112

Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

  • Upload
    vandien

  • View
    214

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015
Page 2: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

ii

Shahil Kanji

Master of Applied Science

Graduate Department of Aerospace Science and Engineering

University of Toronto

2015

NORSAT-1 is a multi-payload microsatellite mission funded by the Norwegian Space Center,

with three overall objectives: investigating solar radiation, space plasma research, and

developing improved methods for detection and management of ship traffic. The successful

development of the NORSAT-1 platform aims to lay the groundwork for additional low-cost

microsatellites in the NORSAT series, and expand the Norwegian presence in space and space-

based ship tracking technologies.

This thesis provides some insight into the NORSAT-1 satellite platform design, and focuses

heavily on the mechanical aspects of design, analysis, and testing. The structural design is

detailed from the early conceptual design phases, and follows the development to the

manufacturing, integration, and testing of the flight spacecraft. Validation of the design through

finite element modeling is presented, along with the development and design of two honeycomb

composite solar panels, and two deployable whip antennas.

Page 3: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

iii

First thanks goes out to my supervisor, Dr. Robert E. Zee, for giving me the opportunity to

pursue a MASc. degree at the Space Flight Lab (SFL). Not really knowing what I was getting

into, my position at SFL turned out to be exactly what I was looking for. It has truly been a

rewarding and challenging experience working here, with amazingly knowledgeable people and

on amazingly awesome technology.

Special thanks to my manager Alex Beattie, who never ceased to entrust me with large project

responsibilities, allowing me to truly learn, grow, and develop my engineering skills, and get the

most out of my masters degree. Thanks to all of the staff and students at SFL for providing such

a welcoming and supportive environment, and especially to the many mechanical apt persons,

Cordell Grant, Benoit Larouche, Dumitru Diaconu, Ben Risi, Mike Ligori, and Brent Brakeboer

for providing much guidance and wisdom to many of the engineering problems with which I was

faced. Thanks to the rest of my fellow students at SFL and UTIAS as well for rendering these

past two years manageable, in particular the students of my year.

A small thanks goes out to my roommate at the time of writing this thesis who graciously offered

to proofread this 100 page monster, and then only got halfway. Thanks also to all of my close

friends for dealing with my constant bailing of weekend plans over the summer while I was

writing this. Final thanks goes out to my family, in particular my parents, who have supported

me my entire life, giving me the freedom and encouragement to pursue what makes me happy.

And thanks to my siblings for always pushing me to be better than the curve, and become

something great.

As the saying goes, enjoy the following.

Page 4: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

iv

Acknowledgments ........................................................................................................................ iii Table of Contents ......................................................................................................................... iv List of Tables ................................................................................................................................ vi List of Figures .............................................................................................................................. vii

Introduction .............................................................................................................................. 1 11.1 Space Flight Laboratory ...................................................................................................... 1 1.2 The NORSAT-1 Microsatellite ........................................................................................... 2

1.3 Thesis Objectives and Outline ............................................................................................ 3

The NORSAT-1 Microsatellite................................................................................................ 5 22.1 Mission Overview ............................................................................................................... 5

2.2 Capabilities ......................................................................................................................... 5

2.2.1 Structure .................................................................................................................. 6

2.2.2 Telemetry and Command ........................................................................................ 6 2.2.3 Thermal ................................................................................................................... 6

2.2.4 Attitude and Control ............................................................................................... 7 2.2.5 Command and Data Handling ................................................................................. 7

2.2.6 Power ...................................................................................................................... 7

2.3 Payloads .............................................................................................................................. 8

2.3.1 Compact Lightweight Absolute Radiometer .......................................................... 9

2.3.2 Langmuir Probes ................................................................................................... 11

2.3.3 AIS Receiver ......................................................................................................... 13

Structural Subsystem Design ................................................................................................ 16 33.1 Driving Requirements ....................................................................................................... 16

3.2 Design Concept ................................................................................................................. 19

3.2.1 Initial Project Status .............................................................................................. 19 3.2.2 Current Design and Layout ................................................................................... 20

3.2.3 Electromagnetic Interference Mitigation .............................................................. 27 3.3 Design for Thermal ........................................................................................................... 29

3.3.1 Solar Panel Wings ................................................................................................. 30 3.3.2 Solar Cell Isolation ............................................................................................... 30

3.3.3 Thermal Surface Additions ................................................................................... 31

3.4 Payload Accommodations ................................................................................................ 33 3.4.1 CLARA ................................................................................................................. 33

3.4.2 AIS Receiver ......................................................................................................... 34

3.4.3 Langmuir Probes ................................................................................................... 35

3.5 Design for Assembly and Disassembly ............................................................................ 35 3.6 Wiring Harness Development ........................................................................................... 37

3.6.1 Solid Model Wiring .............................................................................................. 39

3.7 Materials Selection ............................................................................................................ 41 3.8 Mass Budget ...................................................................................................................... 42

3.9 Solid Modeling .................................................................................................................. 42 3.10 Design Evolution ............................................................................................................. 43

3.11 Manufacturing ................................................................................................................. 44

Finite Element Analysis ......................................................................................................... 45 44.1 Modeling ........................................................................................................................... 45

4.1.1 Primary Structure .................................................................................................. 45

Page 5: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

v

4.1.2 Solar Panel Wings ................................................................................................. 46 4.1.3 Components and Connections ............................................................................... 47

4.2 Boundary conditions ......................................................................................................... 48

4.2.1 Constraints ............................................................................................................ 48

4.2.2 Loads ..................................................................................................................... 48 4.3 Results ............................................................................................................................... 49

4.3.1 Stress Analysis ...................................................................................................... 49

4.3.2 Displacement Analysis .......................................................................................... 51 4.3.3 Modal Analysis ..................................................................................................... 52

Honeycomb Solar Panel Wings ............................................................................................. 54 55.1 Requirements .................................................................................................................... 55

5.2 Proposed Design Concept ................................................................................................. 57

5.3 Failure Modes ................................................................................................................... 58

5.3.1 Sandwich Panel Failure ......................................................................................... 58 5.3.2 Insert Failure ......................................................................................................... 61

5.4 Honeycomb Composite Panel Design .............................................................................. 65

5.4.1 Panel Procurement ................................................................................................ 67

Deployable Components ........................................................................................................ 68 66.1 VHF Antennas .................................................................................................................. 68

6.1.1 Requirements ........................................................................................................ 69

6.1.2 Research ................................................................................................................ 69

6.1.3 Design ................................................................................................................... 73 6.2 Langmuir Probes ............................................................................................................... 78

Ground Support Equipment ................................................................................................. 79 77.1 Assembly and Handling GSE ........................................................................................... 79

7.2 Protective Enclosure ......................................................................................................... 82 7.3 Mock-Up Wings ................................................................................................................ 83

7.4 Radio Frequency Testing GSE .......................................................................................... 84 7.5 Deployment and XPOD Loading GSE ............................................................................. 86

Integration and Testing ......................................................................................................... 88 88.1 Structural Fit Checks and Preparation .............................................................................. 88 8.2 Wiring Harness Fit Check ................................................................................................. 89

8.3 Dirty-Sat Integration and Testing ..................................................................................... 90 8.4 Dirty-Sat EMC Testing ..................................................................................................... 92

8.5 Antenna Pattern Testing .................................................................................................... 93 8.5.1 Uplink/Downlink and GPS Antennas ................................................................... 93

8.5.2 VHF Antennas ...................................................................................................... 94

8.6 Deployment Testing .......................................................................................................... 94 8.6.1 Results ................................................................................................................... 95

8.7 Structural Bake-Out .......................................................................................................... 97 8.8 Flight Integration .............................................................................................................. 99

Conclusions ........................................................................................................................... 100 9References .................................................................................................................................. 101 Appendix A ................................................................................................................................ 103

Page 6: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

vi

Table 1: SFL satellite platform specifications [1] ........................................................................... 2 Table 2: Driving requirements that affect the structural design [8] [9] ........................................ 16 Table 3: Component layout constraints for NORSAT-1 .............................................................. 22 Table 4: Stress and displacement analysis results summary ......................................................... 51 Table 5: NORSAT-1 honeycomb panel design failure modes summary ..................................... 66

Table 6: List of main manufacturing defects found through inspection and fit checks ................ 88

Page 7: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

vii

Figure 1: The NORSAT-1 mission patch [1] .................................................................................. 3 Figure 2: NORSAT-1 microsatellite with overall dimensions ....................................................... 5 Figure 3: CLARA payload mechanical design ............................................................................. 11 Figure 4: Langmuir Probe payload mechanical design ................................................................. 13 Figure 5: Norwegian coastal regions [6] ....................................................................................... 14

Figure 6: AIS Receiver payload .................................................................................................... 15 Figure 7: XPOD-Duo deployment system, Vertical mounted (A), Horizontal (B) ...................... 18 Figure 8: GNB bus AISSat-3 (left), GHGSat-D (middle), and NEMO-AM (right) ..................... 19 Figure 9: Initial NORSAT-1 structural design proposal (Scott Armitage) ................................... 20 Figure 10: Exploded view of NORSAT-1 primary structure ........................................................ 21

Figure 11: NORSAT-1 external component layout ...................................................................... 23 Figure 12: NORSAT-1 internal component layout ....................................................................... 24

Figure 13: Reaction wheel sub-assembly, CAD model (left), clean room assembly (right) ........ 24

Figure 14: NORSAT-1 panel component layouts (front/back) ..................................................... 25 Figure 15: NORSAT-1 battery pack design, Exploded (left), Assembled (right) ........................ 27 Figure 16: Division of avionics and payloads in NORSAT-1 ...................................................... 28

Figure 17: Separation plate sub-assembly design and layout ....................................................... 29 Figure 18: Worst-case hot attitudes (as viewed from the sun) ...................................................... 31

Figure 19: Worst-case cold attitudes (as viewed from the sun) .................................................... 32 Figure 20: +Y worst-case cold dimension increase due to thermal surfaces ................................ 33 Figure 21: +Z and –Z thermal surface additions ........................................................................... 33

Figure 22: CLARA integration sequence ..................................................................................... 34

Figure 23: AIS Receiver (left) and AIS Antenna (Right) accommodations ................................. 35 Figure 24: Langmuir Probe electronics (left) and Langmuir Probe cassette (Right)

accommodations ........................................................................................................................... 35

Figure 25: Wiring harness manufacturing drawing example, Payload cable ............................... 39 Figure 26: NORSAT-1 solid model wiring .................................................................................. 40

Figure 27 Panel wiring for +Y panel (left) and +X panel (right).................................................. 41 Figure 28: NORSAT-1 design evolution ...................................................................................... 43

Figure 29: Majority of the NORSAT-1 flight structural parts ...................................................... 44 Figure 30: Large attachment bracket part (left), idealized part (middle), meshed (right) ............ 46 Figure 31: FEM fastener modeling ............................................................................................... 47 Figure 32: NORSAT-1 finite element model with applied boundary conditions ......................... 48

Figure 33: Screen capture of –Z tray FEM stress results .............................................................. 50 Figure 34: Screen capture of simulation showing first mode of NORSAT-1 at 144Hz, and

subsequent mode frequency values ............................................................................................... 53

Figure 35: Relative stiffness and weight of sandwich panels compared to solid panels. Note that

the numbers shown are normalized to the solid material numbers [13] ....................................... 54 Figure 36: NEMO-AM (A) and NEMO-HD (B) honeycomb panels ........................................... 55 Figure 37: NORSAT-1 solar array concept design ....................................................................... 57 Figure 38: Wing attachment bracket design ................................................................................. 58

Figure 39: Sandwich panel failure modes [15] ............................................................................. 59 Figure 40: Simplified loading scenario for sandwich panel failure mode calculations [14] ........ 61 Figure 41: Insert loading scenarios [18] ....................................................................................... 62

Figure 42: Partially potted insert under tensile load [18] ............................................................. 63

Page 8: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

viii

Figure 43: Face sheet in-plane insert failure modes [18] .............................................................. 64 Figure 44: NORSAT-1 solar panel wing design ........................................................................... 65 Figure 45: AISSat-2 pre-deployed VHF antenna approximate dimensions ................................. 68 Figure 46: Tape spring bend radius .............................................................................................. 70

Figure 47: Tape spring materials explored ................................................................................... 71 Figure 48: Hold down configurations for two tape spring antennas ............................................. 72 Figure 49: NORSAT-1 antenna base exploded view .................................................................... 74 Figure 50: NORSAT-1 in the vertically mounted XPOD-Duo .................................................... 74 Figure 51: VHF antenna stowage system ..................................................................................... 76

Figure 52: NORSAT-1 side view of stowed AIS antennas on spare structure ............................. 76 Figure 53: Expected deployment volume of VHF antennas ......................................................... 77 Figure 54: Langmuir Probe cassette deployment .......................................................................... 78

Figure 55: MSGE assembly tray “legs” in various orientations of use (black) ............................ 80 Figure 56: NORSAT-1 GSE support stand ................................................................................... 81 Figure 57: NORSAT-1 GSE handle assembly .............................................................................. 82

Figure 58: NORSAT-1 protective enclosure design ..................................................................... 83 Figure 59: NORSAT-1 mock-up Wings (left), mock-up wings fitted on structure (right) ........... 84

Figure 60: RF testing GSE blocks ................................................................................................ 86 Figure 61: NORSAT-1 deployment jig design details .................................................................. 87 Figure 62: XPOD-Duo loading (XPOD-Duo GSE designed by Mike Ligori) ............................. 87

Figure 63: NORSAT-1 structural fit checks, CLARA (engineering model) installed (left), GSE

enclosure installed (right) ............................................................................................................. 89

Figure 64: Freshly built (left) and untangled (right) flight Main wiring harness ......................... 90 Figure 65: Payload wiring harness, 3D model (left), fit check in structure (middle), and untangled

flight harness (right) ...................................................................................................................... 90 Figure 66: NORSAT-1 Dirty-Sat integration ............................................................................... 91

Figure 67: NORSAT-1 EMC testing in SFL’s anechoic chamber ............................................... 93 Figure 68: Antenna pattern test setup ........................................................................................... 94 Figure 69: Deployment testing test setup ...................................................................................... 95

Figure 70: Deployment test still shot, just before antennas deploy .............................................. 96 Figure 71: Deployment with mounting plate, no contact ............................................................. 96

Figure 72: Deployment with plate, contact made ......................................................................... 97 Figure 73: NORSAT-1 flight structure bake-out setup ................................................................. 98

Figure 74: NORSAT-1 flight integration progress ....................................................................... 99 Figure 75: NORSAT-1 flight battery pack ................................................................................... 99 Figure 76: Insert strength as a function of core height for a 11mm diameter insert in a 3/16-5052-

.001 honeycomb core with aluminum face sheets under tension [19] ........................................ 103 Figure 77: Insert strength as a function of core height for a 11mm diameter insert in a 3/16-5052-

.001 honeycomb core with aluminum face sheets under compression [19] ............................... 104

Page 9: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

1

1

Microsatellites are paving the way for a new era in space technologies. Their low cost and quick

development time attracts a large new customer base, offering easier access to space. This easier

access to space promotes innovation and discovery at far greater rates than previously observed.

The Norwegian Space Center is taking full advantage of this opportunity, funding over five

microsatellites in the past ten years. While first limiting the funded satellites to maritime traffic

monitoring such as AISSat-1, AISSat-2, and AISSat-3, they now look to further leverage the

capabilities of small satellites through the NORSAT-1 mission, as part of the national space

program. NORSAT-1 is a multi-payload mission, and has three objectives: Investigating solar

radiation, space plasma research, and developing improved methods for detection and

management of ship traffic. The successful development of the NORSAT-1 platform aims to lay

the groundwork for additional microsatellites in the NORSAT series, and expand the Norwegian

presence in space and space-based ship tracking technologies. This thesis provides some insight

into the NORSAT-1 platform design, focusing heavily on the mechanical aspects of design,

analysis and testing.

1.1

The Space Flight Laboratory (SFL) is a company affiliated with the University of Toronto

Institute for Aerospace Studies (UTIAS) established in 1998, with one clear objective in mind: to

render space accessible to companies, government organizations, and end users like never

before. By lowering the high cost barrier to space, they can make way for far more opportunities,

with potential for the next generation of human presence in space. To date, SFL has developed

numerous low cost Nanosatellite (spacecraft mass <10kg) and Microsatellite (spacecraft mass

<100kg) platforms through the unique approach of training graduate students on real-customer

projects. With a “Micro-space” philosophy in mind, SFL leverages the latest technology

advances, off the shelf components, and in-house innovations in order to develop highly capable

spacecraft from conceptualization to launch with drastically reduced development time and cost

of typical spacecraft. With more than a dozen spacecraft currently in low Earth orbit (LEO), and

another nine under development or waiting for launch at the time of writing, SFL is a clear leader

Page 10: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

2

in Canada for its spacecraft capabilities and is pushing the boundaries on LEO space access. The

current director is Dr. Robert E. Zee, and at the time of writing, the company consists of

approximately 35 full-time employees, and 12 full-time graduate students. Table 1 below

outlines the specifications of some of the current SFL spacecraft platforms, and some of the

projects on-orbit and under-development.

Table 1: SFL satellite platform specifications [1]

Typical Specifications – Customization Possible – Specifications Subject to Change

CanX-2 NTS GNB NEMO NEMO-150

Spacecraft Mass 3.5 kg 6.5 kg 7 kg 15 kg up to 150kg

Spacecraft

Volume

10 x 10 x

34cm

20 x 20 x 20

cm

20 x 20 x 20

cm

20 x 20 x 40

cm

60 x 60 x 60cm

Peak Power

25ºC,BOL

2-7 W 4 -7 W 7 - 9 W 50 W up to 500W

Payload Mass 1 kg 2 kg 2 kg 6 kg up to 70 kg

Payload Volume 1,000 cm3 1,700 cm

3 1,700 cm

3 8,000 cm

3 up to 108,000

cm3

Payload Power

@ duty cycle

1-2 W

@100%

2 W @20-

30%

3 - 4 W

@100%, 6 W

max

45 W @ 40%

min, 65 W

max

50W or higher

ACS stability ~2° (1)

~5-10° ~2° (2)

~60" (3)

~2° (2)

~60" (3)

~2° (2)

~10-

20" (3)

Downlink 32 k - 1

Mbps

32 k - 1 Mbps 32 k - 2 Mbps 32 k - 2 Mbps 32k - 50Mbps

Examples CanX-2,

CanX-7

NTS AISSat-1, 2,

3, BRITE

Constellation,

EV9, CanX-

4&5

NEMO-AM,

GHGSat-D,

NORSAT-1

NEMO-HD

1. With magnetometer, sun sensor and one reaction wheel. 2. With magnetometer, fine sun sensor and

three reaction wheels. 3. With star-tracker and three reaction wheels.

1.2

NORSAT-1 is a multi-payload satellite under development at the University of Toronto Institute

for Aerospace Studies – Space Flight Laboratory (UTIAS-SFL) for the Norwegian Space Center

(NSC). The satellite will investigate solar radiation, space weather, and detect ship traffic by

means of three separate payloads: a Compact Lightweight Absolute Radiometer (CLARA), four

Langmuir Probes, and an Automatic Identification System (AIS) receiver, respectively. SFL has

been contracted to design the spacecraft platform to house these payloads into low Earth orbit.

NORSAT-1 is one of SFL’s third generation satellite platforms, the Next-Generation Earth

Monitoring and Observation (NEMO) class, which leverages the experience gained through the

Page 11: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

3

successful development of the Generic Nanosatellite Bus (GNB). It is currently under

development alongside two similar sized NEMO class missions at SFL, GHGSat-D and NEMO-

AM, who both aim to monitor levels of greenhouse gas in the Earth’s atmosphere.

Although NORSAT-1 is not NSC’s first satellite endeavor, nor is it SFL’s first time

collaborating with NSC to develop a small spacecraft (Ex. AISSat-1, 2 and 3), it does represent

Norway’s first satellite project with scientific purpose. The official mission patch for the satellite

is shown below in Figure 1, listing the main partners involved with the project.

Figure 1: The NORSAT-1 mission patch [1]

1.3

With the NORSAT-1 mission, SFL has the opportunity to extend its knowledge and experience

to designing a slightly larger spacecraft, accommodating multiple payloads. This thesis follows

the accomplishments of the author while working at SFL towards all mechanical aspect of

design, analysis and testing for the structural subsystem of NORSAT-1. Given that the

NORSAT-1 development was proposed as a two-year contract from kick-off in July 2013, this

thesis will follow the majority of the development cycle of the spacecraft. The structural design

is outlined from the mere early stages of its proposal concept, through the various design

iterations past the Preliminary Design Review (PDR), and Critical Design Review (CDR), until

the final (current) design. This design is currently undergoing flight level assembly, integration

and system level testing. As the main structural engineer for the project, the author was in charge of

Page 12: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

4

the full structural design of the spacecraft, designing the mechanical interfaces between everything

inside and on the exterior of the spacecraft, as well as the assembly, integration, and mechanical

aspects of testing on the fully built system level spacecraft. The contributions presented are vital to

the success of the NORSAT-1 mission, given the dependence of all included spacecraft components

on the ability of the spacecraft to maintain structural integrity.

