You are on page 1of 75

Brigham Young University

BYU ScholarsArchive

Theses and Dissertations

2019-07-01

Joint Analysis of and Applications for Devices with Expanding


Motions
Kendall Hal Seymour
Brigham Young University

Follow this and additional works at: https://scholarsarchive.byu.edu/etd

Part of the Engineering Commons

BYU ScholarsArchive Citation


Seymour, Kendall Hal, "Joint Analysis of and Applications for Devices with Expanding Motions" (2019).
Theses and Dissertations. 7725.
https://scholarsarchive.byu.edu/etd/7725

This Thesis is brought to you for free and open access by BYU ScholarsArchive. It has been accepted for inclusion
in Theses and Dissertations by an authorized administrator of BYU ScholarsArchive. For more information, please
contact scholarsarchive@byu.edu, ellen_amatangelo@byu.edu.
Joint Analysis of and Applications for Devices with Expanding Motions

Kendall Hal Seymour

A thesis submitted to the faculty of


Brigham Young University
in partial fulfillment of the requirements for the degree of

Master of Science

Larry L. Howell, Chair


Spencer P. Magleby
Anton E. Bowden

Department of Mechanical Engineering


Brigham Young University

Copyright © 2019 Kendall Hal Seymour


All Rights Reserved
ABSTRACT

Joint Analysis of and Applications for Devices with Expanding Motions

Kendall Hal Seymour


Department of Mechanical Engineering, BYU
Master of Science

Origami has been extensively studied by engineers for its unique motions and ability to col-
lapse to small volumes. Techniques have been studied for replicating origami-like folding motion
in thick materials, but limited practical applications of these techniques have been demonstrated.
Developable mechanisms are a new mechanism type that has a similar ability to collapse to a low
profile. The cylindrical developable mechanism has the ability to emerge from and conform to
a cylindrical surface. In this work, a few practical applications of devices with novel expanding
motions are presented. The design and testing of an origami-inspired deployable ballistic barrier,
which was designed by combining and modifying existing thickness accommodation techniques,
is discussed. The properties of cylindrical developable mechanisms are examined and two devices
designed for use with minimally invasive surgical tooling are presented. Various hinge options for
small-scale cylindrical developable mechanisms are then reviewed and discussed. A planar mod-
eling assumption for curved lamina emergent torsional joints in thin-walled cylinders is then ana-
lytically and empirically validated. Conclusions are drawn and recommendations for future work
are given.

Keywords: Origami ballistic barrier, origami inspired, developable mechanisms, joint design, de-
ployable mechanisms, surgical devices, compliant mechanisms
ACKNOWLEDGMENTS

I first and foremost acknowledge my wife, Elle, for her extraordinary efforts to raise our
three children while I finished my graduate work. She sacrificed her time, energy, and personal
educational goals during the many years of my schooling to properly care for them, for which I
will be eternally grateful.
I would like to thank Dr. Larry Howell for accepting me into the Compliant Mechanisms
Research group and connecting me to numerous opportunities that taught me invaluable life lessons
and raised me to a higher level of engineer. His efforts to be a mentor first and administrator second
do not go unnoticed. I thank Dr. Spencer Magleby for his guidance in helping me to see the big
picture, find my research path, and realize that work and fun should not be mutually exclusive. I
also thank Dr. Anton Bowden for sharing his spiritual viewpoint of the wondrous design of the
human body, and for inspiring me to always seek the greatest good in my engineering work.
I also thank the many individuals in the Mechanical Design Collective and Compliant
Mechanisms Research group who helped complete the work presented in this thesis, including
Alex Avila, Terri Bateman, Pietro Bilancia, Dakota Burrow, Jacob Sheffield, and David Morgan,
who co-authored various papers with me. There were many students involved in summer projects
who laid the groundwork for my research by using their skills to show the mechanisms developed
by the lab had hundreds of potential applications. The camaraderie and friendship developed with
members of the MDC will never be forgotten.
I lastly thank my Heavenly Father and Jesus Christ for the guidance I have received through-
out my educational career. Without revelation and personal direction from God, I would not be half
of the husband, father, friend, or engineer I am today.
The material from Chapter 2 is based on work supported by the U.S. National Science
Foundation and the U.S. Air Force Office of Scientific Research through NSF Grant No. EFRI-
ODISSEI-1240417. The material from Chapter 3 and 4 is based upon work supported by the Na-
tional Science Foundation under Grant No. 1663345.
TABLE OF CONTENTS

