Design, Analysis, Assembly, Integration and Testing of Mechanical … · 2015-04-24 · Design,...

126
Design, Analysis, Assembly, Integration and Testing of Mechanical Systems for Micro-Satellites and Micro-Satellite Separation Systems by Jamie Fine A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto © Copyright by Jamie Fine 2014

Transcript of Design, Analysis, Assembly, Integration and Testing of Mechanical … · 2015-04-24 · Design,...

Design, Analysis, Assembly, Integration and Testing of Mechanical Systems for Micro-Satellites and

Micro-Satellite Separation Systems

by

Jamie Fine

A thesis submitted in conformity with the requirements for the degree of Master of Applied Science

Graduate Department of Aerospace Science and Engineering University of Toronto

© Copyright by Jamie Fine 2014

Design, Analysis, Assembly, Integration and Testing

of Mechanical Systems for Micro-Satellites and

Micro-Satellite Separation Systems

Jamie Fine

Master of Applied Science

Graduate Department of Aerospace Science and Engineering

University of Toronto

2014

Abstract

This document summarizes the development activities completed for the Exoadaptable Pyroless

Deployer (XPOD) system, and the MiniMags, EV9 and NORSAT-1 missions. The focus is on

the mechanical design, computer modelling, and assembly integration and testing of mechanical

systems. The XPOD work was associated with a re-analysis and testing of the XPOD Triple

engineering model such that a flight model could be produced. The MiniMags work involved

creating a preliminary bus design, which was ultimately used to determine that the MiniMags

payload could feasibly be flown in a microsatellite. The EV9 work included taking the EV9 bus

design from a mature design stage to flight assembly. Finally, work for the NORSAT-1 mission,

which is a microsatellite mission with several different payloads, took a proposal level bus

mechanical design to a preliminary design such that future work could be continued in later

stages of the mission.

Acknowledgments

The last two years would not have been possible without the help of so many people. Whether

the help was through a casual guiding conversation, moral support and encouragement, daily

management, or finances, it was all needed to work through the challenges that were

encountered.

I would like to thank Dr. Robert Zee for the opportunity that he provided me with by allowing

me to complete my degree at the Space Flight Laboratory. Nathan, for keeping me busy and

providing me with direction in my work, along with all the epic Frisbee games at lunch. Freddy,

for all of his guidance with the XPOD system. Mike for not only his help with the XPODs, but

also for his teachings on how to get ideas from a computer onto a lab bench. Laura, for helping

me learn how to integrate a flight spacecraft. Brad (“The Smurds”... haha), Tom, Josh and John

for all of their moral support, helpful conversations, and great/hilarious times in and out of the

lab.

To my parents I would like to thank you for supporting me through my second degree. To

Kseniya, thank you for always helping me with getting through the tough days no matter how

busy or tired we both were.

Finally, I would like to thank everyone else who made the hard days a little less difficult, the

tiring days a little less tiring, and the good days a little better. I learned so much in the past two

years about not only engineering, but about who I am, and about what is important to me in my

life.

Table of Contents

Acknowledgments .......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................ vii

List of Acronyms ......................................................................................................................... viii

List of Figures ................................................................................................................................ ix

Introduction .................................................................................................................... 1

1.1 The Exoadaptable Pyroless Deployer ................................................................................. 1

1.1.1 Separation Systems Overview ................................................................................ 1

1.1.2 Exoadaptable Pyroless Deployer Operation Method .............................................. 3

1.2 The MiniMags Feasibility Study ........................................................................................ 7

1.3 The EV9 Mission ................................................................................................................ 9

1.4 The NORSAT-1 Mission .................................................................................................. 10

1.5 Thesis Objectives .............................................................................................................. 12

Mechanical Systems Design ......................................................................................... 14

2.1 Derivation of Mechanical Requirements .......................................................................... 14

2.2 XPOD Mechanical Design ................................................................................................ 17

2.2.1 Mechanical Design of the XPOD Mechanism, Internal Preload and Main

Spring .................................................................................................................... 17

2.2.2 XPOD Mechanism Overview ............................................................................... 17

2.2.3 Pusher Plate Preload Design ................................................................................. 19

2.2.4 Designing the Main Spring ................................................................................... 26

2.2.5 Determining the Required Tension in the Mechanism Cord ................................ 31

2.2.6 Determining the Mechanism Jamming Conditions ............................................... 36

2.3 XPOD Tip Off Rate Analysis ........................................................................................... 38

2.3.1 Problem Definition ................................................................................................ 38

2.3.2 Assumed Ejection Geometry for Limiting Angles ............................................... 39

2.3.3 Relevant Equations ............................................................................................... 50

2.3.4 Implementation of the Solution Method ............................................................... 58

2.3.5 Tip-Off Rate Analysis Results .............................................................................. 63

2.4 EV9 Mechanical Design ................................................................................................... 65

2.4.1 EV9-A vs. EV9 vs AISSat-2 Comparison ............................................................ 65

2.4.2 EV9 Mechanical Design Requirements ................................................................ 66

2.4.3 Design Process ...................................................................................................... 67

2.5 NORSAT-1 Mechanical Design ....................................................................................... 79

2.5.1 NORSAT-1 Design Requirements and Starting Point .......................................... 79

2.5.2 NORSAT-1 Design Iterations ............................................................................... 81

Finite Element Modelling ............................................................................................. 89

3.1 MiniMags Finite Element Model ...................................................................................... 89

3.2 Finite Element Model Setup ............................................................................................. 90

3.2.1 Boundary Conditions ............................................................................................ 90

3.2.2 Modelling Methodology ....................................................................................... 91

3.2.3 Material Selection ................................................................................................. 92

3.3 Results ............................................................................................................................... 92

3.3.1 Natural Frequencies .............................................................................................. 93

3.3.2 Stress Results ........................................................................................................ 94

3.4 Conclusions ....................................................................................................................... 97

Assembly Integration and Testing ................................................................................ 98

4.1 XPOD Triple Vibration Testing ........................................................................................ 98

4.1.1 Axis Definition and Mounting Location ............................................................... 98

4.1.2 Accelerometer Placement ..................................................................................... 99

4.1.3 Vibration Levels .................................................................................................. 101

4.1.4 Inspection Procedure ........................................................................................... 102

4.1.5 Vibration Test Procedures ................................................................................... 108

4.2 EV9 Horizontal Deployment Test .................................................................................. 109

4.2.1 Spring Constant Determination Procedure ......................................................... 110

4.2.2 Deployment Test Procedure ................................................................................ 111

Conclusion .................................................................................................................. 114

References ................................................................................................................................... 116

List of Tables

Table 1: Overview of Qualified XPOD Designs ............................................................................ 1 Table 2: Launch Vehicle Mechanical Environment Summary ..................................................... 14

Table 3: List of Input Variables Used in Calculations and MATLAB Code ................................ 49 Table 4: Summary of Tip-Off Rate Analysis Verification ........................................................... 64 Table 5: Bus Subsystems Summary .............................................................................................. 65 Table 6: List of NORSAT-1 Relevant Mechanical Requirements ............................................... 81 Table 7: Summary of Material Properties Used in the MiniMags FEM ....................................... 92

Table 8 - List of Accelerometers ................................................................................................ 100

Table 9: List of Images ............................................................................................................... 104 Table 10: List of Measurements ................................................................................................. 107

List of Acronyms

SFL Space Flight Laboratory

XPOD eXoadaptable PyrOless Deployer

GNB Generic Nanosatellite Bus

RAL Rutherford Appleton Laboratory

CSA Canadian Space Agency

PCW Polar Communications and Weather

HEO Highly Elliptical Orbit

MiniMags Mini-Magnetosphere Shield

LV Launch Vehicle

FEM Finite Element Model

List of Figures

Figure 1.1-1: Images of Currently Qualified XPODs ..................................................................... 2 Figure 1.1-2: XPOD with Labelled Internal Components .............................................................. 3

Figure 1.1-3: XPOD External Components with Labels ................................................................ 4 Figure 1.1-4: Parts of the XPOD Mechanism with Labels ............................................................. 5 Figure 1.1-5: XPOD Arming Steps ................................................................................................. 6 Figure 1.2-1: Van Allen Belt Illustration [4] .................................................................................. 8 Figure 1.4-1: NORSAT-1 PDR Stage Bus Design Overview ...................................................... 11

Figure 2.2-1: Isometric View of XPOD Triple Mechanism before Deployment ......................... 18

Figure 2.2-2: Simplified XPOD Mechanism Actuation Sequence ............................................... 18 Figure 2.2-3: Section View of XPOD Pusher Plate with Labels .................................................. 20

Figure 2.2-4: XPOD GNB Bellville Stack .................................................................................... 21 Figure 2.2-5: Exploded View of XPOD GNB Stack .................................................................... 22 Figure 2.2-6: Bellville Washer Diagram with Labels [14] ........................................................... 22 Figure 2.2-7: Bellville Stacking Arrangement Example ............................................................... 24

Figure 2.2-8: Simplified View of a XPOD GNB Door ................................................................ 31 Figure 2.2-9: Door Free Body Diagram ........................................................................................ 31

Figure 2.2-10: FBD of Mechanism without Internal Forces ......................................................... 33 Figure 2.2-11: FBD of Mechanism with Internal Forces .............................................................. 34

Figure 2.2-12: FBD of XPOD Door ............................................................................................. 35 Figure 2.2-13: FBD of Left Clamp ............................................................................................... 35

Figure 2.3-1: Geometry Overview for Tip-Off Rate Analysis ..................................................... 39 Figure 2.3-2: Geometry Overview for Tip-Off Rate Analysis after Pusher Plate Separation ..... 40 Figure 2.3-3: Stage One Ejection Geometry Overall View .......................................................... 41

Figure 2.3-4: Stage 1 Ejection Zoom 1 ......................................................................................... 41 Figure 2.3-5: Stage 1 Ejection Zoom 2 ......................................................................................... 41

Figure 2.3-6: Stage Two Ejection Geometry Overall View ......................................................... 42 Figure 2.3-7: Stage 2 Ejection Zoom 1 ......................................................................................... 42 Figure 2.3-8: Stage 2 Ejection Zoom 2 ......................................................................................... 43

Figure 2.3-9: Stage Three Ejection Geometry Overall View ....................................................... 44

Figure 2.3-10: Stage 3 Ejection Zoom 1 ....................................................................................... 44

Figure 2.3-11: Stage 3 Ejection Zoom 2 ....................................................................................... 45 Figure 2.3-12: XPOD Deployment Steps with Pusher Plate ........................................................ 46 Figure 2.3-13: Spacecraft in Armed Configuration in Y-P Plane ................................................. 47

Figure 2.3-14: Spacecraft in Armed Configuration in X-P Plane ................................................. 48 Figure 2.3-15: XPOD Launch Rail Taper Geometry .................................................................... 48 Figure 2.4-1: +Z Tray Comparison ............................................................................................... 68 Figure 2.4-2: -Z Tray Reaction Wheel Hole Relocation ............................................................... 69 Figure 2.4-3: -Z Tray Internal Components Comparison ............................................................. 70

Figure 2.4-4: -Z Tray Sun Sensor Comparison 1 .......................................................................... 70

Figure 2.4-5: -Z Tray Sun Sensor Comparison 2 .......................................................................... 71

Figure 2.4-6: +X Panel Internal Comparison ................................................................................ 71 Figure 2.4-7: +X Panel External Comparison............................................................................... 72 Figure 2.4-8: -X Panel Internal Comparison ................................................................................. 72 Figure 2.4-9: -X Panel External Comparison ............................................................................... 73 Figure 2.4-10: +Y Panel Internal Comparison .............................................................................. 74

Figure 2.4-11: +Y Panel External Comparison............................................................................. 74

Figure 2.4-12: -Y Panel Internal Comparison ............................................................................... 75 Figure 2.4-13: -Y Panel External Comparison ............................................................................. 75 Figure 2.4-14: UHF Antenna Cutout Comparisons (Dimensions in Millimeters)........................ 76 Figure 2.4-15: +Z Panel Internal Comparison .............................................................................. 77

Figure 2.4-16: +Z Panel External Comparison ............................................................................. 77 Figure 2.4-17: +Z Panel External Isometric View ........................................................................ 78 Figure 2.4-18: -Z Panel Internal Comparison ............................................................................... 78 Figure 2.4-19: -Z Panel External Comparison .............................................................................. 79 Figure 2.5-1: NORSAT-1 Initial Bus Design ............................................................................... 79

Figure 2.5-2: Design of Payloads at Proposal Phase .................................................................... 80

Figure 2.5-3: Modified CLARA and ASR x50 Payload Volumes ............................................... 82

Figure 2.5-4: NORSAT-1 Bus Design, H27 Form Factor ............................................................ 82 Figure 2.5-5: Views of GHGSat-D during the NORSAT-1 Preliminary Design Phase [20] ....... 83 Figure 2.5-6: NORSAT-1 PDR Bus Design Exterior Views and Rough Dimensions ................. 84 Figure 2.5-7: NORSAT-1 Exterior PDR Design View 1 ............................................................. 85

Figure 2.5-8: NORSAT-1 Exterior PDR Design View 2 ............................................................. 85 Figure 2.5-9: NORSAT-1 Interior PDR Bus Design .................................................................... 86

Figure 2.5-10: Langmuir Probe Clearances from PDR Design .................................................... 87 Figure 2.5-11: NORSAT-1 Modified PDR Design ...................................................................... 88 Figure 3.2-1: Overall Top View of FEM ...................................................................................... 90

Figure 3.2-2: Overall Bottom View of FEM ................................................................................ 90

Figure 3.2-3: Bottom View of FEM with Constraints Shown ...................................................... 91 Figure 3.3-1: Image of First Natural Frequency ........................................................................... 93 Figure 3.3-2: Image of Second Natural Frequency ....................................................................... 93

Figure 3.3-3: Image of Third Natural Frequency .......................................................................... 94 Figure 3.3-4: Overall Nodal Displacement of Bus Components for -Z Loading Case ................. 95

Figure 3.3-5: Overall Panel Stress Distribution Image from -Z Loading Case ............................ 95 Figure 3.3-6: Bottom View of -Z Panel Stress Distribution from -Z Loading Case .................... 96 Figure 3.3-7: Image of Highest Stress Component from -Z Loading Case .................................. 96

Figure 4.1-1: XPOD Triple Vibration Test Mounting Holes and Axes ........................................ 98 Figure 4.1-2: Accelerometer Placement Image 1 ......................................................................... 99

Figure 4.1-3: Accelerometer Placement Image 2 ......................................................................... 99 Figure 4.1-4 – 50g Shock Test Waveform .................................................................................. 102 Figure 4.1-5: XPOD From +Z View Example Image ................................................................. 103 Figure 4.1-6: Measurement Example Image ............................................................................... 103

Figure 4.2-1: Overall Test Setup ................................................................................................. 109 Figure 4.2-2: Cord Looping Example ......................................................................................... 112 Figure 4.2-3: Spacecraft in XPOD Orientation ........................................................................... 113

1

Introduction

The work that was completed for this thesis focused on developing the mechanical aspects of

nanosatellites and microsatellites, along with developing their separation systems. Specifically,

work was completed for the Space Flight Laboratory’s in-house separation system (i.e. the

XPOD), the MiniMags feasibility study, the EV9 mission, and the NORSAT-1 mission. This

chapter will serve to introduce each of these projects along with introducing the objectives that

were set out in relation to these projects.

1.1 The Exoadaptable Pyroless Deployer

1.1.1 Separation Systems Overview

The eXoadaptable PyrOless Deployer (XPOD) is the nano-satellite and micro-satellite separation

system that has been developed at the Space Flight Laboratory (SFL). There are currently five

flight qualified XPOD designs at SFL. Table 1 gives an overview of each of the qualified XPOD

designs at SFL.

Table 1: Overview of Qualified XPOD Designs

XPOD Designation Qualified Maximum Mass Designed Spacecraft Dimensions

Single 1.33 kg 100 mm x 100 mm x 113.5 mm

Double 2.66 kg 100 mm x 100 mm x 227 mm

Triple 3.5 kg 100 mm x 100 mm x 340.5 mm

GNB 7.5 kg 200 mm x 200 mm x 200 mm

H27 10 kg 270 mm x 270 mm x 270 mm

2

Images of each of the currently qualified XPODs are shown in Figure 1.1-1:

Figure 1.1-1: Images of Currently Qualified XPODs

The XPOD uses the push-out deployment method, which simply means that the separation

system fully contains its associated spacecraft and upon receiving the deployment command a

door is allowed to open and a spring pushes the spacecraft out. This method is typically used for

nano-satellites and micro-satellites because their structures do not have much available surface

area or the structural rigidity to allow for more discretized hold down methods. Other examples

of separation systems that use the push-out method are the Poly-Picosatellite Orbital Deployer

(P-POD) from the California Polytechnic State University and the Tokyo Picosatellite Orbital

Deployer (T-POD) from Tokyo University [1]. Most of the variations between these deployers

stem from their available form factors, and the door actuation mechanisms that are used.

Finally, as mentioned above, discretized deployment methods are also used for spacecraft

interfaces with launch vehicles. These methods are typically for larger spacecraft since total

spacecraft containment would require an impractically large and massive push-out deployer

3

structure. These methods will either have features built into the main structure of the spacecraft,

or an additional separation system adapter component mounted to the spacecraft structure, which

are used as the interface points between the launch vehicle and spacecraft. The actuation method

for these systems may be pyrotechnics, or a mechanically actuated system. An example of a

system that uses and separation system adapter component, and is mechanically actuated, is the

Motorized Lightband Mark II from Planetary System Corporation [2].

1.1.2 Exoadaptable Pyroless Deployer Operation Method

As previously mentioned, the XPOD uses the push-out deployment method. This means that to

arm the XPOD, a spacecraft will be inserted, which causes an internal spring to be compressed

and to store potential energy. This spring is attached to the bottom panel of the XPOD and to a

“pusher plate” on which the spacecraft rests. Figure 1.1-2 details several of the internal XPOD

components.

