 Directivity and Power Gain

Directivity is a function that expresses how much the AUT prefers to radiate power in one direction

versus another. Before generating an equation for directivity, we must determine how power

is calculated. Poyntings Theorem provides us with a method of computing the power radiated

through a surface as a result of electric and magnetic fields. We take the dot product of Maxwells

51
Equations (3.19) and (3.18) with H and E respectively to get

( E) H + i| H |2 = 0 and ( H) E i0 | E |2 = J E.

Subtracting these equations gives

( E) H ( H) E + i| H |2 + i0 | E |2 + J E = 0.

We apply the vector identity of Equation (1.10) and find

(E H ) + i| H |2 + i0 | E |2 + J E = 0

where the superscripted star indicates complex conjugation. Integration over a region B enclosed

by the surface S leaves

ZZ Z ZZ

(E H ) n
dS + (i| H |2 + i0 | E |2 + J E) d V = 0 (3.42)
B B

where the Divergence Theorem has been applied to the first integral. The integrand E H is

known as the Poynting Vector. Poynting asserts that the surface integral in Equation (3.42) is the

power flowing out of the region B while the volume integral is the power within the region.

Since we are interested in the power obtained far from the antenna, we can take the surface S

in Equation (3.42) to be a sphere of radius R > r F . In this case, the power radiated through any

particular solid angle i is

!
1
ZZ
Pi = Re (E H ) r r 2 d
2 i

where d = sin d d. We assume that S is in the far field of the antenna, so the Plane Wave

Equation (3.40) applies. Hence,

!
1 1
ZZ ZZ
Pi = Re
E (r E ) d = | E |2 r 2 d. (3.43)
2 i
2 i

52
The integrand in Equation (3.43) is the radiation intensity for any given direction ( , ), i.e.,

| rE |2
U( , ) = . (3.44)
2

Radiation intensity per unit solid angle has units of watts / steradian. Typically, the maximum

radiation intensity occurs in the boresight direction, or ( , ) = (0, 0). The total power radiated

by the antenna is Z Z 2
Pr = max(U) U( , ) sin d d,
0 0

and the average power radiated per steradian is given by

Z 2
1
Z
Pr
U = U( , ) sin d d = .
4 0 0
4

It is often convenient to consider the radiation intensity normalized. This allows one to compare

how much radiation the antenna reflects to directions other than boresight. It is important to note

that radiation not radiated in the boresight direction is considered loss since the intended observer

is not at any other angle in the far field.

Definition 3.3.2. The far field pattern of an antenna is a normalized radiation intensity function, F

given by
U( , )
F ( , ) 2 =

.
max(U)

Note that F = 1 in the direction of maximum radiation intensity (usually the boresight direction).

Definition 3.3.3. Directivity is the ratio of radiation intensity in a certain direction ( , ) to the

average radiation intensity of the antenna. Hence,

U( , )
D( , ) = .
U

Definition 3.3.4. The maximum power one can receive from an antenna is known as the ideal an-

tenna gain and is expressed numerically as

4A
GI = (3.45)
2

53
where A is the area of the aperture [33, Pg. 433].

Equation (3.45) yields a calculation of the antenna gain assuming that the antenna is 100%

efficient. As mentioned at the start of this section, no aperture antenna can achieve an efficiency

of greater than 82%, so we will seek an alternative definition.

Definition 3.3.5. The effective antenna gain is 4 times the ratio of power radiated by the antenna

in a certain direction to the power transmitted by the feed. Mathematically, effective gain is

4U( , ) D( , )Pr
G( , ) = = ,
P0 P0

where P0 is the power transmitted by the antenna.

The effective antenna gain can also be defined as the ratio of the radiation intensity in a certain

direction to the radiation intensity of the isotropic radiator with power P0 . The isotropic radiator

will take the input power, P0 , and radiate over all solid angles. This justifies the division by 4 in

the denominator which presents itself in the numerator of Definition 3.3.5. If a discussion of gain,

G and/or directive gain, D ensues without any reference to an angular direction, it is presumed

that

G = max G( , ) and D = max D( , ).

Since power gain is a power ratio, it is often expressed in decibels:

GdB = 10 log G.

Example 3.3.1. Suppose we have an antenna which takes in a total power of 100 watts at the

electrical input terminals. Let the antenna have a parabolic reflector with an aperture radius of 0.25

meters and a radiation frequency of 20 GHz ( = 0.015 meter). The ideal antenna gain is therefore

4A 4 (0.25 meters)2
GI = = = 10966 = G I dB = 40.401 dB.
2 (0.015 meters)2

Remark: The significance of a 0 dB gain measure is that the antenna is behaving like an isotropic

radiator. In particular, the antenna performs 0 times better than the feed radiating without any

reflector.

54
Definition 3.3.6. The efficiency of an antenna is equal to the ratio of the effective power gain to the

ideal power gain, or


G
e = .
GI

In practice, a measurement of gain will reflect all power losses and so the efficiency as defined

in Definition 3.3.6 is the true efficiency of the antenna. In this report, we will be calculating

a theoretical power gain whose value is dependent only on the power input and the reflectors

surface distortions.

3.3.3 Antenna Characterization

Characterization of an antenna involves calculating the resulting far field magnetic and electric

fields that are reflected by the antenna. The distribution of the signal strength as a function of

angle to the far field is of interest. We present here the computation of the far field pattern for

an ideal paraboloid and then show the method we use to compute the pattern for a deformed

faceted surface. For our computations, we assume that the reflector has a reflectivity constant of

1 and neither absorbs nor refracts any radiation. We also assume that there are no obstructions or

changes in media between the feed and the receiving probe.

Ideal Parabolic Reflector

In this section, we calculate and plot the far field pattern for an ideal axisymmetric parabolic

reflector. From Equation (3.30), we can determine the magnetic vector potential in the far field.

In this case, we integrate over the aperture of the antenna, A . It is sufficient to integrate over

A since the entire radio signal passes through the aperture. We could also integrate the current

density over the reflector surface, R as shown in Equation (3.30). Integration over A , however is

less complicated when we assume a perfect parabolic reflector. We have,


ei| rr |
ZZ

A(r) = Ja (r ) dA (3.46)
A
4| r r |

where Ja is the current density at the aperture, r A is an aperture point and the far field

point is denoted r. We assume, without loss of generality, that the aperture is located in the plane

55
z = 0 and is centered at the origin. From far field approximation of Equation (3.39), we have

| r r | = r r r and | r r | r. Hence, Equation (3.46) can be expressed as

ei r
ZZ

A(r) = Ja (r )e irr dA .
4r A

Note that the field A is essentially obtained by a Fourier Transform of the current density field Ja

at the antennas aperture. From the boundary conditions of Maxwells Equations, we have

a HINC
Ja = n a

where n
a = k is the outward normal of the aperture and HINC
a is the incident magnetic field arriving

at the aperture. A key property of a perfect parabolic reflector is that the reflected fields E and H

are constant in the aperture. We may therefore assume that Ha = H0 j. Then, we have

ei r H0 ei r
ZZ ZZ
irr
A = a H0 j
n e dA = i e irr dA . (3.47)
4r A
4r A

The direction to the far field is r = (cos sin , sin sin , cos ). Recalling that r is in the circular

aperture, we can write r = (r cos , r sin , 0). Hence,

r r = r sin cos( ).

Integrating over the circular aperture, Equation (3.47) becomes,

2
Ra
!
H0 ei r
Z Z

sin cos( )
A = i e i r d r d r
4r 0 0

where R a is the radius of the aperture. The and r integrations are performed using the two

Bessel Function relations [33, Appendix F .3]

2
1
Z Z
i x cos
J0 (x) = e d and x J1 (x) = x J0 (x) d x,
2 0

respectively.

56
Then we obtain the magnetic vector potential

H0 R a ei r
A = i J1 (R a sin )
2r sin

The electric field in the antennas far field is then recovered by Equation (3.41). The details have

been worked out in Stutzman and Thiele [33, Chapter 8]. The radiation integral of Equation (3.46)

is quite arduous to compute, in general. Only in rare cases (such as a perfect paraboloid), is it even

possible to find closed form solutions; it is common to use numerical techniques for reflectors with

variable dimensions or distortions [30]. We will ultimately obtain a far field pattern of the form

|E| 2J1 (R a sin )


F ( , ) = = (3.48)
max | E | R a sin

which is equivalent to the result we would obtain using Definition 3.3.2. For the ideal parabolic

reflector, F ( , ) is axisymmetric since there is no dependence in Equation (3.48). The far field

pattern is plotted in a logarithm plot with units of decibels as in Figure 3.7. In particular, we plot

10 log(F ) versus .