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Motivation and Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

Chapter 2 Origami-Inspired Ballistic Barrier . . . . . . . . . . . . . . . . . . . . . . 4


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.1 Ballistic Barrier Requirements . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.2 Adaptation of Origami Crease Patterns . . . . . . . . . . . . . . . . . . . . 5
2.2.3 Surrogate Folds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.4 Thickness Accommodation . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.5 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.6 Ballistics Protection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.7 Prototyping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3.1 Full-Scale Ballistic Resistant Prototype . . . . . . . . . . . . . . . . . . . 14
2.3.2 Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.3 Ballistics Demonstration . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Final Barrier Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

Chapter 3 Cylindrical Developable Mechanisms for Minimally Invasive Surgical In-


struments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2.1 Compliant and Lamina Emergent Mechanisms . . . . . . . . . . . . . . . 24
3.2.2 Developable Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2.3 Developable Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3 Properties of Cylindrical Developable Mechanisms . . . . . . . . . . . . . . . . . 26
3.3.1 Behaviors of Developable Mechanisms . . . . . . . . . . . . . . . . . . . 27
3.3.2 Characteristics of Developable Mechanisms . . . . . . . . . . . . . . . . . 27
3.3.3 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.4 Adaptation to Surgical Devices in Shafts . . . . . . . . . . . . . . . . . . . . . . . 30
3.4.1 Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5.1 Internal Scissors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5.2 Multiplying Cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

iv
Chapter 4 Hinges and Curved Lamina Emergent Torsional Joints in Cylindrical De-
velopable Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.2 Joint Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.3 Evaluation of Various Hinge Candidates . . . . . . . . . . . . . . . . . . . . . . . 41
4.3.1 Classical Pin Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3.2 Small-Length Flexural Pivots and Initially Curved Cantilever Beams . . . . 42
4.3.3 Membrane Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3.4 Lamina Emergent Torsional (LET) Joints . . . . . . . . . . . . . . . . . . 46
4.4 Modeling Curved LET Joints With a Planar Approximation . . . . . . . . . . . . . 47
4.4.1 Evaluating the Planar Approximation . . . . . . . . . . . . . . . . . . . . 47
4.4.2 Comparison of Planar LET Joint Adaptation Prediction with Physical Pro-
totype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.4.3 Discussion on LET Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.5 At-Scale Laparoscopic Surgical Device in Nitinol . . . . . . . . . . . . . . . . . . 55
4.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

Chapter 5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

v
LIST OF FIGURES

2.1 The crease pattern selected for the barrier design. . . . . . . . . . . . . . . . . . . . . 7


2.2 Profile view of the membrane fold (a), the ”turn of cloth” fold (b), and the center-panel
gap fold (c). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Drawing of final barrier topology. Each larger nesting fold encloses the two adjacent
panels in the folded position. White sections are rigid panels, and the yellow area is
the aramid fabric covering both sides of the panels. . . . . . . . . . . . . . . . . . . . 10
2.4 Top-down view of barrier pattern deployed states. (a) Stowed state (b) Deployed state,
’X’ marking center of gravity (c) Fabrication state (scaled) . . . . . . . . . . . . . . . 11
2.5 Prototype sequence, top left to bottom right: (1) Thick-folding prototype inspiring bar-
rier design, using 2.5 cm-thick foam board for panels and fabric tape for surrogate folds
(2) Initial paper prototype (3) Wood and felt (4) Canvas and fiberglass (5) Weighted
paper model (6) Full-scale model with ballistic-grade aramid fiber fabric . . . . . . . . 13
2.6 Police officer deploying final barrier prototype . . . . . . . . . . . . . . . . . . . . . . 14
2.7 All 15 layers in barrier construction, labeled . . . . . . . . . . . . . . . . . . . . . . . 16
2.8 Rear view of deployed ballistic barrier. Gas springs are placed horizontally to assist
in actuation. Over-center toggle supports, in lower left and right corners of the barrier,
constrain parasitic motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.9 (a) 9mm round striking top center of barrier (b) Panel deflection 20 ms after impact (c)
Permanent panel deflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.10 Fold pattern of the final bullet-resistant barrier design. . . . . . . . . . . . . . . . . . . 19
2.11 Left: Front-facing side of ballistic barrier V3. Right: Back/User-facing side of V3.
Note that no external hardware is needed for the barrier to be self-stable. . . . . . . . . 20
2.12 Diagram comparing the three different patterns from the full-scale barrier prototypes.
Each line drawing is to scale with the other five line drawings. . . . . . . . . . . . . . 21