The main objective of this thesis is to design the structural subsystem for the NORSAT-1

mission such that it properly accommodates all satellite subsystems, including the payloads, and

provides a safe and secure housing through all mission environments. The success of the design

will be based on its ability to satisfy the mission, system, and structural subsystem requirements.

A secondary thesis objective is to detail the novel aspects of mechanical design for NORSAT-1

to serve as a design reference for future microsatellites having similar requirements.

While the current chapter outlines a general introduction and the necessary background

information for the work to be presented, Chapter 2 gives a closer look at the NORSAT-1

mission, overviewing the three different payloads on-board, as well as the overall capabilities of

the spacecraft itself. Chapter 3 describes the author’s work on the structural design for

NORSAT-1, describing how the design was inspired through the need to meet several necessary

requirements as well as adhere to a low cost and stringent timeline. Chapter 4 illustrates how the

structural design was validated analytically through the use of finite element analysis (FEA)

software, in order to ensure the design would stand up to the harsh environments presented

during a rocket launch. Chapter 5 details the design of NORSAT-1’s large solar array, through

the use of two honeycomb composite sandwich panels. Chapter 6 presents the design and testing

of two deployable antennas aboard the spacecraft to receive very high frequency (VHF) signals

for the AIS receiver payload. Chapter 7 then outlines the mechanical ground support equipment

(MGSE) that was designed in order to facilitate the assembly, integration, and testing (AIT) of

the spacecraft, and Chapter 8 depicts the integration and testing of the spacecraft that the author

was directly involved with, including the dirty and clean assembly, antenna pattern testing, and

deployment testing. Lastly, Chapter 9 presents concluding remarks on the experience through the

development of the structural subsystem for NORSAT-1, and outlines the current status of the

project.

Page 13: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

5

2

This section provides a closer look at the NORSAT-1 mission, detailing the three payloads on

board, as well as the spacecraft capabilities. Figure 2 displays an external view of NORSAT-1

with approximate outer dimensions of the satellite bus.

Figure 2: NORSAT-1 microsatellite with overall dimensions

2.1

The NORSAT-1 satellite is currently slated to launch in the first quarter of 2016 into a dawn-

dusk orbit. It is required that the spacecraft remain fully functional for at least one year in orbit,

however, the design goal reaches for at least three years. The mission will allow for simultaneous

operation of all three payloads on-board, and will be operated from a ground station in Norway.

The performance of the satellite is designed to meet the specific requirements of the on-board

payloads set forth by each of the payload providers, as well as the Norwegian Space Center.

2.2

The NORSAT-1 spacecraft is comprised of numerous subsystems that are all vital to the

spacecraft performance. Each of these subsystems for the NORSAT-1 mission is described in

brief below, in order to provide a full overview of the spacecraft design and capabilities.

Page 14: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

6

2.2.1

The structural subsystem comprises of all of the physical components needed to properly house

the satellite avionics, payloads, and any other necessary equipment, and keep them safe through

all expected environments. It is comprised of a heritage design concept from the GNB, utilizing

two intricately designed loadbearing aluminum trays, housing much of the avionics, enclosed by

six aluminum body panels. The structure of the satellite is the main topic of this thesis and is

more thoroughly detailed in Chapter 3.

The total estimated mass of the satellite (at the time of writing) is approximately 16kg. When

loaded inside its separation system, it has a total launch mass under 30kg, and an approximate

launch volume of 300mm x 200mm x 500mm.

2.2.2

The Telemetry and Command subsystem on NORSAT-1 provides a full-duplex, bi-directional

radio communications system between the satellite and the Earth station. It incorporates a two-

radio system: a UHF receiver for uplink communications, and an S-Band transmitter for

downlink. The uplink UHF receiver is supplemented with a cavity band-pass filter to provide

rejection of the spacecraft’s transmitter emissions, a down-converter that provides frequency

translation between S-band and ultra-high frequencies (UHF), and a UHF receiver that includes

demodulation, bit synchronization, and descrambling functionality [2].

Both of these radios are heritage designs used on previous SFL missions, and utilize a dual-patch

antenna system, with the antennas for each link located on opposite sides of the spacecraft; a

total of four S-Band patch antennas are used for this, and provides close-to omnidirectional

coverage.

2.2.3

A fully passive thermal design for NORSAT-1 has been designed by the thermal engineer for the

majority of the considered orbits. The techniques used involve controlling the overall bus

temperature using various thermal control tapes on the outside of the spacecraft, as well as

controlling the internal component temperatures through materials selection and specific

mounting methods. This type of implementation allows for a low-risk and robust satellite thermal

design. In addition to these passive techniques used, an active heater is present in the battery

pack in order to ensure the battery cells are kept above their minimum charging temperature

(0°C). The thermal design will be updated once the orbit of NORSAT-1 is confirmed.

Page 15: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

7

2.2.4

Attitude and control of the satellite, while in orbit, is achieved through various spacecraft

mounted sensors and attitude hardware. Six SFL-designed sun sensors are present on the

spacecraft pointing in the six principle directions for attitude determination, along with a three-

axis magnetometer to determine the local magnetic field. A three-axis rate-sensor is also

incorporated to provide additional attitude information when in eclipse. Three-axis control of the

satellite is achieved using three orthogonally mounted reaction wheels and magnetorquers. A

dedicated on-board computer is present on the satellite to operate the necessary attitude

algorithms. The satellite with the above mentioned attitude hardware, is capable of achieving

pointing within +/-5°. However, with the addition of a precision sun sensor mounted directly on

the CLARA instrument, NORSAT-1 can achieve a fine pointing mode, enabling pointing within

+/- 0.5 degrees (mean plus 3-sigma) while the sun is visible [2]. Additionally, a GPS receiver is

included to provide positioning and timing data, as well as support the payload activities.

2.2.5

Three identical SFL designed on-board computers comprise the Command and Data Handling

(C&DH) subsystem on NORSAT-1. These on-board computers are heritage designs from

previous SFL missions, and represent highly mature hardware and associated software for

controlling all of the spacecraft functions, as well as communications.

The Housekeeping Computer (HKC) is typically dedicated to performing housekeeping tasks on

the spacecraft, such as collecting regular telemetry from each component. The Attitude

Determination and Control Computer (ADCC) is typically dedicated to performing attitude and

determination related processing, such as reading the necessary attitude sensors and issuing

commands to the attitude actuators. The third computer on board is the Payload On-Board

Computer (POBC), dedicated to interfacing with the payloads; the Serial Interface Board (SIB) is

also considered a portion of this computer.

2.2.6

The power subsystem on NORSAT-1 has three main functions: power generation, energy

storage, and power distribution.

Power generation is done by means of externally mounted triple junction solar cells, grouped in

eight-cell strings, with a beginning of life efficiency of roughly 28%. The solar cells are mounted

on the satellite such that at least one string is visible in each of the six primary directions, to

Page 16: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

8

allow for safe-hold power generation in any state. The main sunward facing side of the satellite

however has a total of six strings of solar cells for additional power generation during payload

activities. A total of 96 solar cells exist on NORSAT-1, capable of over 40W of power

generation.

Energy storage on NORSAT-1 is done by means of a three-series, two-parallel lithium-ion

battery pack, with an integrated Battery Interface Module (BIM). The BIM provides battery cell

protection, and provides the battery telemetry that operates the battery heater.

Power distribution on the spacecraft is achieved through an in-house developed Modular Power

System (MPS). It is comprised of a passive backplane, on which multiple Micro Switched Power

Node (uSPN) cards are connected - each load on the spacecraft is connected to an individual

switch. A 5V supply, Solar Array/Battery Regulator (SABR), Solar Array/Bus Interface Node

(SABIN), and Interface Node (IFN) card are also mounted onto this backplane to provide the

remaining functionalities to the power system.

2.3

NORSAT-1 is a collaborative mission with three separate payloads each being designed by

different companies in different countries or cities; therefore effective communication between

the spacecraft and payload design teams is essential. The use of Interface Control Documents

(ICDs) is crucial to freeze the interrelated parts of the payload and spacecraft design early on to

allow for a fully parallel design path. These documents specify and manage every interface

between the payload and the spacecraft, including the mechanical mounting interface, volume

and mass allotments. Efforts are made not to deviate from these documents in order to minimize

conflicts in schedule due to design changes. Once these documents are finalized late in the

design phase, parts can begin to be manufactured. By proceeding with the payload and spacecraft

design concurrently, the overall platform development time is significantly reduced. In this type

of approach, the spacecraft design team must be fairly involved with each of the payload designs,

in order to ensure efficient and compatible designs. Each payload has their own unique design

challenges associated with them, which often times carry over to challenges within the spacecraft

design. In the following sub-sections, the design, motivation, and goals of each of the three

payloads will be discussed, along with the key challenges that affect the spacecraft design.

Page 17: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

9

2.3.1

The Compact Lightweight Absolute Radiometer (CLARA) instrument is the primary payload on

the NORSAT-1 mission. The CLARA payload is being designed by Physikalisch-

Meteorologisches Observatorium Davos / World Radiation Center (PMOD/WRC) in

Switzerland, and its purpose is to measure Total Solar Irradiance (TSI) with high precision and

low noise. It is a compact radiometer consisting of three digitally heat powered regulated

cavities. Each aperture gets aligned directly at the sun in order to measure TSI when the

according shutter is open. Total Solar Irradiance is one of the fundamental parameters in climate

research, and operational TSI monitors in space are crucial for climate forecast and

reconstruction [3]. The CLARA payload is being designed as compact and lightweight as

possible in order to maximize its flight opportunity on a multi-payload satellite such as

NORSAT-1. CLARA has four main science objectives that are briefly explained below:

Absolute Radiometry Validation

The CLARA payload will allow for validation of laboratory results in space that provide an

explanation for some discrepancies measured by PMO6-, DIARAD-, and ACRIM-type

radiometers compared to the American TIM/SORCE experiment [4].

Space Weather

Through successfully modeling the TSI variations, correlations to the Ultra Violet (UV) radiation

variations can also be made [2]. Thus, the long-term stability of UV variations can also be

assessed.

Climate Research

Large amounts of evidence suggest that Total Solar Irradiance (TSI) has an influence on the

Earth’s climate [5]. The CLARA payload aims to extend the TSI data record for solar

atmosphere and climate modelers through monitoring the TSI variations with great accuracy and

sensitivity. Continuous monitoring of the TSI levels is needed in order to reduce uncertainties,

and cover the 11-year solar cycle. The launch of NORSAT-1 provides a suitable timeline for

avoiding any possible gaps in data due to current missions nearing an end.

Helioseismology

The CLARA payload will be the highest-cadence radiometer in space to-date, allowing the

assessment of the TSI variability at very high frequencies. The higher frequency of

Page 18: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

10

measurements allows for helioseismology investigations of the solar atmosphere in order to

assess the acoustic energy carried into the solar atmosphere by high frequency sound waves [6].

2.3.1.1

The measurement principle of CLARA uses black body cavities to absorb the incident solar

energy through a precise aperture, which can be closed with a shutter. The absolute irradiance

can then be measured by referencing the incident irradiance of an open cavity to the measured

irradiance of a closed cavity. This measurement technique is thermal based, and relies directly on

knowing, to a great degree of certainty, the thermal resistance of a fragile flexure structure in

each of the black body cavities. As such, all thermal variability in and around these sensitive

components will directly distort the results. Rather than trying to predict accurately the

(constantly changing) radiated and conductive thermal paths from each of the cavities in the

spacecraft on orbit, a seemingly simpler approach was taken, whereby the cavities and sensitive

components were physically separated as much as possible from the rest of the payload, in

efforts to thermally isolate them from the thermal environment. This approach led the payload

provider to the current mechanical design, as described in the following sub-section.

In addition to this thermal design challenge, the CLARA payload also requires a high level of

pointing accuracy during measurements while pointing at the sun. In order to meet the required

pointing accuracy (mean plus 3-sigma, of +/- 0.5 degrees), an additional precision sun sensor

with adequate space heritage is incorporated into the spacecraft design.

A third design challenge of the CLARA payload is the required level of cleanliness, higher than

would normally be required for even an optical telescope mission. The two areas of concern are

particle contamination inside the apertures, and hydrocarbons settling on the external thermal

control surfaces. The former is primarily a ground handling concern, whereas the latter is

primarily a spacecraft outgassing concern. To cope with this, a special handling and cleanliness

protocol has been implemented for both the payload, and the entire spacecraft. Part of the

protocol involves baking out all spacecraft components, including the other two payloads, as

well as the full spacecraft after integration, in order to reduce some of the risk of outgassing

materials later on orbit. The instrument will also have the means to be constantly purged with

nitrogen, and include a protective cover to minimize contamination.

Page 19: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

11

2.3.1.2

The CLARA payload is separated into two separate aluminum enclosures, connected together via

four large titanium studs, and a couple of wires to transfer the measurement data. The thermally

isolated section contains the sensitive measurement components, and is wrapped with a multi-

layer insulation (MLI) blanket to further the thermal isolation.

In order to minimize any mechanical and thermal pointing misalignments between the sun sensor

and the CLARA apertures, the precision sun sensor is directly mounted on the payload via a

dedicated bracket, extending from the rear, “less thermally sensitive”, half. The solid model of

the CLARA payload is detailed below in Figure 3.

Figure 3: CLARA payload mechanical design

2.3.2

Plasmas are by far the most common phase of ordinary matter in the universe, and in our solar

system, interplanetary space is filled with the plasma of the solar wind that extends from the Sun

out to the heliopause [7]. The main purpose of the Langmuir Probe payload is to measure

electron plasma from the sun, detectable from low Earth orbit (LEO), in an effort to study and

define the plasma parameters. In LEO, at an altitude of a few hundred kilometers, the spacecraft

will primarily be submersed in the dense plasma known as the ionosphere, which is produced

from the ultraviolet radiation from the sun [7]. The instrument is being developed by the

University of Oslo (UiO), and consists of four individual probes each mounted at the end of a

Page 20: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

12

boom (four booms total), and can assess several defining features of the plasma, such as electron

temperature, electron density, and electric potential. The system is a new concept Langmuir

Probe system capable of high-resolution measurements of space plasma density, and can cover

the density range from – . The system consists of two or more cylindrical

needle probes, and is therefore referred to as UiO’s multi-Needle Langmuir Probe system (m-

NLP). A key feature of the m-NLP technique is the ability to determine the electron density

without the need to know the spacecraft potential and electron temperature.

2.3.2.1

One of the main design challenges of the Langmuir Probe system aboard NORSAT-1 is getting

the probe tips in an area of undisturbed space plasma. As the spacecraft is in orbit, it creates a

‘plasma wake’ in the opposite direction of travel. Because this “plasma wake” is poorly

understood and hard to predict, the probe tips are placed on long protruding booms, in an effort

to place the tips as far out into the undisturbed plasma as possible. In doing this, the booms

become quite long, quickly increase in complexity, and increase the launch volume of the

spacecraft quite significantly. To minimize this, the probes must be made deployable. Having to

have a deployment system carries its own set of challenges and requirements as well, such as the

necessity of ground handling equipment and testing methods.

A second design challenge that the Langmuir Probe payload carries is the need to have sufficient

conductive surface area (coupled to the spacecraft chassis reference ground) available on each

side of the spacecraft in order to “close” the measurement circuit. As the probe tips collect high-

mobility electrons from the space plasma, the spacecraft will inevitably charge up, and could

potentially begin to repel incoming electrons if a sufficient charge is achieved. The conductive

surface area on the spacecraft sides provide a path for the less mobile ions to hit the spacecraft in

the direction of velocity, in order to offset the spacecraft charge build-up. The required surface

area is made as a spacecraft requirement, and is simply verified by inspection.

A third design challenge is that the probe tips must be free of contaminations that could impact

the measurements. For example, finger oils can create an insulating layer on the probe, and

would reduce electrons flow through the probes and affect the resulting measurements. A reliable

mitigation plan for this concern is thus required for the design, such as protective covers,

replacement tips, or repeated cleaning.

Page 21: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

13

2.3.2.2

The Langmuir Probe tips are placed on large deployable booms, whose overall size is limited to

the length of the spacecraft itself. Two identical Langmuir Probe Cassettes are included on

NORSAT-1, each housing two Langmuir Probes. The booms are held down by a uniquely

designed mechanism that, through spring preloads, forces the booms to stay stowed. Using a

commercial shape memory alloy pin-puller, the pre-loaded spring mechanism can be released on

orbit, allowing it to perform a half-turn, and consequently push both booms out with a large

enough force to reach their fully deployed positions. Once fully deployed, each of the booms is

able to lock in place via a locking pin. This cassette design is shown below in Figure 4; also

depicted is the electronics box that accompanies the cassettes on the spacecraft.

Figure 4: Langmuir Probe payload mechanical design

2.3.3

Kongsberg Seatex, in Norway, is designing the AIS receiver payload. The AIS receiver will

detect and track maritime traffic in Norwegian and international waters via the Automatic

Identification System (AIS). The AIS system is a line-of-sight, self-organized, time division

multiple-access messaging system that provides situational awareness to a large number of

maritime vessels at sea. It allows the exchange of information such as ship identification,

position, course and speed, allowing governmental organizations to monitor and direct the ship

traffic. Mandated by the International Maritime Organization, all vessels over 300 gross tonnes

are obligated to carry the AIS system. Monitoring and collecting AIS data from space has

Page 22: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

14

recently proven to be effective, and of great interest to the Norwegian government, due to the

large portions of open water that is currently impossible to monitor via coast-based AIS stations

(Figure 5). The icons seen near the mainland coast of Norway represent the extent of the coast-

based AIS monitoring, while the blue shaded areas represent the relevant Norwegian and

international waters.

Figure 5: Norwegian coastal regions [6]

The AIS payload consists of a dual antenna very high frequency (VHF) receiver supporting four

VHF channels each. The technology of the receiver will be similar to the previous AIS receivers

designed as the main payloads on previous SFL satellites, such as AISSat-1, which was launched

in 2010, as well as the recently launched AISSat-2. This new AIS receiver will be more

advanced and will have the opportunity to test out new detection algorithms [6]. The motivation

for ship detection via AIS is fairly clear and proven from previous AIS missions, in that

detection from space allows for a more complete picture of the activity in the waters and can

better prevent ship collisions. NORSAT-1 is also intended to add to the on-going capability of

the Norwegian government of space-based AIS systems, marking this as the fourth satellite in

Page 23: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

15

the constellation of AIS satellites, the others being the already on orbit AISSat-1 and AISSat-2,

as well as the soon to be launched AISSat-3 – for all of which the spacecraft platform was

designed by SFL. The AIS receiver payload is the sole non-science payload aboard the

NORSAT-1 mission.

2.3.3.1

Two main challenges exist with the AIS payload. First, because of the low frequency of

operation (VHF), the antennas have to be quite large relative to the overall spacecraft size. On

previous SFL designed AIS satellites, single pre-deployed antennas were used, however, since

NORSAT-1 intends to use two orthogonal antennas, the volume that the antennas consume

would be significant. Because of this, similar to the Langmuir Probe instrument, the antennas

must be made deployable. In reducing the launch volume of the satellite, the amount of launch

vehicles able to accommodate the spacecraft increases considerably.

The second main design challenge associated with the AIS payload is that the receiver is highly

sensitive to electromagnetic radiation at its frequency of operation (VHF). At this relatively low

frequency, it is not uncommon that many electronics generate noise, and could easily interfere

with the payloads data collection. As such, a fairly strict requirement on platform generated noise

to the AIS payload is placed on the spacecraft design.

2.3.3.2

The overall dimensions of the AIS receiver payload are shown below in Figure 6. The total mass

of the instrument is approximately 1.5kg. The two accompanying deployable VHF antennas are

designed by the author, and are detailed in Chapter 6.

Figure 6: AIS Receiver payload

Page 24: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

16

3

The main objective for this thesis is to design the structural subsystem for the NORSAT-1

mission such that it properly accommodates all satellite subsystems, including the payloads, and

provides a safe and secure housing through all mission environments. The success of the design

will be based on its ability to satisfy the mission, system, and structural subsystem requirements.

Some of these driving requirements that affect the structural design are outlined in the following

subsection.

Most of the previous, and ongoing, satellites designed by SFL leverage off of a Generic

Nanosatellite Bus (GNB) design that has proven effective through multiple missions and requires

minimal structural alterations for different projects. However, due to the larger payloads on the

NORSAT-1 mission, this design cannot be re-used, and a larger satellite design must be realized.

SFL is also in the final stages of development of a larger evolution of the GNB, the NEMO

(Next-Generation Earth Monitoring and Observation) bus. The first spacecraft to use this new

bus technology is the NEMO-AM (Aerosol Monitoring) spacecraft, which is set to launch in

2016. While catering to its own mission requirements, NORSAT-1 will make an effort to use

heritage GNB and NEMO technology wherever possible in order to reduce risk and cost.