Remarks:

1. As long as the boresight direction is that for which U is maximum, the pattern F will always

have a maximum of 0 dB at = 0 .

2. The far field pattern, F , is plotted in this manner so that we can observe how many dB

down the antenna radiates for other directions ( 6= 0).

3. More negative dB levels for nonboresight directions indicate that a lower portion of the

total radiated power is sent in the nonboresight direction. Hence, the antenna performs

more optimally.

4. It is common to measure the dB level of the highest peak after boresight in the F pattern.

This peak is known as the side lobe and it is typically the direction where the most power is

radiated other than boresight. For Figure 3.7, the side lobe level is 17.6 dB.

57
Radiation Intensity 1000
U() Watts/Str

500

0
15 10 5 0 5 10 15

0
Far Field Pattern

10
F () dB

20

30

40
15 10 5 0 5 10 15
Angle to the Far Field, Degrees

Figure 3.7: Far field patterns for an ideal axisymmetric parabolic reflector antenna.

Faceted Surface

The deformed reflector in our model is a faceted surface. In this section we give an overview of

the method used to calculate the far field pattern for such a surface. The far field pattern that we

find can then be used to calculate the efficiency and gain of the antenna as shown in Section 3.3.2.

Let us first establish some notation for the discretized reflector. For each triangle Tn R , we

denote the centroid of Tn by pn . The deformed reflector has the configuration which contains the

triangles Tn = x(Tn ) . We denote the centroid of the deformed triangle by x(pn ) = (x n , yn , zn ).

See Figure 3.8 for an illustration.

We model the antenna feed with a simple dipole having current I0 = 200 Amps oriented along

the yaxis with dipole length y = 2 cm. In this case, the incident magnetic field to Tn is

I0 y
HINC = i ei r rn j , (3.49)

n
4rn

58
where rn is the vector from the focal point to x(pn ). The radiation integral of Equation (3.30)

is calculated over the deformed faceted surface for a selected set of directions to the far field,

( f , f ).

For our results, we consider only antenna diameters less than 20 meters. Also, the radiation

wavelength, is always greater than 0.003 meters. With these conditions, the least possible

distance to satisfy the far field approximation, by Definition 3.3.2, is 2.6105 meters. Choosing

all the sample far field points to be at a distance of 107 meters is sufficient for all of our cases. We

then calculate
Nt /2
exp(i107 ) X
A( f , f ) = Jn exp(ir f rn ) |Tn |
4 107 n=1

where r f is the unit direction vector corresponding to the direction ( f , f ). Also, Jn is the current

density on Tn as determined by facets normal vector n


n and the boundary condition

Jn = 2 n .
nn HINC

We then use Equation (3.41) to calculate E in the far field.

3.4 Geometric Error Analysis

In addition to the efficiency computations presented in Section 3.3, we also seek some surface ac-

curacy measurements that relate purely to the geometry of the deformed reflector and not on the

reflected radiation. While the efficiency computations based on the antennas reflective properties

are certainly important for understanding the operation of the antenna, the following measure-

ments will help identify regions where shape deformation is more pronounced.

3.4.1 RMS Calculation Methods

For both the molded and flat construction models, we can vary parameters that control the loading

forces to determine what effect these parameters have on the surface accuracy of the reflector.

Our results will assist in developing a cutting pattern, support structure or other constructions

that may reduce the surface distortions. The geometric configuration of the deformed reflector

can be evaluated by considering the global RMS surface errors as determined by both vertical

59
Aperture

an
n
n fn

Focal Point
F
zn 4F z = x 2 + y 2

Figure 3.8: Deformed triangle Tn with centroid (x n , yn , zn ). The vertical displacement, z, is


labeled and the pathlength for this triangle is n = f n + an .

displacement of the reflector and path length error.

A reflector in an ideal position should have each triangles centroid nearly satisfying the gen-

erating equation 4F zn = x n2 + yn2 . The centroids will likely not satisfy the generating curve exactly

since it was the vertices of the triangles that were positioned using the parabolic form. For a fine

mesh, however, this error is negligible as will be demonstrated in Section 3.4.2. The Euclidean

measure of RMS, EUC , is obtained by the integral [16]

ZZ
2 2 dA
EUC () = z (3.50)
A
|A |

where z is the vertical displacement between the deformed reflector and an ideal paraboloid,

| A | is the area of the circular aperture, and dA is Lebesgue area measure. We calculate EUC by

considering all of the reflector facets and then computing Equation (3.50) discretely by

Nf
2
1 X 2
EUC () = z n |Tn |(
nn k) (3.51)
|A | n=1

where |Tn |(
nn k) is the projected area of Tn onto the aperture.

Another RMS calculation includes the pathlength error of a radiation wave traveling from the

focal point to the focal plane via triangle Tn . For an ideal paraboloid, this distance should be twice

the focal length of the parabolic reflector for each triangle (recall Section 3.2). Having denoted

60
the total path length of a ray striking the centroid of triangle Tn by n , we have

ZZ 2


2
dA
RMS () = (3.52)
A
2 |A |

where = n 2F is the pathlength error of a wave reflected at a given point on the reflector

surface and F is the focal distance of the reflector [16]. As we did for the Euclidean RMS, we also

compute the integral of Equation (3.52) discretely by

Nf
2
1 X 2
RMS () = n 2F |Tn |(
nn k). (3.53)
4| A | n=1

3.4.2 Numerical Accuracy of RMS

The order of error in the surface accuracy computations depends on the discretization of the tri-

angular mesh we choose. The two primary controlling factors for the mesh size are the number of

rings of triangles, Nr and the number of gores, N g .

In Equation (3.50), we see that the surface accuracy is determined by the vertical discrepancy

of the facets centroid from a paraboloid. An ideal configuration for our model is one whose vertices

lie exactly on the parabolic mold. This configuration, however, will not have facet centroids that lie

on the actual paraboloid and so the ideal configuration would not have EUC = 0. To understand the

error in these computations, we calculate EUC (R ) where R is the molded reference configuration.
y
Let Tn R be a typical facet with vertices denoted by v j = (v jx , v j , v zj ) for j = 1, 2, 3. Then,

the centroid can be defined based on these three vertices as

y y y
v1x + v2x + v3x v1z + v2z + v3z

v1 + v2 + v3
pn = , , . (3.54)
3 3 3

The vertical discrepancy, zn between pn and the ideal paraboloid 4F z = x 2 + y 2 is

(pnx )2 + (pn )2
y
zn = pzn
4F

y
where pnx , pn and pzn are the coordinates of pn . Using Equation (3.54), we can write the vertical

61
discrepancy as a function of the vertices of Tn . We have,

2 y y y 2
v1z + v2z + v3z 1 v1x + v2x + v3x

v1 + v2 + v3

zn = + .
3 4F 3 3

After expanding the trinomials and performing some algebra, we can obtain

1  
(v1x v3x )(v1x v2x ) + (v2x v3x )2 + (v1 v3 )(v1 v2 ) + (v2 v3 )2 .
y y y y y y
zn =
18F

i R2 to be the projection of the point vi to the plane z = 0. Also, define P(Tn ) to


Now define v

the projection of the triangle Tn to the plane z = 0. Then we can write

1  
zn = ( 3 ) (
v1 v 2 ) + |
v1 v 3 |2 .
v2 v
18F

The definition of dot product leaves

1  
zn = | 3 | |
v1 v 2 | cos 1 + |
v1 v 3 |2
v2 v
18F

where 1 is the interior angle of P(Tn ) at vertex 1. Let hn be the length of the longest edge of

P(Tn ). Then, we can bound zn by

1 + | cos n,1 | h2n h2n




z
n .
18F 9F

From Equation (3.50), we obtain

Nf Nf
2
1 X 2 1 X h4n
EUC (R ) = z n | Tn |(
nn k) | Tn |(
nn k).
|A | n=1
|A | n=1
81F 2

Let H = max hn . Then, we can reduce this further to


Tn R

Nf
2
H4 X
EUC (R ) | Tn |(
nn k).
81F 2 | A | n=1

Recall that | Tn |(
nn k) is the area of P(Tn ). Therefore, the sum over all reflector facets will cancel

62
with the aperture area | A |. This leaves

H2
EUC (R ) .
9F

We now compute a reasonable value for H with the knowledge that the triangles P(Tn ) that

are adjacent to the vertex are always the largest by construction of the grid. The longest side of

one of these triangles is

Rf Rf p 
 

H = max , 2 1 cos 2 / N g
Nr Nr

where Nr is the number of rings of triangles on the reflector and N g is the number of gores. It is

straightforward to show that


p  
2 1 cos 2 / N g <1

whenever N g > 5. The grid parameters for all our experiments use a value of N g 7. Therefore,

we can fix the quantity


Rf
H = .
Nr

Notice that the error bound no longer depends on N g as long we keep N g > 5. The bound on

EUC (R ) becomes
R2f D2
EUC (R ) = (3.55)
9F Nr2 36F Nr2

where D, R f and F are the diameter and radius and focal length of the reflector aperture, respec-

tively. We compare some of the models computations for error on the reference configuration R

to these bounds. The results in Chapters 4 and 5 typically use Nr = 36.