3.1 Three types of developable surface classes. Generalized cylinder, generalized cone,
and tangent developable. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Three developable mechanisms (DMs). From left to right, a cylindrical DM with a
planar four-bar linkage, a conical DM with a spherical four-bar linkage, and a tangent
DM with a Bennett four-bar linkage. . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3 Behaviors and characteristics of cylindrical DMs and possible functions of developable
medical devices. The functions list is not comprehensive but is included to show po-
tential tasks. Functions marked with a (p) can be considered processes. . . . . . . . . . 28
3.4 A prototype demonstrating emergent motion behavior. Embedded mechanisms can
cleanly emerge from and conform to surfaces to preserve surface function and ap-
pearance, potentially including the water-tight behavior of solid cylinders. . . . . . . . 29
3.5 A Roberts’ straight line mechanism as a developable mechanism on a cylinder, demon-
strating the conforming characteristic. The first and third links are conforming and the
coupler link is partially conforming. . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.6 A prototype demonstrating the anchoring function. Feet on the coupler link could pro-
vide stability or prevent a cylinder from rolling. . . . . . . . . . . . . . . . . . . . . . 30
3.7 A chart showing the functions, characteristics, and behaviors needed to complete the
design of an example device. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

vi
3.8 Rendering of an Internal Scissors design embedded in a shaft. . . . . . . . . . . . . . . 34
3.9 Kinematic skeleton diagram model of the scissors mechanism, the developable mech-
anism version, and a 3D-printed prototype with altered link shapes. Note: The ground
link is not physically present in the prototype, and blue link is the inner cylinder, actu-
ated by rotating with respect to the outer cylinder. . . . . . . . . . . . . . . . . . . . . 35
3.10 A 3D-printed internal scissors prototype with a forceps tool inserted. . . . . . . . . . . 35
3.11 Kinematic skeleton diagram model of the multiplying cylinder mechanism, the devel-
opable mechanism version, and a 3D-printed prototype with altered link shapes. . . . . 36
3.12 Multiplying Cylinders mechanism with ISI forceps inserted. LED lights included to
demonstrate potential application as a dual-source flashlight or a vehicle for stereo-
scopic vision cameras. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.13 Rendering of multiplying cylinder superimposed on open source image from MIS ap-
pendectomy procedure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.14 Secondary channel created when an elongated version of the multiplying cylinder
mechanism is opened. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4.1 Left: The pseudo-rigid-body model (PRBM) demonstrated for a small-length flexural
pivot. Right: The PRBM applied on an initially-curved beam with a curved rigid link
attached on the end. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2 A 3D printed prototype of the expanding cylinder mechanism with one initially curved
flexure, highlighted in white. The path the right end of the flexure is forced through
creates a deflected shape that suggests end forces and moments, similar to a traditional
fixed-guided beam. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.3 Top: The wall thickness reduction method (reduces h in beam) for creating SLFPs or
pre-curved flexures. Bottom: The width reduction method (reduces w in beam), which
is only viable when the wall thickness is less than the beam width. . . . . . . . . . . . 44
4.4 Parallel guiding cylindrical developable mechanism with membrane joints. The cylin-
der is 38mm diameter polycarbonate tubing, cut on an Epilog M2 CO2 laser cutter,
and the membrane is red acrylic tape. . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.5 Left: Skeleton diagram showing a parallel guiding mechanism aligned to a reference
surface (dotted line) in the middle of the cylinder walls. Right: The mechanism on the
left, when joints are pushed to the location of the membrane hinges, is forced to change
its ground link length. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.6 Plot showing relative error between torsional constants of annular cross section and
rectangular cross section beams. The finite mesh in cross sections input to ANSYS
may account for the non-smooth pattern from low to high t/R ratio. . . . . . . . . . . 48
4.7 Spring diagram for an inside LET joint. . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.8 Schematic describing the parameters used in the MATLAB analysis and elsewhere in
this chapter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.9 Schematic of the software framework . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.10 Plot showing a comparison between curved LET joint stiffness in FEA and planar
assumption models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.11 Experimental test setup. Torque gauge is connected to white wire, laser encoder is
connected to colored wires. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