3.1

A list of driving requirements that largely affect the structural design of NORSAT-1 is detailed

below in Table 2. These requirements are compiled from various sources, including overall

mission programmatic desires, payload specific constraints, lessons learned from previous SFL

missions on the design, assembly, and testing, as well as requirements directly from the launch

vehicles. The success of the structural design of NORSAT-1 hinges on its ability to satisfy the

listed requirements.

Table 2: Driving requirements that affect the structural design [8] [9]

# Requirement Comments/Rationale

General Requirements

1

The structure shall support and ensure survival of all spacecraft components through integration, transport, handling, and launch.

Basic definition of the structural system. Cannot be verified until post-launch.

Page 25: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

17

2 The mission should use GNB and/or NEMO heritage components to the extent practical.

Programmatic desire to minimize Non Recurring Engineering (NRE) and cost in general.

3 The spacecraft dimensions, including appendages, shall be compatible with a qualified SFL satellite deployment system.

In order to leverage an already existing SFL deployment system, such as the XPOD-Duo.

4 The spacecraft mass shall be less than 20 kg. Maximum spacecraft mass supported by the XPOD-Duo.

5

The subsystems and components used in the construction of the spacecraft shall be composed of materials that exhibit a total mass loss of no more than 1% of the component’s initial mass, and that contain no more than 0.1% collected volatile condensable material.

Desire to prevent material degradation and minimize deposit build-up on sensitive surfaces.

6

To allow air to easily escape from the satellite during launch, all volumes containing air shall be vented using an aperture with area (mm2) no smaller than 7x10-6(mm-1) x V, where V (mm3) is the volume of air.

Desire to prevent stresses induced by pressure differentials in vacuum.

Payload Requirements

7

The CLARA payload shall be accommodated such that its sensor apertures see the sun during nominal operations, and such that its aperture is the furthest protruding face of the spacecraft in its line of sight direction.

Needed in order to make solar measurements. Avoids any significant thermal impact from other components during measurements.

8 The Langmuir Probes (qty. 4) shall be accommodated externally parallel to each other and orthogonal to the CLARA Line of Sight.

Need to be as far away from satellite as possible, in order to be more submersed in the plasma environment.

9

The AIS antennas (qty. 2) shall be accommodated externally and be pointed orthogonal to each other and orthogonal to the Langmuir Probe booms.

In order to utilize polarization discrimination to improve the AIS message detection rate.

10

The platform shall limit platform-generated noise at the input to the AIS payload to -124 dBm measured in a 25 kHz bandwidth, within the band 156.025-162.025 MHz.

As measured by the AIS payload. Desire to limit platform noise propagation to the AIS receiver so as to not affect payload measurements

Launch Vehicle Requirements

11

The spacecraft must be capable of surviving, with positive safety margins, expected quasi-static launch loads of the launch vehicles under consideration, which is defined as the 5-sigma acceleration value of the composite random vibration spectrum.

In order to ensure the spacecraft will survive all launch loads without yielding. The 5-sigma value is used to instill a high level of confidence in the design. Due to launch uncertainty, all launch vehicles are considered.

12

All spacecraft components must have a first natural frequency (FNF) in excess of that required by the launch vehicles under consideration.

In order to prevent dynamic coupling between the spacecraft and launch vehicle. Due to launch uncertainty, all launch vehicles are considered. The PSLV imposes the most severe requirement, where the spacecraft must have a FNF greater than 90Hz.

Page 26: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

18

13

In its flight configuration the satellite shall be subjected to an acceptance-level vibration test at levels specified by the launch provider and must pass this test without failure.

Required by the launch provider to prove the spacecraft can handle the expected loads.

Assembly and Testing Requirements

14 All mechanisms shall be testable in a 1g environment, with suitable GSE.

To verify their functionality in a more stringent environment.

15 Threaded inserts (Helicoils) shall be used in all non-steel materials requiring threading (aluminum, magnesium, etc.) where possible.

Prevents thread damage to expensive custom components.

16 The use of nuts should be avoided in favor of mounting bosses with threaded holes.

Reduces the number of components in the spacecraft and reduces risk of small components coming lose during launch. Also simplifies the assembly/disassembly process.

17 The spacecraft structure should allow access to any subsystem component without requiring full system disassembly.

Desire to simplify the assembly/disassembly process and to enable rapid de-bugging.

18 The spacecraft should not require custom tools to assemble or disassemble.

Custom tooling is expensive and its necessity complicates integration and testing activities, especially if done off-site.

A major defining constraint for the overall geometry and mass of NORSAT-1 is the desire to use

an existing SFL designed deployment system (Requirement #3 and #4). This greatly reduces

costs associated with Non-Recurring Engineering (NRE) because the design of the spacecraft can

then leverage large amounts of design work from previous projects. The XPOD-Duo is presented

below in Figure 7; its design is largely based off the successful XPOD deployment system that

was developed for the GNB class of nanosatellites at SFL, and it was chosen for the NORSAT-1

mission due to its larger capacity.

Figure 7: XPOD-Duo deployment system, Vertical mounted (A), Horizontal (B)

The XPOD-Duo houses the satellite aboard the launch vehicle into orbit, and upon command,

ejects the satellite by means of a compressed spring. Four points of contact are required by the

spacecraft to mate with the deployment system (known as the satellite “feet”), as well as four

Page 27: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

19

launch rails to provide a smooth ejection. The XPOD-Duo also features two open faces to

accommodate spacecraft external appendages such as antennas and solar panels. During the

course of this thesis, the XPOD-Duo has been slightly redesigned from Figure 7A to Figure 7B –

with the main change being that it will be horizontally mounted on the launch vehicle instead of

vertically. Due to this change midway during the design of NORSAT-1, some minor changes

were made to ease the accommodation and will be discussed later.

3.2

The design builds on the design methodology of the GNB nanosatellite, whereby two trays are

used to mount large and/or massive components (e.g. reaction wheels, batteries, radios, etc.). The

satellite is then enclosed using metallic panels onto which solar cells or other light

deployable/pre-deployed components can be attached; these panels also provide a degree of

additional structural support. Each panel is 2mm thick with additional cross braces machined

directly onto the inward facing surface to increase panel natural frequencies, reduce deflections

under acceleration and increase stiffness during machining. The trays are positioned on opposing

sides of the satellite, leaving a relatively large volume between them at the center of the satellite

available for the internal payloads. This concept has been successfully implemented on the GNB

bus for multiple missions, and has been extended to the NEMO bus; therefore, NORSAT-1 will

continue to implement the proven design concept. This concept can be seen below in Figure 8,

showing the 3D solid model for various GNB and NEMO bus designs.

Figure 8: GNB bus AISSat-3 (left), GHGSat-D (middle), and NEMO-AM (right)

3.2.1

Upon the authors joining of the NORSAT-1 project in October 2013, SFL had just undergone the

Preliminary Design Review (PDR) for the project; placing NORSAT-1 in the detailed design

phase of development. Two fellow SFL workers, Scott Armitage and Jamie Fine, had worked on

Page 28: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

20

the preliminary structural design prior to the PDR, including generating an initial structural

design proposal of the satellite bus, shown below in Figure 9.

Figure 9: Initial NORSAT-1 structural design proposal (Scott Armitage)

Initially, large volumes were allotted for the payloads due to the uncertainty of their designs,

which largely drove the size of the concept at this initial stage. These volumes then significantly

decreased as the payload designs progressed. One large solar array was included to generate the

required power during operations while the CLARA payload is directed at the sun.

3.2.2

The primary structure of NORSAT-1 (Figure 10) consists of a pair of aluminum trays (+Z and –

Z) on which most of the avionics and payloads are attached, a set of panels (+X, -X, +Y, -Y, +Z,

-Z) and risers (+X and –X) that form the outer structure, and an internal separation plate that

separates the avionics from the payloads. The risers serve to extend the bus volume through one

of the openings on the XPOD-Duo for added capacity. The rails and feet that interface with the

XPOD-Duo are integral to the trays, which form the main load-bearing structure. A separate

dedicated bracket for the three reaction wheels is included, and attaches directly to the –Z tray,

along with a number of attachment brackets for the large solar array wings.

The entire primary structure is machined out of aluminum 6061-T6, commonly used in aerospace

applications, in order to obtain the required tolerances and mechanical properties. Numerous M3

sized stainless steel screws are used to fasten everything together, with the majority of the

threaded hole features machined directly into the two trays, in compliance with Requirement #16

in Table 2.

Page 29: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

21

Figure 10: Exploded view of NORSAT-1 primary structure

Table 3 below lists all of the main components that must be positioned in the NORSAT-1

structure. In it, the table briefly lists the main constraint or subjective requirement regarding its

placement in the satellite, in addition to those specifically highlighted in Table 2. For each of the

components, some general guidelines that helped shape the overall layout and were implemented

where possible include the following:

- Position heavier components towards the geometric center of the bus.

- Position antennas to allow for omni-directional coverage.

- Minimize the number of electronic boards in a single stack.

- Avoid mounting any components on the solar panel wings besides solar cells.

- Minimize wiring unrelated to the payloads in the payload volume.

Page 30: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

22

Table 3: Component layout constraints for NORSAT-1

Component Layout Constraint or Requirement

Solar Panel Wings Must be positioned such that they do not interfere with integration into the XPOD, and must face in the same direction as the CLARA apertures.

Reaction Wheels (3) Must be aligned with the principle body axes and mounted orthogonal to each other.

Sun Sensors (6) Must be able to see in all six principle directions.

Magnetorquers (3) Must be aligned with the principle body axes and mounted orthogonal to each other.

Magnetometer Should be mounted as far away as possible from the reaction wheels, large current sources, and large magnetic dipoles. Must be aligned with the principle body axes.

Rate Sensors Must be aligned with the principle body axes.

House-Keeping Computer (HKC)

Locate as close as possible to the ADCC to minimize wiring.

Attitude Determination and Control Computer (ADCC)

Locate as close as possible to the HKC to minimize wiring.

Payload On-Board Computer (POBC)

Locate as close as possible to the HKC and SIB to minimize wiring.

Power Avionics/Modular Power System (MPS)

Locate as close as possible to the batteries to minimize wiring.

Batteries All cells should be located as close together as possible and as close to the power avionics as possible to minimize wiring.

Serial Interface Board (SIB) Should be placed close to the payload connectors and POBC to minimize wiring.

S-band Transmitter Should be enclosed in a similar fashion to GNB with coaxial connections near +Y end of spacecraft for ease of access during assembly.

S-band Down-converter Should be close to the UHF receiver to minimize wiring.

S-band Combiner Should be close to the S-band Cavity Filter to minimize wiring.

S-band Cavity Filter Should be close to the S-band Down-converter and Combiner to minimize wiring.

UHF Receiver Should be enclosed in a similar fashion to GNB.

GPS Receiver Should be mounted near the GPS Antenna to minimize wiring.

GPS Antenna Must be aligned such that the Solar Array and main body do not interfere with coverage. Should not be mounted to the solar panel wings.

S-band Patch Antennas Must be aligned such that the Solar Panel Wings and main body do not interfere with omni-directional coverage. Should not be mounted to the solar panel wings.

CLARA Payload Must be pointed out the main sun-facing side.

Precision Sun Sensor Should be attached to the same mounting structure as CLARA to minimize thermo-elastic distortions between them.

AIS Receiver Payload Should be as close as possible to the VHF antennas to minimize wiring, and should be in a RF noise reduced area.

AIS antennas Should be as close as possible to the AIS receiver to minimize wiring. Must be outside the spacecraft “Faraday cage”.

Langmuir Probe Electronics Should be positioned as close as possible to the Langmuir probe cassettes to minimize wiring.

Langmuir Probe Cassettes Positioned externally, with connectors as close as possible to the Langmuir Probe electronics to minimize wiring.

Page 31: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

23

In addition to these component layout constraints, a number of external surface area

requirements also exist that reserve amounts of area on each side of the spacecraft. For example,

an area sufficient for one solar cell string of eight cells shall be reserved on the structure in each

six directions.

From the above-mentioned constraints and requirements, a component layout for NORSAT-1

could be realized. The final external and internal component layout is depicted in Figure 11 and

Figure 12 respectively. Much of the internal layout design of the spacecraft electronics was

leveraged from another SFL on-going project of similar size and requirements (GHGSat-D) in

order to minimize costs and time due to Non-Recurring Engineering (NRE) during the design

phase. The payload volume used in this satellite is large enough to house the three internal

payloads for NORSAT-1; therefore much of the avionics component layout could be left as is.

Figure 11: NORSAT-1 external component layout

Page 32: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

24

Figure 12: NORSAT-1 internal component layout

3.2.2.1

The –Z tray sub-assembly houses all three on-board-computers (HKC, ADCC, POBC), the rate

sensor, the –Z sun sensor, the S-band Transmitter, and the UHF Receiver. The three computer

boards are stacked together using aluminum spacers, and are recessed into the tray via a

machined-in housing. The rate sensor has its own housing to mount the sensors in each axis, and

the housing is directly mounted on the –Z side of the tray; on this same side is mounted one of

the sun sensor boards for the –Z panel. The radio enclosure on the +Z side of the tray

incorporates a nearly identical housing for the S-band transmitter and UHF receiver to that which

is used on the GNB bus. This was done in order to keep a heritage design and avoid potential

compatibility issues between the already designed receiver/transmitter boards and its mechanical

enclosure.

A dedicated reaction wheel bracket is designed to house all three reaction wheels in their

required orientation, in order to provide a compact and modular integration of the wheels into the

satellite. This bracket is mounted as a sub-assembly onto the –Z tray, and it is shown below in

Figure 13. In order to facilitate the wheel integration with the wiring harness, the wheels are not

included at the sub-assembly level of the –Z tray, but are integrated as their own sub-assembly.

A similar approach is taken with the CLARA payload.

Figure 13: Reaction wheel sub-assembly, CAD model (left), clean room assembly (right)

The avionics housed on the +Z tray sub-assembly include the Modular Power System (MPS),

battery pack, GPS receiver, S-band down-converter, S-band combiner, S-band cavity filter, as

well as the AIS receiver and Langmuir probe electronics payloads on the lower half. The +X side

of the tray is slightly recessed in order to compactly accommodate some of the taller components

such as the modular power system and battery pack. These components are directly fastened onto

the tray via machined-in features at the mounting points, which also raise them off the surface in

Page 33: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

25

order to provide clearance for the board mounted components, while also providing an area to

strategically place thermal control materials as required by the thermal design.

Some of the lighter components, such as the sun sensors and magnetorquers, are directly

mounted on the back of each of the panels, as well as all of the antennas and solar cells on the

front. Each panel is fitted with a panel connector for its sub-assembled components, which then

mates to the main wiring harness of the bus when the panel is integrated. Each of the panels’

sub-assemblies is depicted below in Figure 14. Note that there are some multiples of similar

components on different panels (six sun sensors, three magnetorquers, etc.) and have only been

labeled once in the figure for simplicity. The different colors seen on the front of some of the

panels represent a proposed thermal tape scheme based on the thermal design for a certain orbit.

Each panel is also fitted with a temperature sensor, which is not shown in the figure.

Figure 14: NORSAT-1 panel component layouts (front/back)

As seen in the above figure for the +Z, +X, and –X panels, the spacecraft avionics are kept in the

top (+Y) half of the panel, in order to keep these components in the avionics bay of the satellite.

Page 34: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

26

The only non-payload components in the payload bay of the satellite are the –Y sun sensor,

temperature sensor, solar cells, and uplink/downlink antennas. Polycarbonate screws are used to

mount the uplink/downlink antennas and the magnetometer in order to not affect their

performance. A single string of eight solar cells are included on the +Y, -Y, +X, and –X panels.

For the +Z and –Z sides of the spacecraft, solar cells are placed on the front and back of the solar

array wings. For the Y panels, the solar cells are directly laid on the aluminum substrate, and

alternatively on the X panels, they are laid on a separate aluminum coupon, and then mounted on

the X panel over stainless steel spacers. The outer appendages, such as the AIS antennas,

Langmuir probe cassettes, and solar panel wings are designed to be installed after the entire

spacecraft is integrated, via external mounting points and panel mounted skin connectors.

The battery pack design incorporates six SAFT rechargeable lithium ion battery cells, two

parallel strings of three cells in series. The battery cells have a much narrower temperature range

than most of the other spacecraft electronics; therefore efforts are made to control the

temperature of the battery pack separately from the spacecraft as much as possible. The cells are

arranged in a set of Delrin acetal homopolymer resin collars, which allow the heat produced by

the cells to be somewhat isolated from the spacecraft due to the materials’ low thermal

conductivity, and allows for the control of the amount of conductivity from the pack to the +Z

tray using varying amounts of thermal control materials below the pack. A heater is also needed

for each string of cells in order to keep the batteries at a safe temperature in colder scenarios;

polyimide film insulated flexible heaters are used for this. Since the Delrin material provides an

inefficient medium to spread the heat of the heater to the cells, an aluminum heater plate is

incorporated to more efficiently spread the heat – the cells are each thermally strapped together,

as well as to this heater plate via various thermal control materials (Gap Pad, Pyrolytic Graphite

Sheets (PGS)). Clearance for each cell in the collars and between adjacent cells is carefully

selected in order to allow for expected battery swelling after long-term use. The battery pack

design is displayed below in Figure 15. Note that in the flight battery pack (not shown in the

figure), the cells are thermally strapped together using PGS, and the entire exposed cell area of

the pack is wrapped in thermal tape in order to prevent thermal exchange through radiation in the

spacecraft. Temperature sensors are also included on the middle cell of each string, but are also

not shown in the figure. The fully integrated flight battery pack is shown in the later Section 8.8

Flight Integration.

Page 35: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

27

Figure 15: NORSAT-1 battery pack design, Exploded (left), Assembled (right)

The Battery Interface Module (BIM) is directly mounted on the battery pack, but its thermal path

is separated via the BIM sink plate. The sink plate allows for the BIM to sink its thermal energy

to the spacecraft +Z tray, rather than directly to the battery cells in order to prevent thermal

gradients across the individual battery cells. A custom terminal block was also incorporated into

the battery pack design, for the sole purpose of ease of assembly. This allows for the battery cells

and BIM to be wired and prepared separately, and then assembled together in a clean room

environment. Wire tie mounts are used on the top collar for battery cell wire management.

3.2.3

As mentioned in the previous Section 2.3 Payloads, the AIS receiver payload is greatly sensitive

to Radio Frequency (RF) noise produced in and around the satellite because of its relatively low

frequency operating levels. Because of this, efforts are directly made in the structural design to

reduce the amounts of potential electromagnetic interference (EMI) propagation to each of the

payloads in order to satisfy Requirement #10 in Table 2. The strategy was to attempt to isolate all

of the noise producing spacecraft avionics from the payloads, allowing minimal RF leakage out

of the isolated avionics bay, without requiring a large redesign of the leveraged bus layout; this

was done in a number of ways:

A “separation plate” was added internally to the spacecraft, physically separating roughly

90% of the satellite electronics from the payloads, resulting in an “Avionics bay” and a

“Payload Bay” as shown in Figure 16.

EMI reducing gaskets were incorporated in the design, in an effort to render the avionics

bay close to a “Faraday Cage”, reducing RF leakage to the payloads.

Page 36: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

28

Additional screws were used around the volume of the avionics bay to both ensure

sufficient gasket compression, as well as reduce the gap lengths between screws where

noise can potentially leak through.

EMI filtered connectors were used at any interface from the avionics bay to the payload

bay or outside the spacecraft.

Figure 16: Division of avionics and payloads in NORSAT-1

Since much of the design was leveraged from heritage projects in order to reduce development

time, there was limited time to create a dedicated design for isolating the avionics. A “best

effort” approach was taken, whereby gaskets were introduced into the design wherever they

could be easily accommodated, with minimal effect on the overall design and assembly

procedure. A gasket is placed fully around the Main and Small separation plate, on the +X and –

X sides of the +Y panel, on the +Y and –Z side of the –Z tray, on the +Z panel under the AIS

antenna mounting plate, and on the +Y panel below the GPA antenna. Other areas where it was

too difficult to introduce a gasket feature were either left as is, or the screw spacing in those

areas were decreased by means of introducing additional screws – these areas include the –Z

panel, +X and –X panel upper section on the avionics side, the +Z panel, and the risers.

Efforts were not made to render the payload bay of the spacecraft as isolated as the avionics due

to the difficulty in mounting the CLARA payload. Due to its thermal sensitivity, physically

contacting the payload anywhere other than the dedicated mounting points could skew the

scientific results - sealing something that cannot be contacted proves to be a difficult task.

Therefore, an opening is left surrounding the CLARA payload in the payload bay, and efforts are

focused on shielding the noise from escaping the avionics and potentially entering the payload

Page 37: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

29

bay from this gap. In addition, similar low noise emitting requirements exist for each of the

payloads themselves; therefore the noise produced in the payload bay is expected to be minimal.