Reflector Size Eqn. (3.55) Model Calculations


Nr
D (m) F (m) EUC (mm) EUC (mm) RMS (mm)
36 2.130 0.914 0.1064 0.0805 0.0665
36 10.650 4.570 0.5320 0.4027 0.3325
44 10.650 4.570 0.3561

Table 3.3: Comparison of predicted versus actual errors for various grid sizes.

63
Chapter 4

Demonstration Cases

The results we present in this chapter demonstrate the behavior of the reflector surface under

various conditions. In Section 4.2, we model the antenna reflector using the molded reference

configuration. Some of the results of the parametric studies shown here will therefore demonstrate

predictions that this model gives for the surfaces that are already constructed and tested. To assess

the actual accuracy of the reflector and the degree to which it may be useful in operation, we

apply the analysis developed in Chapter 3. In Section 4.1, a mylar balloon is also considered as

a test case of the model. The construction methods for the faceted surface that we presented in

Chapter 2 can be easily adapted to model the construction of such a balloon.

4.1 A Test Case: The Mylar Balloon

The design of the triangular mesh and the reference configuration pattern are appropriate for

modeling the shape of a mylar balloon. We make some predictions for the geometry of an inflated

mylar balloon and compare them to an idealized mylar balloon as discussed in [24]. The mylar

balloon is constructed by seaming together two flat disks of material and then inflating the region

in between. The equilibrium configuration for this balloon will exhibit wrinkling which is large

relative to the balloons dimensions. In Figure 4.1, we show the generating curve of an idealized

inflated mylar balloon whose reference configuration has radius a.

We label the semimajor radius of the inflated mylar balloon R M and the depth M . Assuming

that the mylar does not stretch, the arclength along the profile curve must be the same as the

64
z

North Pole
z(x)

M
RM x

South Pole

Figure 4.1: Cross section of the Mylar Balloon with generating curve z(x).

radius of the uninflated circular membrane, i.e.,

Z RM p
1 + z (x) d x = a (4.1)
0

where a is the radius of the uninflated membrane. We can also construct an equation for the

volume of the balloon using the fact that it is a surface of revolution. Implementing the shell

method gives Z RM
V = 4xz(x) d x. (4.2)
0

Using Equations (4.1) and (4.2), one can apply Calculus of Variations and obtain the correspond-

ing EulerLagrange Equations. From that, the profile shape is ultimately found to be the elliptic

integral
t2
Z RM
z(x) = d t,
R4M t 4
p
x

where 0 x R M [24]. Relationships between the semimajor radius, depth and reference

configuration radius can be obtained in terms of gamma functions. In [24], Mladenov and Oprea

derive the ratios for these geometric properties. In Table 4.1, we compare Opreas inelastic analytic

results for the ratios R M /a, M /a and M /(2R M ) to those obtained by our elastic model.

It is interesting to note that there is good agreement between our results that include wrin-

kling and the idealized problem of Oprea that neglects elastic effects. In our model, we have

a = 1.065 meters and we can obtain the generating curve as shown in Figure 4.2. Though this

65
R M /a M /a M /(2R M )
Oprea [24] 0.7627 0.9139 0.5991
Model for p0 = 125 Pa 0.7720 0.9273 0.6009
Model for p0 = 500 Pa 0.7756 0.9506 0.6128

Table 4.1: Geometric results of mylar balloon test for two p0 values.

curve is smooth and convex, in reality there are many wrinkled regions in the mylar balloon. For

certain cross sectional cuts of the mylar balloon, the profile shape will locally collapse.

0.4
0.3
zaxis

0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
xaxis
Figure 4.2: Generating curve of modeled mylar balloon.

The stress resultants for an axisymmetric balloon surface are derived in Section 5.3 of [20]

and are given by


 
p0 r2 r2 p0 r2
= 2 and =
2 r1 2

where is the radial direction and is circumferential direction. A mylar balloon is assumed to

have zero circumferential stress, . This is not, however, physically plausible in a doubly curved

surface with p0 > 0. Also, the curvatures are related by

8
k1 = 2k2 and K= H2
9

where k1 and k2 are the principal curvatures and K and H are the Gaussian and mean curvatures,

respectively [24]. For this reason, the mylar balloon is an example of a Weingarten Surface. Actual

mylar balloons exhibit crimpling of the inflated surface which is responsible for the large folds

that can be seen around the rim of the balloon in Figure 4.3.

66
(a) A real mylar balloon [38]. (b) Top view of model. (c) Side view of model.

Figure 4.3: Fully inflated mylar balloon (actual and modeled).

4.2 Analysis of Molded Reflectors

In this section, we investigate the response of the reflector shape and its surface accuracy to

changes in the boundary support tendons, gravity and symmetry of the loading conditions. These

were discussed in [10] using a coarser mesh.

We use the material properties of Kapton to serve as the reflector material and apply the di-

mensions of an actual prototype. Table 4.2 contains the parameter values of Kapton that are used

in our numerical experiments as well as the dimensions of the reflector geometry.

Property Variable Value


Kapton Thickness h 12.7 microns
Kapton Mass Density 1419.98 kg/m3
Kapton Youngs Modulus E 25.9 108 N/m2
Kapton Poissons Ratio 0.34
Reference Tendon Length m,R 0.06 meters
Reflector Diameter D 2.13 meters
Reflector Focal Length F 0.914 meters
Triangular Facets N 20,736
Reflector Facets Nf 10,368
Supporting Tendons Nt 72

Table 4.2: Mechanical properties of Kapton, reflector dimensions and grid size.

For the remainder of this dissertation, we fix the parameter g = 9.80665 meters/sec2 for the

acceleration due to gravity. In each case, the gravity vector g will be defined with a direction. The

only case for which g will not have the aforementioned magnitude is for a 0g environment.

67
The surface accuracy results will include the RMS measurement, the antenna power gain

and the reflector efficiency as they were defined in Chapter 3. We use the radiometric RMS,

RMS as defined (3.52). Expressions for the antenna gain and efficiency can be found in Defini-

tions 3.3.5 and 3.3.6. We will denote efficiency by e.

The tendon foreshortening parameter, is important in many of the experiments of Chap-

ters 4 and 5; it was briefly mentioned in Section 2.1.2. The reference length of the tendons are

foreshortened by the percentage where 0 < 1. This parameter enables us to vary the force

on the reflectors initial configuration.

4.2.1 Reflector Size Variation

The effects on a larger reflector are of concern since a large radiating aperture is desired for long

range missions. In particular, we are interested in the relationship between the reflectors size and

the surface accuracy of the antenna when other significant parameters are held constant.

We perform two case studies; one with a tendon foreshortening of = 0.03 and the other with

= 0.06. For each case study, we increase the size of the reflector while holding the ratio F /D

constant. The reflector sizes in Table 4.3 are obtained by taking the reflector dimensions of Ta-

ble 4.2 and increasing the dimensions by factors of 1.0, 2.5, 4.0, 5.5 and 7.0. The tendon lengths

are not changed by the size factor. Gravity is held at g = gk for these tests and p0 = 10 Pa.

The equilibrium configuration is determined by our model for these parameters and we report the

surface accuracy data in Table 4.3.

Reflector Dimensions Case I: = 0.03 Case II: = 0.06


D (m) F (m) RMS (mm) e (40 GHz) RMS (mm) e (40 GHz) Gain dB
2.130 0.914 1.4156 98.59 % 4.2595 91.35 % 58.61
5.325 2.285 4.7162 93.47 % 2.5938 97.94 % 66.88
8.520 3.656 15.5799 61.91 % 14.5017 66.85 % 69.30
11.715 5.027 31.0087 48.34 % 28.1322 53.22 % 71.07
14.910 6.398 50.8096 36.48 % 52.2116 40.84 % 72.02

Table 4.3: Comparison of energy components for various antenna sizes.

By comparing the results of Cases I and II in Table 4.3, we see that the size of the reflector

having greatest efficiency for Case I is different from that of Case II. This suggests that the force

68
in the tendons is dependent on the size of the antenna and should be considered when choosing

a certain diameter for an antenna reflector. Additional analysis is needed to assess the benefits or

drawbacks of certain parameter values which can be combined to help improve a large antenna

reflectors surface accuracy.