vii
4.12 Schematic showing the deflection path of the end of the LET joint arm (in red) and
the corresponding estimated circular axis of rotation (dashed line) used to design the
experimental setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.13 Comparison of experimentally derived 3D printed curved LET joint stiffness to FEA
and planar LET models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.14 FEA results of the Expanding Cylinder mechanisms with LET joints in nitinol. The
rigid beams, pinned to the cylinder, were rotated about the center of the cylinder to
mimic the circular displacement of the rotating inner cylinder actuator. . . . . . . . . . 56

viii
CHAPTER 1. INTRODUCTION

1.1 Motivation and Objective

The art of origami has been developed for centuries, with unique and interesting folding
patterns and motions being created along the way. Engineers have studied origami extensively
in the recent past to learn from its history and adapt these patterns and motions to mechanical
applications. One interesting behavior of many origami patterns is the ability to expand to a specific
shape and size and then collapse into a very small volume. This behavior is desirable in many
engineering applications, as space is a valuable and scarce resource in many designs. However,
most origami patterns were originally designed to be made with paper, a material that can readily
crease to create hinge-like motion. Research has been done in the past to show how origami patterns
can be adapted to engineering design, but the adapted designs often have various characteristics
that make the device impractical for application in the real world.
Developable Mechanisms (DMs), a new mechanism type that conforms to curved surfaces,
also have a similar ability to collapse to a small profile. These mechanisms show promise for en-
abling motion and force transfer around cylinders, cones, and tangent developed surfaces while
preserving the shape and profile of the curved surface when the mechanism is collapsed. Cylindri-
cal DMs may be of particular interest to the laparoscopic or minimally invasive surgical field, as the
majority of laparoscopic surgical devices are cylindrical in nature. These devices are consistently
pushed to smaller dimensions to become less invasive to patients and to enable new procedures. At
these small scales, the design process becomes difficult due to the nature of manufacturing, part
assembly, and reduced grip strength due to weak miniaturized pin joints and relatively high friction
forces. The emphasis on smaller devices suggests the need for compliant designs, which can have
much lower part counts and are more scalable than traditional pin and link mechanisms. The traits
of DMs could facilitate small-scale, compliant, and potentially monolithic cylindrical mechanisms.

1
The objective of this work is two-fold. First, this work seeks to demonstrate that origami
and DMs can advance various disciplines by showing practical and promising applications of these
technologies. Second, to increase efficiency of the analysis and design process for real-world de-
velopable mechanisms by presenting developments in the techniques available for modeling and
creating hinge motion in these mechanisms.

1.2 Thesis Outline

The outline of this dissertation is as follows. The background necessary for understanding
the subject matter will be presented at the forefront of each chapter.
Chapter 1 introduces origami and developable mechanism based design by presenting the
motivation and research objective for this work.
Chapter 2 presents the design of an origami-inspired deployable ballistic barrier. Previous
research on the folding of thick origami pattern enabled the creation of this device. To complete
the goal of creating a collapsing bullet-resistant shield, a number of folding patterns and thickness
accommodation techniques were collected and evaluated. Two thickness accommodation methods
were combined and then modified to enable a thick aramid fiber membrane with rigid interior
panels to fold into a compact shape. The barrier was tested at a firing range. This work demonstrates
that practical, manufacturable, and effective origami-inspired devices can be designed with proper
hinge design. This work was published in the Proceedings from the 7th International Meeting on
Origami in Science, Mathematics and Education, Volume 3, 2018.
Chapter 3 analyzes the properties of cylindrical developable mechanisms (DMs) and dis-
cusses their application to minimally invasive surgical tooling. DMs are a novel collapsing, con-
forming device type whose origins stem from origami and compliant mechanisms design. Two
example cylindrical DMs and prototypes are presented that have potential for adaptation to mini-
mally invasive surgical tooling. This work was published in the Proceedings from the International
Design Engineering Technical Conferences and Computers and Information in Engineering Con-
ference (IDETC/CIE), 2019.
Chapter 4 is an extension of the work in Chapter 3, and presents a study of the hinge
options available for small-scale cylindrical DMs. One promising option is lamina emergent tor-
sional (LET) joints. A MATLAB and ANSYS routine is developed to analytically validate the