Figure 1.1-2: XPOD with Labelled Internal Components

4

The door of the XPOD is then closed, which causes the spacecraft to push against it, and then the

actuation mechanism is locked into the closed position. To lock/arm the mechanism two clamps

are placed around top and bottom "wedges" that provide the holding force needed to keep the

door closed until the clamps are removed. A cord with one end fixed to a feature on the XPOD

and the other end tied to an eye-bolt is wrapped around the two clamps along with the XPOD

heater, where the heater is the device used to cut the cord. The cord is then tensioned by placing

the eyebolt through a fixed hole in the XPOD mechanism and tightening a nut onto the eyebolt.

A stack of spring washers is placed on the eyebolt as well such that the tensile force can be

measured by measuring the compression in the stack of spring washers. Finally, once armed the

XPOD is ready for flight on the launch vehicle (LV) that will be used for the mission. A signal

can then be sent from the LV to the XPOD through the XPOD-LV electrical interface, which

then will allow the actuation mechanism to be activated. Figure 1.1-3 illustrates the location of

the XPOD electronics, electrical interfaces, and the actuation mechanism.

Figure 1.1-3: XPOD External Components with Labels

Once the deployment signal has been received by the XPOD the cord used to hold the two

clamps onto the top and bottom wedges is cut using the heater. This allows the clamps to move

5

off of the wedges and the door to open. Figure 1.1-4 points out the parts of the mechanism that

are of importance for this step in the deployment process.

Figure 1.1-4: Parts of the XPOD Mechanism with Labels

Once the door is open the stored spring energy that was generated when the spacecraft was

placed into the XPOD pushes the spacecraft out through the open door area. The energy from the

spring is converted into the kinetic energy of the spacecraft, which dictates what the final relative

velocity of the spacecraft will be. An illustration of the arming events is shown in Figure 1.1-5

and for more information on the origination of the design of the XPOD see [3]. Finally, the side

panel of the XPOD is hidden in Figure 1.1-5 to aid in illustrating the arming steps, but in reality

it is always on the XPOD.

6

Figure 1.1-5: XPOD Arming Steps

The XPOD work for this thesis was completed mostly for the XPOD Triple. This XPOD had

previously and successfully been flown for the CanX-2 mission. At the start of the thesis work,

7

the CanX-7 mission, which uses a 3U form factor bus like CanX-2, began looking into using the

XPOD Triple as their separation system. However, the CanX-7 anticipated launch mass was

approximately four kilograms, which was greater than the mass that the XPOD Triple had

previously been qualified for. Therefore, the work completed for this thesis was to determine if

this increased mass was acceptable for the XPOD Triple.

1.2 The MiniMags Feasibility Study

The MiniMags feasibility study was a project that was funded by the Canadian Space Agency

(CSA) and carried out at SFL in conjunction with the Rutherford Appleton Laboratory (RAL).

The overall objective of this partnership was to produce and test a spacecraft subsystem that

produces active space radiation shielding. This subsystem, pending the successful demonstration

of its performance, would then be used on the CSA funded Polar Communications and Weather

(PCW) mission.

The PCW mission is specifying a Molniya orbit because of its observation requirements, and it

has a total lifetime requirement of 20 years. Using metal shielding on the PCW spacecraft, the

CSA predicts an average satellite lifespan of approximately five years based on the total

radiation dose they expect and what they have deemed tolerable for their parts. The MiniMags

payload would be used to increase the lifespan of each of the spacecraft in the PCW mission by

reducing the total radiation dose per unit time, such that the overall cost associated with the

mission would be decreased.

The goal of this portion of the mission, the feasibility study, was to determine the feasibility of

flying the MiniMags payload in a microsatellite. RAL was responsible for designing the payload

for this mission, where SFL was responsible for the satellite bus that could be used to support

their payload. Following the feasibility study phase, if this mission were to proceed, the goal

would then be to demonstrate, in flight, the effectiveness of the RAL radiation shield. The

purpose of demonstrating the RAL payload is such that other large satellite missions could use

the technology to increase the lifespans of their busses, without excessive metal shielding.

According to [4] the Van Allen belts are radiation belts that exist around the Earth in two bands.

The inner Van Allen belt exists from an altitude of approximately 100 km, in some areas, to

10,000 km and contains trapped electrons along with high energy protons. The outer Van Allen

8

belt exists from approximately 13,000 km to 16,000 km and contains mostly trapped electrons.

Figure 1.2-1 shows a representative illustration of the Van Allen belts around the Earth.

Figure 1.2-1: Van Allen Belt Illustration [4]

The radiation in these belts tends to cause both short term and long term issues with spacecraft

[5]. The two main short term issues that are realized are bit flips (aka single event upsets

(SEU’s)) and latch-ups (aka single event latch-ups (SEL’s)), which are both known as single

event effects (SEE’s). A bit flip is when a radiation particle changes the state of a bit in memory.

This causes errors in memory and operation of the spacecraft. A latchup is when a radiation

particle encounters a gate in a circuit and causes it to latch in a state where it allows the flow of

current when current is not supposed to flow. This can permanently damage electronics and

careful design must be implemented to counteract these effects, along with the errors that can

occur in memory from bit flips.

The long term effects of radiation are that electronics tend to degrade over time when exposed

[6]. Typically, an electronic component can endure a certain total exposure to radiation before it

is not reliable anymore. Testing of this effect on the ground is difficult because the true radiation

environment in space is not very reproducible, so instead of producing absolute correlations

between radiation and part life, comparative testing is carried out [7]. Comparative testing does

not necessarily reveal how long a part will last in space, but it can determine how well,

relatively, two parts work when exposed to the same type of radiation. Special materials can also

9

be used to increase resistance to radiation degradation, but typically these parts are more

expensive and can be out of reach of some missions.

The MiniMags mission was unique compared to other SFL missions because of its planned orbit,

along with the different type of payload. Typically, SFL satellites operate in Low Earth Orbit

(LEO), but the MiniMags demonstrator would either operate in a Highly Elliptical Orbit (HEO)

or Molniya Orbit. These orbits then require different launch vehicles in some cases because not

all LEO launch vehicles can deliver to these more energetic orbits. Also, the thermal and

radiation environments in these orbits are different than LEO because of their distances from

Earth and different eclipse periods [8].

SFL missions typically have either optical payloads, or communications based payloads. For

example, the AISSat satellites at SFL all carry Automatic Identification System (AIS) receivers.

These busses can receive the AIS signals from ships around the world, and then transmit them

back to Earth for terrestrial use. The BRITE satellites at SFL carry optical instruments for

observing stars. Both of these missions carry unique requirements for pointing accuracy, power,

mass, and operational characteristics. The magnetic shield payload therefore also carries unique

requirements in these areas when compared to optical and communications payloads.

Lastly, the MiniMags feasibility study kicked off after the work for this thesis began and was

finished before the thesis work completed. Therefore, this thesis covered all of the mechanical

design aspects that were completed for the MiniMags feasibility study.

1.3 The EV9 Mission

The EV9 mission is an Automatic Identification System (AIS) signal detection mission that is

being carried out by SFL along with exactEarth Limited. Originally, the EV9 mission scope was

such that it had two separate payloads, along with an original set of attitude requirements and

volume constraints. However, due to contractual changes the original scope was changed, which

led to the EV9 mission being partially re-designed.

The original mission, now called EV9-A, was designed to contain two payloads related to AIS

signal detection and processing. It was also designed to use hysteresis rods along with permanent

10

magnets for attitude control. Deployable UHF antennas were the required antennas for the uplink

from a ground station due to the volume constraints imposed on the EV9-A mission by the

launch provider.

The new mission, EV9, has had changes to the payload, attitude control system, and UHF

antennas compared to EV9-A. EV9 has only one payload and has a three-axis attitude control

system, which contains three reaction wheels along with three orthogonal magnetorquers. The

UHF antennas on EV9 are fixed, which means that they are in their flight configuration during

ascent on the launch vehicle. Both missions were scoped to use a GNB satellite bus design.

Aside from comparing EV9-A and EV9, one difference between both EV9 missions and other

SFL AIS satellites is that EV9 uses a deployable VHF antenna, where the others use a fixed VHF

antenna for AIS signal detection. The reason for this difference is due to launch provider volume

requirements imposed on the EV9 mission, which requires a deployable VHF antenna solution.

When work for the EV9 mission began, as part of this thesis, the requirements for the new EV9

bus were already established. Along with that, other GNB bus designs that met portions of these

requirements had already been designed for other mission. Therefore, the EV9 mission was able

to draw from these other missions in creating its bus design and this reduced the analysis that

was required since the designs that were drawn from had already been qualified. Finally, by the

end of the thesis work, the flight EV9 bus had been fully assembled and testing was completed

that verified the bus met all of its design requirements.

1.4 The NORSAT-1 Mission

Norwegian Satellite – 1 (NORSAT-1) is a mission being carried out at SFL to design a bus that

will contain three different Norwegian payloads. The three payloads are: 1) The Compact

Lightweight Absolute Radiometer (CLARA), 2) Langmuir probes, and 3) an AIS signal

detection payload. In what follows, the requirements for each of the payloads can be found in

[9]. An exterior view of the bus design from the Preliminary Design Review (PDR) stage of the

mission is shown in Figure 1.4-1:

11

Figure 1.4-1: NORSAT-1 PDR Stage Bus Design Overview

The CLARA payload is a scientific instrument that will be used to determine the total solar

irradiance of the Sun. This instrument is the primary payload in the NORSAT-1 mission and will

take operational precedence over the other payloads. The CLARA payload contains

measurement cavities that must be exposed, unshadowed, and must be pointed at the Sun with

±0.5 degree, 3σ, to carry out its scientific measurements. These cavities must also be held at a

constant temperature, with a maximum drift of 0.1 Celsius per hour, while the measurements are

being taken.

The Langmuir probe instrument is the secondary payload for the NORSAT-1 mission. This

instrument will measure the plasma around the Earth at a higher resolution compared to other

Langmuir probe instruments that have been flown in space. This instrument uses four probes that

are held at different electrical potentials outside of the bus, and measures the electrical

characteristics of the plasma near the spacecraft as the bus moves through the plasma around the

Earth. Most other Langmuir probe instruments use a probe that sweeps through different

voltages, but due to the time it takes for this sweep and the high velocities of orbiting satellites,

spatial resolutions of these measurements tend to be on the order of one kilometer. Since the

probes in this mission are held at constant voltages, sampling rates are much higher and give

12

spatial resolutions of the measurements on the order of one meter. These probes must be held

outside of the plasma sheath that forms around the bus as it goes through its orbit, which places

constraints on the orientation of the probes relative to the orbit.

The AIS signal detection payload (the ASR x50) is the tertiary payload for the NORSAT-1

mission. It is similar to the payloads used in the AISSats and on EV9, however it is of higher

performance. This payload is expected to be able to detect the AIS signals from more ships than

the AIS receivers in AISSat and EV9 in high ship density areas. This payload requires a

minimum of two VHF antennas for its AIS detection algorithms to work properly, but can use up

to four antennas. These antennas must be orthogonal to each other when mounted to a spacecraft,

and they must be orthogonal to the Langmuir probes in the NORSAT-1 mission. The baseline for

the VHF antennas is to use the deployable VHF antennas that are used in the EV9 mission.

The NORSAT-1 portion of the thesis work began at the kick-off of the mission. At this point

there was an existing bus design that was submitted as part of the proposal for SFL to obtain the

mission. This design was worked with, along with accounting for the evolving designs of each of

the payloads, to lead into the design of the bus that was generated for the thesis work.

1.5 Thesis Objectives

The objectives for the thesis work were associated with developing the mechanical aspects of

XPODs, the MiniMags feasibility study, the EV9 mission, and the NORSAT-1 mission. These

objectives were approached with the use of engineering design, computer simulation, and

physical testing such that they could be successfully fulfilled.

The objectives for XPODs were to first determine, with the use of analysis and testing, if the

XPOD Triple was compatible with a spacecraft mass of four kilograms. Once compatibility was

determined, then procurement of the XPOD Triple structure was to be completed such that the

CanX-7 mission had a separation system for their satellite.

The objectives for the MiniMags feasibility study were to determine, from a mechanical

standpoint, if the MiniMags payload could be supported in a nanosatellite or microsatellite. This

involved determining the mechanical requirements for the payload by consulting with the

13

payload manufacturer. Following this, plausible bus designs were proposed to the manufacturer

and they selected the most suitable design. Iterations were then completed to come up with a

design that was more tailored to the specific MiniMags payload requirements. Documentation of

the final design was created and the study was completed.

For the EV9 mission, the objectives were to merge the existing bus designs into the new EV9 bus

design, procure the newly created design, assemble the bus, and then test the assembled system.

These objectives were all completed by the time the thesis work was completed.

Finally, for the NORSAT-1 mission, the objective was to iterate on the bus deign that was used

for the mission proposal. This was to be completed based upon the evolving designs of the

payloads and as the bus design matured. The work for this mission was then handed off to

another SFL student such that the design could be further matured.

14

Mechanical Systems Design

2.1 Derivation of Mechanical Requirements

Since a major part of the mechanical requirements for a spacecraft are associated with the

loading that the spacecraft must withstand during launch, an investigation into what the loading

will be was required. The XPOD Triple, MiniMags, and NORSAT-1 did not know what launch

vehicle they would be flown on when they were being designed. Therefore, they were designed

to be compatible with a range of possible launch vehicles.

To be compatible with a launch vehicle from a mechanical standpoint a spacecraft and its

separation system must both survive the expected loading of the LV without failure, along with

exhibiting a certain degree of stiffness, which is quantified by the first natural frequency of the

spacecraft. Table 2 summarizes some of the relevant launch vehicle expected loading conditions,

along with their stiffness requirements:

Table 2: Launch Vehicle Mechanical Environment Summary

Vehicle Name

PSLV SOYUZ Ariane 5

Reference Document [10] [11] [12]

Dynamic Loading

Requirement

7g compression/

3g tension

6.5g

compression/

2.34g tension

2g compression/

2g tension

Dynamic Loading Factor of

Safety Requirement

1.25 1.3 1.1

Maximum Power Spectral

Density for Random

Vibration

6.7 GRMS 11.2 GRMS Not Listed

15

Random Vibration Factor of

Safety Requirement

Not Listed 2.25 Not Listed

Shock Loading Requirement 2 millisecond

half sine pule of

amplitude 105 g

Dependent on

separation system

Dependent on

separation

system

Shock Loading Factor of

Safety Requirement

Not Listed Not Listed 1.41

First Natural Frequency

Requirement

≥ 90 Hz ≥ 35 Hz ≥ 31 Hz

From the table above, the most mechanically demanding load is the shock load from PSLV,

along with their stiffness requirement. Therefore, a spacecraft must survive a 105g two

millisecond half sine pulse, and the spacecraft must have a first natural frequency above 90 Hz.

However, although it is not shown in Table 2, there is an underlying requirement that states that

all mechanical systems must be tested and function after experiencing a 50g two millisecond half

sine pulse, while only surviving a 105g pulse. Therefore, when designing mechanical systems at

SFL the following mechanical requirements are typically used for performing structural analysis:

1) The system shall survive a static load of 105 g without component failure with respect to the

ultimate strength of the materials comprising the component. If a 105g shock anomaly does

occur the components must not break apart and possibly damage other spacecraft on the

vehicle.

2) The system shall survive a load of 50 g without component failure with respect to the yield

strength of the materials comprising the component. If a 50g shock loading condition occurs

the spacecraft shall still be operable.

3) The system shall have a first natural frequency above 90 Hz.

16

When validating these requirements, both a Finite Element Model (FEM) and physical tests are

used. The 105g loading condition is only analyzed with the use of a FEM because the PSLV

provider believes this loading condition is a highly unlikely event on the launch vehicle, which

can lead to a high degree of overdesign of spacecraft components. Therefore, since this load will

likely permanently damage the structure during physical testing, it is not required. However, both

a FEM and physical testing with the 50g load are completed since this loading condition is said

to be much more likely to occur.

Thorough inspections along with accelerometers are used to determine the performance of the

test article during physical tests. To determine the natural frequencies of the structure, following

a prediction with the use of a FEA, a low amplitude sine wave is input into the structure using a

vibration table, and then accelerometers measure the accelerations of different points on the

structure. When these measured accelerations exhibit resonance with respect to the input load,

then the natural frequencies are determined and compared against the requirement. The results

are also compared against the results from the FEA that was previously completed for validation

purposes.

Physical tests are also completed to ensure that the dynamic and random vibration loads from

launch vehicles are tolerated by the structures that are tested. However, because their amplitudes

are less severe than both the 50g and 105g case they are not the main design drivers for the

structure. More information on the random and dynamic loading tests can be found in [13].

17

2.2 XPOD Mechanical Design

2.2.1 Mechanical Design of the XPOD Mechanism, Internal Preload and Main Spring

The XPOD Triple was originally designed to carry a 3.5 kg spacecraft, but the CanX-7 satellite

that will use a XPOD Triple has a currently predicted mass between 3.5 kg and 4 kg. Therefore,

a decision was made to increase the mass capacity of the XPOD Triple to 4 kg. This requires that

there be additional investigation into the design of the mechanism for the XPOD, along with its

main spring. The expected modification to the mechanism will be that the preload in the cord

that is used to hold the mechanism shut must be modified to be appropriate for the four kilogram

spacecraft. The main spring must also be made to contain more energy since the same ejection

velocity is desired and the spacecraft will be more massive, which will require more kinetic

energy.

The mechanism preload and main spring designs are also important to other XPODs that were

developed during this thesis activity. One other aspect that was not mentioned above is the

design of the pusher plate preload. This preload is designed to allow for practical machining

tolerances when manufacturing an XPOD, along with helping to maintain contact between a

spacecraft, the XPOD door, and the pusher plate when the system experiences loading in the

XPOD deployment direction.

Sections 2.2.2 through 2.2.6 will give an overview of the XPOD mechanism working principal

along with the details of how the other mechanical aspects of the XPOD are designed.