For Case II, we include the antenna power gain. We see that one can still attain higher levels of

power gain for larger diameter antennas despite the reduction of efficiency. From Equation (3.45),

we can find that the antenna in case (D, ) = (14.910, 0.06) performs like an antenna operating

at 100% efficiency with diameter 9.53 meters. While larger aperture antennas are desired for

increased power transmission, higher accuracy will be required to make use of the full size of the

reflector at hand.

Gravitational Variation for Large Reflectors

Earlier results reported in [10] and [11] suggested that changes in the gravitational environment

have lesser effects on the surface accuracy of the reflector than the supporting tendons. These

conclusions were based on results from our model for antennas of diameter up to 2.13 meters. In

this section we will demonstrate that larger reflectors are somewhat more sensitive to changes in

the gravitational environment. The pressure is set to p0 = 12.5 Pa for this experiment.

Case I: D = 2.13 meters Case II: D = 14.91 meters


g/g
RMS (mm) e (40 GHz) RMS (mm) e (40 GHz) Gain dB
k 1.4156 98.59 % 50.8096 36.48% 71.53
p
( j + k )/ 2 1.3568 98.51 % 44.9216 47.29% 72.66
j 1.6830 98.35 % 41.6859 46.23% 72.56
0 1.6802 98.41 % 40.8349 48.31% 72.75

Table 4.4: Gravitational effects on large reflectors.

Case I of Table 4.4 shows that the 2.13 meter axisymmetric reflector is not particularly sensitive

to a rotation in the gravity vector. The RMS values have a range of less than 0.3 mm. In [10],

we showed that a smaller size 0.3 meter offaxis reflector had even less sensitivity to changes in

the gravitational environment. In Case II of Table 4.4, however, we see that the larger reflector

is affected particularly when the gravity is in the direction k. The RMS ranges from 40.83 mm

(for a 0g environment) to 50.80 mm (for a 1g environment). Larger weight and minimal vertical

69
support are likely the key factors for this observation. The affect of gravity on the deformation

and overall displacement of the entire reflector should be considered when working with antenna

reflectors on the order of 10 meters.

4.2.2 Boundary Support Variation

Initial findings for the inflatable antennas that we are investigating suggest that deflection near the

rim is a cause of efficiency loss. In this section, we investigate the reflector surface accuracy under

the impact of both low and high boundary support tension. In [10], we found that if the tension

was sufficiently high, there can be significant deflection near the rim. Similarly, we found that the

reflector displaces vertically near the vertex of the reflector. These effects can lead to severe loss

of the radiation reflected by the antenna.

Symmetric Boundary Forces

Assume that each tendon has the same stiffness constant K t,m and has the same foreshortening

factor, . Then, the force applied to the antenna is uniform around the entire rim. We also as-

sume gravity is g = gk so that totally symmetric loading applies. In this case, the computed

equilibrium configurations of the reflector are axisymmetric although the solver does not assume

an axisymmetric geometry. In Figure 4.4, we plot the radial and vertical components of the facet

displacements as a function of radial distance from the symmetric axis. It is clear that the dis-

placements are smaller and hence less severe for the lower parameter value of = 0.02. Surface

accuracy values of these results are presented in Table 4.5.

Most of our results do not contain a W shaped error curve. The W shape has a maximum

vertical error displacement at about 3/4R f , where R f is the radius of the reflector. See [21] for an

example where a circular pressurized membrane with a fixed edge is studied. In Figure 4.4, both

vertical error plots are monotonic and reveal maximum displacement at the rim of the reflector.

This is significant because a substantial vertical displacement of the reflectors vertex indicates a

general vertical shift of the reflectors position with respect to the feed. A loss of power gain is

expected since the optical properties of the paraboloid and its focus are disturbed.

70
Profile
Profile ofof Facet
Facet Displacements
Displacements (mm)
in millimeters

==0.02
0.02
Displacement
Radial Disp1.5
==0.04
0.04
Radial

0.5

0
0 0.2 0.4 0.6 0.8 1
Disp

3
Displacement
Vertical

2
Vertical

0
0 0.2 0.4 0.6 0.8 1
Radial Distance
Radial Distancefrom
fromVertex (meters)
Vertex (meters)

Figure 4.4: Reflector Distortion Pattern for High and Low Tendon Tension.

Nonsymmetric Boundary Forces

In practice, the reflector may be subject to nonsymmetric loading forces. The asymmetry can be

modeled by setting the tendon stiffness constants K t,m in Wt,m of Equation (2.14) to

  km 
K t,m = K t,m 1 + cos (4.3)
Nt

where K t,m = 250 N is the average tendon stiffness and k is a positive integer. For k = 4, the

tendon force is maximum at the i positions on the rim and 0 in the j positions. Hence, the

entire reflector responds by straining more in the x direction and less in the y direction. As shown

in Figure 4.5, this results in a large variation in the separation angles between the upper and lower

shells of the clam shell. Also note that the variation in this angle corresponds to the peaks and

troughs of K t,m in Equation (4.3).

71
Shell Separation for Deformed Antenna Reflector
Angle Between Upper & Lower Shells (Degrees) 105

100

95 Deformed Configuration
of Case 3 in Table 4.5
90

85

80

75
Molded Reference
70 Configuration

65

60

55
0 50 100 150 200 250 300 350
Angular Position on Reflector Rim (Degrees)

Figure 4.5: Effect of tendon force distribution for Case 3 of Table 4.5.

Figure 4.6 shows the shell separation angles for a tendon stiffness distribution with a sudden

jump at the direction +i. The first and seventysecond tendons (lying next to each other and on

either side of +i) exert a maximum and minimum force, respectively. This jump generates a non

symmetric distribution of force along the rim of the reflector.

Case Figure No. Stiffness, K t,m (N) Foreshortening RMS (mm)


1 4.4 K t,m = 250 = 0.02 0.5513
2 4.4 K t,m = 250 = 0.04 2.3190
3 4.5 Eqn (4.3) with k = 4 = 0.03 4.2592
4 4.6 Eqn (4.3) with k = 3 = 0.03 6.0464

Table 4.5: Summary of RMS values for various physical testing conditions.

72
Shell Separation for Deformed Antenna Reflector
120
Angle Between Upper & Lower Shells (Degrees)
110
Deformed Configuration
of Case 4 in Table 4.5
100

90

80
Molded Reference
Configuration
70

60

0 50 100 150 200 250 300 350


Angular Position on Reflector Rim (Degrees)

Figure 4.6: Effect of tendon force distribution for Case 4 of Table 4.5.

4.2.3 Offaxis Reflector

The offaxis antenna we consider here is substantially smaller in size than the axisymmetric re-

flector considered in Section 4.2.1. The diameter for the offaxis antenna is 0.3048 meters with

a focal distance of 0.1524 meters. While this antenna holds the advantage of not having the feed

and its mounting device overshadow the reflector, it does not hold its parabolic shape as well as

its axisymmetric cousin. This is most likely due to the fact that, unlike the axisymmetric case, a

nonconstant edge force needs to be applied. The pressure is set to p0 = 12.5 Pa and the tendon

foreshortening is = 0.03 for this experiment.

In Figure 4.7, we present the lower and upper shell separation angle around the rim of the

antenna for two offaxis antennas. A summary of the dimensions for these two antenna reflectors

and the surface accuracy analysis is presented in Table 4.6. Recall that the offaxis antenna is not a

symmetric shape on its own (refer back to Section 1.1), therefore we see an ideal separation angle

73
Shell Separation for Deformed Antenna Reflector

Angle Between Upper & Lower Shells (Degrees) 50

45
Molded Reference
40 Configuration

35

30

25

20 Deformed Configuration
of Case 1 in Table 4.6
15

10

5
0 50 100 150 200 250 300 350
Angular Position on Reflector Rim (Degrees)

Figure 4.7: Shell separation for small offaxis antenna reflector.

which is periodic around the reflector rim.

In Table 4.6, we see that the RMS is much larger for Case 1 than for Case 2. Since the size

of the reflector is far larger for Case 2, but the translation parameter, a see Equation (2.1) is


left unchanged, the overall shape is close to axisymmetric. Hence, we find a much lower RMS as

we were able to attain in Case 1 of Table 4.5, for example. Case 3 of Table 4.6 uses a translation

parameter which was increased proportionally with the diameter and focal lengths. A much higher

RMS value is obtained in Case 3, yet the efficiency remains at 93.43%. The reflector was deformed

in a manner that caused severe deflection near the rim. This deflection is primarily responsible for

the RMS calculation of 28.5148 since the RMS calculation restricted to the last ring of facets is 120

mm. The remainder of the antenna was moved to a position where its pathlength error over the

majority of the surface was actually quite low. The median contribution to the RMS for the interior

rings was only 3.18 mm.