Filtered connectors for each of the payloads are directly mounted on the main separation plate, in

order to transfer the power and data signals from the avionics to each payload. A filtered

connector is also included on the plate for the precision sun sensor, and the –Y panel avionics

since they reside in the payload bay. The RF signal for the –Y panel uplink/downlink antennas is

syphoned through the separation plate via a pair of SubMiniature Verson A (SMA) connector

adaptors. Filtered skin connectors are also directly mounted on the –Z tray for the solar array

wings, and on the +Z panel for the Langmuir probe cassettes. The design of the separation plate

is shown below in Figure 17. The Serial Interface Board (SIB) was directly mounted on the

avionics side of the separation plate in order to facilitate its wiring, since it connects directly to

each of the payload connectors on the plate. Numerous tie mounts are epoxied to the main

separation plate for use in wire management, which will be discussed in Section 3.6 Wiring

Harness Development.

Figure 17: Separation plate sub-assembly design and layout

The separation plate is separated into a Main and Small plate in order to minimally affect the

existing tray designs. The plates assemble together via five M3 fasteners while sandwiching the

+Z tray. Together, the plates provide structural mounting points for the –Z tray, +X Panel, -X

panel, +Z panel, and both risers, while also compressing the gasket between them.

3.3

As seen during the battery pack design description in the previous section, the thermal design of

a satellite is closely linked to the structural design. Materials selections, component layout,

component mounting, and overall geometry have a large effect on the thermal results. The details

Page 38: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

30

of each component mounting connection and layout are discussed with the thermal engineer for

NORSAT-1, Vincent Tarantini, during the detailed design phase, in order to understand the

thermal point of view of each design decision. Over the course of the project, various thermal

concerns were raised regarding the overall design and proposed working orbit, which amounted

in structural design modifications in order to better accommodate the fully passive thermal

design – these modifications are explained in this section.

3.3.1

Initially, the proposed NORSAT-1 structural design included one large solar panel to achieve the

necessary surface area for all of the solar cells during payload operations, similar to the NEMO-

AM microsatellite project (seen in Figure 9). Soon after the Preliminary Design Review (PDR),

it was discovered that this proposed large solar panel could potentially be a thermal concern and

alternative solutions were considered. The problem exists because the large solar array is able to

fully shadow the entire spacecraft bus. In the scenario where only the large solar array points at

the sun continuously, the spacecraft would need to be well thermally coupled to the array in

order to keep warm; however, if it is well coupled in other attitudes, the solar array would act as

a large radiator, and large thermal swings would be seen by the spacecraft. A solution was found

by splitting the large solar array into two separate solar array “wings”, whereby a direct heat path

to the spacecraft exist in the middle through the –Z panel for thermal control. In addition, the

solar array wings would be thermally isolated from the bus, in order to have more control over

the bus temperatures in all attitudes.

3.3.2

During the detailed design phase of NORSAT-1, the thermal simulations of the structural design

found that the satellite bus would get too hot in the “worst-case hot situations”. The worst-case

hot situation describes an orientation of the spacecraft on orbit where the average temperature is

maximized. This often occurs in an orientation where the maximum possible surface area of the

spacecraft is directed at the sun continuously, resulting in maximum heat absorption. For

NORSAT-1, this happens at the orientations seen in Figure 18.

Page 39: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

31

Figure 18: Worst-case hot attitudes (as viewed from the sun)

Having the solar cells mounted directly on the spacecraft bus leads to a convenient mechanical

design, however each solar cell then acts as a large heat absorber into the bus. Due to the number

of solar cells in view in the worst-case hot orientations, the incoming heat produced becomes too

significant for the passive thermal design to regulate. The solution to this problem was to isolate

some of the solar cells from the bus by mounting them off the spacecraft, in order to achieve the

correct balance of incoming heat. This was done in two ways: the solar cells originally mounted

on the +Z panel were moved to the rear of the solar panel wings, and the solar cells on the X

panels were moved onto dedicated aluminum coupons, which are mounted to the X panels with

stainless steel spacers at the mounting points. The solar cells on the Y panels were left directly

mounted on the spacecraft panels in order to maintain a thermally safe worst-case cold situation,

described below.

3.3.3

In addition to getting too hot, the thermal simulations were also finding that the satellite would

get too cold in the “worst-case cold situations”. The worst-case cold situation describes an

orientation of the spacecraft on orbit where the average temperature is minimized. This often

occurs in an orientation where the minimum possible surface area of the spacecraft is directed at

the sun continuously, resulting in minimum heat absorption. For NORSAT-1, this happens at the

orientations seen in Figure 19.

Page 40: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

32

Figure 19: Worst-case cold attitudes (as viewed from the sun)

Due to the numerous components mounted on the Y panels (GPS antenna, eight solar cells, two

s-band patch antennas, sun sensor, antenna hold down hitch), there is minimal room for thermal

control tapes that would allow additional heat from the sun to be absorbed. Because of this, the

amount of heat able to be absorbed in these orientations was not sufficient to maintain the

spacecraft within the temperature limits. After exploring several options, including relocating

some of the Y panel mounted components and imposing on-orbit attitude constraints, it was

decided that the most simple and minimal impact solution would be to add additional surface

area to the satellite in these cold attitudes. This was done by extending the –Y panel, and adding

a pair of flat aluminum “thermal surfaces” to the mid-plane of the satellite, which would

essentially extend the useable thermal surface area of the satellite in the worst-case cold

orientations by about 180%. By mounting these on the mid-plane of the satellite, there is an

added benefit of having the incoming energy heat up the satellite more effectively from the

middle of the spacecraft, rather than from one end. Another large benefit to this solution is that

while a large surface area increase is seen in the worst-case cold orientations, virtually no surface

area is added in the worst-case hot orientations. The inclusion of both these additional parts

would add less than 100 grams to the mass of the spacecraft, and was deemed a low-risk, and

satisfactory thermal solution. Figure 20 and Figure 21 display the design and placement of these

added thermal surfaces.

Page 41: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

33

Figure 20: +Y worst-case cold dimension increase due to thermal surfaces

Figure 21: +Z and –Z thermal surface additions

3.4

The way in which each of the payloads has been mechanically accommodated in the satellite in

order to satisfy the relevant requirements is discussed in this section.

3.4.1

The CLARA payload is mounted in the payload bay of the satellite, such that it is the most

protruding face in the –Z direction – this is to directly satisfy Requirement #7 in Table 2. It has a

Precision Sun Sensor mounted on a bracket that is directly attached to the rear section of the

payload in order to minimize calibration offsets through thermal and mechanical distortions. The

payload is uniquely integrated into the satellite at a late stage in order to provide minimal

handling of the sensitive payload and exposed MLI. This is done by the use of slots in the

spacecraft –Z tray, whereby the feet of the CLARA payload can slide through into the spacecraft.

Once the feet are fully inside, the payload is shifted towards the +X direction into some

Page 42: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

34

machined-in slots, and then four M5 titanium blots can be inserted from the outside of the

spacecraft into the payload mounting feet. Finally, an aluminum side plate is installed from the -

X opening of the satellite to shield the majority of the opening left beside CLARA. This

sequence of installation is depicted in Figure 22.

Figure 22: CLARA integration sequence

Some advantages to this method of mounting CLARA are that in order to remove CLARA from

the satellite assembly after full integration, only removal of the –X panel and -X wing is needed.

This reduces much of the risk of de-integrating the satellite if removal of CLARA is needed at a

late stage. It also allows CLARA to be installed at a very late stage in the assembly procedure –

minimizing potential contaminations through handling.

3.4.2

The AIS receiver is mounted on a set of structural cross braces machined into the +Z tray using

eight M4 fasteners and is located in the noise-reduced payload bay; connector access is achieved

through the –X side of the spacecraft. Both of the VHF antennas are placed on the +Z panel,

orthogonal to each other (Requirement # 9 in Table 2), and 45 degrees to the satellite face. These

accommodations are shown in Figure 23, with the relevant mounting holes highlighted in red.

Page 43: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

35

Figure 23: AIS Receiver (left) and AIS Antenna (Right) accommodations

3.4.3

The Langmuir Probe electronics are mounted on the opposite side of the cross braces supporting

the AIS receiver on the +Z tray using four M3 fasteners. The two cassettes of probes themselves

are mounted in parallel on either Riser using eight M3 screws each (Requirement #8 in Table 2).

They connect to the Langmuir probe electronics box by means of an intermediate wiring harness,

which allows each cassette to be plugged directly into the +Z panel from the outside. These

accommodations are shown in Figure 24, with the relevant mounting holes highlighted in red.

Figure 24: Langmuir Probe electronics (left) and Langmuir Probe cassette (Right)

accommodations

3.5

Considerations are made during the design to ensure assembly and disassembly is both possible,

and convenient (Requirement #17 in Table 2). The design is kept modular to allow a fairly

smooth integration process. Each sub-assembly is first fully integrated individually, and then all

of the sub-assemblies are integrated together. The large wiring harnesses are first integrated to

the +Z tray sub-assembly, followed by the joining of both tray assemblies via the separation plate

and some ground support equipment (GSE); then begins panel integration. Due to the loss of

access once panel assembly onto the trays begins, a specific order is derived for the assembly

Page 44: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

36

procedure to ensure that the necessary access is available at all stages of integration. For

example, access is needed from the –Y end of the satellite in order to mate the Langmuir probe

cassette connectors to the Langmuir probe electronics box, therefore, the +Z panel must be

integrated prior to the –Y panel. The full detailed procedure for the sub-assembly and system

level assembly and integration for NORSAT-1 has been developed by the author in [10], and has

been summarized at a high level in the list below.

1) Integrate all sub-assemblies

2) Install the large wiring harness on the +Z tray

3) Connect the –Z tray to the +Z tray sub-assembly

4) Connect and route all of the wiring harnesses

5) Integrate the +Z panel sub-assembly

6) Integrate the –Y panel sub-assembly

7) Integrate the Reaction Wheel sub-assembly

8) Integrate the +Y panel sub-assembly

9) Integrate the +X panel sub-assembly

10) Install the +X and –X riser

11) Integrate the –Z panel sub-assembly

12) Integrate the CLARA payload

13) Integrate the –X panel assembly

14) Integrate both Langmuir Probe Cassettes

15) Install both AIS antennas in stowed position

16) Integrate both solar panel wings

The connectors on most electronics are placed such that connector access is always achieved

from the X sides of the spacecraft, allowing most connections to be made during panel

integration from the large open sides, and renders the X panels to be the last panels integrated.

The main bulkheads from the Modular Power System, CLARA, the AIS receiver, the separation

plate connectors, the UHF receiver, the S-band transmitter, the SIB, and the OBC connectors are

all accessible from the +X or –X side of the spacecraft during integration. This also allows for a

large amount of debugging to be done during the testing phase with the removal of just one panel

of the spacecraft.

Skin connectors are incorporated for the solar array wings and Langmuir probe cassettes on the

–Z tray and +Z panel respectively. This allows for these large protruding elements to be installed

at the final stage of integration to simplify handling during the rest of the assembly. The same

strategy is applied to the AIS antennas, which are designed to be mounted from the outside, and

can thus be integrated last. The AIS antenna mounting plate is designed as a separate piece from

the +Z panel for ease of assembly as well. This allows for the antenna connectors to be soldered

to the gold plated antenna mounts as their own sub-assembly, and then mounted to the +Z panel

sub-assembly separately.

Page 45: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

37

The proposed solar cell layout on the solar array wings minimizes the size of the panels; however

it forces the panels to have a larger Y dimension than the satellite itself, which can cause certain

assembly complications. To accommodate this, the wings are purposefully not centered about the

Y-axis of the satellite; rather, they are offset, and biased towards the +Y end. This is done in

order to allow for the fully integrated satellite to rest on the –Y satellite “feet”, and ease the

integration of the wings in this orientation, and not limit the potential resting attitudes of the fully

integrated satellite. This was favored over enlarging the panels in the X direction, which would

force the center of mass of the cantilevered panels further outward.

3.6

While designing the spacecraft layout and structure, careful consideration must be made for the

planning of wire management. The structural design can easily incorporate simple inclusions to

aid wire routing and management, through means of tie down points and wire path cutouts, and

thus should not be left to a late stage in the design. A clean, robust wiring harness in the fully

assembled satellite is sought, in order to ease the burden of accessing components after

integrations, and to avoid loose wire bundles shaking violently during launch vibrations,

potentially de-mating connections. Some general guidelines that were followed for the wire

harness development are explained below.

Design for Assembly – The harness layout should be optimized for assembly efficiency during

integration.

Minimize lengths of wires –To save mass, to minimize signal power loss along the wires, and to

avoid turning them into Radio Frequency (RF) antennas or current loops (magnetic fields).

Minimize connectors – Based on the already designed hardware, namely the Modular Power

system (MPS), certain signal levels are limited to specific outputs from the MPS, which then

may connect to several different components, causing large, interconnected, harnesses with

numerous connectors. These large harnesses are difficult to build, manage, and integrate.

Clean routing – Common tie down points should be used for multiple components, which can

simplify the assembly and integration process, minimize tie down points, and prevents the

harness from becoming a “web”, blocking access to components after integration.

Tie down locations – Wires should be tied down around every 10cm to prevent large amounts of

loose wiring and each connector should have a tie down point immediately before its mating

connector. All wires should have the necessary strain relief as well as wire minimum bending

Page 46: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

38

radius accounted for. The free wires next to their connectors should always provide a

compressive force, so as to counteract de-mating.

The wiring harness interconnect diagram, generated by the system engineer, depicts all the

satellite electronic components and their specific pin-to-pin assignments. This is generated early

in the project with only electrical and power requirements considered for a system level

architecture. This was used to determine all wire paths in the satellite for initial wire routing and

component layout design. This diagram was then converted into graphical harness manufacturing

drawings for each individual harness on NORSAT-1. Including each panel harness, this

amounted to 23 individual wiring harnesses with over 100 total connectors, each harness ranging

from having just one connector, to one large one having 53 connectors.

The large harness with 53 connectors posed a concern for the project because of the numerous

points of failure, and difficulty to build without error. In an effort to reduce the number of

connectors on this large harness, the author was tasked to investigate rearranging pin

assignments on some connectors. Due to the limited outputs levels on the MPS, no amount of

rearranging of pin assignments would allow the harness to be broken into smaller pieces.

However, after some pin reorganizing, a point in the harness was found where only four wires

connected two large sides. It was then decided to incorporate an “in-line” connector between

these two halves of the harness, adding two connectors overall, but splitting the large 53

connector harness into one 39 connector harness, and one 16 connector harness. The smaller of

these two harnesses was dubbed the Payload (and sun sensor) harness, connecting the power and

data lines to all of the payloads and sun sensors, while the larger one was dubbed the Main

harness, connecting all the remaining spacecraft avionics.

Each graphical manufacturing drawing, created by Payam Mehradnia, was then supplemented by

the author with additional information and instruction for the build, such as specific wire lengths,

colors, gauge, twisted pairs, splice orientations, connector part information, and pictures when

necessary. The SFL technicians meant to be building the physical harnesses were heavily

consulted with in order to ensure a clear and easy to understand presentation of information to

ensure a smooth harness build with minimal mistakes. This marked the first time a graphical

approach to building the wiring harness had been used at SFL. Using the Altium Designer

software, the graphical manufacturing drawings could be easily navigated digitally during the

Page 47: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

39

build. Figure 25 below shows an example of one of the NORSAT-1 manufacturing drawings, for

the Payload cable.

Figure 25: Wiring harness manufacturing drawing example, Payload cable

The length information added to the manufacturing drawings is largely determined by the

proposed routing in the satellite, and was obtained in a number of ways, explained in the

following subsections.

3.6.1

In order to consider the wire routing at an early stage, before the satellite structure had been

ordered, the routing was virtually designed into the 3D solid model using the “Harness Design”

application in Solid Edge. All component connectors were ensured to be included in the model,

and wire paths were then introduced from connector to connector, and the specific routing could

be manipulated in 3D to the user’s choice. In order to simplify the model, bundles of wires

coming from the same connectors were modeled as a single wire of larger diameter.

Through this process, potential locations for wire tie down points were determined, and were

implemented either by including a pair of holes for a zip-tie to wrap through, or providing a

screw down point for a wire zip-tie mount or P-clip. Due to the relatively low impact of

including a pair of holes for zip-ties, additional tie down points were included in a number of

places (for example along the –Z tray where numerous wires pass) in order to allow some

flexibility on the wire routing when being implemented. The size and location of cutouts needed

in the structure in order to pass through wires could also be determined through this process. The

Page 48: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

40

solid model internal wire routing of all the wiring harnesses is shown below in Figure 26,

including the coax cables for the radio equipment.

Figure 26: NORSAT-1 solid model wiring

Having the entire wiring harness of the satellite incorporated into the 3D solid model provides a

number of benefits to the overall development of the project. It allows for changes to the wiring

to easily be realized and implemented, it allows for a graphical depiction of each wiring path,

that can be used to facilitate and check the physical wire routing during integration, and also

provides accurate lengths and mass properties for the included wires. In order to account for

inconsistencies of modeling the wiring, such as not accounting for the presence of other wires, a

length margin is added to each of the predicted lengths, as a percent slack compensation - this

adds some margin to each length for later refinement.

The panel harnesses, being much more simple, were not incorporated into the solid model, and

the routing design was done in 2D. The routing for the +Y and +X panel is shown as examples

below in Figure 27. Tie down points are incorporated directly next to each connection point in

order to ensure a compressive force exists at the mated connector. Note that the yellow markings

indicate the use of Kapton tape, “T.S.” indicates the placement of the temperature sensor, and the

thick blue wires indicate coax cables.

Page 49: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

41

Figure 27 Panel wiring for +Y panel (left) and +X panel (right)

Once the satellite structure was procured and arrived in house, the proposed solid model wire

routing could be validated using the physical structure and physical wires. It was found that the

lengths assumed from the solid model were quite accurate when the volume of wires in the area

was low, while the predictions in the dense wire areas tended to be slightly short. This was due to

the fact that the wiring in the solid model does not account for the physical presence of other

wires that it might have to detour in reality. However, with the added length margin to all of the

solid model estimates, this could easily be accounted for.

3.7

The main considerations for materials selection in the spacecraft include: low cost, low density,

low outgassing properties, non-magnetic, and desired strength. Some of these considerations are

alluded to in the driving requirements of Table 2, through Requirements #2, #4, and #5. The

majority of the spacecraft structure is made from aerospace grade aluminum 6061-T6. This is

commonly used in aerospace applications due to its lightweight, good thermal conductivity,

relatively high strength and low-cost characteristics. An alternate material was also considered,

magnesium ZK-31, which is used in the XPOD separation system due to its lower density and

similar characteristics to AL 6061-T6. This was however dismissed due to its higher cost, and

difficulty to machine.

Various other materials are also found in the spacecraft when different characteristics are

desired, usually for thermal purposes. Delrin acetal and G-10/FR4 are used when thermal or

Page 50: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

42

electrical isolation is desired, or when a non-abrasive surface is needed. Delrin is used for the

battery pack top and bottom collars, and the AIS antenna guides/hitch. G-10/FR4 was chosen for

the solar array wing spacers to achieve the necessary thermal isolation, and was chosen over

Delrin due to its greater performance at higher temperatures.

All of the fasteners used in the spacecraft are Stainless Steel 316, while some of the other

stainless steel components, such as spacers for the X solar cell coupons, are 18-8 stainless steel.

Helicoil inserts are used at every mounting point in the spacecraft structure (Requirement #15 in

Table 2). They are installed in-house, and are Nitronic 60 stainless steel. These fastener and

helicoil materials selections stems from a desire to only have non-magnetic components in the

spacecraft, so as not to affect the performance of the magnetometer.

3.8

A system level mass budget was kept up to date by the author using an SFL template mass

budget used for previous missions. All components from each subsystem are included in this

budget with the latest mass measurements. When a component is still under design, or is not in

house to physically measure, estimates are used, with added contingency. For example, for all of

the structural parts, mass estimates from the solid model were used in the budget until the

structure had been manufactured and arrived in-house to physically measure. For cables,

connectors or electronic components, mass numbers are taken directly from the datasheet. As

flight components are integrated, their masses are measured and remaining contingency in the

budget is reduced to zero. Each individual screw is accounted for in the Spacer & Fasteners sheet

of the mass budget, where a detailed list of each screw, spacer, and helicoil in the assembly is

kept. This proves useful not only to keep track of the incremental mass of each of these small

components, but also in providing a full materials list of the hardware needed. At the time of

writing, just before flight integration, the mass budget tallies NORSAT-1 to be a total of

~16.1kg, which still includes ~300grams of contingency. This current mass estimate meets

Requirement #4 in Table 2. The remaining contingency is largely due to not having all of the

payloads in house to measure, and the outstanding solar array wings.

3.9

The 3D solid model of the spacecraft and all associated parts were created and managed in the

software package Solid Edge. The solid model of the spacecraft proved to be an invaluable tool

for the structural design, allowing the designer to visualize structural components, keep track of

Page 51: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

43

the bill of materials and mass properties, and allows for seamless modification and iteration. All

the material and mass properties for each component are inputted and maintained in the solid

model in order to make use of the in-software mass and inertia calculations. Through this, a

fairly accurate estimate of the spacecraft’s center of mass and moments of inertia matrix can be

determined – which would be quite difficult to estimate analytically for something as oddly

shaped and mass distributed as NORSAT-1. These estimates are used as initial estimates in

designing the spacecraft attitude and control system, and are also used by the launch providers

when designing the satellite’s secondary payload accommodation.