2.2.2 XPOD Mechanism Overview

The XPOD mechanism is used to both lock, and release the door of the XPOD with the use of a

clamp-wedge interference system (See Figure 2.2-1). The clamps are held together by a cord,

and upon the deployment signal being received by the XPOD electronics a heater is activated

that burns through the cord. A simplified sequence of images shows the general working concept

of the mechanism in Figure 2.2-2.

18

Figure 2.2-1: Isometric View of XPOD Triple Mechanism before Deployment

1) Mechanism Before Deployment 2) Cord cut by heater.

3) Clamps begin to move off of wedges. 4) Clamps fully moved off of wedges.

5) XPOD door allowed to open freely

Figure 2.2-2: Simplified XPOD Mechanism Actuation Sequence

19

The tension in the cord that is used to hold the clamps onto the wedges must be sufficient such

that the clamps do not move when the XPOD and spacecraft experience loading. This tension is

calculated using the expected worse case loading from the launch vehicle, the XPOD internal

forces, and accounting for the geometry of the mechanism components. After the tension that is

required is found then the compression that is required in the preload stack in the mechanism is

calculated.

2.2.3 Pusher Plate Preload Design

Due to tolerances in the XPOD system it is not practical to design an XPOD that when the door

is closed there are no gaps between the parts that lie between the bottom of the XPOD door and

the top of the XPOD base plate. If this was attempted it would likely result in a gap between

these parts if the tolerance stacking of the parts is too short. Alternatively, if the tolerance

stacking is too high, the XPOD door may not be able to close. Therefore to resolve this issue a

compressible, “Bellville Stack”, section is included in the stack of components between the

baseplate and door as shown in Figure 2.2-3.

This compressible section is typically a stack of spring washers that are located inside of the

XPOD pusher plate. However, in the event that the XPOD is for a CubeSat, an off-the-shelf

preload plunger is used instead of spring washers (see [3]). The total height of the components

between the door and base plate is chosen to be a height such that if the compressible section

were solid, would be too tall to fit within the door and base plate. However, once the

compressible section is compressed by a predetermined amount the door can close. After the

door is closed and the deployment mechanism is armed, the compressed section is allowed to

partially decompress, which causes it to force the spacecraft against the door. The force that is

generated is called the “preload force” and the magnitude of this force will be discussed in

Section 2.2.3.1. A section view of an XPOD pusher plate is shown in Figure 2.2-3.

20

Figure 2.2-3: Section View of XPOD Pusher Plate with Labels

2.2.3.1 Calculating the Design Preload Force

The preload force that is used in the XPOD must be sufficient such that if launch vehicle loading

on the XPOD forces the spacecraft towards the pusher plate, separation between the upper

spacecraft – XPOD interfaces does not occur. If the spacecraft and XPOD door were to separate

then there is a chance that damage could result when they come back into contact due to the

impact nature of that event. To determine which values should be used for the launch vehicle

loading an investigation of the applicable launch vehicles must be carried out (see Section 2.1).

Once a value has been selected for the LV loading then the preload stack force is determined

using Equation (2.1):

𝐹𝑝𝑟𝑒 = (𝑚𝑝𝑝 + 𝑚𝑆𝐶 + 𝑚𝑚𝑎𝑖𝑛 𝑠𝑝𝑟𝑖𝑛𝑔)𝑎𝐿𝑉𝑔 (2.1)

where

𝐹𝑝𝑟𝑒 is the required total preload stack force

𝑚𝑝𝑝 is the mass of the pusher plate

𝑚𝑆𝐶 is the mass of the satellite

𝑚𝑚𝑎𝑖𝑛 𝑠𝑝𝑟𝑖𝑛𝑔 is the mass of the XPOD main spring

21

𝑎𝐿𝑉 is the design acceleration from the launch vehicle

𝑔 is the acceleration due to gravity

An assumed mass of the main spring may need to be used since the main spring design may not

have been completed yet. Iteration to include the actual main spring mass may be required if the

assumption is very different than the actual value that will be found in future calculations.

2.2.3.2 Bellville Stack Design

A Bellville stack is comprised of several spring washers, shims, and normal (non-spring)

washers. The shims and non-spring washers are included in the stack to increase the height of the

stack without affecting its spring constant. These stacks are used in both the XPOD mechanism

and inside the XPOD pusher plate to create compressible assemblies that allow for measured

deflections to be converted to calculated compressive forces. A stack from an XPOD GNB is

shown in Figure 2.2-4:

Figure 2.2-4: XPOD GNB Bellville Stack

The Bellville stack along with the top washer are placed over a rod type component for stability

as shown in the exploded view in Figure 2.2-5.

22

Figure 2.2-5: Exploded View of XPOD GNB Stack

To calculate the force that a single Bellville washer exerts at a given compressed height, a

process from [14] was used, which is applicable to Bellville washers with a material thickness of

two millimeters or less. Figure 2.2-6 shows a cross-section of a Bellville washer along with

important quantities:

Figure 2.2-6: Bellville Washer Diagram with Labels [14]

where

𝐷 is the outer diameter of the washer

𝑑 is the inner diameter of the washer

𝑂. 𝐻. is the overall height of the washer

ℎ is the inside height of the washer

𝑡 is the thickness of the washer material

The diameter ratio of a Bellville washer is then given by Equation (2.2):

23

𝛿 =

𝐷

𝑑 (2.2)

where

𝛿 is the diameter ratio

Following the calculation of the Bellville washer diameter ratio, Equation (2.3) is used to find

the following dimensionless calculation constant:

𝑀 =

6

𝜋 × ln(𝛿)×

(𝛿 − 1)2

𝛿2 (2.3)

where

𝑀 is the dimensionless calculation constant

The deflection of a single Bellville washer is then defined by Equation (2.4):

𝑓𝑖 = 𝑂𝐻 − 𝑂𝐻𝑖 (2.4)

where

𝑓𝑖 is the deflection that the washer has undergone

𝑂𝐻𝑖 is the compressed overall height of a spring washer

Now, with the results of Equations (2.2) to (2.4), and with knowledge about the material

properties of the washer, the force being exerted can be calculated:

𝑃𝑖 =

𝐸 × 𝑓𝑖

(1 − 𝜇2) × 𝑀 × (𝐷2)

2 × [(ℎ −𝑓𝑖

2) × (ℎ − 𝑓𝑖) × 𝑡 + 𝑡3] (2.5)

where

𝑃𝑖 is the force exerted by the washer at the given deflection

𝐸 is the Young’s modulus of the washer material

𝜇 is the Poisson’s ratio of the washer material

24

After the force that results from a single spring washer is determined, then depending on how

these washers are stacked one can determine the force of the entire Bellville stack. An example

stacking arrangement is shown in Figure 2.2-7:

Figure 2.2-7: Bellville Stacking Arrangement Example

The stack shown in Figure 2.2-7 has a total of eight spring washers. These spring washers are

arranged with two unidirectional washers in each stack, and there are a total of four “individual

stacks”. In designing a stack one must consider the desired total force, the uncompressed height,

and the compressed height. These will depend on the XPOD geometry along with the desire to

make the deflection as measurable as possible given geometric and force constraints.

To explain the measurability of the stack, typically a total stack compression is between one and

two millimeters and errors of five percent on the final compression magnitude are acceptable.

Therefore, a stack that has a larger absolute deflection to achieve the desired force will be easier

to measure and compress the desired amount.

Once a desired stack design has been selected the overall force that the stack will exert for a

given deflection is found by first calculating the height of an “individual stack”:

𝐻𝑠𝑖= (𝑂𝐻𝑖 + 𝐴 × 𝑡) (2.6)

where

25

𝐻𝑠𝑖 is the resulting individual stack height

𝐴 is the number of unidirectional washers in an individual stack

Following the calculation of the height of the “individual stacks”, the total stack height can be

found:

𝐻𝑡𝑖= 𝐵 × 𝐻𝑠𝑖

(2.7)

where

𝐻𝑡𝑖 is the resulting total stack height

𝐵 is the number of individual stacks

Finally, the total stack force and total stack deflection can be found:

𝑃𝑠𝑡𝑎𝑐𝑘𝑖= 𝐴 × 𝑃𝑖 (2.8)

𝑓𝑠𝑡𝑎𝑐𝑘𝑖= 𝐵 × 𝑓𝑖 (2.9)

where

𝑃𝑠𝑡𝑎𝑐𝑘𝑖 is the total stack force for a given stack arrangement

𝑓𝑠𝑡𝑎𝑐𝑘𝑖 is the total stack deflection for a given stack arrangement

From a practical standpoint the equations above are somewhat difficult to solve for the stack

deflection as a function of force since the force equations are given as a function of deflection.

Therefore, when designing a stack, iterations through deflections are carried out until the correct

total stack force is achieved.

Other means of verifying the stack force, such as a force sensor, may be required to verify the

results of the design. This is because of the manufacturing tolerances for the spring washers,

which lead to variability in the forces that are generated.

26

Finally, there may be situations where two different stacks will be used in series to achieve the

desired uncompressed height, stack force and stack deflection. In this case, since both stacks are

in series, they will both take the same force. Therefore, to determine the total deflection of the

stack, trial and error can be used for each stack such that the force they produce is equal to the

desired stack force. Then the deflections of both stacks can be summed to give the total

deflection.

2.2.4 Designing the Main Spring

The main spring in the XPOD stores the majority of the energy that is used to eject the spacecraft

from the XPOD once the XPOD is signaled to deploy. The process and equations used to design

this spring are taken from [15] with slight modification to the process to be more appropriate for

this application.

Step 1) Determine the energy stored in the pusher plate preload:

𝐸𝑝𝑟𝑒𝑙𝑜𝑎𝑑 =

1

2𝑘𝑝𝑟𝑒𝑙𝑜𝑎𝑑(𝑥2

2 − 𝑥12)

(2.10)

where

𝐸𝑝𝑟𝑒𝑙𝑜𝑎𝑑 is the energy stored in the pusher plate preload stacks

𝑘𝑝𝑟𝑒𝑙𝑜𝑎𝑑 is the total effective spring constant of all pusher plate preload stacks

𝑥2 is the compression of the preload stack in its armed state

𝑥1 is the compression of the preload stack in its unarmed state

Step 2) Determine the energy required to eject the spacecraft at the desired ejection velocity:

𝐸𝑑𝑒𝑝𝑙𝑜𝑦 =

1

2(𝑚𝑆𝐶 + 𝑚𝑝𝑝)𝑣𝑓

2 + 𝜇𝑚𝑆𝐶𝑔𝑙𝑓 − 𝐸𝑝𝑟𝑒𝑙𝑜𝑎𝑑 (2.11)

where

𝐸𝑑𝑒𝑝𝑙𝑜𝑦 is the energy required to deploy the spacecraft

𝑣𝑓 is the desired exit velocity

27

𝜇 is the coefficient of friction between the XPOD rails and the spacecraft

𝑙𝑓 is the length that the spacecraft will be in contact with the XPOD rails

𝑔 is the acceleration due to gravity

𝑚𝑆𝐶 is the spacecraft mass

𝑚𝑝𝑝 is the pusher plate mass

Equation (2.11) assumes that there are frictional losses that are equal to those that would exist on

Earth. This assumption is used to be conservative in the analysis, but is not necessarily

physically representative.

Step 3) Determine the compressed length of the spring (𝐿𝑐), which is a function of the XPOD

geometry. When the spring is compressed it will be the total length that is between the top of

the XPOD base plate, and the bottom of the pusher plate when the pusher plate preloads are

compressed.

Step 4) Select a free length for the spring (𝐿𝑜).This value must be small enough such that the

pusher plate does not rest outside the XPOD when the spring is at its free length.

Step 5) Determine the required spring constant for the main spring to achieve the desired energy

storage for the deployment:

𝑘𝑠𝑝𝑟𝑖𝑛𝑔 =

2𝐸𝑑𝑒𝑝𝑜𝑦

(𝐿𝑜 − 𝐿𝑐)2 (2.12)

where

𝑘𝑠𝑝𝑟𝑖𝑛𝑔 is the required spring constant

Step 6) Based on an assumed wire diameter, calculate the mean diameter of the spring:

𝐷 = 𝑂𝐷 − 𝑑 (2.13)

where

𝐷 is the mean diameter of the spring

𝑂𝐷 is the outer diameter of the spring and is a function of XPOD geometry

28

𝑑 is the assumed wire diameter

Step 7) Determine the force in the spring when it is at its compressed length:

𝐹𝑐 = 𝑘𝑠𝑝𝑟𝑖𝑛𝑔(𝐿𝑜 − 𝐿𝑐) (2.14)

where

𝐹𝑐 is the force exerted by the spring at its compressed length

Step 8) Calculate the spring index and the stress concentration factor:

𝐶 =

𝐷

𝑑 (2.15)

𝐾𝐵 =

4𝐶 + 2

4𝐶 − 3 (2.16)

where

𝐶 is the spring index

𝐾𝐵 is the stress concentration factor

It is best if 4 ≤ 𝐶 ≤ 12 (See [15]). If the spring index is too small the springs are difficult to

manufacture. If the spring index is too large then there may be packaging issues since the springs

tend to easily tangle. However, since low quantities of these springs are purchased for XPODs

the issue of tangling is not of concern and 𝐶 ≥ 4 is the driving constraint.

Step 9) Calculate the number of active coils in the spring (Used in future calculations in the

spring design process):

𝑁𝑎 =

𝐺𝑑4

8𝐷3𝑘𝑠𝑝𝑟𝑖𝑛𝑔 (2.17)

where

𝑁𝑎 is the number of active coils in the spring

𝐺 is the Shear Modulus of the spring material

29

Step 10) Check if the design of the spring is satisfactory with respect to its resulting

geometry and stresses. First calculate the solid length of the spring:

𝐿𝑠 = 𝑑(𝑁𝑎 − 1) (2.18)

where

𝐿𝑠 is the solid length of the spring

The solid length must be greater than the compressed length (𝐿𝑐) determined in Step 3). If it is

not, then a new free length (𝐿𝑜) must be set and a new iteration started. If the design is still

satisfactory, using the solid length, calculate the spring’s solid force:

𝐹𝑠 = 𝑘𝑠𝑝𝑟𝑖𝑛𝑔(𝐿𝑜 − 𝐿𝑠) (2.19)

where

𝐹𝑠 is the force the spring exerts at its solid length

Using the solid force and other geometry, calculate the shear stress in the spring at its solid

length:

𝜏𝑠 = 𝐾𝐵 ×

8𝐹𝑠𝐷

𝜋𝑑3 (2.20)

where

𝜏𝑠 is the shear stress in the spring at its solid length

The factor of safety for the spring at its solid length can then be found:

𝑛𝑠 =

𝑆𝑠𝑦

𝜏𝑠 (2.21)

where

𝑛𝑠 is the factor of safety of the spring with respect to its shear yield strength

𝑆𝑠𝑦 is the shear yield strength of the spring material

A simplification can be made that assumes that the shear yield strength of the spring material is

equal to 45 percent of the spring’s ultimate tensile strength [15].

30

Another important quantity that should be determined is the fractional overrun of the spring. This

value is a measure of how close the compressed force, which is the operational force of the

spring, is to the solid force of the spring. If the solid force is within 15 percent of the compressed

force, with respect to the compressed force, then the spring may behave in a non-linear manner

and should be used as a design constraint [15]. The factional overrun is calculated using:

𝜉 =

𝐹𝑠

𝐹𝑐− 1 (2.22)

where

𝜉 is the fractional overrun

Finally, the critical free length of the spring is calculated. This value must be greater than the

free length of the spring selected in Step 4) or else the spring may be unstable and buckle when

compressed. The critical free length is calculated using:

𝐿𝑐𝑟 =𝜋𝐷

𝛼[2(𝐸 − 𝐺)

2𝐺 + 𝐸]

12

(2.23)

where

𝐿𝑐𝑟 is the critical free length of the spring

𝛼 is the spring end constraint (𝛼 = 0.5) for XPODs

The free length of the spring (𝐿𝑜) from Step 4) is the variable used for iteration for a given wire

diameter. A spreadsheet can be set up that calculates all of the required values for a given spring

for a given free length along with performing all of the checks in Step 10). If a free length for the

given wire diameter passes all of the checks in Step 10) then the spring can be used for the

XPOD. If not, then a different wire diameter should be used and the process repeated until a

suitable design is found.

31

2.2.5 Determining the Required Tension in the Mechanism Cord

The first step in determining how much tension is required in the mechanism cord to prevent the

clamps from coming off is to determine the forces that are trying to open the XPOD door when

the XPOD is in an armed and LV loaded state. Figure 2.2-8 shows a simplified view of an XPOD

door and Figure 2.2-9 shows a free body diagram (FBD) that results from the expected loading

condition.

Figure 2.2-8: Simplified View of a XPOD GNB Door

Figure 2.2-9: Door Free Body Diagram

32

The following variable definitions are required for this analysis:

𝑙1 The distance between the center of the hinge pin and the center of force from the

spacecraft, pusher plate and main spring

𝑙2 The distance between the center of the hinge pin and the center of the clamps in

the X-direction

𝑙3 The distance between the door center of mass and the hinge pin in the X-direction

𝑀ℎ𝑖𝑛𝑔𝑒 The moment generated by the torsion spring in the hinge when the XPOD door is

in the armed configuration

𝐹𝑜𝑝𝑒𝑛 The force that is a result of the acceleration from the launch vehicle on the

combined mass of the spacecraft, pusher plate and main spring. The full mass of

the main spring is not necessarily held by the door since the main spring is

attached to both the pusher plate and XPOD base plate. However, since the

proportion that is held by pusher plate compared to the base plate is not easily

calculated, a conservative assumption for this analysis is made that the entire

force is transferred from the pusher plate to the door.