74
Reflector Dimensions Surface Accuracy
Case
D (meters) F (meters) a (meters) RMS (mm) e (40 GHz) Gain (dB)
1 0.3048 0.1524 0.1778 8.0649 84.53% 41.39
2 3.0480 1.5240 0.1778 0.4223 99.40% 62.09
3 3.0480 1.5240 1.7780 28.5148 93.43% 61.82

Table 4.6: Surface accuracy of offaxis reflectors.

There was no design involved to obtain the fairly efficient antenna of Case 3 in Table 4.6. Our

model would not advocate that a deformation for an antenna of this type will generally deform

to an efficient reflector. Generally, these tests show that the two parabolic sheets are easily forced

together as the separation angle is dramatically reduced around the entire antenna. The offaxis

antenna class will need substantial modifications to reduce the deflection of the rim. This will

likely entail a correct computation of ideal boundary conditions.

75
Chapter 5

Parametric Studies

We present here some results which model an (initially) axisymmetric reflector constructed from

flat panels of material. The results presented in this chapter focus on experiments involving varia-

tions in the tendon supports, the internal pressure and changes to the cutting patterns for the flat

panels. The properties of the triangular grid for the faceted surface will vary slightly depending

on which construction is used. While the faceted surface models an antenna with the same dimen-

sions in each case, the triangular grid will change slightly for the gore construction and require

a change in the number of supporting tendons. The grid specifications are provided in Table 5.1.

The mechanical properties used for the membrane surface are again that of Kapton and can be

found in Table 4.2.

5.1 Parametric Study for RMS versus Ng and p0

This parametric study involves the flat gore construction that was outlined in Section 2.2. We

investigate changes in the surface accuracy when the internal pressure is increased and when the

Flat Gore Construction All Other


Number of . . . Variable
14 Gores 16 Gores 18 Gores Constructions
Triangular Facets N 17,528 20,032 22,536 20,736
Reflector Facets Nf 8,764 10,016 11,268 10,368
Supporting Tendons Nt 70 80 75 72

Table 5.1: Grid parameters and dimensions of the reflector being modeled.

76
number of gores is changed. We foreshorten the tendons by 1% ( = 0.01) and set the gravita-

tional acceleration to g = gk where the unit vector k = 0, 0, 1 is normal to the plane of the

aperture. Recall that g = 9.80665 meters/sec2 . As shown in Table 5.2, we test the reflector for the

differential pressures of p0 = 10 Pa, p0 = 12.5 Pa, p0 = 25 Pa and p0 = 50 Pa and the number of

gores set to 14, 16 and 18. We use our model to compute the equilibrium configurations and the

corresponding surface accuracy values, RMS , which are presented in Table 5.2. If we consider a

fixed pressure of 25 Pa, the RMS value is decreased by 57.7% to 0.7590 mm when we increase the

number of gores from 14 to 18. This suggests that an increase in the number of gores can help to

reduce surface distortions. We can also reduce RMS by fixing the number of gores and increasing

the differential pressure to 25 Pa. For a grid with 18 gores, RMS is reduced by 47.8% to a value of

0.7590 mm by increasing p0 from 10 to 25 Pa. The data in Table 5.2 suggests that for a fixed N g ,

there is a p0 that minimizes RMS . For higher differential pressure, we would expect deformation

of the surface in the normal direction in addition to high film stresses. In the case of p0 = 50 Pa,

the deformations become too large and cause a decrease in surface accuracy.

p0 = 10.0 Pa p0 = 12.5 Pa p0 = 25.0 Pa p0 = 50.0 Pa


Ng RMS (R )
RMS (N g , p0 ) RMS (N g , p0 ) RMS (N g , p0 ) RMS (N g , p0 )
14 4.6484 2.7030 2.5675 1.7947 1.5651
16 3.6082 2.0223 1.9020 1.2119 1.5249
18 2.8808 1.4564 1.3342 0.7590 1.9245

Table 5.2: Comparison of RMS (N g , p0 ) values in millimeters.

We use the deformed reflectors facet normal vectors to predict the path of a sample light ray

from the feed to the aperture; see Figure 3.8. The phase plane in Figure 5.1 is an image of the

paths terminal points on the aperture where the shading indicates the phase of the radiation at

the aperture for that path. For the reflector dimensions we use here (see Table 4.2), the expected

phase of the radiation at the aperture is


= mod 360 = 284.89
c

where the frequency is = 40 GHz, the ideal traveling distance of a radiation wave from the focal

point to the aperture is = 2F and c is the speed of light.

77
AUTAperture
AUT Aperture Phase
Phase Diagram
Diagram = 10
p0 p Pa Pa
= 10
0
360
0.9
300
0.6
240
0.3
yy (meters)
(meters)

0 180
0.3
120
0.6
60
0.9

0
0.9 0.6 0.3 0 0.3 0.6 0.9
x (meters)
(meters)
Figure 5.1: Aperture radiation phase plane at 40 GHz with N g = 16.

Figure 5.1 demonstrates how the radiation pattern at the aperture is disturbed by the cutting

pattern of the reference configuration. Variation in the phase at the aperture is clear along the

lines where the seams of the panels are located. Figure 5.2 shows plots of the phase distributions

at the aperture based on the sampling of the 10,016 light rays (one for each reflector facet) for

two different values of p0 . Each data point plots the number of light rays having a phase within

2 of the corresponding phase value. The peak of the distribution is closer to the expected phase

value of 284.89 for the higher pressure value of 25 Pa. Higher pressure can help reduce the

phase deviation, but one must be careful to consider the loss of surface accuracy for much greater

(relatively) values p0 ; recall Table 5.2.

78
AUT Aperture
Aperture RadiationPhase
Radiation Phase Distributions
Distributions
3000

2500 pp00 =
= 10
10 Pa
Pa
Rays

pp00 =
= 25
25 Pa
Pa
LightRays

2000
of Light

1500
Number of
Number

1000

500

0
180 200 220 240 260 280 300 320
Phase (Degrees)
Phase (degrees)
Figure 5.2: Phase distributions for two values of p0 at 40 GHz.

5.2 Parametric Study for RMS versus and

Since one serious problem with this inflatable antenna design is the deflection of the reflector

near the rim, we conduct the following parametric study that involves modifying the reflector rim

shape. The band construction is wellsuited for this experiment since the rim of the reflector is

contained in one section of the cutting pattern. We modify the slope of this outermost band by

decreasing the radius of the outermost ring of vertices (situated at the reflectors rim) as shown

in Figure 5.3.

The band construction we consider for this parametric study consists of 10 bands with the
D
seams being set at intervals of 0.1R f where R f = 2
is the radius of the aperture. We introduce

a shape modication factor [0, 1] to adjust this outermost band. This factor determines the

79
distance, d, from the symmetric axis of the paraboloid to the ring of vertices at the reflectors rim;

see Figure 5.3. In particular, d = (1 )R f . Note that the nominal reference configuration is

= 0. The result is a reference configuration modeling a paraboloid for the inner 9 bands with a

tapered adjustment for the tenth.

Parabolic Cross Sections Rf

Symmetric Axis
Symmetric Axis

0.3
0.3 d

0.2
0.2

0.1
0.1

00
1
1 0.8
0.8 0.6
0.6 0.4
0.4 0.2
0.2 0
0 0.2
0.2 0.4
0.4 0.6
0.6 0.8
0.8 1
1
Radial Axis (meters)
Radial Axis (meters)

Figure 5.3: Parabolic cross section for a flat ring configuration with = 0.05.

The geometric adjustment to the rim of the reference configuration was introduced to explore

means of adjusting the construction patterns in order to help reduce the deflection due to the

tension in the supporting tendons. We perform a set of tests to determine the reflectors shape

accuracy as a function of the rim radius modification factor, . The differential pressure is fixed to

p0 = 12.5 Pa and the acceleration due to gravity is set at g = gk. Table 5.3 contains RMS values

as calculated by Equation (3.52) for various values of the parameters and .

For each value of , the strain energy in the tendons is determined by using Equation (2.14)

where K t,m = 250 N for all Nt tendons. In Figure 5.4, we plot RMS as a function of for 0

0.044 and for = 0.01, 0.02, 0.03 and 0.04. One can see in Figure 5.4 that the RMS can be reduced

by adjusting the parameter . For a fixed value of , we can find a parameter value = such

that

RMS ( , ) = min RMS (, ) (5.1)


A

where A is the set of values for which we computed reflector configurations. The results of

these local minima are summarized in Table 5.3. The third column contains the data for the flat

band reflector that achieves the minimal RMS value for each . The gain is calculated as in Defi-

nition 3.3.5. In the second column of Table 5.3, we report the RMS attained by both the molded

80
4

3.5

2.5
RMS (, )

1.5

1
= 0.01
= 0.02
0.5
= 0.03
= 0.04
0
0 0.01 0.02 0.03 0.04 0.05
Rim Modification Factor, .