A proposed thermal tape scheme for an intended orbit of the spacecraft was also incorporated

into the solid model. Having the tape scheme in the model allows for the accurate assessment of

the area coverage for each tape, and can be incorporated into the thermal model. A simple tape

scheme is desired, in order to have it easy to implement. Figure 14 of the earlier Section 3.2.2.1

shows a proposed tape scheme on each of the panels in the solid model for NORSAT-1, based on

the desired areas of coverage determined by the thermal engineer in [11]. Each different color

represents a different thermal tape with different optical properties. Note that this tape scheme is

not final, and is subject to change once the final orbit is confirmed.

3.10

The structural design for NORSAT-1 constantly evolved throughout the design process. Figure

28 below presents a look at the overall solid model at various stages of the project. The most

notable changes include the splitting of the large solar array into two solar array wings, moving

both AIS antennas to the same face, the placement of the Langmuir probe cassettes, the solar cell

accommodations, and the (not seen) inclusion of the internal separation plate.

Figure 28: NORSAT-1 design evolution

Page 52: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

44

3.11

The majority of the structure for NORSAT-1 was manufactured by a local machine shop, with a

few small parts being machined in-house by the author. As expected, the two trays of the primary

spacecraft structure were the most expensive due to their complexity, with the remaining parts

being a fraction of their costs. Two sets of each structural component were procured, in order to

have a “flight” structure, as well as a second “spare” structure. If anything were to happen to the

flight components, a spare component could be swapped in. The spare structure is also

strategically used in parallel for various testing while the flight structure is otherwise occupied.

Figure 29 below shows the full set of the flight structure in SFL’s class 10,000 clean room.

Figure 29: Majority of the NORSAT-1 flight structural parts

Page 53: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

45

4

In order to validate the structural design of the spacecraft, a finite element method (FEM) model

of NORSAT-1 was created for the Critical Design Review (CDR) using Siemens NX 8. The

model is continually updated post-CDR in order to reflect any major design alterations before

parts manufacturing, and to point out any unforeseen design flaws during the design process. The

analysis and post-processing of the finite element results are done using the NX Nastran solver,

and the results are then compared against the requirements to ensure a satisfactory design has

been achieved. The requirements at question are specifically #11 and #12 in Table 2, stating the

requirements set forth by the launch vehicle to ensure the design can withstand the expected

launch loads, with some added margin of safety. The details of this analysis for NORSAT-1 are

outlined in this section.

4.1

4.1.1

All of the NORSAT-1 structural components were modeled using hexahedron and tetrahedron

3D elements. Hexahedron elements were favored over tetrahedron elements for the majority of

the components for various reasons related to user setup time and computational costs.

Tetrahedron elements can be used for complex geometries and are generated automatically,

resulting in a very small user setup time. However, due to the often oddly shaped elements, quite

a large number of elements are needed for acceptable accuracy, which results in a high

computational cost, and the irregular triangular shaped elements often lead to undesirable

elements with high aspect ratios, skewing the results. Hexahedron elements are contrarily

manually meshed, and require significant effort in idealizing the part geometry for suitable

swept-meshing. Once properly meshed, the elements form a clean continuous mesh that is easily

customizable, and can provide a comparable accuracy with far fewer elements, resulting in a

smaller computational cost.

As part of the idealization process, non-essential features of each part are suppressed, such as

non-essential holes for tie wraps or components mounting, and non-load bearing rounds included

for part machinability. The remaining geometry is split into multiple sections, allowing

Page 54: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

46

individual swept-meshes to be manually inserted. Options also exist to split the geometry at areas

of interest, and provide meshing controls, to facilitate obtaining a finer mesh at specific

locations. For example, the area around each mounting hole is split, and a mesh control is added,

to allow for a higher density of nodes to be present at these anticipated high areas of stress.

Figure 30 below shows the idealizing and meshing process for the large attachment bracket.

Figure 30: Large attachment bracket part (left), idealized part (middle), meshed (right)

Material properties are specified for each fully meshed part and applied to each created element

to ensure the correct behavior is exhibited in the simulations.

4.1.2

Due to the complexity and orthotropic material behavior of honeycomb sandwich structures used

for the solar array wings, an alternative method of modeling has to be considered. Given the

popularity in the area of modeling composites for finite element analysis, many modeling

methods exist and have been thoroughly researched and implemented for specific situations. A

study on alternative modeling methods for the honeycomb solar panel used on NEMO-AM has

been completed by Dumitru Diaconu in [12], and due to its similarity to NORSAT-1’s solar

panel wings, a similar approach to the concluded best method will be taken. The study compared

three modeling methods to some known data: an accurate three-dimension (3D) model of the

panel using a detailed model of the honeycomb core and a face sheet mesh, a two-dimension

(2D) mesh with composite shell properties, and an equivalent panel method where the sandwich

panel properties are converted to an equivalent solid panel. It was concluded that the 2D mesh

with composite shell properties method was able to produce very similar and accurate results to

that of the accurate model and equivalent panel methods, with just a fraction of the nodes and

computational effort. This was the chosen method used for modeling the NEMO-AM

honeycomb panel, and is thus the chosen method for the NORSAT-1 solar array wings.

Page 55: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

47

The “PCOMP” composite shell implementation in the software package is used to create a 2D

mesh at the mid-plane of the part; the material properties and thickness for each layer of the

sandwich panel, the two face skins and the honeycomb core, are then inputted into the meshing

details in order to exhibit a realistic behavior of the composite. Added mass due to the adhesives,

inserts, and solar cells are added to the panel material properties as non-structural mass, which is

evenly distributed across the full geometry.

4.1.3

Much of subsystem hardware, including but not limited to, the magnetorquers, printed circuit-

boards (PCBs), the battery pack, reaction wheels, and antennas were modeled in the FEM

assembly as manually defined zero-dimensional (0D) lumped mass elements, located at the

component’s center of mass, with one-dimensional (1D) rigid element links connecting them to

their mounting locations in the spacecraft. This is done for the components that are non-load

bearing, and whose geometries would not affect the analysis results.

The payloads, contrarily, were fully modeled with 3D elements to represent their structural

design, using the latest solid model provided by the payload provider. However, in cases where

significant information was still not confirmed at the time of analysis, a similar 0D lumped mass

approach was taken to best represent the payload internal design and mechanical interface with

the satellite.

Fastener modeling was exclusively performed using infinitely stiff 1D rigid links to represent the

shank of the screw and load path between connected parts, as well as less stiff 1D rigid links to

represent the contact area of the screw head. This method has been used to model previous SFL

spacecraft structures and has been deemed to produce acceptable results. An example of this

implementation is shown in Figure 31.

Figure 31: FEM fastener modeling

Page 56: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

48

4.2

4.2.1

The constraints for the finite element simulations aim to represent as closely as possible the

method of restraint of the spacecraft inside the XPOD-Duo separation system under launch

conditions. The ball cup interfaces at each of the eight satellite feet and the XPOD-Duo are

designed to constrain the lateral movement of the satellite, and therefore lead to fixed

translational constraints at each of the satellite feet in the X and Z directions. Since the satellite is

restricted in the Y directions when the XPOD door is closed, translational motion in the

longitudinal direction (Y axis) is fixed on the –Y feet, interfacing with the XPOD door. On the

+Y feet (located at the base of the XPOD-Duo, interfacing the with pusher plate), translational

motion is left unconstrained, in order to allow compression under longitudinal accelerations.

Finally, a compressive load is applied in the Y axis on the +Y feet in order to represent the pre-

load of the XPOD-Duo spring in the pusher plate. These boundary conditions applied to the full

FEM model are shown in Figure 32. Surface contacts were not modeled in order to minimize the

complexity of the simulations. By not including these, stress concentrations are over-estimated,

due to the lack of friction which would otherwise impede motion, leading to more conservative

results.

Figure 32: NORSAT-1 finite element model with applied boundary conditions

4.2.2

The applied loads represent the worst-case loading that NORSAT-1 will likely see while loaded

in the XPOD-Duo inside the launch vehicle. At the time of analysis, a launch vehicle had not

been confirmed; therefore all launch vehicles under consideration are taken into account to

determine the greatest acceleration likely to occur. Each launch vehicle specifies various

+Y

+X

+Z

Page 57: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

49

vibration tests to be done on the satellite for qualification purposes prior to being accepted as a

secondary payload. Of these tests, random vibration is likely to impose the highest accelerations

to the spacecraft, and is therefore used to determine the highest likely quasi-static acceleration

that the spacecraft might see due to vibrations. This can be done by finding the standard

deviation of the accelerations experienced by the satellite if exposed to the composite random

vibration spectrum, which includes the most severe qualifications loads from each considered

launch vehicle. This static load determined for NORSAT-1 is 10.46g root-mean-square (Grms).

Applying a 5-sigma level of margin onto this load in order to account for various uncertainties in

the modeling, apply a level of safety margin, and instill a high level of confidence in the design,

results in a load of 52.3g. This load is applied as a global quasi-static acceleration in the FEM

model acting on the entire spacecraft, and is distinctly applied in each axis for six separate

loading scenarios. In order to reduce the number of simulation cases, the acceleration was

applied in three axes at once, resulting in just two separate cases to analyze with a 90.6g

magnitude load. Using this 5-sigma acceleration value of the random vibration represents a

conservative approach in analyzing the spacecraft structure, and avoids the necessity of a more

complex dynamic analysis.

4.3

4.3.1

Quasi-static stress results from simulations completed in June 2014, shortly after the CDR and

prior to part manufacturing, are compiled in Table 4 for each structural component. The relevant

requirement (#11 in Table 2) states that the spacecraft must have positive stress margins under

expected launch loads at 5-sigma. Stress margins are calculated as a measure of requirements

verification. Since the safety factor is incorporated in the applied load, a positive margin on the

yield stress for each component represents a satisfactory design. These margins of safety are

calculated as follows:

(4.1)

A margin of safety of 0% indicates that the design is just barely capable of withstanding the

applied loads (with safety factors) without failure, while a margin of safety of 100% represents

no loads being applied, and therefore would never result in failure. For example, for the

component +X Riser, the maximum applied load seen in the simulations is 161MPa, and the

Page 58: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

50

failure load is represented by the yield stress of the material used (aluminum 6061-T6, 276MPa).

Using the above equation (4.1), a margin of safety of 42% can be calculated.

From the cases analyzed, all the stress margins are positive, therefore the design is acceptable. A

high level of safety margin is carried through in this analysis in order to instill confidence in the

structural design, since no physical structural tests will be performed on the spacecraft before

launch, other than an acceptance vibration test on the flight spacecraft at somewhat smaller load

levels.

The lowest stress margin is seen on the +Z tray near one of the –Y feet, a screen capture of this

result is shown below in Figure 33. As expected on both trays, relatively large stress

concentrations exist at all eight satellite feet, since these feet are the main load path to the rest of

the structure, and the only points of restraint in the XPOD-Duo. Small fillets were added to both

tray designs at the junction between the feet at the rest of the tray in efforts to minimize this

stress concentration.

Figure 33: Screen capture of –Z tray FEM stress results

Page 59: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

51

4.3.2

The main requirement on the displacements seen under this worst-case load is to ensure that the

panel mounted solar cells would not see significant bending causing them to crack. Each solar

cell is limited to ~0.6mm displacement, which comes from a worst-case geometry based

calculation derived from the cracking radius of the cells (~2 meters). Allowable deflections on

each of the solar cell mounted panels are then determined geometrically based on this cracking

radius. In the analysis, all of the solar panels see a displacement significantly less than their

allowable deflections, with an average (plus 3-sigma) of ~0.37mm. Since all of the

displacements seen yield positive margins, they are acceptable. The displacements are also

monitored to ensure that no harmful contact is made between components themselves and with

the separation system. This is where the displacement requirements stem for the tray rails and

attachment brackets, the amount of displacement that would cause contact with the XPOD-Duo.

For all other components where no specific displacement requirement exists, a maximum value

of 1mm is used, which represents in many places, their distance to neighboring components, and

a conservative upper limit on their deflection. The displacement margins are calculated in a

similar fashion to the stress margins using equation (4.1).

Table 4: Stress and displacement analysis results summary

Component

Yield

Stress

(MPa)

Displacement

allowed

MAX

Displacement

(Absolute)

(mm)

Displacement

Margin (%)

MAX Stress

(MPa)

Stress

Margin

(%)

+Z Tray 276 1.00 0.17 83 223 19

+Z Tray Rail 276 0.41 0.17 59 223 19

-Z Tray 276 1.00 0.28 72 187 32

-Z Tray Rail 276 0.41 0.2 52 187 32

+Y Panel 276 2.82 0.21 93 192 30

-Y Panel 276 2.82 0.13 95 207 25

+X Panel 276 2.82 0.28 90 145 47

-X Panel 276 2.82 0.27 90 170 38

+Z Panel 276 2.28 0.2 91 142 49

-Z Panel 276 1.00 0.18 82 97 65

+X Riser 276 1.00 0.15 85 161 42

-X Riser 276 1.00 0.15 85 161 42

Separation Plate 276 1.00 0.19 81 85 69

Small Separation

Plate 276 1.00 0.14 86 54 80

Page 60: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

52

Reaction wheel

Bracket 276 1.00 0.22 78 89 68

Large brackets 276 1.50 0.75 50 116 58

Half brackets 276 1.50 0.44 71 83 70

VHF Plate 276 1.00 0.21 79 52 81

Radio Cover 276 1.00 0.33 67 83 70

LP Electronics 276 1.00 0.2 80 48 83

LP Cassettes 276 1.00 0.17 83 71 74

CLARA 276 1.00 0.38 62 107 61

CLARA sun

sensor bracket 276 1.00 0.78 22 111 60

X Solar Cell

Coupons 276 1.00 0.25 75 83 70

Solar Panel Wings 325 2.82 1.2 57 157 52

Solar Panel Wing

Skin 591 2.82 1.2 57 359 39

Note that the maximum displacement values shown are absolute and at the system level,

therefore they don’t necessarily represent the amount a component is deformed. For example, the

displacement seen by the solar panel wings is largely due to the cantilevered method on which

they are mounted to the attachment brackets. The applied loads cause the brackets to deflect, and

the panels simply remain attached to the deflecting brackets. Through hand calculations based on

the relevant geometries, it can be found that the actual panel deflection, not due to the brackets,

is roughly 0.13mm.

4.3.3

Requirement #12 in Table 2 states that the spacecraft must have a first natural mode greater than

90Hz in order to comply with the worst-case launch vehicle requirements. Results from the

spacecraft modal analysis indicate that the first natural frequency (FNF) of NORSAT-1 is 144Hz

and is a local mode of the solar panel wing and attachment bracket oscillating back and forth as a

cantilevered mass. The result is shown below in Figure 34; it is above the minimum required

FNF of 90Hz with a 60% margin of safety and is therefore acceptable. The second wing shows

identical behavior at much the same frequency and numerous similar modes of the solar panel

wings and brackets are found around this frequency range. Note that the displacements shown in

the figure is not meaningful since this is a modal analysis with no specified loads.

Page 61: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

53

Figure 34: Screen capture of simulation showing first mode of NORSAT-1 at 144Hz, and

subsequent mode frequency values

Page 62: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

54

5

Honeycomb composite sandwich panels are commonly used in the aerospace industry due to

their numerous advantages over solid materials. Their high specific stiffness-to-weight ratio

allows them to be an extremely low mass solution for large structures where mass is of main

concern. In space structures, a common use for honeycomb panels is as large solar arrays – these

panels have a sole purpose of having numerous solar cells mounted on them for power

generation. The high stiffness of the panels allows them to be very simply supported, further

reducing their mass. Figure 35 below depicts their advantage over solid materials.

Figure 35: Relative stiffness and weight of sandwich panels compared to solid panels. Note

that the numbers shown are normalized to the solid material numbers [13]

In order to satisfy the power requirements required for simultaneous payload operations,

NORSAT-1 needs a total of six strings of eight solar cells (48) able to generate power during

operations of CLARA while pointing at the sun. Due to the spacecraft geometry proposed, the

largest face can accommodate at most two strings of eight solar cells, for a total of 16 solar cells

– less than half of what is required. If the satellite geometry were to grow in order to

accommodate the required cells, it would result in an inefficient mass structure, with large un-

used internal volume, and would no longer be compatible with the XPOD-Duo deployment

system. Because of this, it was evident that a dedicated solar panel would be required to meet the

requirements.

Page 63: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

55

Due to a desire to minimize risks, fixed wing panels are favored over deployable panels for

NORSAT-1. Also, due to the size required to fit the 48 solar cells, it was decided to use a

honeycomb composite structure in an effort to not add significant mass to the spacecraft.

Initially, the proposed design for the large solar array matched closely to the ongoing SFL

project, NEMO-AM, which uses a large honeycomb sandwich panel as their main solar array and

is of similar size to NORSAT-1 (shown in Figure 36A). A second ongoing SFL project, NEMO-

HD, also uses honeycomb composite panels as various structural and solar panels (Figure 36B).

The panel design for NORSAT-1 is able to leverage the experience gained through these two

projects. This section outlines the design of the two large honeycomb composite solar panels

used on NORSAT-1.

Figure 36: NEMO-AM (A) and NEMO-HD (B) honeycomb panels

5.1

For NORSAT-1, there are very few specific requirements for the honeycomb panel design, other

than its main function; however, various programmatic drivers shaped the design approach. After

the initial proposal of a single large array that covers the entire spacecraft similar to NEMO-AM,

some preliminary thermal analysis concluded that it would be more beneficial to have a direct

line of sight to the spacecraft structure. From this, the panel was split into two separate solar

panel “wings”, which left the area of the satellite in between available for thermal control. The

NORSAT-1 “large solar array” refers to all of the solar cells located on the CLARA aperture side

(-Z side), which includes both front surfaces of the wings. Some additional general requirements

of these solar panel wings are explained below:

Page 64: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

56

1) The large solar array shall be capable of successfully mounting 48 Azur 3G solar cells.

o This is the only size-driving requirement on the panels, however they are also

kept as small as possible in order to minimize the overall bus volume.

2) The large solar array shall meet the same structural and modal requirements as set for the

entire spacecraft (positive stress margin, first natural frequency greater than 90Hz).

o This ensures it will meet the system level requirements, however, because the

panels are not entirely independent, their design is largely interrelated with the

entire structure and will be analyzed together as well.

3) The large solar array shall be able to withstand temperatures ranging from -100°C to

+80°C

o Predicted temperature range from the thermal model. This range is large because

the panels are thermally isolated from the spacecraft bus.

4) The large solar array should be made of common components/materials that do not

require a long lead time, and should be composed of similar materials to that used by

NEMO-AM and NEMO-HD

o Schedule is one of the main concerns for the project. Because the panel is a fairly

simple design, the details should not be over complicated, resulting in a long lead-

time. Using heritage components from previous projects can also reduce some of

the risk associated with the foreign materials.

5) The two solar array wings should be identical.

o This is done in order to save money during procurement of the panels, and allows

for a modular design.

6) The large solar array should be designed such that it can be integrated onto the satellite

last.

o This is done to facilitate the integration process of the satellite, and to minimize

the handling of the spacecraft while these expensive/delicate panels are installed.

A large goal in the panel design was to not add significant cost and time to the project; the panels

were kept as simple as possible in order to achieve this. For example, additional components,

such as antennas and sun sensors, were excluded from being mounted on the panels, for further

complication of their design. As was shown in Chapter 4, the mass of these panels directly affect

Page 65: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

57

the first natural frequency of the spacecraft, therefore, extra effort is placed in keeping the mass

of each panel as low as possible and accurately predicted.

5.2

Derived from the above requirements, and fitted to the relatively mature spacecraft structural

design, a design concept for the honeycomb solar array was realized. Two identical “wings”

make up the entire solar array, each supported by five mounting inserts, attached to the satellite

via three aluminum attachment brackets. The size of each wing is 20cm x 50cm, so as to each fit

three full strings of eight solar cells, to make up the entire array of 48 cells; additional area is left

for the wiring of the cells, as well as ground support equipment (GSE) holes for various panel

protectors and handling measures. Although the 50cm dimension is larger than the longest

dimensions of the satellite bus, the panels are mounted slightly off-center in order to allow the

spacecraft to rest on its –Y face, and facilitate the installation of the panels at a late stage. Also,

to assist in the late installation, each panel is fitted with a dedicated connector to carry the wiring

of each panel’s solar cells, and a single temperature sensor, into the spacecraft avionics; this

connects to a filtered skin connector on the –Z tray into the avionics bay of the spacecraft. In

order to keep the panels as small as possible, the mounting inserts used are blind threaded holes;

this allows the front side of the panel to be fully available for the placement of the solar cells. A

depiction of this proposed concept is shown in Figure 37. A single string of eight solar cells is

also placed centered on the back of each panel, to act as the –Z face solar array. An extra insert

was included in order to keep the panel design and solar cell placement/wiring identical on each

panel.