𝐹𝑚𝑒𝑐ℎ The resulting force that the mechanism carries due to 𝐹𝑜𝑝𝑒𝑛

𝐹ℎ𝑖𝑛𝑔𝑒 The resulting force at the hinge due to 𝐹𝑜𝑝𝑒𝑛

𝐹𝑑𝑜𝑜𝑟 The inertial force acting at the door center of mass due to the acceleration of the

launch vehicle on the door’s mass

33

The following equations are used in finding force that the mechanism carries:

𝐹𝑜𝑝𝑒𝑛 = (𝑚𝑠𝑐 + 𝑚𝑝𝑝 + 𝑚𝑠𝑝𝑟𝑖𝑛𝑔)𝑎𝐿𝑉𝑔 + 𝐹𝑝𝑟𝑒𝑙𝑜𝑎𝑑 + 𝐹𝑚𝑎𝑖𝑛 𝑠𝑝𝑟𝑖𝑛𝑔 (2.24)

𝐹𝑑𝑜𝑜𝑟 = 𝑚𝑑𝑜𝑜𝑟𝑎𝐿𝑉𝑔 (2.25)

𝐹𝑚𝑒𝑐ℎ =

𝐹𝑜𝑝𝑒𝑛𝑙1 + 𝑀ℎ𝑖𝑛𝑔𝑒 + 𝐹𝑑𝑜𝑜𝑟𝑙3

𝑙2 (2.26)

where

𝑚𝑠𝑐 is the mass of the contained spacecraft

𝑚𝑝𝑝 is the XPOD pusher plate mass

𝑚𝑠𝑝𝑟𝑖𝑛𝑔 is the mass of the main spring

𝑎𝐿𝑉 is the acceleration of the launch vehicle acting on the contained mass

(assuming worst case orientation of the XPOD).

𝐹𝑝𝑟𝑒𝑙𝑜𝑎𝑑 is the XPOD pusher plate preload force

𝐹𝑚𝑎𝑖𝑛 𝑠𝑝𝑟𝑖𝑛𝑔 is the force due to the XPOD when it is in its armed configuration

𝑚𝑑𝑜𝑜𝑟 is the mass of the XPOD door

𝑔 is the acceleration due to gravity

After the force that the mechanism must resist is found, then determining the forces within the

mechanism can begin. A FBD of the XPOD mechanism is shown in Figure 2.2-10:

Figure 2.2-10: FBD of Mechanism without Internal Forces

34

where

𝐹𝑇 is the required tension in the mechanism cord

The forces that are applied to the clamps are a result of the mechanism cord tension, clamp

springs, and the mechanism opening force (the clamp springs are compressed between each

clamp and the mechanism and are used to push the clamps off once the cord is cut). To

determine the magnitude of the tension required in the cord, all internal forces that are present

must be accounted for. Figure 2.2-11 shows a FBD with these forces labelled. Only the forces on

the left clamp are shown for simplicity, but the forces are assumed to be of a similar nature on

the right clamp due to symmetry.

T

Figure 2.2-11: FBD of Mechanism with Internal Forces

where

𝐹𝑁 is the normal force generated at each clamp face

𝐹𝑓 is the frictional force generated at each clamp face

𝜃1 is the angle between the x-axis and cord at the clamp tops

𝜃2 is the angle between the x-axis and the cord at the clamp bottoms

𝜃𝑤𝑒𝑑𝑔𝑒 is the angle between the x-axis and a wedge/clamp face

𝐹𝑠𝑝𝑟𝑖𝑛𝑔 is the force exerted on each clamp from the clamp spring

35

The overall FBD of the mechanism can then be broken down to find the normal forces required

to prevent the XPOD door from opening. The FBD of the XPOD door is shown in Figure 2.2-12:

Figure 2.2-12: FBD of XPOD Door

Since the door must not be allowed to move in the Y-direction as shown in Figure 2.2-12, the

sum of forces in the Y-direction must be zero. Once this force balance is rearranged for the

normal force the result is:

𝐹𝑁 =

𝐹𝑚𝑒𝑐ℎ

2 cos 𝜃𝑤𝑒𝑑𝑔𝑒 (2.27)

The normal force can then be used in solving for the tension required in the cord to hold the

clamps to the mechanism. Figure 2.2-13 is the FBD of the left clamp and is used to determine the

required relationships.

Figure 2.2-13: FBD of Left Clamp

36

The angles of the tension forces in the FBD can vary for each model of the XPOD. These angles

can also be equal to zero, but they are still included in the following analysis such that the

equations can be applied to a variety of XPODs. One assumption with these angles, which was

made because of the geometry of all XPODs that have been currently designed, is that the angles

formed at the tops of both clamps are the same, and the same applies for the angles at the

bottoms.

Based on the FBD in Figure 2.2-13, and summing the forces in the X-direction, the following

result for the required tension is found:

𝐹𝑇 =

𝐹𝑠𝑝𝑟𝑖𝑛𝑔 + 2𝐹𝑁 sin 𝜃𝑤𝑒𝑑𝑔𝑒 − 2𝐹𝑓 cos 𝜃𝑤𝑒𝑑𝑔𝑒

cos 𝜃1 + cos 𝜃2 (2.28)

One assumption that is made when finding the required tension is that the frictional forces do not

aid in holding the clamps onto the XPOD. This assumption is used to increase the conservative

nature of the analysis and yields Equation (2.29), which is used for finding the tension in the

XPOD mechanism cord:

𝐹𝑇 =

𝐹𝑠𝑝𝑟𝑖𝑛𝑔 + 2𝐹𝑁 sin 𝜃𝑤

cos 𝜃1 + cos 𝜃2 (2.29)

Using the tension force found in Equation (2.29), along with the Bellville stack design processes

detailed in Section 2.2.3.2, a Bellville stack can then be designed for the XPOD mechanism.

Finally, a structural analysis to determine the suitability of the design of the XPOD structural

components, such as the mechanism wedges or hinge, should be completed in a separate

analysis.

2.2.6 Determining the Mechanism Jamming Conditions

Although friction is not considered when finding the required tension to hold the clamps onto the

XPOD during launch vehicle loading, it must be considered when determining an allowable

coefficient of friction between the wedges and clamps once the XPOD is signaled to deploy. For

this analysis it is assumed that the launch vehicle has already reached its final orbit and will no

37

longer be applying loads to the XPOD. This then leaves only the pusher plate preload force and

the force generated by the main spring acting against the XPOD door.

Using the conditions described above, the following equations are used to determine the

maximum allowable coefficient of friction between the clamps and the mechanism wedges.

Assuming: 𝐹𝑇 = 0, 𝐹𝑓 = 𝜇𝐹𝑁 , 𝑎𝑛𝑑 𝑎𝐿𝑉 = 0

where

𝜇 is the coefficient of friction between the clamps and XPOD wedges

From the Door FBD (Figure 2.2-12) the door opening force and the force the mechanism must

resist are found using:

𝐹𝑜𝑝𝑒𝑛 = 𝐹𝑝𝑟𝑒𝑙𝑜𝑎𝑑 + 𝐹𝑚𝑎𝑖𝑛 𝑠𝑝𝑟𝑖𝑛𝑔 (2.30)

𝐹𝑚𝑒𝑐ℎ =

𝐹𝑜𝑝𝑒𝑛 × 𝑙1 + 𝑀ℎ𝑖𝑛𝑔𝑒

𝑙2 (2.31)

Then, from the Top Wedge FBD (Figure 2.2-11), the required clamp normal forces can be found:

𝐹𝑁 =

𝐹𝑚𝑒𝑐ℎ

2(cos 𝜃𝑤𝑒𝑑𝑔𝑒 + 𝜇 sin 𝜃𝑤𝑒𝑑𝑔𝑒) (2.32)

Finally, from the Clamp FBD (Figure 2.2-13) and taking the sum of forces in the X-direction, the

maximum allowable coefficient of friction can be found. This relation is an inequality because

the coefficient of friction must be less than the ratio of the forces pushing the clamps off to the

forces holding them on for them to move:

𝜇 <

𝐹𝑠𝑝𝑟𝑖𝑛𝑔 + 2𝐹𝑁 sin 𝜃𝑤

2𝐹𝑁 cos 𝜃𝑤 (2.33)

The resulting maximum coefficient of friction, along with an acceptable margin, can then be

used as a guideline in selecting the materials and coatings for the clamps and wedges. However,

38

considerations may need to be taken to ensure that other processes of material bonding aside

from friction do not occur as a function of the coatings or materials used.

2.3 XPOD Tip Off Rate Analysis

2.3.1 Problem Definition

When a spacecraft is ejected from an XPOD, there is a resulting angular velocity for the

spacecraft. From a simplified standpoint, the main reason for this is the offset between the

XPOD pusher plate force and the center of mass of the spacecraft. This offset then creates a

torque that spins up the spacecraft, which is then subject to the geometric constraints of the

XPOD rails that limit the maximum angle the spacecraft can be rotated for a given position. It is

probable that there are other causes for the angular velocity, such as unplanned launch vehicle

rotations, or imperfections in the interface between the XPOD and spacecraft. However, these

are neglected in this analysis because they are not easily measurable and are not necessarily

controllable, which would result in the error bounds for the analysis being unpredictable.

One other assumption is that the pusher plate remains in contact with the spacecraft until the

XPOD main spring reaches its free length. Since the pusher plate is the only force acting to push

the spacecraft out of the XPOD, then the spacecraft cannot travel faster than the pusher plate.

Therefore, until the pusher plate begins to slow down relative to the spacecraft, which will only

happen once the main spring reaches its free length, they will remain in contact until that point.

This then allows the force being imparted on the spacecraft to act along the line of the geometric

center of the pusher plate for the duration that the spacecraft and pusher plate are in contact.

Finally, for this analysis, the term “block” will be used to refer to either the pusher plate and

spacecraft assembly while in contact, or to just the spacecraft. This term will be used because

similar equations and processes will be applied to either the pusher plate / spacecraft assembly or

to the spacecraft by itself, depending on the stage of the analysis.

39

2.3.2 Assumed Ejection Geometry for Limiting Angles

When a spacecraft is ejected from an XPOD, there are geometric constraints that limit the angle

that the spacecraft can rotate to, which also depend on its position. Figure 2.3-1 shows an

overview of the important geometric parameters used in determining these limits.

Figure 2.3-1: Geometry Overview for Tip-Off Rate Analysis

While the pusher plate and spacecraft are in contact it is assumed that the top corner of the

pusher plate drags against the “Top XPOD Rail”. It is also assumed that bottom spacecraft edge

drags against either the leading or outer edge of the bottom launch rail taper. This geometry will

be further explained in Sections 2.3.2.1 through 2.3.2.3.

Once the pusher plate and spacecraft have separated, then the upper corner of the spacecraft that

is still inside the XPOD will drag along the top rail. The bottom edge constraint will remain the

same as it was when the pusher plate and spacecraft were in contact. The new spacecraft

constraint is shown in Figure 2.3-2.

40

Figure 2.3-2: Geometry Overview for Tip-Off Rate Analysis

after Pusher Plate Separation

The combination of rail geometry along with the pusher plate separation event gives rise to six

different possible stages for a deployment. Each of these will be further detailed in Section

2.3.2.1 through Section 2.3.2.4.

2.3.2.1 Stage 1 Ejection Geometry

The overall geometry of the first stage of the ejection is show in Figure 2.3-3 where the red block

can represent either the spacecraft, or spacecraft and pusher plate assembly. This stage of the

ejection is characterized by the upper corner of the block dragging along the non-tapered edge of

the top XPOD rail. The bottom edge of the block drags against the leading edge of the bottom

rail taper. These two areas of importance are shown with zoomed images in Figure 2.3-4 and

Figure 2.3-5.

41

Figure 2.3-3: Stage One Ejection Geometry Overall View

Figure 2.3-4: Stage 1 Ejection Zoom 1

Figure 2.3-5: Stage 1 Ejection Zoom 2

Zoom

1

Zoom

2

42

2.3.2.2 Stage 2 Ejection Geometry

The overall geometry of the second stage of the ejection is show in Figure 2.3-6 where the red

block can represent either the spacecraft, or spacecraft and pusher plate assembly. This stage of

the ejection is characterized by the upper corner of block dragging along the non-tapered edge of

the top XPOD rail and the bottom edge of the block dragging against the outside edge of the

bottom XPOD rail taper. This stage begins once the angle that the block has rotated exceeds the

angle of the taper in the XPOD rail and will exist depending on the geometry of the spacecraft,

XPOD rails, and the clearances between these components. These two areas of importance are

shown with zoomed images in Figure 2.3-7 and Figure 2.3-8.

Figure 2.3-6: Stage Two Ejection Geometry Overall View

Figure 2.3-7: Stage 2 Ejection Zoom 1

Zoom

1

Zoom

2

43

Figure 2.3-8: Stage 2 Ejection Zoom 2

2.3.2.3 Stage 3 Ejection Geometry

The overall geometry of the third stage of the ejection is show in Figure 2.3-9 where the red

block can represent either the spacecraft, or spacecraft and pusher plate assembly. This stage of

the ejection is characterized by the upper corner of the block dragging along the tapered section

of the top XPOD rail and the bottom edge of the block dragging against the outside edge of the

bottom XPOD rail taper. This stage begins after the angle that the block has rotated exceeds the

angle of the taper in the XPOD rail, and the upper corner of the block begins to drag on the

tapered section of the top XPOD rail. These two areas of importance are shown with zoomed

images in Figure 2.3-10 and Figure 2.3-11.

44

Figure 2.3-9: Stage Three Ejection Geometry Overall View

Figure 2.3-10: Stage 3 Ejection Zoom 1

Zoom

1

Zoom

2

45

Figure 2.3-11: Stage 3 Ejection Zoom 2

2.3.2.4 Pusher Plate Contact Geometry

Sections 2.3.2.1 through 2.3.2.3 give rise to three phases of the spacecraft deployment and do

not take the separation even between the spacecraft and pusher plate into account. The separation

event occurs once the XPOD main spring has reached its free length during the deployment,

which causes a pulling force on the pusher plate. Since there are no fasteners or adhesive joining

the spacecraft and pusher plate, theoretically the spacecraft will maintain the velocity that it has

at the instant that the spring reaches its free length, while the pusher plate will experience

acceleration back into the XPOD body. This causes the pusher plate to slow down at the instant

that the main spring free length is reached and separation to occur at that instant. This can occur

during any of the three geometric stages, depending on how the main spring is designed, and

gives rise to the remaining three deployment stages. Figure 2.3-12 shows a deployment that is

constrained to limiting angles for phases one through three, with the pusher plate separation

shown as well:

a)

b)

c)

d)

46

e)

f)

g)

h)

i)

j)

Figure 2.3-12: XPOD Deployment Steps with Pusher Plate

In image a) the deployment begins, the pusher plate and spacecraft are still in contact, and the

geometry is that of phase one. The deployment continues through image b), until image c) where

the pusher plate and spacecraft are no longer in contact. The spacecraft continues to travel out of

the XPOD, not in contact with the pusher plate from image d) to image e). Stage two geometry

begins in image f) and continues until image g). Stage three geometry beings in image h) and is

continued until the spacecraft is leaving the XPOD in image j).

The deployment illustrated in Figure 2.3-12 was based primarily on the geometry of a 3U (See

[16]) spacecraft with increased taper angles and clearance between the rails to make the

spacecraft tipping angles more noticeable. This geometric scenario resulted in the separation

event occurring during phase 1 of the rail geometry as described in Section 2.3.2. However, since

this event can occur during any one of the phases, the analysis includes a pusher plate separation

check at all stages of the deployment to correctly calculate the geometric constraints.

It should be noted that Figure 2.3-12 does not include the solid model of the XPOD main spring

since it is not needed to show the separation and deployment events. Also, Figure 2.3-12 does

not show how the pusher plate will be pulled back into the XPOD due to the pulling force of the

XPOD main spring once the separation has occurred. This was not shown because the particular

47

dynamics of that event do not affect how the spacecraft tipping angle is constrained by the rails

after separation since the pusher plate and spacecraft are not in contact.

2.3.2.5 Variable Definitions

Figure 2.3-13 and Figure 2.3-15 are illustrations of the geometric variables that are described in

Table 3. It should be noted that the reference frames used in these images do not correspond

with those of the XPOD or spacecraft. The references frames that are used are solely used for

this solution.

Figure 2.3-13: Spacecraft in Armed Configuration in Y-P Plane

Figure 2.3-13 gives variable definitions based on spacecraft rotation about the X-Axis. To

determine the variable definitions for rotation about the Y-Axis, the variables in Figure 2.3-14

would be used. They are all the same, except for the center of mass offset (𝑐𝑚).

48

Figure 2.3-14: Spacecraft in Armed Configuration in X-P Plane

Figure 2.3-15 shows a zoomed in view of the taper in the launch rail detailing the variables used

to define the length of the taper, and the taper height:

Figure 2.3-15: XPOD Launch Rail Taper Geometry

49

Table 3 defines the input variables to the tip off rate solution along with a brief description of

each variable:

Table 3: List of Input Variables Used in Calculations and MATLAB Code

Variable Name Description

𝑷 Solution “position” coordinate direction

𝑷𝒕𝒂𝒑𝒆𝒓 Position that the rail taper begins in

𝑭𝒔𝒑𝒓𝒊𝒏𝒈 Force exerted on the pusher plate by the XPOD main spring

𝒉𝒕𝒂𝒑𝒆𝒓 XPOD rail taper height

𝒍𝒕𝒂𝒑𝒆𝒓 XPOD rail taper length

𝜽𝒕𝒂𝒑𝒆𝒓 The taper angle of the XPOD rail tapers

𝒉𝒊𝒏𝒔𝒊𝒅𝒆 XPOD rail to rail height

𝒉𝒑𝒑 The dimension of the pusher plate between XPOD rails

𝒍𝒑𝒑 Length of pusher plate in deployment direction

𝒍𝒔𝒂𝒕 Length of satellite in the deployment direction

𝒉𝒔𝒂𝒕 Dimension of satellite between XPOD rails

𝒄𝒎𝒙 Offset of satellite center of mass in the solution X direction

𝒄𝒎𝒚 Offset of satellite center of mass in the solution Y direction

50

2.3.3 Relevant Equations

The following equations were used in calculating the tip-off rates when a spacecraft is ejected

from an XPOD. Some of the equations have variables that change as the spacecraft is ejected,

which are calculated at different steps/iterations.