Figure 5.4: Effect of rim modification factor on RMS for various tendon loading.

reflector construction and the unmodified flat band construction ( = 0), for comparison. The

reduction in RMS (fourth column) compares the values RMS (0, ) and RMS ( , ).

Molded =0 Reflector with = Reduction



RMS RMS RMS ( , ) Gain (dB) in RMS
0.01 0.8082 0.5084 0.002 0.4718 58.97 7.2%
0.02 0.2238 0.4369 0.000 0.4369 58.99
0.03 1.0456 1.3537 0.028 0.8993 57.12 33.6%
0.04 1.9960 2.3340 0.032 1.2478 56.81 46.5%

Table 5.3: Comparison of RMS (, ) values (mm) for various .

Consider, for example, the tendon foreshortening factor = 0.03. The best surface accuracy

attained is RMS = 0.8993 mm for the parameters (, ) = (0.028, 0.03). This RMS value is 33.6%

less than that of the case (, ) = (0, 0.03). The case for (, ) = (0.028, 0.03) also has an RMS

81
value that is 14.0% lower than RMS = 1.0456 mm for the molded reflector. These results advo-

cate for the use of design parameters such as to help reduce surface distortions such as those

generated by the boundary support forces.

5.3 Tendon Supported vs. Fixed Boundary Reflectors

From Figure 5.4, we see that for sufficiently large (in our case > 0.02), RMS ( , ) increases

as a function of . This suggests that the surface accuracy may inevitably suffer for very strong

tendon forces. This motivates an investigation of the antennas RMS when the reflector is not sub-

ject to tendon forces at the boundary. In this section, we model the antenna reflector using the flat

band construction with the boundary fixed to a rigid frame and = 0. This entails eliminating S t

from Equation (2.2) and fixing the vertices of the antennas rim. Equilibrium shapes are calculated

for the differential pressures p0 = 5, 10, 15, 20, and 25 Pa. The shapes are also tested in three

different gravitational fields as listed in Table 5.4.

Gravitational p0 = 5 Pa p0 = 10 Pa p0 = 15 Pa p0 = 20 Pa
Environment RMS (p0 , g) RMS (p0 , g) RMS (p0 , g) RMS (p0 , g)
g=0 0.1923 0.0853 0.2483 0.4454
g = gk 0.1861 0.0879 0.2546 0.4519
g = gj 0.1925 0.0856 0.2484 0.4454
p
g = g(j + k)/ 2 0.1969 0.0807 0.2447 0.4420

Table 5.4: Collection of RMS (p0 , g) values (mm) for a fixed boundary reflector.

Generally, the shape of the reflector is far more accurate for these experiments than for tests

involving the supporting tendons. For the pressure values of 10 and 15 Pa, the fixed boundary

reflector models have surface accuracy range 0.0853 RMS 0.2546 as presented in Table 5.4. In

Table 5.3, we see that the band construction achieves its best surface accuracy of RMS = 0.4369 for

the case (, ) = (0, 0.02). Figure 5.5 compares the vertical discrepancy of the deformed reflector

to an actual paraboloid for both a fixed rim antenna and tendonsupported one. Each group of

data points in Figure 5.5 is the minimum, average and maximum vertical facet displacement for

all the facets in that band. With exception of the boundary support and configuration type, all

conditions for the two tests are identical. The internal pressure is held at 12.5 Pa and gravity is set

82
to g = gk.

The top set of data in Figure 5.5 is for a tendon supported antenna modeled with the band

configuration and (, ) = (0.028, 0.03). The lower set of data is for a fixed boundary reflector

modeled with the band configuration. The tenth band of the tendonsupported reflector has a

large variance in displacement due to = 0.028 for that case.

Distribution of Facet Centroid Displacements


2
Vertical Displacement of Centroid (millimeters)

1.5

0.5 Tendon Supported


Fixed Boundary

0.5
0 0.2 0.4 0.6 0.8 1
Radial Distance from Vertex (meters)

Figure 5.5: Comparison of error for a fixed rim and a tendonsupported reflector.

The low RMS values for the fixed boundary reflector suggest good performance at radio fre-

quencies greater than 40 GHz. We see that the RMS values, despite their growth for larger pressure

values, remain low compared to the RMS values of the studies shown in Tables 5.2 and 5.3. Even

changes to the gravity field have little affect on the RMS values that the experiments in this sec-

tion attain. The radiation and efficiency data in Table 5.5 show that the performance of a fixed

boundary antenna may be able to perform well at frequencies above KaBand.

83
Radiation Frequency (GHz) 40 60 80
Radiation Wavelength (mm) 7.50 5.00 3.75
Intensity at Boresight (Watts/Str) 1282.71 2884.89 5125.71
Antenna Gain (dB) 59.0046 62.5246 65.0208
Side Lobe Level H plane (dB) 17.57 17.59 17.56
Side Lobe Level E plane (dB) 24.19 24.18 24.18
Radiometric RMS (mm) 0.1609 0.1609 0.1609
Euclidean RMS (mm) 0.1707 0.1707 0.1707
Efficiency 99.89% 99.85% 99.79%

Table 5.5: Characterization of fixed rim inflatable antennas; p0 = 12.5 Pa.

At a radio frequency level of 80 GHz, we can see how much more efficient the fixed rim an-

tenna is compared with the tendon supported one; see Table 5.6. Note that the tendon supported

reflector has RMS < 1 mm. For high radio frequency, however, the RMS has far lower tolerance.

If these large reflector antennas are used at these frequencies, one may wish to reexamine the

current tendon support boundary condition.

Tendon Supported Fixed Rim


Flat Band Construction: 0.028 0.000
Tendon Foreshortening: 0.03
Intensity at Boresight (Watts/Str) 3004.45 5125.71
Antenna Gain (dB) 62.7009 65.0208
Side Lobe Level H plane (dB) 17.29 17.56
Side Lobe Level E plane (dB) 21.82 24.18
Radiometric RMS (mm) 0.8993 0.1609
Euclidean RMS (mm) 4.1233 0.1707
Efficiency 58.49% 99.79%

Table 5.6: Characterization (for comparison) at 80 GHz.

84
Chapter 6

Conclusions and Future Research

Inflatable antenna reflectors have been successfully deployed and operated in radio frequencies

below Ka Band in past missions. There is interest in improving the surface accuracy of elastic

deployable antennas to meet the demands of future data transmission in both long and short

range space missions. Deployable antennas of the type discussed throughout this dissertation are

generally sensitive to either one or many of the supporting boundary conditions. The shape of the

deformed membrane varies when the parameters or boundary conditions are adjusted.

In Chapter 4, we considered a number of demonstration cases. For antenna applications, the

geometry is typically a paraboloid or an offaxis section of a parabaloid. To assess validity of our

model of a pressurized membrane formed by sealing together two opposite facing surfaces, we

considered the idealized problem of an inelastic mylar balloon in Section 4.1. A mylar balloon is

not molded, but constructed by sealing together the edges of two disks of equal diameter. The

equilibrium equations for the ideal mylar balloon are set in the deformed configuration, and yield

an axisymmetric smooth surface free of wrinkling. The generating curve can be computed explicitly

using elliptic functions. We find that the height to diameter ratios for the ideal mylar balloon

and our numerical solutions are in good agreement. Furthermore, our model is able to capture

wrinkling, an effect that must occur since a flat surface (like a disk) cannot be mapped smoothly

onto a doubly curved surface.

For an inflatable antenna, we fix the F /D ratio and vary the diameter of the reflector. Our

results find that the support tendon force is a key parameter that significantly influences the

85
efficiency of the reflector. If this support force is fixed, then the antennas RMS increases as

the reflector diameter increases. One particular example is an antenna that has the parameters

(D, ) = (14.91 meters, 0.06) and has the same gain as an ideal D = 9.53 meter antenna. A large

antenna may be desirable, but greater surface accuracy is needed to obtain the corresponding ra-

diating aperture size. In the case of a smaller antenna, a more careful analysis was considered. It

was concluded that tendon forces that are too high deform the reflector in a manner that vertically

displaces the vertex of the reflector. This sort of deformation shifts the position of the reflector

with respect to the feed. In future work, we plan to carry out a dimensional analysis so that we

can better isolate the key parameters for additional trade studies.

Gravitational effects are noticeable for large (10 meter range) antennas but not for smaller

ones. It is particularly clear that large antennas experience less deformation in zero gravity en-

vironments or when they are positioned with the aperture plane containing the gravity vector.