Figure 37: NORSAT-1 solar array concept design

Page 66: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

58

The attachment brackets used stem from the attachment bracket design used on the NEMO-AM

mission, mounting to the spacecraft and the panels in a similar fashion. However, due to the

separation of the large panel into two small panels, there exists a cantilevered mass of each of the

panels on the extended portion of each of the brackets, which will result in a low frequency mode

of the satellite. In light of this, the brackets are reinforced in the direction of bending by using an

I-beam concept to keep the mass as low as possible; this is shown in Figure 38. The appropriate

sizing of these brackets is designed through an iterative process during the finite element analysis

of the entire spacecraft in order to maximize the first natural mode; this was detailed in Chapter

4.

Figure 38: Wing attachment bracket design

The remaining honeycomb panel specifics, such as core/skin thickness and other relevant

honeycomb parameters, are designed based on failure criteria of the sandwich panel and insert

design.

5.3

Due to the complexity of sandwich panels, several failure modes exist as opposed to general

isotropic materials. Separate failure modes exist for the sandwich panel itself, and for the inserts

that are added for fixation purposes.

5.3.1

Sandwich panels have many failure modes as described in various literature, including in [13]

and [14], which were referenced heavily during this study; some of these are briefly described

below and in Figure 39.

Page 67: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

59

a) Face yielding/fracture - Tensile or compressive failure of the face sheets

b) Core Shear – Core material shear failure

c) Face wrinkling - Local face sheet buckling inwards due to core compression failure

d) Delamination – Local face sheet buckling outwards due to delamination

e) General buckling - Panel buckling due to in‐plane compressive loads

f) Shear crimping - When buckling results in additional, localized core shear failure

g) Face dimpling - Multiple inter‐cell buckling due to large core material cell sizes

h) Core Indentation – Failure of the core due to localized pressure

Figure 39: Sandwich panel failure modes [15]

All of these failure modes were investigated for the NORSAT-1 honeycomb panel wings,

however, due to the expected loading on the panels many of these modes are not applicable. For

example, failure modes c), d), e), f) and g) only occur when large compressive loads are applied

to the panels. Due to the proposed method of mounting, the NORSAT-1 honeycomb panels will

be simply supported by some fixation inserts, and will carry no loads other than their own weight

under a directional acceleration, therefore, will see very little to no compressive loads. The

remaining three failure modes, face/core yielding/facture, core shear, and core indentation, were

analyzed as part of the panel design. These failure modes, however, are not expected to be design

drivers. Due to the mounting method of the panels and the expected loads, insert failure is likely

to dominate. The equations for determining the relevant stress via these three failure modes are

displayed below for a simply supported honeycomb panel.

Page 68: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

60

Facing Stress Core Shear Stress Local Compression

(5.1) (5.2) (5.3)

Where:

M = Maximum bending moment

h = Distance between facing skin centers

tf = Thickness of facing skin

b = Beam width

F = Maximum shear force

P = Applied load

A = Area of applied load

The deflection of the panel is also of interest due to having sensitive glass covered solar cells

directly mounted, which can crack if a large enough deflection is seen. The deflection of a

simply supported honeycomb panel “beam” can be expressed as:

(

) (

) (5.4)

Where:

kb and ks = Deflection coefficients from [13]

D = Panel bending stiffness

S = Panel shear stiffness

l = Panel free length between supports

The above equations are used to validate the NORSAT-1 proposed panel make-up, and the

relevant results and margins of safety are displayed in Table 5 of Section 5.4. The boundary

conditions of the panel are simplified to approximate a simply supported beam panel with an

applied load. The fifth center mounting point of the panels is ignored, and the two mounting

points on either end of the panel are approximated as the simple supports, to approximate a

scenario as shown in Figure 40.

Page 69: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

61

Figure 40: Simplified loading scenario for sandwich panel failure mode calculations [14]

5.3.2

A common failure occurrence in sandwich structures is through insert failure. Inserts are the

main method for adding fixation points to sandwich panels. This is often done by a post-cold

process, whereby an insert is implanted into an already laid-up sandwich panel and fixed using a

two-part epoxy resin system, or “potting compound”. Because of this post-process

manufacturing, they often tend to be a weak point of failure, depending on how they are being

used, because of the discontinuity placed in the sandwich panel. Although significant work has

been done to predict the behavior of these insert systems, they continue to be a somewhat fickle

point of failure because of their sensitivity to the small details in the manufacturing process.

Insert pullout tests are sometimes done in order to characterize the insert installation method,

however these are often forgone for small projects due to their expensive nature, and panels are

instead often compared to relevant similar test data. Generally, predicted insert failure through

analytical methods prove to be a conservative approach when compared to actual test data,

barring any manufacturing defects [16].

There are four main types of insert loading scenarios, which can each lead to various failure

modes of the face sheets, potting compound, honeycomb core, or insert itself, due to the added

discontinuity in the panel. These are displayed in Figure 41 below, and are each discussed in this

sub-section. The Shur-Lok and ESA Insert Design handbook ([17] and [18] respectively) were

referenced heavily during this study.

Page 70: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

62

Figure 41: Insert loading scenarios [18]

Loading scenario a) involves an out-of-plane tensile or compressive load. A tensional load leads

to a failure mode commonly known as pullout, which depends greatly on the core thickness of

the honeycomb and has little effect from the face sheets. An expression provided in [18] for

determining the critical pull out force ( ) of a partially potted (blind hole) insert in a

sandwich panel where the core thickness is much larger than the skin thickness is shown below.

A diagram of the scenario is also shown in Figure 42.

( )

(5.5)

Where Pcrit is the static load carrying capability for a fully potted insert under the same

conditions, evaluated by:

(5.6)

And where:

bp = Effective potting radius

hp = Potting height

c = Core thickness

τccrit = Core shear strength

σccrit = Core tensile strength

d = Distance between the face sheet middle surfaces

Page 71: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

63

Figure 42: Partially potted insert under tensile load [18]

An effect of the insert’s distance to the edge of the panel also accounts for an appreciable

reduction in strength of the insert. This factor is incorporated into relevant calculations through

means of amplification on the applied load, as detailed in the Edge Influence section of [18].

For a partially potted insert under out-of-plane compressive load, the critical failure load is

almost always larger than the critical failure load under tensile load, due to the added strength of

the bottom face sheet and the increased compressive over tensile strength of the core, therefore

its calculation is not considered for the NORSAT-1 panels.

Loading scenario b) involves in-plane shear loads on the insert, which can result in various

failure modes of the face sheets and core material. For a sandwich panel with CFRP face sheets

and aluminum honeycomb core, a permissible shear load ( can be determined based on the

combined properties of the core and face sheets, this formula is shown below for potting radius’

(bp) less than 11mm [18]. In addition, four independent failure modes of the CFRP face sheets

can occur: Tensile, Shear-out, Dimpling, and Bearing failure; these are shown in Figure 43.

(5.7)

Where:

bp = Effective potting radius

τWcrit = Core shear strength in (weaker) W direction

σfy = Yield strength of face sheets

f = Thickness of upper face sheet

Page 72: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

64

Figure 43: Face sheet in-plane insert failure modes [18]

The detailed predicting formulae of the above face sheet failure modes are comprehensively

explained in the Insert Design Handbook [18]. Due to the relatively high strength of CFRP, these

failure modes tend to be quite insignificant, but depend heavily on the chosen thickness of the

face sheets. Another large influence is distance from the insert to the edge of the panel – this has

a substantial influence particularly on the shear-out failure mode.

Loading scenario c) involves an in-plane torsional load to an insert, which could lead to shear

failure between the potting adhesive and honeycomb core. This is greatly minimized through the

use of multiple inserts, and thus is not a concern for the NORSAT-1 panels. The only torsional

load that will be seen would be the screw torque, which will be substantially less than the failure

torque.

Loading scenario d) involves an out-of-plane bending moment placed on the insert. Inserts are

typically not advisable to be placed under such loading conditions because of their poor

performance under such conditions. Much like reducing the torsional load, this is commonly

mitigated through the use of coupled inserts, which can convert the bending moment to a

tensional and compressive load. It is further mitigated through the use of a clamping surface (or

washer) that is larger than the insert diameter, capable of spreading the load across the face sheet.

In light of these simple mitigation methods for failure through torsional and bending loads, these

insert failure modes were not considered for the NORSAT-1 panels.

Page 73: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

65

5.4

Given the proposed design concept and the relevant failure modes, the specifics of the

honeycomb panel for NORSAT-1 can be realized. Figure 44 shows a breakdown of the resulting

sandwich panel design.

Figure 44: NORSAT-1 solar panel wing design

The panel specifics were derived through trial and error, in an effort to minimize mass, obtain

positive safety margins, and make use of standard sizes and parts (i.e. stock inserts, common

CFRP ply thicknesses and orientations, standard honeycomb core details). Aerospace grade

honeycomb core material is used, which is commonly either Al 5052 or Al 5056, and

incorporates vented honeycomb cores that make them appropriate for use in vacuum. Carbon

Fiber Reinforced Plastic (CFRP) skins are used due to their low-weight, and minimum distortion

in extreme temperatures – this matches closely with the solar cells, which have a glass cover

component with very low thermal expansion. In minimizing the difference in thermal expansion

of the face sheet and the solar cells, it also minimizes the potential stresses that could occur due

to mismatch in thermal expansions and cause solar cell cracking. Based on these panel details, an

estimated mass of each panel, prior to solar cell laydown, is approximately 300 grams.

In arriving at these panel details, the relevant failure mode margins are shown below in Table 5.

In all cases, the applied gravity load is matched to the load used in the finite element analysis of

the spacecraft (Chapter 4), which covers all expected shock values, taken as 52.3g’s. The load is

applied in the worst-case direction for each failure mode, and is distributed among four of the

inserts, neglecting the middle insert, and assumes each insert has the worst-case edge influence;

Honeycomb Panel Wing Design Details

Honeycomb core material,

cell size, density, foil

thickness

Al 5052, 4.0mm, 3.8pcf,

0.03mm

Honeycomb core

thickness

9.5mm

Skin material, density, #

of plies, ply orientation

CFRP, 1446 kg/m3, 2 ply,

0/90

Skin thickness 0.25mm

Insert material, Type,

diameter, Height

Al 2024-T4, Fully Potted,

11mm, >6mm

Page 74: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

66

combined loading was not considered. These margins of safety are calculated similarly to as

described in Section 4.3.1 using equation (4.1); since the safety factor is incorporated into the

applied load, a positive margin represents a satisfactory design.

Table 5: NORSAT-1 honeycomb panel design failure modes summary

Panel

Failure mode

Facing Stress

Core Shear Stress

Local Compression

Deflection

Margin of Safety

96% 93% 99% 76%

Insert

Failure mode

Critical Shear

Tension Shear-out Dimpling Bearing Pullout

Margin of Safety

87% 98% 64% 30% 94% 83%

As seen, all of the calculated margins of safety are positive, and therefore the panel design is

satisfactory. The most critical and relevant failure modes proved to be out-of-plane insert pullout

and in-plane insert shear-out. The pullout failure is most heavily dependent on the core thickness

and insert size, where having a thicker core, and larger diameter insert would result in a higher

resistance to failure. However, the shear-out failure is most dependent on the edge influence

caused by the location of the insert relative to the panel edge, where a larger diameter insert

would only make this distance smaller, and more prone to failure. Therefore these variables had

to be balanced to ensure acceptable positive margins in all affected failure modes.

All other modes of failure proved to be fairly irrelevant due to the low expected loads on the

panel. The deflection of the honeycomb panel was found to be as low as 0.9mm, which is well

below what would be required to cause the panel mounted solar cells to reach their cracking

radius of two meters. Although the insert dimpling failure mode is seen to be relatively low, this

failure mode is difficult to predict analytically due to its dependence on the specific panel

geometry, type of loading, and boundary conditions and is usually determined through test. As

seen, the panel was not optimized down to minimum margins; this is due to the relevant

uncertainties in the panel manufacturing that cannot be anticipated, thus higher margins are kept

in order to reduce risk and instill additional safety into the design. In addition to this, very small

Page 75: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

67

mass and cost savings are gained through reducing the panel size further, therefore a slightly

over-designed panel has little negative impact on the project.

Also exists is an abundance of data regarding the tensile and compressive strength of post-cold

bonded inserts in [19]. The data is presented in graphical form, where a sandwich panel of

particular make-up and insert size is considered, and shows how the minimum and average insert

strength values vary as a function of the core height. The data is produced by means of an

analytical method, though has been compared and validated through empirical testing [20].

According to [19], the minimum and average tensile strengths displayed are applicable to

honeycomb panel design without further investigation, provided that the core properties match,

the manufacturing of the inserts were performed in accordance with their specified quality

assurance, and the load is carried by multiple inserts. Although not certain that the quality

assurances of the NORSAT-1 panel will match, and the different facing material used, given that

the main parameter of the pull out strength is the core properties, these values can be regarded as

ballpark values. The graphs for tensional and compressive load on a similar honeycomb core and

similar sized insert are shown in Appendix A. As seen, from these graphs, the minimum pull out

strength capability is above 500N in both cases, and the typical values are near 1000N. These

values relate fairly close to the predicted worst-case pullout load for the NORSAT-1 panel

inserts, which was found to be approximately 450N per insert.

5.4.1

Because the NORSAT-1 panel isn’t very sensitive to small changes in the honeycomb make-up

details, procurement of the panels involved compromising on certain details in favor of

items/materials that were in stock, which would result in quicker panel procurement. For

example, the panel manufactures suggested switching the insert material from Al 7075-T73 to

AL 2024-T4 due to them finding the latter material in stock. In each case, the relevant change

did not significantly affect the failure modes of the panel, and would only make the panel more

robust. Some of these changes would result in additional mass over the original panel design or

modifications to the attachment brackets; however it was deemed that the schedule would take

priority. The NORSAT-1 honeycomb panels are currently being manufactured and discussions

regarding various material/design compromises are on-going with the manufacturer.

Page 76: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

68

6

In an effort to reduce the overall launch volume of NORSAT-1, many of the required large

protruding components were made deployable. By reducing the overall volume via deployable

components, NORSAT-1 becomes easier to accommodate as a secondary payload aboard a third-

party launch provider. On NORSAT-1, two very high frequency (VHF) receiving antennas, and

four Langmuir probe booms have been made deployable, for a total of six deployed components.

6.1

The Space Flight Laboratory (SFL) and the Norwegian Space Centre (NSC) have collaborated

on a number of Automatic Identification System (AIS) satellites. Numerous previous satellites

on-orbit of the GNB size, including AISSat-1 and AISSat-2, use a single pre-deployed VHF

antenna as shown in Figure 45. The pre-deployed method being favored because of the reduction

of risk over deployable components (i.e. a deployable component always has a single point of

failure – if it does not deploy properly). The size of the GNB satellite bus is less than half the

volume of the NEMO bus type, causing the added volume of the antenna to be less influential on

the restrictive nature of the launch volume.

Figure 45: AISSat-2 pre-deployed VHF antenna approximate dimensions

In the case of NORSAT-1, however, two orthogonal antennas are required, rendering the volume

consumed by pre-deployed antennas to be significantly larger. The added volume by two

orthogonal antennas of each ~50cm length would restrict the number of potential launch

opportunities, and would place additional constraints on the launch vehicle in order to

Page 77: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

69

accommodate the satellite. Because of this, deployable antennas were favored over pre-deployed

for the two VHF antennas on NORSAT-1.

6.1.1

Tape springs are a common solution to deployable components in space applications, due to their

lightweight, energy storage, and stiff characteristics. An on-going GNB sized satellite project to

be launched in the near future was already at a mature design stage at the time of decision, and

uses a tape spring system as a deployable VHF antenna (EV-9). Because of SFL’s heritage and

experience in their design and use as VHF receiving antennas, tape springs are used as the

deployable VHF receiving antennas for NORSAT-1 as well. No other deployable antenna

systems were considered, because of the believed favorable simplicity of the tape spring

approach.

Only three main requirements exist for the antenna design:

1) The antenna shall be a simple 50-Ohm quarter-wavelength (~46 cm) monopole with no

specific beam width.

2) The two antennas shall be mounted orthogonal to each other (Requirement #9 in Table

2).

3) The deployment shall be able to be performed and tested in a 1g environment

(Requirement #14 in Table 2).

The first requirement limits the antennas to materials that can be used as antennas (conductive

materials), and also sets a length to achieve the necessary frequency (162 MHz), while the

second requirement specifies their orientation. In addition to these, the antennas should also be

easily stow-able, and able to be installed last, in order to simplify ground handling and testing.

An additional desire for the deployment design is to have the antennas deploy passively, not

requiring additional equipment such as burn wires or on-orbit commanded deployment systems –

this can significantly simplify the design and reduce the risk of failure.

6.1.2

Research was conducted to better understand tape springs and use them as intended, which

ultimately shaped many of the design decisions. These activities are listed and explained below.

Most of these activities were explored using Commercial Off The Shelf (COTS) carbon steel

tape measure tape springs.

Page 78: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

70

Width

Typical tape measure tape springs come in two widths: one-inch and half-inch. The difference in

width have little effect on the antenna performance, however the larger width tape spring has a

larger mass, causing gravity to have a larger effect on it during testing on the ground.

Machining

Machining a tape spring can be a difficult task, due to its extremely thin and curved cross

section. For cutting the tape spring to size, it was found that a sharp pair of shears would do the

best job, while a duller pair may cause the corners of the cut to become slightly deformed, and

would affect the continuity of the tape spring. When making holes, a drill press was found to be

the preferred method, over other methods such as hole punching. Best results were achieved

when drilling into the concave side of the tape spring, and ensuring a hard piece of material was

below the tape spring. Clamping of the tape spring to the hard material as close as possible to the

location of the drilled hole was also necessary to ensure the tape spring would not get caught in

the drill bit blades. Punching the hole-location with a center punch prior to drilling also helps a

great deal to ensure a smooth drilling process. An improperly drilled hole in the tape spring can

easily distort the continuity, and will cause the tape spring to perform undesirably.

Critical Bend Radius

Since the method of deploying tape springs often involves bending them in some fashion, and

allowing them to store the energy needed to deploy, critical bending limits were explored.

Through testing on COTS half-inch tape measure tape springs, the following bending limits were

found. Bending was performed with the concave side of the tape spring facing inwards to the

bend, as shown in Figure 46.

Figure 46: Tape spring bend radius

Page 79: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

71

< 4mm radius = plastic deformation.

< 6mm radius = No visible deformation or decrease in performance, however the tape spring has

a preferred kink spot when bent in that area.

> 6mm radius = No change in performance.

Materials

All COTS tape measures are made of carbon steel, and often have a yellow plastic coating on

them for annotations - these materials aren’t ideal for space applications. The plastic coating can

easily be removed using a paint remover and/or through abrasion. The carbon steel material is of

slight concern, due to its magnetic properties. The dipole created can affect the readings from the

on-board magnetometer used for attitude sensing. It is because of this concern, that another

ongoing SFL project, CanX-7, has opted to create a custom tape spring in-house, out of non-

magnetic Copper Beryllium (CuBe), shown below in Figure 47. Their use for the tape springs,

however, is slightly different than for NORSAT-1. On CanX-7, these tape springs are used for

deployment of a large drag-sail, and are stowed in a very small containment unit, having the tape

spring wrapped around itself several times. Once deployed, the booms are more than double the

length of the planned NORSAT-1 VHF antennas, and thus, if steel tape springs were used, they

would cause a larger dipole to be seen which affects the performance of the limited attitude

determination hardware on board.

Figure 47: Tape spring materials explored

Though the same curvature radius was created, the CuBe tape springs were found to have

approximately 20% less stiffness than their steel counterparts. This causes the antennas to be

more influence by gravity during ground testing. Additional leftover CuBe tape spring material

was only available in the one-inch width size, and it was not desired to go through the effort of

time/money to manufacture thinner ones.

Stowage

A similar stowage technique to that used on EV-9 (using a tooling ball) was desired, however, in

NORSAT-1’s case, there were two antennas to hold down rather than just the one. The

Page 80: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

72

difficulties of holding down multiple tape springs were explored in order to potentially hold the

antennas together using one single technique. Considerations were to ensure a relatively stiff free

length of antenna, to avoid accidental contact with the spacecraft, and rattling between the two

antennas during vibrations. Four options existed when wrapping the antennas together,

depending on the direction of each antenna’s curved face – these are depicted below in Figure

48.

Figure 48: Hold down configurations for two tape spring antennas

It was found that scenario 1, with each tape spring nestled together, with the concave face facing

downward, was the best solution. By having both antennas nestled in one-another, there is little

chance for rattling/slippage between them during vibrations, and both antennas can act as one. It

also allows the antennas to bend in their preferred direction, curved face down, around the

satellite, minimizing chances of plastic deformation when bending. Having the curved face

downward also allows the antennas to be flattened out at the hold down point, causing the

stiffness of the free lengths to seemingly increase.