2.3.3.1 Linear Ejection Profile Equations

The effective spring constant of the pusher plate preloads is found using the following (See

Section 2.2.1 for more information):

𝑘𝑝𝑟𝑒 =

𝐹𝑝𝑟𝑒𝑚𝑎𝑥 − 𝐹𝑝𝑟𝑒𝑚𝑖𝑛

∆𝑃𝑝𝑟𝑒

(2.34)

where

𝑘𝑝𝑟𝑒 is the effective spring constant for the pusher plate preload stacks

𝐹𝑝𝑟𝑒𝑚𝑎𝑥 is the force that the preload typically exerts when the XPOD is armed

𝐹𝑝𝑟𝑒𝑚𝑖𝑛 is the force that the preload typically exerts when the XPOD is unarmed

∆𝑃𝑝𝑟𝑒 is the typical deflection of the pusher plate preload

The compression of the pusher plate preloads when the XPOD is unarmed is found using the

following:

𝑃𝑝𝑟𝑒1

=𝐹𝑝𝑟𝑒

𝑘𝑝𝑟𝑒

(2.35)

where

𝑃𝑝𝑟𝑒1 is the compression of the preload stack when in its unarmed state

The compression of the pusher plate preloads when the XPOD is armed is found using the

following:

𝑃𝑝𝑟𝑒2= 𝑃𝑝𝑟𝑒1

+ ∆𝑃𝑝𝑟𝑒 (2.36)

51

where

𝑃𝑝𝑟𝑒2 is the compression of the preload stack in its armed position

The velocity that the spacecraft and pusher plate will have as result of the energy contained in

the pusher plate preloads is found using the following and assuming no losses:

𝑣𝑝𝑟𝑒 = √𝑘𝑝𝑟𝑒(𝑃𝑝𝑟𝑒2

2 − 𝑃𝑝𝑟𝑒12 )

𝑚𝑡

(2.37)

where

𝑚𝑡 = 𝑚𝑝𝑝 + 𝑚𝑠𝑎𝑡

and

𝑣𝑝𝑟𝑒 is the velocity that will result for the pusher plate and spacecraft from the

preload stack energy

𝑚𝑡 is the combined mass of the spacecraft and pusher plate

𝑚𝑝𝑝 is the mass of the XPOD pusher plate

𝑚𝑠𝑎𝑡 is the mass of the satellite being ejected

The main spring force (𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖) that acts on the pusher plate can be found using Equation (2.38)

when the pusher plate and spacecraft are still in contact.

𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖= 𝑘𝑠𝑝𝑟𝑖𝑛𝑔 (𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑓

− 𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑐− 𝑃𝑖) (2.38)

where

𝑘𝑠𝑝𝑟𝑖𝑛𝑔 is the spring constant of the XPOD main spring

𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑓 is the free length of the XPOD main spring

𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑐 is the length of the XPOD main spring when it is armed

𝑃𝑖 is the position of the spacecraft at iteration “i”

52

The time between the positions of two consecutive iterations is found using Equation (2.39). This

duration of time can also be thought of as the time it takes to pass through an iteration.

∆𝑡𝑖 =

∆𝑃

𝑣𝑖 (2.39)

where

∆𝑡𝑖 is the time it takes to pass through the current iteration

∆𝑃 is the distance between the positions of two consecutive iterations. This is

deemed the position stepping size and is held constant.

𝑣𝑖 is the spacecraft linear velocity during the current iteration

The time that it has taken for the spacecraft to reach the current solution iteration can be then be

found using:

𝑡𝑖 = 𝑡𝑖−1 + ∆𝑡𝑖−1 (2.40)

where

𝑡𝑖 is the time that it has taken the spacecraft reach the current iteration

𝑡𝑖−1 is the time that it took the spacecraft to reach the previous iteration

∆𝑡𝑖−1 is the time it took the spacecraft to pass through the previous iteration

Finally, the linear velocity of the spacecraft can be determined using:

𝑣𝑖 = 𝑣𝑖−1 +

𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖−1

𝑚𝑡∆𝑡𝑖−1 (2.41)

where

𝑣𝑖−1 is the velocity at the previous iteration

𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖−1 is the main spring force acting on the pusher plate during the previous

iteration

∆𝑡𝑖−1 is the time it took the spacecraft to pass through the previous iteration

53

2.3.3.2 Geometric Constraint Equations

The limiting angles that the XPOD rails constrain the rotating assembly to must also be

calculated. The taper angle of the XPOD rails is found using:

𝜃𝑡𝑎𝑝𝑒𝑟 = tan−1 (

ℎ𝑡𝑎𝑝𝑒𝑟

𝑙𝑡𝑎𝑝𝑒𝑟)

(2.42)

The distance between the point where the spacecraft assembly touches the top rail and where the

spacecraft assembly touches the bottom rail, measured in the axis perpendicular to the

deployment axis, is required in the solution. This value, which will hereon be referred to as the

“total height”, will change as the spacecraft is ejected and must be calculated at every position.

When constrained to phase one geometry, the relationship is given by:

𝐻𝑡𝑖= ℎ𝑖𝑛𝑠𝑖𝑑𝑒 (2.43)

where

𝐻𝑡𝑖 is the total height at position “i”

When the spacecraft assembly passes into the phase two constraint scenario, the total height is

calculated using:

𝐻𝑡𝑖= ℎ𝑖𝑛𝑠𝑖𝑑𝑒 + ℎ𝑡𝑎𝑝𝑒𝑟 (2.44)

Once the spacecraft assembly has passed into the phase three constraints scenario, the total

height is given by:

𝐻𝑡𝑖= ℎ𝑖𝑛𝑠𝑖𝑑𝑒 + ℎ𝑡𝑎𝑝𝑒𝑟 + ℎ𝑝𝑖

(2.45)

where

ℎ𝑝𝑖 is the partial height added by the top rail taper at position “i”

The partial height arises from the top spacecraft corner dragging along the taper on the top rail.

As the corner drags along the taper, height is added that ranges from no additional height at the

54

leading edge of the taper, to a height equal to the taper height (ℎ𝑡𝑎𝑝𝑒𝑟) when the spacecraft is at

its last position inside the XPOD.

The equation used to calculate the amount of partial height that exists depending on the current

position of the spacecraft is:

ℎ𝑝𝑖= (𝑃𝑖 − 𝑃𝑡𝑎𝑝𝑒𝑟)tan (𝜃𝑡𝑎𝑝𝑒𝑟)

(2.46)

Now, the limiting angle for the spacecraft can be found as a function of its position. When the

spacecraft is in the first phase of its deployment, Equation (2.47) should be used. When the

spacecraft is in phase two or three of its deployment, then Equation (2.48) should be used. (See

[17] for derivation)

0 = ℎ𝑠𝑎𝑡 cos 𝜃𝑔𝑒𝑜𝑖+ (ℎ𝑠𝑎𝑡 sin 𝜃𝑔𝑒𝑜𝑖

+ 𝑙𝑝𝑝 cos 𝜃𝑔𝑒𝑜𝑖+ 𝑃𝑡𝑎𝑝𝑒𝑟 − 𝑃𝑖) tan 𝜃𝑔𝑒𝑜𝑖

− 𝐻𝑡𝑖

(2.47)

0 = ℎ𝑠𝑎𝑡 cos 𝜃𝑔𝑒𝑜𝑖+ (ℎ𝑠𝑎𝑡 sin 𝜃𝑔𝑒𝑜𝑖

+ 𝑙𝑝𝑝 cos 𝜃𝑔𝑒𝑜𝑖+ 𝑙𝑠𝑎𝑡 − 𝑃𝑖) tan 𝜃𝑔𝑒𝑜𝑖

− 𝐻𝑡𝑖 (2.48)

where

𝜃𝑔𝑒𝑜𝑖 is the limiting angle at position “i”

These equations would typically be solved using MATLAB where all variables are known

except for 𝜃𝑔𝑒𝑜𝑖.

Determination of the angular velocity that would result if the spacecraft followed the limiting

55

angles between two iterations (a.k.a. the “geometric angular velocity”) can then be found using:

𝜔𝑔𝑒𝑜𝑖

=𝜃𝑔𝑒𝑜𝑖

− 𝜃𝑔𝑒𝑜𝑖−1

∆𝑡𝑖−1

(2.49)

where

𝜔𝑔𝑒𝑜𝑖 is the geometric angular velocity

𝜃𝑔𝑒𝑜𝑖−1 is the limiting angle for the assembly at the previous iteration

Once the geometric angular velocity has been found, all of the information available from purely

geometric constraints has been found. Now, the consideration of the forces acting on the

spacecraft can begin such that physically relevant rotation angles and rates can be found.

2.3.3.3 Ejection Analysis Considering Forcing

Up to now the presented equations used to determine rotation angles only took the geometric

boundary conditions into account. These boundaries, while still giving important information

since the spacecraft will be constrained within the boundary angular limits may not be realistic.

This is because the forces required to have the spacecraft follow these boundaries may be much

higher than the actual rotation inducing forces that are experienced based on the assumptions

given in Section 2.3.1. The following equations are then used to calculate the spacecraft rotation

profile as a function of both the boundary conditions and rotation inducing forces.

The sum of torques acting on the block during the current iteration, assuming no impact occurs

between the block and the rails, is found using:

∑𝜏𝑖 = 𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖𝑐𝑚 (2.50)

where

∑𝜏𝑖 is the sum of torques acting on the block

The axis in which the torques are acting about will depend on the plane that the solution is being

found in. For example, if the solution is being found in the X-P plane, then the torques will be

56

about the Y-Axis, and the center of mass offset between the P-Axis and the main spring force

will be 𝑐𝑚𝑦. On the other hand, if the solution is being found in the Y-P plane, then the torques

will be about the X-Axis, and the center of mass offset between the P-Axis and the main spring

force will be 𝑐𝑚𝑥.

Using the sum of torques found in Equation (2.50), the current block angular velocity is found

using:

𝜔𝑖 =

∑𝜏𝑖−1 ∆𝑡𝑖−1

𝐼+ 𝜔𝑖−1 (2.51)

where

𝜔𝑖 is the block actual angular velocity in the current iteration

𝜔𝑖−1 is the block actual angular velocity in the previous iteration

∑𝜏𝑖−1 is the sum of torques acting on the block over the previous iteration

𝐼 is the moment of inertia of the block

The axis of rotation for the spacecraft in this equation will correspond to the axis that the

moment of inertia is calculated from, along with the axis about which the torques are calculated.

The moment of inertia must also be that of the spacecraft and pusher plate assembly, if they are

still in contact, since they are assumed to rotate together.

The actual angle that the block has rotated to is found using:

𝜃𝑖 = 𝜃𝑖−1 + 𝜔𝑖−1∆𝑡𝑖−1 (2.52)

where

𝜃𝑖 is the actual angle that the block is at, at the current iteration

𝜃𝑖−1 is the actual angle that the block was at, at the previous iteration

If the angle that the spacecraft has rotated to is greater than the allowable geometric angular limit

(ie 𝜃𝑖 > 𝜃𝑔𝑒𝑜𝑖) then impact has occurred between the block and rail. The current angle must then

be set to the allowable geometric angle (ie 𝜃𝑖 = 𝜃𝑔𝑒𝑜𝑖) and the effective angular velocity over the

previous time step must be recalculated using Equation (2.53), which is a modified version of

57

Equation (2.52):

𝜔𝑛𝑒𝑤𝑖−1

=𝜃𝑖 − 𝜃𝑖−1

∆𝑡𝑖−1

(2.53)

where

𝜔𝑛𝑒𝑤𝑖−1 is the effective angular velocity during the previous iteration

The acceleration that the block experienced as a result of the impact between the block and the

rails is found using:

𝛼𝑖−1 =

𝜔𝑛𝑒𝑤𝑖−1− 𝜔𝑖−1

∆𝑡𝑖−1

(2.54)

where

𝛼𝑖−1 is the angular acceleration during the previous iteration and will be negative

Equation (2.54) uses values that are all taken from a previous iteration (ie. 𝑖 − 1) because the

angular velocity at the current iteration must have already been found to determine if impact

occurred. Therefore, as part of the calculation process of the current iteration, you may still need

to find results that occur at the previous iteration’s position.

Since an acceleration occurred during the previous iteration, the torque that was originally found

for that iteration using Equation (2.50), which assumed that no impact occurred, is no longer

correct. The sum of torques that acted during the previous iteration can now be found using:

∑𝜏𝑛𝑒𝑤𝑖−1= 𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖−1

𝑐𝑚 + 𝐼𝛼𝑖−1 (2.55)

where

∑𝜏𝑛𝑒𝑤𝑖−1 is the corrected sum of torques that acted during the previous

iteration

58

2.3.4 Implementation of the Solution Method

This section will detail the implementation of the solution method used in predicting the tip off

rates of spacecraft that are deployed from XPODs. For iterative sections a single iteration will be

shown, but the iteration method will still be described.

2.3.4.1 Set XPOD Dependent Variables

The first step in solving for the tip off rate is to determine all of the following variables, which

are determined by direct measurement, that are associated with the XPOD being used:

1. ℎ𝑡𝑎𝑝𝑒𝑟 2. 𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑓

3. 𝐹𝑝𝑟𝑒𝑚𝑖𝑛 4. 𝑚𝑝𝑝

5. 𝑙𝑡𝑎𝑝𝑒𝑟 6. 𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑐

7. ℎ𝑖𝑛𝑠𝑖𝑑𝑒 8. 𝑥𝑝𝑟𝑒

9. 𝑙𝑝𝑝 10. 𝐹𝑝𝑟𝑒𝑚𝑎𝑥

11. 𝑘𝑠𝑝𝑟𝑖𝑛𝑔

2.3.4.2 Set Spacecraft Dependent Variables

The second step is to determine all of the following variables, which are determined by direct

measurement, that are associated with the spacecraft being ejected:

1. 𝑙𝑠𝑎𝑡 2. 𝑚𝑦

3. 𝑚𝑠𝑎𝑡 4. 𝑐𝑚𝑥

5. ℎ𝑠𝑎𝑡 6. 𝐼𝑥𝑥

7. 𝐼𝑦𝑦

59

2.3.4.3 Set Iteration Stepping Distance and Calculate Preliminary Values

Following the determination of all of the properties that are used in the solution, which are

derived from XPOD geometry, spacecraft geometry, and spacecraft physical properties, the

following initial values can be obtained.

The position step size, which is a selected value and determines the linear position separation

between iterations in the P-axis:

∆𝑃

The coordinate, in the P-axis, of the pusher plate to spacecraft interface once the XPOD main

spring has reached its free length:

𝑃𝑠𝑝𝑟𝑖𝑛𝑔 = 𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑓− 𝑙𝑠𝑝𝑟𝑖𝑛𝑔𝑐

(2.56)

The maximum position allowable in the solution coordinate system:

𝑃𝑚𝑎𝑥 = 𝑙𝑠𝑎𝑡 (2.57)

The coordinate, in the P-axis, of the leading edge of the XPOD rail taper:

𝑃𝑡𝑎𝑝𝑒𝑟 = 𝑃𝑚𝑎𝑥 − 𝑙𝑡𝑎𝑝𝑒𝑟

(2.58)

The total mass of the spacecraft and pusher plate:

𝑚𝑡 = 𝑚𝑝𝑝 + 𝑚𝑠𝑎𝑡 (2.59)

After the determination of the above constant values, the XPOD rail taper angle can be found

using Equation (2.42). Then the pusher plate preload spring constant (𝑘𝑝𝑟𝑒), initial compression

(𝑃𝑝𝑟𝑒1), armed compression (𝑃𝑝𝑟𝑒2

) can be found using Equations (2.34), (2.35), and (2.36)

respectively. Finally, the resulting velocity that the pusher plate preload imparts on the spacecraft

and pusher plate (𝑣𝑝𝑟𝑒) can be found using Equation (2.37).

60

2.3.4.4 Calculation of the Linear Ejection Velocity Profile

The calculation of the linear ejection velocity profile can now be completed, which neglects

frictional effects and energy loss in other areas. This assumption is made such that the ejection

velocity is as high as possible. This will result in the shortest time differences between iterations,

and therefore between angular positions, which should give higher angular velocities. Therefore,

since higher spacecraft tumble rates are of more concern when considering recovery to a stable

attitude, this should also yield a more conservative analysis.

2.3.4.4.1 Set and Calculate Initial Conditions

The following initial conditions for the block can be set and calculated, starting with the starting

time for the solution (a.k.a. the time at iteration zero):

𝑡0 = 0 𝑠

The velocity at iteration zero, which is equal to the velocity that the pusher plate preloads impart

onto the block, can also be set:

𝑣0 = 𝑣𝑝𝑟𝑒

The spring force acting during iteration zero (𝐹𝑠𝑝𝑟𝑖𝑛𝑔0) is found using Equation (2.38) and the

time it takes to pass through iteration zero (∆𝑡0) is found using Equation (2.39).

2.3.4.4.2 Calculate the Ejection Profile for Remaining Positions

While the spacecraft is still inside the XPOD (i.e. 𝑃𝑖 < 𝑃𝑚𝑎𝑥), the following iterative section

should be repeated such that the linear ejection profile of the spacecraft is determined:

First, calculate the time (𝑡𝑖) at the start of the current iteration using Equation (2.39). Then,

calculate the linear velocity (𝑣𝑖) that the block is travelling at during the current iteration using

Equation (2.41). Next, if the XPOD main spring has not reached its free length (i.e. 𝑃𝑖 <

𝑃𝑠𝑝𝑟𝑖𝑛𝑔), then calculate the main spring force (𝐹𝑠𝑝𝑟𝑖𝑛𝑔𝑖) using Equation (2.38). If the main spring

has reached its free length(𝑖. 𝑒. 𝑃𝑖 ≥ 𝑃𝑠𝑝𝑟𝑖𝑛𝑔), then set the main spring force to zero. Finally,

using the current iterations velocity, along with the position step size that was chosen at the

61

beginning of the solution, find the time it takes to pass through the iteration (∆𝑡𝑖) using Equation

(2.39).