Gravitational forces that are normal to the aperture plane can displace a large reflector away from

the feed. Hence, the surface accuracy of a large reflector is better when the rim is oriented verti-

cally.

We also probed the effects of varying the boundary support system. In our results, asymmetric

loading of the symmetric antenna system can dramatically influence the radiometric RMS. While

we only considered a sinusoidal variation of K t,m with period (2)/(k/Nt ) = 2Nt /k and k = 3, 4,

tendon failure (i.e., K t,m = 0) in a number of contiguous locations could be catastrophic, rendering

the reflector inoperable. This strengthens the argument for attaching the reflector edge directly to

a rigid structure and eliminating the tendons altogether. However, if an offaxis antenna system is

desirable, then the edge forces will need to be applied in an appropriate nonsymmetric manner

and attaching the reflector edge to a rigid structure could be problematic. A backup system of

tendons could eliminate some of the tendon support concerns.

In Chapter 5, we discovered that reflectors constructed from flat panels of elastic material can

be pressurized to nearly parabolic shape. This removes the necessity of the molding process (which

is expensive) and also opens the possibility of altering the construction pattern. Either geometric

or physical parameters can be analyzed to determine what affect they may have on the reflectors

surface distortions. It was shown that the parameters can be implemented to improve antenna

performance.

86
In Section 5.1, we considered parametric studies where we varied the differential pressure

p0 and the number of gores N g . As one might expect, increasing the number of gores leads to

an increase in surface accuracy. Similarly, we found that the differential pressure can be used to

reduce surface distortions. If the pressure is too low, then it is impossible for the antenna to hold

the parabolic shape. However, if the pressure is too high, surface distortion increases. For a given

set of design criteria, there should be an optimal value of p0 .

We also explored how the radiation pattern at the aperture is perturbed by the cutting pattern

of the reference configuration. In our investigations, we found that most distortion was concen-

trated along the gore seams, but these effects could be mitigated by increasing the pressure. The

gore structure induces the wrinkling to take place along the seams. However, these distortions

could also be reduced by altering the cutting pattern. For example, one might curve the bound-

aries of i in Figure 2.6. This approach has been followed in the fabrication of high altitude large

scientific balloons and spherical pressure vessels.

In Section 5.2, we investigated the effects of varying the tendon foreshortening and the rim

modification factor . The results in Section 5.2 advocate for the use of design parameters such

as to help reduce surface distortions such as those generated by boundary support forces. Our

parametric studies showed that surface accuracy degrades with very strong applied tendon forces.

This suggested an investigation of a parabolic reflector that eliminated the tendons altogether. For

this reason, we attached the edges of the paraboloid to a rigid structure. In general, we found

that the shape of the reflector for this boundary support is far more accurate than the shapes

supported by tendons. Moreover, the low RMS values for the fixed boundary reflector suggest

good performance at radio frequencies greater than 40 GHz. The results in Section 5.3 show the

performance of a fixed boundary antenna should perform well at Ka Band. Moreover, for high

radio frequency applications (beyond Ka Band), RMS had a far lower value than the 1 mm value

found for the tendon supported system.

We altered the for the outer band, but there is no reason why similar parameters 1 , . . . , k

cannot be introduced for other bands. Certainly, this is worth exploring if it will further reduce

surface distortions. In Appendix A, we present metrology data that was collected in June 2008 at

NASAs Glenn Research Center.

Our general approach to compute the reflectors configuration can be applied to a wide vari-

87
ety of construction types and large range of parameter values. Combined with the more accurate

surface error computations (compared to the Ruze Equation), these methods should be able to pre-

dict the performance capabilities of larger antenna reflectors subject to a host of different factors.

This model has shown that properly selected design parameters and construction types can help

increase the performance of inflatable antennas which will open new applications where greater

surface accuracy and reliability are required.

Directions for Future Research

There are a number of projects that naturally follow from this research. There is interest in devel-

oping large aperture antennas that are deployable and operable for sustained periods in potentially

extreme temperatures or harsh environments. Many avenues of research can be pursued to explore

the shape retention of inflatable antennas even in controlled environments.

If it is desirable to use an offaxis system, then one needs to derive the correct tendon loads

that will yield the appropriate parabolic geometry. If the reference geometry of the reflector is

molded, calculation of the boundary conditions should be straightforward. However, if we are

considering a large reflector of flat panels, then the cutting pattern may also need to be adjusted

in order to achieve the appropriate inflated geometry.

For our preliminary investigations, it was sufficient to use an isotropic membrane model. How-

ever, it may be desirable to exact more precise control over certain regions of the inflated mem-

brane. This could be done using orthotropic materials and so it would be desirable to develop an

orthotropic material model in order to carry out the necessary trade studies.

In the future, it would be useful to carry out a dimensional analysis of the equilibrium equations

in order to isolate key parameters to better understand how the inflated membrane structure scales

as the aperture diameter increases. Once this is done, one could carry out trade studies involving

the deployed to packaged volume ratio, a critical parameter in space applications.

Because the physical models that were made available to us had an F /D ratio of 0.4291, we

restricted our attention to this class of parabolic reflectors. However, it would be revealing to study

the radiometric RMS, antenna power gain, and reflector efficiency as a function of K t,m , , , p0

for different F /D ratios. It is not too difficult to imagine that a pressurized large aperture deep

88
parabolic antenna in a 1g gravity field would be less susceptible to gravity effects than a shallow

parabolic antenna of the same size. How might the design engineer trade the reduced mass of

the shallow antenna for the more robust, albeit heavier, deep antenna? A shallow antenna would

require higher support tendon tension than a comparably sized deep antenna, since more of the

membrane tension will be transferred across the interface between the opposite facing paraboloids

in a deep antenna. How significant is this, if the antenna is to operate in zero-gravity? Additional

trade studies would be able to provide partial answers to these questions.

The inflatable antenna system has a number of features that make it very appealing for ap-

plications where surface accuracy, efficiency, low mass, ease of deployment, and high packaging

efficiency are important. It is not a simple task to find a solution to such a multiparameter opti-

mization problem, let alone the optimal one. However, our findings as supported by our analytical

modeling and numerical simulations suggest that these questions are tractable and can be probed

with modest computing power and an appropriate mathematical model that captures the physics

of the problem with sufficient fidelity.

89
References

[1] Antman, Staurt S. Nonlinear Problems of Elasticity. Second Edition. New York: Springer Ap-

plied Mathematical Sciences 107, 2005.

[2] Baginski, Frank E., Michael C. Barg and William Collier. Existence Theorems For Tendon

Reinforced Thin Wrinkled Membranes Subjected to a Hydrostatic Pressure Load. Mathemat-

ics and Mechanics of Solids 13 (2008): 532 570.

[3] Baginski, Frank E. and Willi W. Schur. Structural Analysis of Pneumatic Envelopes: Vari-

ational Formulation and OptimizationBased Solution Process. AIAA Journal, Volume 41

Number 2 (2003): 304 311.

[4] Baginski, Frank E. and William Collier. Modeling the Shapes of Constrained Partially Inflated

HighAltitude Balloons. AIAA Journal, Volume 39 Number 9 (2001): 1662 1672.

[5] Beck, A. and H. Grabmller. Wrinklefree Solutions in the Theory of Curved Circular Mem-

branes Journal of Engineering Mathematics, Volume 27 Number 4 (1993): 389 409.

[6] Chou, Pei Chi and Nicholas J. Pagano. Elasticity: Tensor, Dyadic and Engineering Approaches.

New York: Dover Publications Incorporated, 1967.

[7] Chow, Shui Nee and Jack K. Hale. Methods of Bifurcation Theory. New York: SpringerVerlag,

1982.

[8] Ciarlet, Philippe G. Mathematical Elasticity, Volume I. New York: North Holland, 1988.

[9] Ciarlet, Phillipe G. Mathematical Elasticity, Volume III: Theory of Shells. New York: North

Holland, 2000.

90
[10] Coleman, Michael J. and Frank E. Baginski. Modeling the Shape of Deployable Large Aper-

ture Antennas in Nonsymmetric Loading Environments. AIAA Student Region MAI Con-

ference. April 2009.

[11] Coleman, Michael J. and Frank E. Baginski. A Shape Deformation Study of Large Aperture

Inflatable Elastic Parabolic Antenna Reflectors. Eleventh AIAA Gosammer Systems Forum.

Orlando, Florida. AIAA 2010 2501 (2010).

[12] Courant, R. and D. Hilbert. Methods of Mathematical Physics, Volume II: Partial Differential

Equations. New York: Interscience Publishers, 1962.

[13] Fichter, W.B. Some Solutions for the Large Deflections of Uniformly Loaded Circular Mem-

branes. NASA Technical Paper 3658 (1997).