Relative Placement

It was desired to have both antennas be held down at a common point to allow for the simplicity

of a single deployment mechanism; this may require the antennas to be placed quite close to one

another, and could cause some negative antenna coupling effects. In order to assess the

performance of the antennas at various locations on the satellite and distances from each other,

Clement Ma, communications engineer at SFL, performed a number of simulations of the

orthogonal antennas at various placements on the spacecraft. It was concluded that a favorable

configuration has the antennas mounted on a common face (+Z panel), with a distance of 100mm

between them, and angled 45° from the spacecraft.

Page 81: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

73

Mounting

As learned from the EV-9 project, when mounting the tape spring to a flat surface, the tape

spring will flatten out at the mounted point, and cause it to lose stiffness and not perform as

expected. To solve this problem, the tape spring can be mounted to a curved surface, matching

the tape spring’s curvature; this causes no loss in continuity. However, it was later found that

when tightening the mounting screws a top the tape spring onto the curved surface, if the

direction of the mounting screws do not match the curved surface, the screw head will not make

full contact, and may deform the tape spring under the screw head. This would affect the angle of

the mounted antenna. Having the mounting screws at an angle matching the curve would solve

this problem, but would introduce more complicated machining of the antenna mounting piece.

6.1.3

The design of the deployable AIS antennas is derived from a previous SFL designed AIS antenna

used on the GNB. Tape spring antennas are used, are wrapped around the satellite and held down

at the top +Y face by a spherical tooling ball attached directly to the XPOD. Upon release, the

antennas deploy themselves by using their own stored energy. This technique can be seen in

Figure 50, which shows the satellite installed in the vertically mounted XPOD-Duo.

In light of some of the conclusions from the research done using the tape springs, each tape

spring is mounted differently, such that when wrapped together their curved sides are concentric

and can rest together. The concave side of the tape spring is faced towards the satellite to avoid

accidental contact to the spacecraft (this is the stiffer direction), to allow the tape springs to wrap

in their favorable direction, and be stiffer at the hold down point; as explained in the research

section above. Minimum bending radii are limited to 7mm to avoid any damage to the antennas

during stowage.

COTS carbon steel tape measure tape springs are used, and are machined in-house by the author.

It was deemed that the relatively small dipole introduced by the magnetic steel is of not large

concern, and it was preferred to have the stiffer, lower mass half inch steel tape spring to

facilitate ground testing. This is also common to the material of the antenna used on EV-9.

SubMiniature Version A (SMA) connectors are used as the input for each antenna, and are

soldered to each VHF mount through a seal feed through component. Once soldered, the sub-

assembly of the VHF mounts and the mounting plate remain together. Each VHF mount is

coated in gold to increase its conductive ability. The antennas are mounted 100mm from each

Page 82: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

74

Stowed AIS

Antennas

other, in accordance with the antenna simulations, and at 45° to the satellite, using uniquely

designed VHF mounts, with either a concave or convex mounting surface for the antenna. The

design of the specific components of the antenna base is largely based on the one used on EV-9,

however the components were miniaturized and simplified where possible. The antenna base

design for NORSAT-1 is shown below in Figure 49.

Figure 49: NORSAT-1 antenna base exploded view

Originally, when using the old vertically mounted design of the XPOD-Duo deployment system,

the antennas would be held down at the top +Y face by a spherical tooling ball attached directly

to the XPOD door. A separate small bracket, as seen in Figure 50, would attach to the XPOD

door via two screws, and this bracket would house the tooling ball. This method allowed for

minimal modifications to be made to the XPOD (just requiring two tapped holes to be drilled),

and matches closely to the deployment technique for EV-9. The antennas would then deploy

upon release of the XPOD door, and would deploy outward, away from the launch vehicle.

Figure 50: NORSAT-1 in the vertically mounted XPOD-Duo

Page 83: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

75

However, when the XPOD-Duo design was adjusted to mount horizontally, this deployment

technique had to be reevaluated. Due to the large solar array wings, the orientation of NORSAT-

1 in the horizontally mounted XPOD-Duo is restricted, which renders the VHF antennas slated to

be deployed downward, towards the launch vehicle, assuring that some sort of contact would

likely be made (depending on the geometry of the mounting surface). In talking with the launch

providers, having the antennas potentially contact and scrape the launch vehicle mounting

surface would not be a large concern, and would also not be a large concern for damaging the

antennas, due to their extremely low mass and force of deployment.

Two possible options were explored: leaving the orientation as is, with the hold down point on

the XPOD door, or reorienting the satellite 180° in the XPOD and having the hold down point on

the XPOD pusher-plate rather than the door. With the former, the antennas would be travelling in

the opposite direction of the satellite, and would therefore have a chance of catching on

something, and potentially obstructing the deployment or damaging the antenna. With the latter,

the antennas would deploy in the same direction as the satellite, and would tend to simply slide

on the bottom surface in a favorable direction. In both cases, there is still a large chance that the

antennas will make contact with the launch vehicle. Because it was deemed acceptable to have

the antennas contact the launch vehicle, the latter option was favored, in order to ensure a smooth

deployment.

A small bracket was designed with Michael Ligori, one of the XPOD-Duo designers, which

would house the tooling ball, and attach directly to the pusher-plate via two attachment screws.

The bracket is mounted on the exterior of the plate, and is accessible on one of the open faces of

the XPOD – this way, deployment stowage can easily be inspected, and the bracket can also be a

late addition to the assembly. The bracket proposed is shown below in Figure 51. The tooling

ball fits into a carefully sized hole on the antenna hitch of the +Y panel on the satellite, while

securing both antennas through similar sized holes. The use of a spherical tooling ball is ideal to

ensure no obstructions can be made whilst releasing, also, Antenna 2 is fitted with a slightly

larger hole than Antenna 1, in order to ensure contact with the tooling ball and avoid slippage.

The antenna hitch and guides are made of Delrin plastic so as to minimize potential damage

during vibrations due to friction.

Page 84: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

76

Figure 51: VHF antenna stowage system

A simple stowage technique is also applied to the antennas, whereby a standoff can be threaded

through both of the antennas into the antenna hitch, locking the antennas in place until

integration with the XPOD. Once loaded in the XPOD, the standoff can be removed by hand

from the opening, and the antennas are then held solely by the pusher-plate. A large, noticeable,

“REMOVE BEFORE FLIGHT” tag will be added to this standoff to ensure that it is not

forgotten to be removed before launch. An additional antenna guide is also placed on the +Z

panel, in an effort to reduce the free length of the antennas while stowed. This guide is taller than

the one shown in Figure 51, in order to ensure that the antennas physically bend over it,

increasing their stiffness. Figure 52 shows the antennas stowed on the spare structure of the

spacecraft, and the relevant antennas guides.

Figure 52: NORSAT-1 side view of stowed AIS antennas on spare structure

Page 85: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

77

Thermal considerations are still underway for these VHF antennas, however, the current thermal

solution is to strip the yellow paint off the antenna, and apply Kapton tape directly to the steel

faces – this is to keep the antennas from becoming too hot in orbit due to their extremely small

thermal mass, and also prevent the antennas from rusting on the ground.

The ideal length of the quarter-wavelength antennas can be derived by using the wavelength

formula, where the wavelength = the speed of light ÷ the frequency. Using the speed of light in a

vacuum, and the frequency of interest (162Hz), 462.6mm is obtained. The antennas for

NORSAT-1 are cut to a length slightly longer than this (482.6mm), in order to allow for tuning

to be done on the flight spacecraft before launch, or during VHF antennas pattern testing – this

way they can be easily cut shorter at a late stage if necessary.

Deployment tests were performed in order to better understand the behavior of the tape spring

antennas in the proposed deployment method, and confirm the assumptions being made. The

conclusions from these tests are explained in Section 8.6. A depiction of the expected behavior

of the antennas is shown below in Figure 53, as the spacecraft is being ejected from the XPOD-

Duo. Having the constant pushing force of the pusher plate upon deployment, the antennas are

expected to remain stowed until the deployment spring of the pusher plate has reached its free

length (which is about three-quarters of the length of the XPOD), after which the antennas will

be able to slip out and deploy.

Figure 53: Expected deployment volume of VHF antennas

Page 86: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

78

6.2

The Langmuir probe deployable mechanism was designed and tested by the payload provider,

the University of Oslo. Brief details of the design are mentioned in Section 2.3.2. Two separate

cassettes house two Langmuir probes each, each mounted on a long boom. The booms are

deployed using a Shape Memory Alloy commercial pin puller on orbit via a ground command,

and are each locked in place with a locking pin once fully deployed. A simple position sensor is

included to confirm a successful deployment. Figure 54 below is a depiction of the probes

deploying from their cassette.

Figure 54: Langmuir Probe cassette deployment

Page 87: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

79

7

Ground support equipment (GSE) includes anything that must be used in order to facilitate the

development of the satellite from beginning to finish. This can include anything from software

for testing, to shipping containers used to transport the satellite to the launch site. In accordance

with the micro-space philosophy, one of the main driving requirements for all GSE designs is

that they be kept simple and low cost. The author was mainly involved with the mechanical GSE

(MGSE) needed to support assembly, handling, and some of the mechanically related testing –

this MGSE developed for NORSAT-1 is detailed in this section.

7.1

In order to reduce the risk of damaging components during assembly and integration of the

satellite, various MGSE is designed to ease the assembly process. Some requirements for these

designs are:

1) The Assembly and Handling GSE should be multi-use wherever possible, in order to

minimize cost and complexity

2) The Assembly and Handling GSE should interfere with the final assembly or be distinctly

labeled, in order to avoid forgetting to be removed

3) The Assembly and Handling GSE should use GSE-specific holes wherever possible, in

order to reduce the number of cycles on flight screw holes

4) The Assembly and Handling GSE should be lightweight, in order to not add significant

weight during manipulation of the satellite

5) The Assembly and Handling GSE should be easy to install/remove, and be low risk of

damaging sensitive components

6) The Assembly and Handling GSE should not load the honeycomb panels

One example of this is a set of four Delrin machined tray “legs” that were designed to allow

clearance for mounted components in multiple orientations of the +Z tray during assembly, as

well as to ease the process of connecting the +Z tray to the –Z tray prior to any panel integration

– this is depicted below in Figure 55. Additional GSE was designed to facilitate the installation

Page 88: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

80

of the +Z panel while complying to the order of assembly, as well as support stands for all sub-

assemblies (such as panels), during sub-assembly integration.

Figure 55: MSGE assembly tray “legs” in various orientations of use (black)

Due to the many external protrusions of NORSAT-1, finding a comfortable resting orientation

that eases assembly and disassembly proves difficult. In order to have the large solar array wings

away from harm, as well as both the deployed AIS antennas and deployed Langmuir Probes free

from obstruction, the natural resting position has the +Z face towards the ceiling, and the –Z face

with the solar array wings resting on the ground or work bench. This position can be easily

achieved prior to integrating the protruding wings and CLARA payload by resting the satellite on

the solar array wing attachment brackets as seen in Figure 55 above. Once these components are

integrated however (near the end of the assembly procedure), we can no longer use this resting

technique, and additional GSE is needed.

A support stand was designed to solve this problem and is shown below in Figure 56. Using two

end-platforms and a pair of cross braces, it delivers a sturdy stand to rest the satellite on

throughout the assembly, integration, and testing. The satellite rests directly on the end platforms

using the solar array wing attachment brackets, and is secured to the platform using three M4

sized thumbscrews on each side. The support stand raises the satellite ~12cm of the ground

(work surface), and is also used for deployment testing, and integration into the XPOD-Duo

deployment system (detailed in section 7.5). Copper straps are added to all the Delrin handling

Page 89: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

81

GSE to ensure grounding of the satellite during integration; an example of the copper strap is

seen in Figure 61 and Figure 66 on the support stand and assembly legs.

Figure 56: NORSAT-1 GSE support stand

Due to the awkward geometry and relatively large mass of NORSAT-1, handling of the satellite

during and after assembly can be a difficult task and must be done with the utmost care to avoid

potential mishaps and damage to interior and exterior components. A set of GSE handles was

designed to facilitate the transport and manipulation of the satellite during/after assembly and is

shown below in Figure 57. The overall strength requirement for this handle system is that it be

capable of supporting the entire mass of the satellite, and any protective or supporting GSE under

a load factor of 5g.

The handle design is adapted from a design used for the NEMO-AM project, and employs a

thick aluminum support bar that directly attaches to the wing attachment brackets into GSE-

specific threaded holes. This provides a reliable pick-up point on either side of the satellite on

which the remaining handle assembly is then mounted, consisting of a half-inch aluminum bar

and three aluminum brackets. This handle assembly can be quickly installed and removed from

the satellite when needed using the three M4 thumbscrews, and provides an unobtrusive two-

person system for maneuvering and handling the satellite. The support bar is also used to provide

pick up points for thermal vacuum testing and craning via eyebolts. An alternate handling system

is also incorporated directly on the support stand, using commercial offset handles, to provide

additional pick-up points that would be favored in certain scenarios – these can be seen in Figure

61, and were mainly added to facilitate loading the spacecraft into the XPOD- Duo, and

deployment testing.

Page 90: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

82

Figure 57: NORSAT-1 GSE handle assembly

7.2

Some of the most sensitive and expensive components on a typical microsatellite are mounted

directly on the exterior surfaces – these are mainly the solar cells. These cells have a glass top

layer that can be easily shattered with a sharp impact. Also on the exterior are several delicate

thermal tapes that could be damaged through constant unnecessary contact. Because of these

reasons, among many others, such as keeping the satellite as clean as possible, a protective

enclosure is designed to house the satellite after assembly. With access to the testing port of the

satellite, much of the pre-flight testing can still be done on the satellite whist the protective

enclosure is installed, minimizing risks of physical damage to the satellite up until integration

into the deployment system and onto the launch vehicle. Some requirements for the protective

enclosure design include:

1) The protective enclosure shall be clear, in order to allow visual inspection with the

enclosure installed.

2) The protective enclosure should need minimal tools to install/disassemble.

3) The protective enclosure shall provide a barrier over all solar cells and delicate exterior

components.

4) The protective enclosure shall allow for access to the testing port of the satellite.

5) The protective enclosure shall be composed of static dissipative materials.

The protective enclosure design for NORSAT-1 is based on the protective enclosures previously

developed for SFL’s GNB class of satellites. It consists of multiple clear Lexan panels that are

connected to four Delrin machined rails. The rails rest on the satellites launch rails and are the

sole point of contact to the satellite. Figure 58 shows this enclosure installed on the fully

Page 91: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

83

assembled satellite, along with additional protective panels on the solar array wings – note that

the satellite is still compatible with the assembly jig stand whilst in the enclosure. Because the

solar array wings are externally protruding, their protective panels will remain installed up until

the satellite is mounted on the launch vehicle, even while being loaded into the deployment

system. They are outfitted with easy-to-remove thumbscrews and handles to facilitate the

removal at that stage. Meanwhile, the rest of the enclosure will be removed prior to integration

with the XPOD-Duo deployment system, and will then be shipped to the launch site, with

additional protective panels around the XPOD-Duo.

Figure 58: NORSAT-1 protective enclosure design

7.3

The large solar array wings on NORSAT-1 are a unique addition to the microsatellite. While the

rest of the structure is obtained from local machine shops, made in-house, or sourced from

commercial off-the-shelf (COTS) distributors, the solar array wings are manufactured by a third

party company who specializes in honeycomb composite materials for the space industry.

Because of the highly specialized field, relatively new technology, and amounts of qualification

testing required, these panels become an important economical consideration for the project. It

was decided that while two full sets of structure was procured as a “flight” and “spare”, only one

set of solar array wings would be procured, purely for economic reasons. Still needed, however,

was a set of geometrically similar wings to simulate these large solar arrays during various stages

of testing, in order to avoid damage to the expensive composite ones. The main design

requirements for these panels include:

Page 92: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

84

1) The Mock-Up Wings shall be similar to the intended flight wings in geometry.

2) The Mock-Up Wings shall be electrically similar to the intended flight wings.

While not fully weight representative, the mock-up wings serve as acceptable placeholders on

the satellite to achieve the correct geometry for various testing such as structural fit checks,

electromagnetic compatibility, uplink/downlink antenna, GPS antenna, VHF antenna, and

deployment testing. These mock-up wings are comprised of two thin aluminum sheets cut to the

correct geometry, and then sandwiched over spacers using epoxy adhesive to achieve the correct

thickness - the open edges are covered with Kapton tape to keep the panel easy to clean. They

are shown below in Figure 59.

Figure 59: NORSAT-1 mock-up wings (left), mock-up wings fitted on structure (right)

7.4

In order to characterize the performance of the communications subsystem design coupled with

the antenna placement on the satellite, and overall geometry for NORSAT-1, various antenna

pattern testing must be done in an anechoic chamber. To facilitate this testing, MGSE is needed

to help support the satellite in various orientations to point the antennas under test. Because this

GSE will be involved with the test itself and could potentially affect the results, some specific

requirements are derived for their design:

1) The Radio Frequency (RF) testing GSE shall be made of materials with a relative

permittivity (dielectric constant) that is as close to 1 (vacuum) as possible in order to

minimize its influence on test results

Page 93: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

85

2) The RF testing GSE should be lightweight and easy to transport to off-site testing

facilities

3) The RF testing GSE design should leverage previous designs wherever possible to

minimize time, cost, and effort spent

4) The RF testing GSE shall provide four stable orientations at 45° increments around the

Y-axis – capable of being placed on a rotating platform without obstruction. This

provides all the necessary testing orientations for the antennas

The RF testing GSE will be used to perform testing on three different spacecraft antennas: The

S-Band Uplink/Downlink antennas, the GPS antenna, and the VHF antennas. Testing of these

three different antenna types will be performed in two separate off-site facilities – the test setup

and methods are further detailed in Section 8.5. In both facilities, a platform capable of rotating

360° is provided, on which the NORSAT-1 RF testing GSE is mounted. The GSE is composed

of a wooden baseplate that attaches to the rotating platform, on which is then stacked a large

polystyrene (EPS) block of foam to achieve a height similar to the source antenna in the facility.

Additional foam blocks are then stacked on top to achieve the appropriate orientations of the

satellite in 45° increments as seen in Figure 60. Each foam block is aligned and held in place by

using several wooden dowel pins, and the satellite rests directly on these foam blocks – wiring

for the antennas is routed through a hole at the center of the large foam block down to the

rotating platform. Note that the satellite is not fully integrated for these tests (only the outer

protruding elements are needed: the metallic structure, the LP Cassettes, and the antennas), and

is therefore relatively straightforward to re-orient to the required positions with the reduced

mass.

A large amount of this RF testing GSE was leveraged from previous SFL antenna pattern tests.

In fact, only the top foam blocks as seen in Figure 60 for the 45° and 90° orientations needed to

be slightly modified by the author to accommodate NORSAT-1. An additional foam block was

used under the 90° foam block in order to achieve the necessary clearance for the Langmuir

Probes. Note that the 45° GSE foam block doubles as the -45° orientation by simply rotating

the block under the satellite 180°.

Page 94: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

86

Figure 60: RF testing GSE blocks

Further detail and results of the specific RF testing that was done for NORSAT-1 using this GSE

is detailed in Section 8.5.

7.5

In the past, with the smaller sized GNB satellites (~7kg), the satellite could easily be picked up

via handles on one side of the spacecraft by a single person and manually loaded into its XPOD

deployment system with the aid of gravity and no additional GSE. NORSAT-1 is more than

double the mass and size of a GNB, so loading into the XPOD-Due in a similar way could not be

done in a reliable manor. Some design requirements for this deployment GSE are:

1) The XPOD loading GSE shall allow the spacecraft to be loaded into the XPOD and be

able to be removed once fully loaded in the XPOD

2) The XPOD loading GSE should double as the Deployment GSE since their purpose

entails fundamentally the same process

3) The Deployment GSE shall allow for horizontal deployment of the spacecraft from the

XPOD – Horizontal deployment minimizes risks associated with the large mass’ and

gravity

4) The Deployment GSE shall allow for full deployment of the VHF antennas

By adding a base plate and some retractable nylon ball transfers to the GSE support stand, the

satellite is able to slide freely on the support stand when needed. This provides a method to load

the satellite into the XPOD, and doubles as a method to perform deployment testing. This

deployment jig is designed to have all the necessary clearances to properly load into the XPOD-

Duo, and allow the XPOD door to fully open for a representative deployment. The height of the

Page 95: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

87

jig is sized to mate appropriately with the XPOD-Duo GSE, as shown in Figure 62, holding the

deployment system horizontally in order to perform this loading and deployment technique.

Commercial offset handles were also incorporated to easier handle the spacecraft and GSE

during this test. The deployment jig design is depicted below in Figure 61.

Figure 61: NORSAT-1 deployment jig design details

Figure 62: XPOD-Duo loading (XPOD-Duo GSE designed by Mike Ligori)

Page 96: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

88

8

All of the mechanically related integration and testing in which the author was involved for

NORSAT-1 is outlined in this section. Note that due to some schedule delays, not all of the

planned testing was complete at the time of writing this thesis.