Once all iterations have completed, a list of times, velocities, and spring forces as a function of

position will be generated. These values will be used in the remaining sections of the solution.

2.3.4.5 Calculate the Limiting Spacecraft Angles as a Function of Position

Here the limiting angular positions that satisfy the geometric constraints shown in Section 2.3.2

are calculated. Forcing is not considered at this point since this is only a geometric consideration.

The results of this section can then be used to check for impact and sliding when the torques on

the block are considered in Section 2.3.4.6.

The calculation begins at iteration zero with the block under stage one geometric constraints.

First, the total internal height (𝐻𝑡𝑖) must be found using Equation (2.43). Then, a determination

must be made as to if the pusher plate and spacecraft are still in contact. If they are in contact

(i.e. 𝑃𝑖 < 𝑃𝑠𝑝𝑟𝑖𝑛𝑔) then the length of the pusher plate (𝑙𝑝𝑝) must be included when solving for the

limiting angle (𝜃𝑔𝑒𝑜𝑖), and if they are not (i.e. 𝑃𝑖 ≥ 𝑃𝑠𝑝𝑟𝑖𝑛𝑔), then it can be ignored. The limiting

angle can now be calculated using Equation (2.47). These phase one limiting angle calculations

should be repeated while the phase one geometric conditions are met (i.e. 𝜃𝑖 < 𝜃𝑡𝑎𝑝𝑒𝑟 and 𝑃𝑖 <

𝑃𝑡𝑎𝑝𝑒𝑟).

Next, once the block has passed into phase two geometry the calculation begins with finding the

total internal height using Equation (2.44). Then the same pusher plate contact rules as in the

phase one scenario are applied, such that the pusher plate length will only be included in

calculating the limiting angle if it and the spacecraft are still in contact. Then the limiting angle

can be calculated using Equation (2.48). This process should be repeated while phase two

geometric conditions are met (i.e. 𝑃𝑖 < 𝑃𝑡𝑎𝑝𝑒𝑟).

The block is now subject to phase three geometry. First, the partial height (ℎ𝑝𝑖) must be

calculated for the given iteration using Equation (2.46). The partial height can then be used in

finding the total internal height using Equation (2.45). The same pusher plate contact rules as in

the phase one and two scenarios are applied , such that the pusher plate length will only be

62

included in calculating the limiting angle if it and the spacecraft are still in contact. Then the

limiting angle can be calculated using Equation (2.48). This process should be repeated while for

the remaining positions (i.e. 𝑃𝑖 < 𝑃𝑚𝑎𝑥).

Now that the limiting angles have been calculated at all positions of interest, the geometric

angular velocity (𝜔𝑔𝑒𝑜𝑖) for each position can be found using Equation (2.49).

2.3.4.6 Calculate the Spacecraft Tipping Profile

This section of the analysis calculates the angle and angular velocity of the spacecraft that is due

to the torque generated by the offset between the spring force and the spacecraft center of mass.

It also takes the rail geometry into account such that impacts between the rails and rotating body

constrain the rotation of the spacecraft.

This analysis assumes the spacecraft has no initial angular velocity or displacement. It also

assumes planar motion of the spacecraft for the solution, and that the angular displacements are

small. Therefore, to find the overall magnitude of the angular velocity of the spacecraft, two

orthogonal angular velocities are first calculated. The resulting magnitude is then found from the

RSS value of the two results.

First, set the initial conditions of the spacecraft and calculate the initial torque acting on the

spacecraft (∑𝜏0) using Equation (2.50). These are the only values required for the first iteration.

𝜃0 = 0

𝜔0 = 0 [1

𝑠𝑒𝑐]

Now, the next iteration may begin, and there are two scenarios for each new iteration. The first is

when no impact occurs between the block and the XPOD rails, and the second is when impact

does occur. When impact occurs the angle that the spacecraft has rotated to at the current

position (i.e. 𝜃𝑖), as a result of the angular velocity from the previous iteration,

exceeds the limiting angle for the current position (i.e. 𝜃𝑖 > 𝜃𝑔𝑒𝑜𝑖). Therefore, each new iteration

must begin with calculating the angle that the block has rotated to as a result of the previous

iterations angular velocity using Equation (2.52), followed by checking if this angle causes

63

impact. Then, depending on if impact did or did not occur, a slightly different process is carried

out. Both processes will be given in what follows.

Assuming that impact has occurred, the current angle must be set to the limiting angle for the

current position (i.e. 𝜃𝑖 = 𝜃𝑔𝑒𝑜𝑖) and the angular velocity for the previous iteration must be

updated to the result of Equation (2.53) (i.e. 𝜔𝑛𝑒𝑤𝑖−1). The angular acceleration that acted during

the previous iteration as a result of the impact (𝛼𝑖−1) is found then using Equation (2.54). The

sum of torques that acted during the previous iteration (∑𝜏𝑛𝑒𝑤𝑖−1) can now be found using

Equation (2.55), which is used to update the original sum of torques acting on the block during

the previous iteration. Now, the current angular velocity (𝜔𝑖) is found using Equation (2.51) and

the torques acting on the block in the current iteration are found using Equation (2.50).

Assuming that impact has not occurred, the angular displacement of the block at the current

position is kept as the result of Equation (2.52). Then, the current angular velocity is found using

Equation (2.51), and the torques acting on the block in the current iteration are found using

Equation (2.50).

After solving for the block angle and angular velocity at each position, then the solution is

complete. This process must be completed for both the X-P plane and the Y-P plane. The results

can then be combined to give the magnitude of the angular velocities for each iteration using:

𝜔𝑚𝑎𝑔𝑖

= √𝜔𝑥𝑖2 + 𝜔𝑦𝑖

2 (2.60)

The magnitude of the angular velocity at the final iteration, which is when 𝑃𝑖 = 𝑃𝑚𝑎𝑥 , is deemed

to be the tip-off rate for the spacecraft. The results of three different spacecraft solutions will be

given in Section 2.3.5.

2.3.5 Tip-Off Rate Analysis Results

Originally, this tip off rate analysis was created to determine the expected tip off rate for the

CanX-7 satellite. However, since CanX-7 has not flow there is no true way to verify if this

64

method produces valid results for that mission. On the other hand, the CanX-2 mission has

flown, was ejected from the same XPOD type that CanX-7 will be ejected from, and has the

same form factor as CanX-7. Therefore, for verification of the 3U form factor version of this

solution, the data from CanX-2 was used for comparison.

On the other hand, although not the original motivation for this work, this analysis was run for

both the AISSat-1 and UNIBRITE busses. These busses have both already been flown, have

measured bus properties, and have on-orbit sensor data for their initial angular rates. Therefore,

the true verification of this method was completed by comparing the analysis results with the

sensor data for these two missions. The results of these comparisons along with the comparison

between the CanX-7 analysis and CanX-2 on orbit data are shown in Table 4:

Table 4: Summary of Tip-Off Rate Analysis Verification

Spacecraft Predicted Tip Off Rate (𝑑𝑒𝑔𝑟𝑒𝑒

𝑠𝑒𝑐𝑜𝑛𝑑) On-Orbit Tip Off Rate (

𝑑𝑒𝑔𝑟𝑒𝑒

𝑠𝑒𝑐𝑜𝑛𝑑)

CanX-2 CanX-7: 3.96 3-5

AISSat-1 6.13 ~6

UNIBRITE 7.82 8-9

Based on the results summary shown in Table 4, it seems that the solution method produces

predictions within two degrees per second of the tip-off rates measured on orbit. This has

allowed the CanX-7 attitude control system designer to have insight into the expected tipoff rates

for their mission, which was the goal of this analysis. However, these results do not include

quantification of the errors on the input variables. Therefore, future work should be completed

for the algorithm to determine error bounds for the results by quantifying the input errors.

65

2.4 EV9 Mechanical Design

This section details the changes that took place from the original EV9-A mission to the current,

EV9, mission. These changes were carried out because of the new contractual agreements that

changed the scope of the mission. Originally, EV9-A did not have 3-axis attitude control, and

had requirements from the launch provider that resulted in the use of all deployable appendages

instead of fixed appendages. The EV9 mission now has a requirement for 3-axis attitude control,

along with relaxed launch vehicle requirements which allow for the use of fixed UHF antennas.

There is also a different payload in the EV9 mission when compared to the original EV9-A

mission.

2.4.1 EV9-A vs. EV9 vs AISSat-2 Comparison

The following table gives a comparison between the components that are part of the EV9-A,

EV9 and AISSat-2 busses:

Table 5: Bus Subsystems Summary

System Name EV9-A EV9 AISSat-2

AISSat Payload Yes No Yes

NTS-Payload Yes No No

EV9 Payload No Yes No

Deployable VHF Antenna Yes Yes No

Deployable UHF Antennas (4) Yes No No

Power Board Yes Yes Yes

Battery Packs (2) Yes Yes Yes

House Keeping Computer Yes Yes Yes

Attitude Control Computer Yes Yes Yes

Payload Computer No No Yes

66

VHF Antenna Switch Yes No No

Hysteresis Rods Yes No No

S-Band Patch Antennas (2) Yes Yes Yes

Sun Sensors (6) No Yes Yes

Magnetometer No Yes Yes

GPS Antenna No No Yes

GPS Radio No No Yes

UHF Radio Yes Yes Yes

S-Band Radio Yes Yes Yes

Rate Sensor No Yes Yes

Reaction Wheels (3) No Yes Yes

Magnetorquers (3) No Yes Yes

Many of the changes from EV9-A to EV9 are a result of the addition of a 3-axis attitude control

system and removal of the secondary payload. The differences compared to AISSat-2 are mostly

a consequence of launch vehicle volume restrictions and not requiring GPS antennas.

2.4.2 EV9 Mechanical Design Requirements

The following list of requirements was generated for the design of the EV9 bus:

1) The VHF antenna load bracket from EV9-A shall be used for the VHF antenna in EV9.

2) Compatibility of the bus with both SFL deployable and fixed UHF antennas shall be

maintained.

67

3) The EV9 design shall be based upon the AISSat-2 design when similar components are

used.

4) The spacecraft shall be oriented in its XPOD with the deployable VHF antenna being

held down by the XPOD door.

5) Compatibility for both an internal and external magnetometer shall be maintained.

The requirements above were used to make the changes from the original EV9-A structure to the

modified version that was required for EV9. There were other requirements that were followed

in the design of the original GNB, but they were not necessarily specific to the EV9 mission so

they were not repeated in the list above. More information on GNB requirements can be found in

[18].

2.4.3 Design Process

The starting point for the EV9 structure was the previously existing EV9-A solid model. From

this solid model, a separate version was created that could be used for modelling the structure

that would be used for EV9. The modifications that were made are detailed in the following lists

and images:

+Z Tray:

A rate sensor was added because the attitude control system for EV9 required a rate sensor,

while the attitude control system for EV9-A did not. The mounting features for the rate sensor

were already part of the EV9-A design since the tray’s design was from a more general GNB

tray design. The added rate sensor can be seen in Figure 2.4-1.

68

Figure 2.4-1: +Z Tray Comparison

-Z Tray:

The mounting holes for the Z-Axis reaction wheel were modified in the EV9 design when

compared to a typical GNB. This was done because the Z-Axis reaction wheel had to be shifted

in the –Z direction by four millimeters such that clearance for the screw heads for the VHF

antenna mounting bracket screws could be made. The Z-Axis wheel was stood off from the –Z

Tray by spacers, and the reaction wheel mounting holes that were in the Y-direction were shifted

in the -Z direction by the required amount. Typically this shifting is not required in the GNB

design because the VHF antenna bracket usually has through holes, and the –Z tray has threaded

holes in an extended boss for their mating. However, since the original EV9-A VHF bracket had

threaded holes for this connection, and not the –Z-tray, the screws heads had to remain inside the

–Z tray, and force the movement of the Z-Axis reaction wheel. Figure 2.4-2 shows a comparison

between the hole locations for the EV9-A structure and the EV9 structure.

69

Figure 2.4-2: -Z Tray Reaction Wheel Hole Relocation

After the holes were moved, the three reactions wheels could be added. These wheels were not

part of the EV9-A design since they were not required for the original attitude control system,

but the modified attitude control did require the wheels. These added reaction wheels are shown

in Figure 2.4-3

The EV9-A structure also had a VHF antenna switch, which was removed in the updated design.

Unlike how the EV9-A mission had two separate payloads that were supposed to use the same

VHF antenna, the EV9 mission only has one payload that requires the use of the VHF antenna.

Therefore, the ability to switch which payload has access to the antenna is no longer required and

the hardware that was designed to implement this capability was removed. The VHF switch

removal is also shown in Figure 2.4-3.

70

Figure 2.4-3: -Z Tray Internal Components Comparison

Sun sensors had to be added to all faces of the satellite because of the updated attitude control

system. The sun sensors that look though both Y Panels are mounted to the –Z Tray and are

shown in Figure 2.4-4 and Figure 2.4-5.

Figure 2.4-4: -Z Tray Sun Sensor Comparison 1

71

Figure 2.4-5: -Z Tray Sun Sensor Comparison 2

+X Panel:

The +X Panel had a Sun sensor added to it, along with a magnetorquer and the magnetorquer

mounting hardware. These components were all added because of the updated attitude control

system and can be seen in Figure 2.4-6 and Figure 2.4-7.

Figure 2.4-6: +X Panel Internal Comparison

72

Figure 2.4-7: +X Panel External Comparison

-X Panel:

The design of EV9-A required hysteresis rods for its attitude control system, but the updated

EV9 design did not require them. A magnetometer was also added to the EV9 design because of

the updated attitude control system. However, when EV9 was being designed the orientation that

the bus would be placed inside the XPOD separation system was unknown, which has impacts

on the placement of the magnetometer. This resulted in multiple mounting locations for the

magnetometer being included in the design of EV9. Therefore, on the –X panel, the hysteresis

rods were removed, a mounting location for the magnetometer was added, and a Sun sensor was

added. These modifications are shown in Figure 2.4-8 and Figure 2.4-9.

Figure 2.4-8: -X Panel Internal Comparison

73

Figure 2.4-9: -X Panel External Comparison

+Y Panel:

The power analysis of the EV9 mission showed that the mission had low power margins. This

necessitated the addition of a solar cell coupon on the +Y face. Another modification on the +Y

Panel was that the VHF antenna guides were moved 90 degrees relative to their original

positions. This was done because in the EV9-A mission the spacecraft was oriented in the XPOD

with its X-Panels on the open faces of the XPOD. This then allowed the deployable VHF

antenna to fold over the –X face. However, because fixed UHF antennas were to be used for

EV9, the Z-Panels had to be on the open sides of the XPOD, which meant the spacecraft had to

rotate 90 degrees. Therefore, the stowed configuration of the VHF antenna had to change such

that the antenna was folded over the –Z face, and therefore facilitated moving the antenna guides

to their new position. A magnetorquer and Sun sensor were also added to the updated +Y Panel.

These changes are all shown in Figure 2.4-10 and Figure 2.4-11.

74

Figure 2.4-10: +Y Panel Internal Comparison

Figure 2.4-11: +Y Panel External Comparison

75

-Y Panel:

Similar to the +Y Panel, a Sun senor was added to the –Y Panel and the VHF antenna guides

were shifted 90 degrees. These changes are shown in Figure 2.4-12 and Figure 2.4-13.

Figure 2.4-12: -Y Panel Internal Comparison

Figure 2.4-13: -Y Panel External Comparison

76

+Z Panel:

The EV9-A mission had deployable UHF antennas, which had coaxial cable feed-through

cutouts that were different than the cutouts for fixed style UHF antennas, which were also used

in AISSat-2. Therefore, since it was unknown as to what style of antenna was going to be used

when the EV9 structure was procured, the cutouts were sized to match the larger dimension of

both style of antenna cutout as shown in Figure 2.4-14. A Sun sensor along with internal

magnetometer mounting feautures were also added. These modificaiton are shown in Figure

2.4-15 and Figure 2.4-16. Finally, an external magnetometer boom was added, which was the

primary magnetometer mounting location for EV9. This is shown in Figure 2.4-17.

Figure 2.4-14: UHF Antenna Cutout Comparisons (Dimensions in Millimeters)

77

Figure 2.4-15: +Z Panel Internal Comparison

Figure 2.4-16: +Z Panel External Comparison

78

Figure 2.4-17: +Z Panel External Isometric View

-Z Panel:

The UHF coaxial cable feed-through holes had to be enlarged on the –Z panel as well. A sun

sensor was added, hysteresis rods were removed, and a magnetorquer was added. These

modifications are shown in Figure 2.4-18 and Figure 2.4-19 and conclude the major modification

summary between EV9-A and EV9.

Figure 2.4-18: -Z Panel Internal Comparison

79

Figure 2.4-19: -Z Panel External Comparison

2.5 NORSAT-1 Mechanical Design

2.5.1 NORSAT-1 Design Requirements and Starting Point

The NORSAT-1 mission mechanical design work began with an existing design of the bus that

was from the proposal stage of the mission. Exterior and interior views of the satellite showing

the proposed layout of components are shown in Figure 2.5-1:

Figure 2.5-1: NORSAT-1 Initial Bus Design

80

The design drivers at this stage of the mission were the dimensions of the payloads along with

their required relative orientations. There were also programmatic requirements that had effects

on the overall design of the bus, such as the requirement to use a qualified SFL separation

system. The total volume that was required to accommodate the bus subsystems and payloads led

to the selection of a bus that would be compatible with the XPOD DUO separation system. The

initial dimensions of each of the payloads are shown in Figure 2.5-2 and the relevant mechanical

requirements that were used at this stage of the mission are shown in Table 6. For a derivation of

the requirements see [9]:

Figure 2.5-2: Design of Payloads at Proposal Phase

81

Table 6: List of NORSAT-1 Relevant Mechanical Requirements

Requirement

Number Requirement Description

1. The spacecraft dimensions, including appendages, should be compatible

with a qualified SFL satellite deployment system.