[14] Freeland, R.E., G.D. Bilyeu, G.R. Veal and M.M. Mikulas. Inflatable Deployable Space

Structures Technology Summary. International Astronautical Federation, Paper IAF98I.5.01

(1998).

[15] Freeland, R.E. and G. Bilyeu. INSTEP Inflatable Antenna Experiment. International Astro-

nautical Federation, Paper IAF920301 (1992).

[16] Greschik, G. et al. Sensitivity Study of Precision Pressurized Membrane Reflector Deforma-

tions. AIAA Journal, Volume 39 Number 2 (2001): 308 314.

[17] Halliday, David and Robert Resnick. Physics Part II. New York: John Wiley and Sons, 1960.

[18] Hencky, Ing. H. ber den Spannungszustand in kreisrunden Platten mit verschwindender

Beigungssteifigket. Zeitshcrift fr Mathematik und Physik, Volume 63 (1915): 311 317.

[19] IEEE Standard Letter Designations for RadarFrequency Bands. IEEE Aerospace and Electronics

Systems Society. IEEE Standard 5212002. New York: IEEE, 2003.

[20] Irvine, Max. Cable Structures. New York: Dover Publications, Incorporated, 1981.

[21] Jenkins, C.H. and D.K. Marker. Surface Precision of Optical Membranes with Curvature.

Optics Express. Volume 1 Number 11 (1997): 324 330.

91
[22] Leifer, Jack, David C. Jones and Adam M. Cook. GravityInduced Wrinkling in SubScale,

SinglyCurved Parabolic Gossamer Membranes AIAA 20071819 (2007).

[23] Milford, Frederick J. and John R. Reitz. Foundations of Electromagnetic Theory. Second Edi-

tion. Reading, Massachusetts: AddisonWesley Publishing Company, 1967.

[24] Mladenov, Ivalo M. and John Oprea. The Mylar Balloon Revisited. The American Mathe-

matical Monthly, Volume 110 Number 9 (2003): 761 783.

[25] Pinsky, Mark A. Partial Differential Equations and BoundaryValue Problems with Applications.

Third Edition. International Series in Pure and Applied Mathematics. New York: McGraw

Hill, 1998.

[26] Pipkin, Allen C. Relaxed Energy Densities for Large Deformations of Membranes. IMA Jour-

nal of Applied Mathematics, Volume 52 (1994): 297 308.

[27] Romanofsky, Robert R., Kevin Lambert, Brian W. Welch and Irene Bibyk. The Potential for

Gossamer Deployable Antenna Systems in KaBand Exploration and Science Commuica-

tions Architectures. Twelfth Ka and Broadband Communications Conference. Napoli, Italy.

September 2006.

[28] Romanofsky, Robert R. Antenna, Microwave and Optical Systems Branch. NASA Glenn Re-

search Center. Cleveland, Ohio.

[29] Rudin, Walter. Principles of Mathematical Analysis. International Series in Pure and Applied

Mathematics. New York: McGrawHill, Incorporated, 1976.

[30] Scott, Craig. Modern Methods of Reflector Antenna Analysis and Design. Boston: Artech House,

1990.

[31] Smail, Lloyd L. Analytic Geometry and Calculus. Appleton Century Mathematics Series. New

York: AppletonCenturyCrofts Incorporated, 1953.

[32] Sneddon, Ian N. Elements of Partial Differential Equtions. International Series in Pure and

Applied Mathematics. New York: McGrawHill Book Company, 1957.

92
[33] Stutzman, Warren L. and Gary A. Thiele. Antenna Theory and Design. New York: John Wiley

and Sons, 1981.

[34] Thomson, Mark W. Astromesh Deployable Reflectors for Ku and Ka Band Commercial Satel-

lites AIAA 20022032 (2002).

[35] Weinberger, Hans F. A First Course in Partial Differential Equations. New York: John Wiley

and Sons, 1965.

[36] http://www.lgarde.com/images/ldp.jpg

[37] http://www.abc.net.au/science/news/img/inflate.jpg

[38] http://www.partyballoonsbears.com/browseproducts/MetallicGoldRoundMylar

Balloons.html

93
Appendix A

Surface Metrology Data

Some surface metrology data on inflatable antennas was collected at NASAs Glenn Research Center

in June 2008. The data was obtained using a high precision Leica LR200 Laser Scanner. This

scanner can measure the position of a point on a surface with a precision of 20 microns at a

distance of up to 48 meters away. For our experiments, the surface of the antennas were measured

at a distance of not more than 10 meters away. The scanner defines a particular point in space as

the origin and data is measured based on that reference point in a Cartesian coordinate system.

The units of the metrology data are inches, but we convert to meters in our presentation here.

Since the coordinate frame of the laser scanner may not put the antenna near the origin, we

will need to rotate and translate the metrology data so that it corresponds to a paraboloid with

a vertex at the origin and directrix plane perpendicular to the zaxis, as in Equation (2.1). The

original data points (x n , yn , zn ) can be transformed to (x n , y n , z n ) = T (x n , yn , zn ), where T is a

matrix of translations and rotations. The translation parameters are x 0 , y0 and z0 ; the rotation

parameters are 0 (a rotation about the zaxis) and 0 (a rotation about the xaxis).

For the given metrology data set, we aim to determine a least squares fit paraboloid. In other

words, we seek the set of parameters (x 0 , y0 , z0 , F0 , 0 , 0 ) that minimizes

N  2 2 !1/2
X x n + y 2n
(x 0 , y0 , z0 , F0 , 0 , 0 ) = zn
n=1
4F0

where N is the number of data points and F0 is the focal length. A MATLAB code was written to

solve this minimization problem and it utilizes the MATLAB function lsqnonlin.

94
For several data sets, we present the results of the least squares fit paraboloid in Table A.1 and

show some corresponding scatter plots in Figures A.1, A.2 and A.3.

Manufacturers Properties Results of lsqnonlin


Case N
F (meters) R (meters) F0 (m) R0 (m) (mm)
1 32,830 0.914 1.065 0.9143 1.0841 1.0051
2 83,466 0.914 1.065 0.9161 1.0805 1.3203
3 9,353 0.152 0.152 170.17 0.1670 8.6509
4 7,077 0.152 0.152 0.1445 0.1501 2.4556

Table A.1: Least squares paraboloid parameters for surface metrology data.

The following are the key physical conditions that were present during each of the data sampling

cases in Table A.1.

Case 1. The antenna reflector was situated with the gravity polar angle = /2. The differen-

tial pressure was held at p0 = 39.75 Pa (0.159 inAq). The scanner was set to measure

the position of surface points with a rectangular discretization of 1.02 cm 1.02 cm

over the surface.

Case 2. The antenna was leaning so that = 76 . We set p0 = 15 Pa (0.06 inAq) and the

discretization of the surface points was 0.63 cm 0.63 cm.

Case 3. Small offaxis antenna which is positioned upside down ( = ). Differential pressure

was held within the range 13.25 to 14.25 Pa (0.053 to 0.057 inAq). The discretization

of the surface points for this case was 0.25 cm 0.25 cm. Severe deflection at the rim

results in a least squares fit paraboloid which is far from the actual intended geometric

dimensions; see Table A.1 and note Figure A.3.

Case 4. Small offaxis antenna with = /2 and p0 = 14.75 Pa (0.059 inAq). The dis-

cretization is 0.25 cm 0.25 cm. Several tendons are removed such that are only 7

of the total 32 are supporting. These 7 tendons are distributed so that the reflector

is supported in three principal directions. While these tendons and the pressure offer

physical support for the reflector, substantial wrinkling is present. The least squares

fit paraboloid for this case is more representative of the intended geometry than in

Case 3. This is due to less deflection at the rim and a resulting shape which is globally

nearer to the desired geometry, despite wrinkling.

95
In Figures A.1 and A.2, gravity is oriented in the x direction. In Figure A.3, gravity is oriented

in the z direction. The units are meters and the axes are equally scaled within each figure.

1.0

0.5

0
x axis

0.5

1.0
1.0
0.5
0
0.5 y axis
3.2 1.0
3.8
z axis

Figure A.1: Surface metrology plot of axisymmetric reflector for = 90 .

96
1.0

0.5

0
x axis

0.5

1.0
1.0
0.5
0
3.2 0.5
3.8 1.0 y axis
z axis

Figure A.2: Surface metrology plot of axisymmetric reflector for = 76 .


z axis

0.96
0.98
1.00 0
2.1 0.1
2.2 2.3 0.2 x axis
y axis 2.4

Figure A.3: Surface metrology plot of offaxis reflector; exhibits substantial rim deflection.

97

Potrebbero piacerti anche