8.1

Structural fit checks are a necessary step in the development process in order to avoid possible

surprises at a later stage in the spacecraft integration that could cause large delays. Upon

receiving all of the manufactured structural parts from the machine shop, each part is

individually inspected for any detrimental machining defects and errors. The parts are then fit

checked with their mating components and payloads in an effort to reveal any further defects in

the parts or design. The third step is to install all of the threaded inserts into the parts, in order to

allow a full spacecraft structural fit check to be performed, including all GSE parts (shown in

Figure 63). A path forward to resolve any defects found is determined at each step, and this

entire process is performed twice, for both the flight and spare set of structures. A list of the main

machining defects found during these fit checks is outlined below in Table 6, along with the path

that was taken for its rectification.

Table 6: List of main manufacturing defects found through inspection and fit checks

Defect Affected Rectification

+Y, -Y, -Z panels outer

dimensions slightly too large Their fit between the trays In-house filing

+Z tray had rounds machined on

the X panel mounting bosses Flush placement of the X panels In-house filing

Missing or wrong size holes in

+Z and –Z tray Mounting of various equipment

Returned to manufacturer for

modifications

Reaction wheel bracket machined

with some angled surfaces

Orthogonal placement of reaction

wheels

Returned to manufacturer for

rectification, use of copper shims

to adjust the height to account for

the missing material

Vent hole missing on risers, and

certain spots on the trays Sufficient venting of volumes In-house vent hole drilled

+X, -X, +Z, and solar cell

coupons were slightly warped

Installment of panels and

Laydown of solar cells on

coupon

Deemed acceptable and left as is

Protective panel holes all

misaligned

Installation of the protective

enclosure

Holes were enlarged in-house for

extra clearance

Page 97: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

89

Once the structural fit checks have been completed, and all errors rectified, the metallic parts are

sent for Iridite chemical conversion coating, in order to provide a layer of protection over the

bare aluminum surface.

Figure 63: NORSAT-1 structural fit checks, CLARA (engineering model) installed (left),

GSE enclosure installed (right)

8.2

SFL technicians build the flight and spare wiring harnesses for NORSAT-1. Once built, they are

each fit checked into the satellite structure to confirm the routing assumptions and lengths. In

many cases, the added length margin to the manufacturing drawings caused some of the

harnesses to be slightly longer than necessary, however this is not generally a problem – the

extra length can be accounted for by routing the wires in an “S” form between tie down points,

consuming much of the extra length. After the spare harnesses were built and integrated to the

Dirty-Sat (discussed in Section 8.3 Dirty-Sat Integration), certain lengths were updated for the

second flight build.

The larger harnesses, the Main and Payload ones, were slightly more complicated to fit check.

Due to the vast number of interconnecting connectors, and no easy way to sequence the order

that each connector be built in the harness, the resulting harnesses are seemingly tangled, causing

them to not fit in the satellite verbatim. During this fit check stage, the author proceeded to group

together common routed wires, while fit checking it into the structure. In some cases, this

involved removing crimped wires from certain simple connectors and reinserting them once

untangled in order to have them route as intended. Once untangled, and fully routed in the

structure, the harness is neatly bundled together in order to prevent it from intertwining, and to

Page 98: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

90

keep its form to facilitate its integration to the satellite during assembly. Figure 64 below shows

a before and after of this process for the Main wiring harness and Figure 65 shows the fit check

process for the Payload harness, showing the 3D model of the harness for comparison.

Figure 64: Freshly built (left) and untangled (right) flight Main wiring harness

Figure 65: Payload wiring harness, 3D model (left), fit check in structure (middle), and

untangled flight harness (right)

8.3

In order to mitigate much of the risk associated with a fairly new satellite design, a Dirty-Sat is

created as a fully functioning platform for testing. The Dirty-Sat is a close-to-fully representative

model of the flight spacecraft, using all flight representative (spare) equipment, including the

spare structure; however, it is assembled and integrated outside of a clean room (in a typical lab

environment) hence the name. This is done to make the integration and testing process simpler

and quicker to troubleshoot than performing everything in a clean room environment.

Structurally, the Dirty-Sat integration allows for a representative test and run-through of the

assembly procedure, allowing it to be further refined, and it also serves as a test-fit for the wiring

harness routing and design.

Page 99: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

91

Dirty-Sat integration for NORSAT-1 began at SFL in February 2015, was completed on March

12th

2015 and was followed by system level spacecraft functional testing. The integration was

performed semi-clean, in that parts were cleaned prior to integration, and gloves were used to

handle all equipment – this was done in order to reduce the risk of contaminating the spare parts.

Spare equipment was not however needed for all of the equipment, and alternatives were used

for the Dirty-Sat. Spare solar cells were not procured, and were simply not included in the Dirty-

Sat – instead, external power supplies are used to properly simulate them. The GSE Mock-up

Wings were used in the place of the solar array wings, and only one spare reaction wheel was

procured, therefore two flight reaction wheels were used in the Dirty-Sat. Some pictures of the

Dirty-Sat integration are shown in Figure 66 below. An engineering model (EM) of the battery

pack was also assembled for this integration.

Figure 66: NORSAT-1 Dirty-Sat integration

The Dirty-Sat integration and testing was able to reveal several insights to the satellite design

that would serve as refinements for the flight integration, these are briefly listed below:

Page 100: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

92

o The size of the Langmuir probe connectors on the electronics box were much larger than

anticipated, which affected the installation of the –Y panel. The payload provider has since

changed these connectors for flight.

o The data signals of the -Y sun sensor seemed to be “over-filtered” by the filtered MicroD

connector through the separation plate, causing the signals to be diminished or lost. These

data lines were since removed from the filtered MicroD connector, and alternatively

syphoned through the separation plate using individual, less intensive, feed through filters.

o The assembly procedure was refined, and in some cases modified. The integration of the

reaction wheels was accommodated at a very late stage in the integration in order to

minimize possible harm and contamination. This allowed for a much easier installation of the

wiring harness while connecting the trays together, providing extra room in the satellite to

route the wires without risk of damaging the wheels.

o Lessons were learned about the “Hand-Flex” coax cables that were used in the satellite. It

was found that they would get damaged quite easily when trying to form the path of the cable

once connected. Efforts were made to form the cable as best possible prior to installation to

avoid this. This is particularly of concern for the shorter cables (under 7 inches) that require

precise routing.

o The EMI gaskets used in various places in the satellite were found to fall out of their gasket

groove quite easily during installation. The gaskets were glued in place with very small

amounts of Room-Temperature Vulcanization (RTV) at their tips to avoid the gaskets falling

out during integration.

o The final location of all tie-down points for the wiring harness was confirmed. Many wire

lengths that were either slightly too short or too long were updated for improved routing in

the flight harness.

8.4

A large milestone for the NORSAT-1 project was electromagnetic compatibility (EMC) testing

of the Dirty-Sat. This testing began on March 16th

2015, was completed on March 24th

2015 and

was performed inside SFL’s in-house anechoic chamber (Figure 67). The purpose of this test is

to ensure that there is no significant interference that affects the performance of any of the on-

board electronics, and is particularly of importance due to the sensitivity of the AIS payload to

Page 101: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

93

very low frequency emissions, levels at which much of the spacecraft avionics is known to

produce noise. The effective signal strengths are measured at the receiver inputs of interest as

each of the spacecraft avionics are activated, in order to assess any produced interference. A

large external battery pack was used in order to fully simulate the power generation of the absent

solar cells; this ensures that the full functionality of the power system is active during testing.

Figure 67: NORSAT-1 EMC testing in SFL’s anechoic chamber

Overall, it was found that the NORSAT-1 platform provided a very “quiet” Radio Frequency

(RF) environment in the payload bay, exceeding the demanding requirement set forth by the AIS

receiver (#10 in Table 2). This successful result mitigated much of the risk of the novel design of

the platform, and confirmed the performance of the EMI reduction techniques that were

implemented in the structural design (Section 3.2.3).

8.5

In order to validate the spacecraft’s communication subsystem design and simulations, various

antenna pattern tests are done using spare structure of the satellite at various RF testing facilities.

8.5.1

S-Band and GPS antenna pattern testing was performed between June 17th

and June 19th

2015 at

the University of Toronto St. George Campus. The purpose of this testing is to characterize the

antenna performance, coupled with their mounting location on the full geometry of the spacecraft

bus. The tests involve measuring parameters of the spacecraft transmitted and received signals

Page 102: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

94

with respect to a stationary source antenna inside an anechoic chamber. In order to achieve a

final spherical distribution of results around the satellite, measurements are taken while

performing 360° rotations of the spacecraft on an RF positioner platform, at four different tilt

angles of the spacecraft; details of these test orientations were discussed in Section 7.4. This

allows for the results to be plotted as a spherical distribution, and characterizes the

communication link of the satellite as a function of its orientation. A depiction of the test setup is

shown below in Figure 68. The results were post-processed by SFL’s Communications Engineer,

Clement Ma, and were all found to be acceptable.

Figure 68: Antenna pattern test setup

8.5.2

Because of the much lower frequency of operation, an alternate facility is needed to test the VHF

antennas. At the time of writing, the facility has not been finalized.

8.6

As per Requirement #14 in Table 2, all spacecraft deployable mechanisms must be tested in a 1g

environment. This requirement stems from a reliability standpoint – where if the deployment

mechanism is designed for a 1g environment, then it should certainly work in a 0g environment.

The main purpose of performing deployments tests of the satellite was to assess the performance

of the deployable VHF antennas. This became more necessary when the XPOD-Duo was slightly

modified to deploy horizontally instead of vertically. The areas of interest for this testing are to

Page 103: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

95

determine when the antennas start to deploy, the dynamics and path of their deployment, and

their behavior upon contact of a potential obstruction such as the launch vehicle.

The test setup for the deployment testing is shown in Figure 69. Note that the test setup has the

mounting legs of the XPOD facing upward, and will thus have the antennas deploying upward as

well. An engineering model (EM) version of the XPOD-Duo is used, along with the empty spare

spacecraft structure. The antenna hold down piece is attached to the XPOD pusher plate via

epoxy for these tests for convenience. The total mass of the spacecraft with the deployment GSE

for these tests is ~12kg. Large sheets of plastic are laid down as a track for the deployment in

order to reduce the friction as much as possible for a representative deployment.

Figure 69: Deployment testing test setup

8.6.1

Four deployments were performed in order to examine the behavior of the tape spring antennas

under various conditions. The tape springs are expected to stay stowed under the force of the

pusher plate upon deployment until the spring has reached its full length, however, a possible

scenario could exist whereby the antennas could slip out early due to the shock created when the

door opens, or due to the counteracting force of the tape springs themselves.

The first two tests are done in the configuration shown above in Figure 69, where the top of the

setup is all clear, allowing clear inspection of the antenna behavior from above. After these two

tests, it was confirmed that the antennas behave as expected, and do not move at all until the

spring has reached its full length. Figure 70 is a still shot during one of these deployments, at the

point just before the antennas start to deploy.

Page 104: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

96

Figure 70: Deployment test still shot, just before antennas deploy

For the subsequent two deployments tests, a large plate is placed atop the mounting feet of the

XPOD-Duo in order to simulate the mounting surface of the launch vehicle and examine the tape

spring behavior if contact is made. Due to lack of information of the exact mounting surface for

NORSAT-1 on its launch vehicle, the size of this “simulated launch deck” is based on an

expected mounting surface aboard a PSLV rocket for the NEMO-AM mission.

In the first of these two tests, the speed of the spacecraft deploying is fast enough that no contact

is made of the antennas on the plate, and the antennas are able to deploy unobstructed; a still shot

from this deployment is shown in Figure 71.

Figure 71: Deployment with mounting plate, no contact

For the second test, an extra 4kg is added to the spacecraft in order to slow down the

deployment, and ensure the antennas would contact the mounting plate; a still shot from this

deployment is shown in Figure 72 at the point of antenna contact. Both antennas sequentially hit

the mounting plate, causing them to bounce downward at first, and then continue to routinely

deploy.

Page 105: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

97

Figure 72: Deployment with plate, contact made

The speeds obtained during these deployment tests were ~1m/s, whereas in orbit, a much higher

speed of ~1.6m/s is expected. The lower speeds are caused by the various added frictional forces

in the test setup, largely due to gravity, and also due to the non-ideal condition of the EM XPOD-

Duo being used. As seen in the deployment shown in Figure 71, with a higher deployment speed,

the antennas have a better chance of deploying later, and avoiding potential obstructions.

However, in 0g, the deployment of the antennas themselves would also be much faster, because

they would not have to combat their own weight under gravity. Due to these numerous non-ideal

testing conditions, no clear conclusion can be made about the exact contact, if any, that would be

made by the antennas on the launch vehicle during deployment of NORSAT-1 in orbit; however,

this testing gives an idea of how the antennas could behave if contact is made, and can be used to

predict what kind of contact might be made once a clear mounting surface on the launch vehicle

is defined.

8.7

Due to the high sensitivity of the CLARA payload to particle contaminations, extra measures of

cleanliness are taken at all stages of the spacecraft integration. A bake-out of the flight structure,

prior to any avionics integration, was performed in July 2015 as an extra measure of cleanliness,

specifically for the CLARA payload. Due to having no electronic equipment in the flight

structure for this activity, a higher bake-out temperature can be achieved without risk of

damaging electronics, and a better result can be achieved. Included in the bake-out are all of the

flight structural components, mounting hardware, gasket material, and wire tie mounts. A bake-

out involves using heat in a vacuum environment in order to release particle contaminations from

something, as a method of cleaning, and forces materials to outgas at an accelerated rate. A “cold

Page 106: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

98

wall” system is used; where-by a plate in a thermal vacuum chamber is kept extremely cold, and

allows a place for the removed particle contaminates to settle.

The bake-out was performed in the in-house large thermal vacuum chamber at the Space Flight

Laboratory, which is able to closely simulate space environments of a vacuum, and extremely

hot/cold temperatures. Seven temperature sensors are placed in various locations on the structure

in order to monitor and achieve a uniform temperature of interest. For this activity, a temperature

of 80°C was chosen based on the limits of the various included materials, and this was

maintained for approximately 38 hours. Figure 73 below shows the setup of this activity in the

large thermal vacuum chamber. Infrared lamps are directed towards the satellite in all directions

in order to obtain the high temperatures, and the satellite is suspended in the center via four steel

cables. The power of each lamp is controlled individually, allowing for some tuning to be done

while achieving a stable average temperature at each of the temperature sensors.

Figure 73: NORSAT-1 flight structure bake-out setup

Similar bake-outs will be performed at the component level of some of the other satellite

components, including the wiring harnesses, the solar panel wings, and the payloads. Finally, a

system level bake-out will also be performed on the integrated flight spacecraft with all avionics

prior to the CLARA payload integration.

Page 107: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

99

8.8

Flight integration is currently on going for NORSAT-1 at SFL in the class 10,000 clean room. At

the time of writing, only the AIS receiver payload has been delivered and has been integrated

with the on-going assembly. Once the remaining flight payloads are received, they will be

immediately integrated in the flight satellite to complete the flight assembly in September 2015

(expected). Other outstanding items to be integrated include the power system, the spacecraft

panels and all solar arrays. Figure 74 below shows the current integration progress.

Figure 74: NORSAT-1 flight integration progress

A full assembly procedure for the battery pack has also been created by the author [21], and was

built by an SFL technician and engineer for flight. The fully integrated flight battery pack for

NORSAT-1 is shown below in Figure 75.

Figure 75: NORSAT-1 flight battery pack

Page 108: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

100

9

NORSAT-1 represents Norway’s first scientific satellite, being a microsatellite carrying three

distinct and separately developed payloads. The main objectives of the payloads on-board are to

measure Total Solar Irradiance (TSI) levels from the sun, investigate space plasma

characteristics, and provide maritime ship tracking information through the Automated

Identification System (AIS). Much of the mechanical aspects of design, analysis, and testing for

the NORSAT-1 satellite bus has been completed and presented in this thesis. Subject to funding,

interest exists in developing subsequent multi-payload satellites using the NORSAT-1 design as

a platform; the development of NORSAT-2 has already begun, and is able to conserve

significant time, effort, and cost in the early stages by leveraging the presented design.

The structural design of NORSAT-1 has been developed from an early conceptual stage to a

detailed design currently under assembly and flight preparation. Verifications of all of the design

related requirements, including the driving requirements listed in Table 2, have been shown

through inspection or analysis, and in some cases, will be verified through testing in the near

future. The designs of two honeycomb composite solar array panels and two deployable whip

antennas mounted on the satellite have also been detailed to their current states. Various testing

and integration activities, including their test-specific mechanical ground support equipment

designs, are lastly presented with any significant results. Moving forward, several mechanical

aspects of development still exist for the NORSAT-1 satellite prior to launch. Completion of the

flight assembly, system level thermal vacuum testing, and system level vibration testing

represent some of these predominant tasks.

In thoroughly documenting the design and specific design decisions with their reasoning in this

thesis, the author has contributed a valuable reference for future microsatellite mechanical

designs. Efforts were made to design a very capable and modular satellite bus, for direct use on

future missions having similar requirements.

Page 109: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

101

[1] Space Flight Laboratory, University of Toronto Institute for Aerospace Studies Space Flight

Lab, January 2014. [Online]. Available: www.utias-sfl.net.

[2] A. Beattie, "NORSAT-1 CDR System Design Document v1.0," SFL Internal Document #

SFL-NS1-SYS-D001, Toronto, 2014.

[3] J. D. Haigh, "Climate Variability and the Influence of the Sun," AAAS - Science, London,

2001.

[4] A. Fehlmann , "Metrology of Solar Irradiance," Zurich , 2011.

[5] J. D. Haigh, " Climate - Climate variability and the influence of the Sun," SCIENCE, vol.

294, pp. 2109-2111, 2001.

[6] P. Brekke and M. Osmundsen, "NORSAT-1: Total Solar Irradiance, Space Weather and

Ship Detection," in SOURCE Science Conference, Cocoa Beach, 2014.

[7] A. B. Donald A. Gurnett, Introduction to Plasma Physics: With Space and Laboratory

Applications, Cambridge University Press, 2005.

[8] A. Beattie, NORSAT-1 System Requirements and Verification Matrix v1.0, Toronto, ON:

SFL Internal Document # SFL-NS1-SYS-R001, 2014.

[9] S. Kanji, NORSAT-1 Structural Requirements and Verification Matrix, Toronto: SFL

Internal Document # SFL-NS1-STR-R001., 2013.

[10] S. Kanji and M. Chaumont, NORSAT-1 Assembly Procedure - Issue 0.4, SFL Internal

Document # SFL-NS1-MEC-G001, 2015.

[11] V. Tarantini, NORSAT-1 Thermal Design v2.0, Toronto: SFL Internal Document # SFL-

NS1-THM-D001, 2015.

[12] D. Diaconu, Mechanical Aspects of Design, Analysis and Testing of the Nanosatellite for

Earth Monitoring and Observation – Aerosol Monitor (NEMO-AM), Graduate Department

of Aerospace Science and Engineering University of Toronto, 2014.

Page 110: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

102

[13] Hexcel Composites, HexWeb Honeycomb Sandwich Design Technology, 2000.

[14] T. N. Bitzer, Honeycomb Technology - Materials, Design, Manufacturing, Applications and

Testing, 1 ed., Springer, 1997.

[15] E. Greene, Failure Modes, Eric Greene Associates, 2013.

[16] S. Heimbs and M. Pein, "Failure Behaviour of Honeycomb Sandwich Corner Joints and

Inserts," Elsevier - Composite Structures, December 2008.

[17] Shur-Lok Corporation, Design Manual - Fasteners for Sandwich Structure, 1996.

[18] ECSS‐E‐HB‐32‐22 Working Group, Space Engineering - Insert Design Handbook,

Noordwijk: ESA Requirements and Standards Division, 2011.

[19] Structures and Mechanisms Division - European Space Research and Technology Centre,

Insert Design Handbook, European Space Agency, 1987.

[20] G. Bianchi, G. S. Aglietti and G. Richardson, "Optimization of Bolted Joints Connecting

Honeycomb Panels," in 1st CEAS - 10th European Conference on Spacecraft Structures,

Materials and Mechanical Testing, Berlin, 2007.

[21] S. Kanji, NORSAT-1 Battery Pack Assembly Procedure, Toronto: SFL Internal Document #

SFL-NS1-MEC-G002, 2015.

Page 111: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

103

Figure 76: Insert strength as a function of core height for a 11mm diameter insert in a 3/16-

5052-.001 honeycomb core with aluminum face sheets under tension [19]

Page 112: Shahil Kanji - University of Toronto T-Space · PDF fileii Shahil Kanji Master of Applied Science Graduate Department of Aerospace Science and Engineering University of Toronto 2015

104

Figure 77: Insert strength as a function of core height for a 11mm diameter insert in a 3/16-

5052-.001 honeycomb core with aluminum face sheets under compression [19]