2. The spacecraft shall include a design feature that can prevent it from

turning on while inside the separation system / launch vehicle.

3. The total launch mass of the spacecraft and its separation system shall be

less than 30 kg.

4. The satellite platform shall accommodate the payloads internally, except

where payload requirements explicitly state otherwise.

5. The CLARA payload shall be accommodated such that its sensor apertures

see the sun during operations.

6. The Langmuir Probes (qty. 4) shall be accommodated externally, parallel to

each other, and orthogonal to the CLARA line of sight.

7.

The AIS antennas (qty. 2) shall be accommodated externally, and be

pointed orthogonal to each other and orthogonal to the Langmuir Probe

booms.

2.5.2 NORSAT-1 Design Iterations

Following the proposal stage of the mission there were updates to the required volumes for both

the CLARA and ASR x50 payloads. The modified dimensions of these two payloads are shown

in Figure 2.5-3 along with their original volumes for comparison:

82

Figure 2.5-3: Modified CLARA and ASR x50 Payload Volumes

The reduced sizes of these payloads allowed for different designs of the bus to be possible since

the payloads were smaller. The first iteration on the design made use of the bus form factor that

would be compatible with the XPOD H27 separation system (See [19]). This design allowed for

an overall volume decrease of the bus and is shown in Figure 2.5-4:

Figure 2.5-4: NORSAT-1 Bus Design, H27 Form Factor

83

Some of the benefits of this design are that the overall volume of the bus is decreased, which

results in lower mass and lower launch cost. However, some of the downfalls of this design are

that there is not much remaining volume in the bus that would allow for the payloads to grow as

the mission progressed, along with the fact that SFL has not yet designed its own structure that is

compatible with the XPOD H27. This would result in more risk in the mission given the

uncertainty of the details of a bus with this form factor, along with the reduced volume margins

in the bus.

While the H27 form factor was being developed, a second option from the proposed bus design

was being looked into. This option was based off the GHGSat-D bus design that was also under

development at SFL. Figure 2.5-5 shows an external and internal view of the GHGSat-D bus

design that was used as the starting point for the second modified option of the NORSAT-1 bus

design:

Figure 2.5-5: Views of GHGSat-D during the NORSAT-1 Preliminary Design Phase [20]

The blue volume that is shown represents the internal payload volume that was available to be

used by the NORSAT-1 mission for payloads. Some of the other modifications that had to be

made were the removal of the star tracker, removal of the UHF antennas, addition of two more

S-Band antennas, and the addition of the exterior solar panel. The design of the NORSAT-1 bus

that resulted from these changes and was the design presented for the Preliminary Design

84

Review for the NORSAT-1 mission is shown in Figure 2.5-6, Figure 2.5-7, Figure 2.5-8, and

Figure 2.5-9:

Figure 2.5-6: NORSAT-1 PDR Bus Design Exterior Views and Rough Dimensions

85

Figure 2.5-7: NORSAT-1 Exterior PDR Design View 1

Figure 2.5-8: NORSAT-1 Exterior PDR Design View 2

86

Figure 2.5-9: NORSAT-1 Interior PDR Bus Design

The external solar panel used in this design was specified to use the same material and supplier

as the external solar panel for the NEMO-AM mission at SFL. There were other design iterations

for the NORSAT mission that attempted to not use an external solar panel, but they resulted in

the bus having faces that would be larger than the PDR version of this design. This then lead to

fairly excessive amounts of extra volume in the bus to exist, which was not beneficial for the

mission due to the extra structure that would be needed. These iterations also left little room on

the solar cell covered panel for thermal tapes, which was anticipated to be an issue from a

thermal standpoint.

Another design choice that was made for the PDR stage of the design was that the CLARA

instrument would be responsible for its own internal thermal control. The CLARA payload,

while taking measurements, requires a maximum temperature drift of 0.1 Celsius per hour for the

sensors in the instrument, which is different than thermal requirements for other SFL satellites.

Typically, the goal of a thermal designer at SFL is to keep all components in a bus within their

survival temperature limits, plus at least a five degree margin, in both a worst case hot and worst

case cold scenario. Therefore, a design that would allow the CLARA payload to have its entire

front face used for the thermal control of its internal sensors, having the sensors thermally

isolated from the bus, and then having the rear portion of the CLARA structure thermally

coupled to the rest of the bus was chosen. This is the reason why the CLARA payload has its

87

front face exposed, and it is not contained within the bus. The details of this design were not

fully finalized at the PDR stage of the mission, but are being worked out as the NORSAT-1

mission continues by the mission’s new mechanical designer.

One final consideration that was made was trying to keep the tips of the Langmuir probe

instruments as far away from the bus as possible. This was required since as the bus goes through

its orbit, it disturbs the plasma that it is travelling through. The probes, which attempt to measure

undisturbed plasma, want to be outside of the volume of plasma that is disturbed by the bus. The

placement that was decided upon was that they come out one of the X-faces since preliminary

mission operation analysis showed this location would result in them typically not being in the

ram or wake direction of the bus in its orbit. They then needed to be as far away from the bus as

possible, and the clearances that were achieved in the design from PDR are shown in Figure

2.5-10:

Figure 2.5-10: Langmuir Probe Clearances from PDR Design

88

There were concerns over the distance between the probe tips and the external solar panel during

the PDR meeting, which then lead to a modified PDR design. This design had both probes

moved toward the +Z panel on both the X panels, and was the final design that was a result of

the thesis work. An image of this design is shown in Figure 2.5-11:

Figure 2.5-11: NORSAT-1 Modified PDR Design

89

Finite Element Modelling

3.1 MiniMags Finite Element Model

The MiniMags mission required a preliminary analysis to determine if the proposed structure for

the bus was a feasible option for the expected payload. To create this model a simplified

approach was used that implemented the use of simplified geometry, lumped masses and face

gluing to reduce the model’s complexity.

Simplified geometry allowed the meshes that were created in the Finite Element Model (FEM) to

neglect small features, which reduced meshing time and reduced simulation time since less

elements were part of the solution. Lumped masses allowed for some of the non-structural

components in the bus to be neglected from a geometric standpoint, and to only consider the

mass and inertia from these components. This also reduced the time it took to create the model

and reduced simulation time. Finally, face gluing is a feature of the NX 8.0 FEM package that

allows individual and dissimilar meshes to be joined together with the use of stiff spring

elements that couple both the rotation and translation of connected elements (See [21]). The

stiffness of the elements that are created is a function of the two meshes being joined together

and will be similar to the average of the two. Using face gluing reduces the amount of time

required to mesh the structure because individual fasteners do not need to be considered,

however it can artificially increase in the overall stiffness of the joint.

All of the above simplification techniques reduce the accuracy of the model, and therefore

caution was used when comparing the results of the analysis against requirements. Higher factors

of safety were used and critical components would need to be modelled with a more

representative model when the design is more mature and less likely to have large changes.

Finally, since the design of the MiniMags bus was very preliminary and could likely experience

fairly large changes if the mission were to continue, a simplified analysis was deemed

appropriate for the feasibility study.

90

3.2 Finite Element Model Setup

Figure 3.2-1 and Figure 3.2-2 show the FEM that was made for the MiniMags structure:

Figure 3.2-1: Overall Top View of FEM

Figure 3.2-2: Overall Bottom View of FEM

3.2.1 Boundary Conditions

Based on the investigation of the launch vehicles from Section 2.1, the following boundary

conditions were derived for use in the finite element model:

1. A 105 g static load applied in any principal axis, not simultaneously. A minimum factor of

safety of 1.41 was considered when reviewing the results to determine feasibility, which was

derived from the shock loading factor of safety requirement of the Ariane 5 launch vehicle

2. A fixed translation and rotation constraint placed on the launch adapter ring, which is shown

in blue in Figure 3.2-3

91

Figure 3.2-3: Bottom View of FEM with Constraints Shown

3.2.2 Modelling Methodology

The FEM that was created for this structural analysis was based on the simplified model of the

spacecraft as previously described. This was done because the purpose of this analysis was to

prove the feasibility of the overall design and not to focus on the fine details associated with the

individual components in the design. The payload, along with other major components such as

the on-board computers, power subsystem, and attitude control system, were modelled as lumped

masses. The overall mass of the modelled spacecraft was made to replicate that of the mass given

by the preliminary mass budget for the spacecraft. This budget was based on the solid model that

was created for MiniMags, along with measured masses of components that were previously

used at SFL.

The FEM was made in the program NX 8.0 using NASTRAN based elements (See [21] for more

information on the following elements that are mentioned). All side panels of the bus were

modeled using 2D-PCOMP elements since they are used for representing honeycomb materials.

The solar cell coupons were modeled using 2D-PCOMP elements as well since they were also a

honeycomb structure. The threaded spacers in these structures were not modeled, but their

masses were accounted for by adding non-structural mass to the applicable structural

components. The adapter ring and plate were modeled using 3D-HEX8 elements. Finally, the

brackets that connect the panels together were modelled using 3D-HEX 8 elements as well.

92

Connections between the solar coupons and the bus were created using rigid RBE-2, 1D

elements, between nodes near the locations of the screws that would be used in the real world.

Connections between the panel joints and panels were created using “face gluing” between the

two interfacing surfaces. Finally, the connections between the adapter ring, adapter plate and bus

were all made using “face gluing” between their interfacing surfaces.

3.2.3 Material Selection

The materials selected for the structural components in the model were derived from the original

solid model of the MiniMags structure. This solid model was generated as part of the earlier

thesis work in the MiniMags feasibility study and was based on the existing NEMO-HD bus.

The physical properties for these materials are given in the table below:

Table 7: Summary of Material Properties Used in the MiniMags FEM

Material Density

(kg/m3)

Eorthogonal

(MPa)

Eparallel

(MPa)

Tensile Yield

Strength (MPa)

Compressive

Yield Strength

(MPa)

Shear

Yield

Strength

(MPa)

5056 Aluminum

Honeycomb Core [22]

97.7 2.03 2030 0.06 5.7 2.1

Aluminum 7075-T6 [23] 2,810 71,700 71,700 503 503 317

Carbon Fiber Reinforced

Plastic [22]

1,800 240,000 240,000 4000 4000 N/A

3.3 Results

After solving the model for the natural frequencies of the bus and for the bus behavior given the

loading described in Section 3.2.1, the following results were obtained.

93

3.3.1 Natural Frequencies

The first natural frequency was determined to be 102Hz and is shown in the image below. The

shape of the mode is a rocking of the bus structure about the Y-axis. In this mode the –Z Panel

does not resonate with the side panels and +Z Panel, and is a partially localized mode since the

entire bus does not resonate as a whole.

Figure 3.3-1: Image of First Natural Frequency

The second natural frequency was determined to be 112Hz and is shown in the image below. The

shape of the mode is a rocking of the entire bus structure about the X-axis. In this mode the –Z

Panel resonates with the side panels and +Z Panel, unlike the first mode, and is not a localized

mode.

Figure 3.3-2: Image of Second Natural Frequency

94

The third natural frequency was determined to be 133Hz and is shown in the image below. The

shape of the mode is a rocking of the entire bus structure about the Y-axis. In this mode the

–Z-Panel resonates with the side panels and +Z Panel, like the second mode, and is again not a

localized mode.

Figure 3.3-3: Image of Third Natural Frequency

3.3.2 Stress Results

The FEM was run using a 105g acceleration in each principal axis, in both the positive and

negative directions, not simultaneously. The worst case stresses were realized when the load was

in the -Z direction. This caused bending of the +Z panel toward the –Z panel, some bulging of

the side panels, and bending of the -Z panel around the separation adapter plate. The image

below shows the resulting displacement plot, where the units are in millimeters:

95

Figure 3.3-4: Overall Nodal Displacement of Bus Components for -Z Loading Case

The following image shows the stresses in the exterior panels from the same loading condition:

Figure 3.3-5: Overall Panel Stress Distribution Image from -Z Loading Case

96

The maximum stress experienced in the panels from this loading case was 135 MPa, which was

found in the skin layer of the –Z panel. A bottom view of this result is shown below:

Figure 3.3-6: Bottom View of -Z Panel Stress Distribution from -Z Loading Case

The maximum stress, and lowest safety factor, in the entire assembly was 171MPa, which was

found in a panel joint that joins the +Z panel and the –X panel. The maximum allowable stress in

this component, with respect to the yield stress is 503 MPa, which results in a factor of safety of

2.94. An image of this is shown below:

Figure 3.3-7: Image of Highest Stress Component from -Z Loading Case

97

3.4 Conclusions

The results of the FEM showed that the feasibility study design of the MiniMags bus met the

stiffness and loading requirements that were derived from the launch vehicle investigation. The

first natural frequency of the bus was found to be 102 Hz, which met the minimum first natural

frequency requirement of 90 Hz. The minimum factor of safety due to the 105g shock load was

found to be 2.94, which met the minimum factor of safety requirement of 1.41. Finally, it should

again be noted that since this modelling effort was at a high level and used coarse meshes along

with simplifications to the structure, that a more detailed analysis should be performed when the

bus design is more mature to ensure that the design requirements are still met.

98

Assembly Integration and Testing

4.1 XPOD Triple Vibration Testing

A “risk reduction” vibration test was performed on an XPOD Triple engineering model to

determine if the original XPOD Triple design was compatible with a spacecraft mass of four

kilograms. This test was deemed to be a “risk reduction” test because the XPOD Triple

engineering model that was used was not made of the same material that would be used for the

flight model. The objective for this test was to test only the Z-Axis shock load cases since

analysis showed that there may be issues with this loading case. A complete testing campaign

will be carried out in the future once the flight models are procured.

4.1.1 Axis Definition and Mounting Location

The XPOD was mounted to the vibration table using the mounting holes shown in Figure 4.1-1.

These holes are the same mounting holes that will be used for the XPOD when it is mounted to

the LV for the CanX-7 mission.

Figure 4.1-1: XPOD Triple Vibration Test Mounting Holes and Axes

99

4.1.2 Accelerometer Placement

The accelerometers for the test were placed in the locations shown in Figure 4.1-2 and Figure

4.1-3. Plastic cubes were used for mounting the accelerometers on the test article such that

orthogonal accelerometers could be mounted easily, at the same locations on the structure. The

axes shown in Figure 4.1-2 are the reference axes used for the remainder of the test description.

Figure 4.1-2: Accelerometer Placement Image 1

Figure 4.1-3: Accelerometer Placement Image 2

100

A description of each of the accelerometers is shown in Table 8.

Table 8 - List of Accelerometers

Type Point Location

1 Control Control Fixture plate

2 Measure Cube_1x XPOD Front Panel

3 Measure Cube_1y XPOD Front Panel

4 Measure Cube_1z XPOD Front Panel

5 Measure Cube_2x XPOD Door

6 Measure Cube_2y XPOD Door

7 Measure Cube_2z XPOD Door

8 Measure Cube_3x XPOD Bottom Panel

9 Measure Cube_3y XPOD Bottom Panel

10 Measure Cube_3z XPOD Bottom Panel

The “control” accelerometer was placed on the vibration table mounting plate, and not on the

XPOD. This accelerometer has a different purpose than measuring test data. It is used in the

vibration table closed loop control system such that the vibration table outputs the correct

vibration loads. The rest of the accelerometers were used to collect test data.

An accelerometer was paced on the Front Panel of the XPOD because the Front Panel is well

coupled with the use of screws to the other XPOD panels and is used to characterize the overall

XPOD body motion. An accelerometer was placed on the Door because the door is not as well

coupled to the rest of the XPOD body and information on its motion was desired. Finally, an

accelerometer was placed on the Bottom Panel because possible separations and impacts

101

between the Pusher Plate and Bottom Panel can occur, and information on this event was

desired.

4.1.3 Vibration Levels

This section contains information on the vibration levels that were used.

4.1.3.1 Low-Level Sine Test Levels

As previously mentioned, a low energy sine wave is used as the input to a system under test to

determine the natural frequencies of the structure. Specifically, a 0.5g sine wave is used, that

sweeps in frequency from 10-2000Hz at a rate of two octaves per minute. This allows resonances

in the structure to be measured by the accelerometers that are placed in the various locations on

the structure. The original determination of the natural frequencies is completed by performing a

low-level sine test at the start of the vibration test campaign.

The low-level sine test is also performed before and after major test profiles, such as the shock

test for this vibration campaign. The reason for this is to compare the response characteristics of

the structure from before and after the major vibrations to aid in determining if the structure was

altered by these major vibrations. Slight changes in the responses are typical since settling of the

structure usually takes place as it is subjected to vibrations. However, if major changes occur,

they would be dealt with on a case by case basis where the selected actions depend on the nature

of the change and if any damage is evident on the structure.

4.1.3.2 Shock Test Levels

The vibration profile for the shock test is depicted by the launch vehicle requirements that were

stated in Section 2.1. The profile that is used is a 50g half sine waveform, with a duration of ten

milliseconds. Figure 4.1-4 shows a plot of the waveform’s amplitude over time. This profile was

input into the XPOD Triple structure in both the positive and negative sense in the Z-axis since

analysis showed this to be the loading direction of concern.

102

Figure 4.1-4 – 50g Shock Test Waveform

4.1.4 Inspection Procedure

When performing vibration testing of XPODs, easily comparable images and measurements are

required that will detail whether or not there has been damage, or movement between components

in the XPOD assembly due to the test loading. This section gives a list of images and measurements

that should be gathered before and after the testing campaign for comparative purposes.

4.1.4.1 Image Description Conventions

Section 4.1.4.3 describes images by specifying both a location on the XPOD that should be

imaged, along with the direction that the image should be taken from. For example, Figure 4.1-5

is described as “Overall XPOD From +Z View”. Therefore, the image should be taken of the

XPOD (the entire assembly), from a +Z location.

0

10

20

30

40

50

60

0 1 2 3 4 5 6 7 8 9 10

Time (ms)

Ac

ce

lera

tio

n (

g)

103

Figure 4.1-5: XPOD From +Z View Example Image

4.1.4.2 Measurement Description Convention

The measurements that should be taken as part of the inspection are specified with a description,

a location, along with the axis in which the measurement should be taken in. Figure 4.1-6

illustrates the measurement that should be taken for “Distance Between Door and Front Panel at

–Y Corner in the Z axis” as an example. Note that in Figure 4.1-6 the distance is exaggerated for

the sake of clarity.

Figure 4.1-6: Measurement Example Image

+Y

+X

104

4.1.4.3 List of Images

The following list of images should be taken before and after an XPOD vibration test. Not all of

the images can be taken for all of the different XPOD types. Therefore, red text indicates images

that are specific to a particular type of XPOD. For examples of each of the following images see

[24].

Table 9: List of Images

Image # Description Image Direction

1 Overall Test Setup +X, +Y, +Z

2 Overall Test Setup +X, -Y, +Z

3 Overall Test Setup +X, +Y, -Z

4 Overall Test Setup +X, -Y, -Z

5 Overall XPOD +X

6 Overall XPOD (GNB/DUO/H27 Only) -X

7 Overall XPOD +Y

8 Overall XPOD -Y

9 Overall XPOD +Z

10 Overall XPOD (Single/Double/Triple Only) -Z

11 Overall Mechanism +X

12 Left Clamp +X

13 Right Clamp +X

14 Top Wedge +X

105

15 Bottom Wedge +X

16 Interface between Top and Bottom Wedges +X

17 Door to Front Panel Interface – Left side +X

18 Door to Front Panel Interface – Right side +X

19 Door to Front Panel Interface -Y

20 Door to Front Panel Interface +Y

21 Door to Rear Panel Interface (GNB/DUO/H27 Only) -X

22 Door to Rear Panel Interface -Y

23 Door to Rear Panel Interface +Y

24 Hinge (GNB/DUO/H27 Only) -X

25 Hinge +Z

26 Hinge -Y

27 Hinge +Y

28 -Y Door Locking Screw Hole Alignment +Z

29 +Y Door Locking Screw Hole Alignment +Z

30 Top Panel +X+Y Corner +Z

31 Top Panel –X+Y Corner +Z

32 Top Panel –X-Y Corner +Z

33 Top Panel +X-Y Corner +Z

34 Heater Block (Single/Double/Triple Only) -Z

106

35 Heater Block (GNB/DUO/H27 Only) +Z

36 Vectran Line on Countersunk Screw (Single/Double/Triple Only) +X

37 Vectran Line on Eyebolt +X

38 Vectran Line on Eyebolt (Single/Double/Triple Only) +Z

39 Vectran Line on Eyebolt (GNB/DUO/H27 Only) -Z

40 Mechanism Preload Stack +X

41 Mechanism Preload Stack +Z

42 XPOD / S.C Interface @ +X+Y+Z Corner (GNB/DUO/H27 Only) -X, +Y, -Z

43 XPOD / S.C Interface @ -X+Y+Z Corner (GNB/DUO/H27 Only) +X, +Y, -Z

44 XPOD / S.C Interface @ +X-Y+Z Corner (GNB/DUO/H27 Only) -X, -Y, -Z

45 XPOD / S.C Interface @ -X-Y+Z Corner (GNB/DUO/H27 Only) +X, -Y, -Z

46 XPOD / S.C Interface @ +X+Y-Z Corner (GNB/DUO/H27 Only) -X, +Y, +Z

47 XPOD / S.C Interface @ -X+Y-Z Corner (GNB/DUO/H27 Only) +X, +Y, +Z

48 XPOD / S.C Interface @ +X-Y-Z Corner (GNB/DUO/H27 Only) -X, -Y, +Z

49 XPOD / S.C Interface @ -X-Y-Z Corner (GNB/DUO/H27 Only) +X, -Y, +Z

50 Overall XPOD +X, +Y, +Z

51 Overall XPOD +X, -Y, +Z

52 Close-up of Front Panel Sensor Cube Depends on Sensor Placement

53 Close-up of Door Sensor Cube Depends on Sensor Placement

54 Close-up of Bottom Panel Sensor Cube (Single/Double/Triple Only) Depends on Sensor Placement

107

55 Close-up of Pusher Plate Sensor Cube (GNB/DUO/H27 Only) Depends on Sensor Placement

56 Close up of control sensor placement Depends on Sensor Placement

Given that sensor placement can vary for each test a specific direction is not given for the

capture direction of images 52-56. As a guideline, the angles that should be used to capture the

images of the sensor cubes should be such that all sensors, their leads, and their serial numbers

are easily visible. If sensor cubes are not used then images that capture the location, leads, and

serial numbers of each sensor should be taken.

4.1.4.4 List of Measurements

The flowing measurements should be taken as part of the XPOD inspection procedure. Not all of

the measurements can be taken for all of the different XPOD types. Therefore, red text indicates

measurements that are specific to a particular type of XPOD.

Table 10: List of Measurements

Measurement # Description Measurement

Axis

1 Bellville Stack Height Y

2 Distance Between Door and Front Panel at –Y Corner Z

3 Distance Between Door and Front Panel at +Y Corner Z

4 Distance Between Top and Bottom Wedges Z

5 Distance Between Door and Rear Panel at –Y Corner Z

6 Distance Between Door and Rear Panel at +Y Corner Z

108

7 Distance Between Flange on Left Clamp and Flange on

Bottom Wedge (Single/Double/Triple Only)

Y

8 Distance Between Flange on Right Clamp and Flange on

Bottom Wedge (Single/Double/Triple Only)

Y

4.1.5 Vibration Test Procedures

The test plan below was used for the XPOD Triple risk reduction test:

1) Mount the XPOD to the slip table in the Z-Axis orientation.

2) Mount the accelerometers on the XPOD and the fixture plate.

3) Take pictures and measurements of the test setup as described in Section 4.1.4.

4) Run the XPOD short form functional test.

5) Run the first Low Level Sine test with an input vibration amplitude of 0.5 g and a

frequency sweep from 10-2000 Hz at a rate of two octaves per minute.

6) Run the 50 g two millisecond half sine test in the +Z direction.

7) Run the 50 g two millisecond half sine test in the -Z direction.

8) Run the second Low Level Sine test with the same parameters as the first Low Level Sine

test.

9) Run the XPOD short form functional test.

10) Take pictures and measurements of the test setup as described in Section 4.1.4.

11) Remove the XPOD from the slip table.

12) Perform the XPOD deployment test.

Following the completion of the XPOD deployment test a visual inspection of the XPOD should

carried out. The photos and measurements that were taken during the test should also be

compared to determine if damage resulted from the test.

109

4.2 EV9 Horizontal Deployment Test

Since the VHF antenna for the EV9 mission has never been flow, or allowed to deploy following

spacecraft deployment from an XPOD, a risk reduction test was completed. The plan for this test

was to use the XPOD ground test vehicle (GTV), a representative mass for the EV9 satellite, and

the EV9 flight XPOD to perform a horizontal deployment of the spacecraft. Typically a flight

XPOD would not be used for risk reduction testing to avoid the risk of damaging flight

hardware. However, because EV9 has slightly different separation system interfaces compared to

other satellites, the EV9 specific XPOD had to be used. The setup can be seen in Figure 4.2-1.

Figure 4.2-1: Overall Test Setup

The representative mass was composed of EV9-A structural components (due to availability of

EV9 components), EV9 structural components, along with several internal GNB component

mass dummies. The magnetometer boom was not included in this test because it would interfere

with the GTV. The UHF antennas were also not included since they are not needed to test the

110

deployment of the VHF antenna and they had not yet been included in a GTV test, which

introduced undesirable risk to the test. However, if the VHF antenna deployment test indicated

possible contact with the UHF antennas during deployment, a subsequent test could be

performed with the UHF antennas included. The VHF antenna that was used was the flight spare

antenna from the EV9 mission and was the same length as the EV9 flight VHF antenna.

One issue with this test was that the flight main spring for the EV9 XPOD was not at SFL for the

test. To work around this a similar spring was used, but the material of this similar spring was

stainless steel and not the same titanium alloy that flight XPOD GNB springs are typically made

of. This caused the stiffness of the spring to be slightly different, but as a representative test, this

was deemed acceptable. The process that was used to estimate the spring constant of the steel

spring is detailed in Section 4.2.1.

The XPOD was actuated without the use of the XPOD firing mechanism for several of the test

deployments such that the test was more easily carried out. However, for the last deployment test

the XPOD was fired using the XPOD actuation mechanism. A more detailed methodology is

given in Section 4.2.2.

4.2.1 Spring Constant Determination Procedure

The procedure that was used to determine the spring constant of the XPOD main spring is given

below:

Step 1) Place the XPOD on a flat surface with the deployment direction facing upwards and with

the door open.

Step 2) Measure the height of the pusher plate above the baseplate at each of the four corners of

the pusher plate. The average of these four values will be the starting height (ℎ𝑠𝑡𝑎𝑟𝑡).

Step 3) Place a known mass (𝑚) in the center on top of the pusher plate to increase the

compression of the main spring.

111

Step 4) Measure the new height of the pusher plate above the base plate at each of the four

corners. The average of these 4 values will be the ending height (ℎ𝑒𝑛𝑑).

Step 5) Estimate the spring constant of the XPOD main spring using:

𝑘𝑠𝑝𝑟𝑖𝑛𝑔 =𝑚𝑔

ℎ𝑠𝑡𝑎𝑟𝑡 − ℎ𝑒𝑛𝑑 (2.61)

where

𝑚 is the mass of the object placed on the pusher plate

ℎ𝑠𝑡𝑎𝑟𝑡 is the height of the pusher plate off of the base plate before the object was

added

ℎ𝑒𝑛𝑑 is the height of the pusher plate off of the base plate after the object was

added

𝑔 is the acceleration due to gravity

4.2.2 Deployment Test Procedure

The procedure that was used to carry out the test deployment of EV9 from its XPOD to

determine the performance of the deployable VHF antenna is given below:

Step 1) Mount the XPOD to the vibration plate. This plate is needed as part of the GTV assembly

to position the XPOD correctly.

Step 2) Attach the GTV to the EV9 representative mass.

Step 3) Attach the GTV XPOD legs to the XPOD as shown in Figure 4.2-1.

Step 4) Clear an area in the lab such that there is enough space to have the spacecraft deploy

Step 5) Lay down the horizontal deployment mat.

112

Step 6) Place the horizontal deployment stand at one end of the mat.

Step 7) Place the vibration plate with the mounted XPOD onto the horizontal deployment stand

such that the horizontal deployment legs are also engaged.

Step 8) Wrap the EV9 VHF antenna around the mass dummy into its “stowed” configuration.

Tape may be needed to temporarily hold the antenna down while sliding the mass dummy

into the XPOD and checking that the spacecraft can move properly in and out of the XPOD.

Step 9) Verify that the EV9 mass dummy slides into and out of the XPOD easily such that the

mass dummy will not get jammed during deployment. It is likely that leveling of the

horizontal XPOD assembly, along with matching the height of the XPOD assembly to the

height of the spacecraft will be necessary to allow the spacecraft to easily slide into and out

of the XPOD. Perform this task by adjusting the 6 adjustable legs in the horizontal XPOD

setup until the spacecraft can easily slide in and out of the XPOD without jamming.

Step 10) Remove the tape from the VHF antenna if it was previously used to hold the antenna

down.

Step 11) Loop a piece of thin Vectran cord through the cutout on the right side of the XPOD

door (See Figure 4.2-2).This cord will be used to hold the door shut.

Figure 4.2-2: Cord Looping Example

113

Step 12) While holding the VHF antenna in its “stowed” configuration slide the spacecraft mass

dummy with the GTV fully into the XPOD in the flight orientation (See Figure 4.2-3) and

close the door. Hold the door shut with the cord that was looped around the door in Step 11).

Figure 4.2-3: Spacecraft in XPOD Orientation

Step 13) Stand behind the XPOD and ensure that when the spacecraft is ejected it will not hit

anything.

Step 14) Set up a recording device such that the deployment can be analyzed. Using a high speed

recording device is preferred since detailed information about the motion of the antenna will

desired after the deployment.

Step 15) Release one end of the looped cord such that the XPOD door can open.

Step 16) After the spacecraft has come to a full stop, stop the recording device.

Step 17) Repeat Step 11) through Step 16) as necessary to collect the required data and to build

up some confidence in the results that are realized. For the final deployment fully arm the

XPOD and deploy the XPOD using the process described in [25].

114

Conclusion

The work that was completed for this thesis included the mechanical design of three

microsatellites and their separation systems. Several missions and design phases were touched

upon while completing this work. Computer modelling along with AIT activities also comprised

a significant portion of the work that was completed.

The XPOD Triple is now at a point where procurement will take place. This was reached by both

analyzing the mechanism in the XPOD, and its main spring. An FEM was also created such that

the structure of the XPOD was deemed acceptable, although not mentioned in this thesis

document. Finally testing of the structure was carried out to ensure that the design was suitable

for the CanX-7 mission, with a bus mass of up to four kilograms.

The MiniMags feasibility study was completed, which involved the creation of a preliminary

structural design that may be used in the future. This design was deemed acceptable by the

payload provider in terms of its mass, power, and volume availability. The bus was also analyzed

such that it was deemed acceptable from a mechanical standpoint with the MiniMags payload.

The EV9 mission was brought from its previous EV9-A design to the current EV9 design. The

tests that were done in ensuring the VHF antenna is compatible with the XPOD ensured that the

risks in using the deployable VHF antenna were acceptable. The flight model of the EV9 bus

was assembled at SFL and testing of the bus to deem that it was built to specification was

successfully completed.

Finally, the NORSAT-1 mission was brought from a proposal level bus design to the PDR level

bus design during the course of the work for this thesis. Trade studies that compared different

bus designs, along with modifications to the layout of the payload appendages took place. The

mechanical design was handed off and is under further development by the current NORSAT-1

mechanical designer.

Overall, appreciable contributions to the mechanical aspects of the above mentioned programs

were made. The mechanical design activities contributed to allowing the structural systems in the

115

above mentioned programs to fulfill their mission requirements. The computer modelling

activities allowed for the verification of these mechanical design requirements before physical

testing could take place, which allowed for iteration of the designs in a time and cost efficient

manner. Finally, the AIT activities that took place brought all of the previously mentioned

missions much closer to being ready for launch and on orbit operation.

116

References

[1] EO Portal, "CubeSat Concept and the Provision of Deployer Services," 2104. [Online]. Available: https://directory.eoportal.org/web/eoportal/satellite-missions/c-missions/cubesat-concept. [Accessed March 2014].

[2] Planatery System Corporation, "Plantery System Corporation Products," 2011. [Online]. [Accessed 14 May 2014].

[3] W. O'Brien, "The Design, Manufacturing, and Testing of eXperimental Push-," University of Toronto, Toronto, 2007.

[4] Space Radiation Analysis Group, Johnson Space Center, "What is space radiation," 22 04 2013. [Online]. Available: http://srag-nt.jsc.nasa.gov/SpaceRadiation/What/What.cfm. [Accessed 07 01 2014].

[5] J. R. Wertz, D. F. Everett and J. J. Puschell, Space Mission Engineering: The New SMAD, Hawthorne: Microcosm Press and Springer, 2011.

[6] C. Dyer, "Radiation Effects on Spacecraft and Aircraft," QinetiQ, Farnborough, 2001.

[7] R. Zee, "Spacecraft Design Presentation," Toronto, 2012.

[8] D. G. Gilmore, Spacecraft Thermal Design Handbook, El Segundo: The Aerospace Press, 2002.

[9] Space Flight Laboratory, NORSAT-1 System Requirements and Verification Matrix v1.0, Toronto: Space Flight Laboratory, 2012.

[10] Antrix Corporation Limited, PSLV User Manual Issue 5, Bangalore, Karnataka, 2007.

[11] Arianespace, Soyuz User's Manual Issue 2 Revision 0, Courcouronnes, 2012.

[12] Arianespace, Ariane 5 User's Manual Issue 4 Revision 0, Courcouronnes: Arianespace, 2004.

[13] S. Mauthe, XPOD Duo Qualificaiton Vibraiton Test Plan, Toronto, Ontario, 2012.

[14] Spaenaur, Spaenaur Catalog 2013, Ontario: Spaenaur, 2013.

[15] R. G. Budynas and J. K. Nisbett, Shigley's Mechanical Engineering Design 9th Edition, New York: McGraw-Hill, 2011.

[16] California Polytechnic State University, CubeSat Design Specification Rev. 12, 2009.

[17] J. Fine, XPOD Tip Off Rate Analysis, Toronto: Space Flight Laboratory, 2014.

[18] B. P. Larouche, Design, Simulation, and Testing of the Structural & Separation System, Toronto: University of Toronto, 2008.

[19] M. Ligori, XPOD H27 Nanosatellite Separation System, Toronto: Space Flight Laboratory, 2011.

[20] B. Brakeboer, GHGSat-D PDR Mechanical Design Document, Toronto: Space Flight Laboratory, 2013.

[21] Siemens, NX 8.0 Help Documentation, 2011.

[22] Hexcel Composites, HexWeb Honeycomb Attributes and Properties, Hexcel, 1999.

[23] MatWeb, "Aluminum 7075-T6; 7075-T651," [Online]. Available: http://matweb.com/search/DataSheet.aspx?MatGUID=4f19a42be94546b686bbf43f79c51b7d&ckck=1. [Accessed 19 February 2014].

[24] J. Fine, "XPOD Inspection Guide for Vibration Testing," Space Flight Laboratory, Toronto, 2013.

[25] M. Ligori, D. Kekez, S. Mauthe and F. Pranajaya, "XPOD GNB Users Guide v 1.4," Space Flight Laboratory, Toronto, 2013.

[26] S. Mauthe, NEMO-HD Mechanical Design Document, Toronto, Ontario, 2013.