2
assumption that LET joints in thin-walled singly-curved surfaces can be modeled with the simple
planar equations developed previously. The planar modeling method is also empirically validated
by characterizing a 3D printed PETG material before printing and testing curved LET joints of
various geometries.
Chapter 5 presents the conclusions drawn from this work and suggests future research
topics that will assist in making expanding devices, including developable mechanisms, viable for
use in the real world.

3
CHAPTER 2. ORIGAMI-INSPIRED BALLISTIC BARRIER

2.1 Introduction

The objective of this work was to demonstrate practical applications of novel mechanisms
that collapse and expand and research techniques for creating hinge motion in these mechanisms.
This paper presents the design of an origami-inspired deployable ballistic barrier, which was ac-
complished by combining and modifying various thick-origami hinge techniques previously re-
searched.
Techniques have recently been developed to assist in adapting origami design to engineer-
ing applications. Origami has been adapted to engineering applications on the macro scale [1] and
micro scale [2] in various disciplines, including biomedical applications [3] and consumer prod-
ucts [4]. Origami has also inspired the design of deployable disaster relief shelters [5]. In this paper,
we demonstrate how recent developments in origami adaptation techniques assisted in the design
of a deployable ballistic barrier.
Ballistic barriers are large shields designed to protect single or multiple users from ballistic
threats. They are often heavy and difficult to transport before or during deployment and use. Bal-
listic barriers have been designed that collapse for ease of transportation, but they have multiple
degrees of freedom, which slows deployment time and can require extensive external hardware to
support the barrier. Adapting origami design to this field can provide multiple benefits, as many
origami patterns are low-degree-of-freedom mechanisms with large deployed surface areas and
large deployment ratios. A barrier design with these characteristics may provide benefits for po-
lice, specialized forces (such as SWAT or SCO19), military, and private security applications. Its
compactness when stowed may also make it attractive for use in schools, corporate offices, and
other security-sensitive public places.
This work demonstrates the design of a light-weight, compact, and deployable ballistic bar-
rier that draws utility from its monolithic construction, large deployment ratio, and rapid actuation.

4
The barrier design utilizes recent developments in crease pattern adaptation to generate deploy-
ment motion, surrogate folds to allow for crease-like motion of ballistic material, and thickness
accommodation techniques to assist with the folding of the thick ballistic membrane.

2.2 Methods

2.2.1 Ballistic Barrier Requirements

The ballistic barrier was conceived as a defense mechanism to be used by law enforcement
officials, so the requirements were shaped to help the barrier be a potential aid in high-stress,
rapid-response situations. After interviewing current and former police officers, members of the
Utah County SWAT team, and Homeland Security armed officers, the following requirements list
was decided upon:

• Ballistics protection consistent with U.S. National Institute of Justice (NIJ)


level IIIa standard
• Deployment time of less than 5 seconds
• Small storage volume (size of a spare tire - 4.74 ft3 )
• Single person operation
• Self-standing once deployed
• Minimum 4 ft x 4 ft frontal area
• Maximum mass of 50 lb
• Prevents injury to user in case of blunt trauma (bricks/rocks/etc)
• Dark, non-reflective color
• Operating temperature range of -5 ◦ F to 120 ◦ F

2.2.2 Adaptation of Origami Crease Patterns

A key advantage of origami-based research is that it builds on a wealth of origami art de-
veloped over many centuries, although systematic selection methods do not yet exist for choosing
an origami pattern given certain criteria. For many practical applications, intuition and experience
can be used to select an origami pattern that will help meet the required function of a particular
device.

5
Rigid-foldable crease patterns are of particular interest when transferring origami models
to engineering materials. Rigid-foldable crease patterns do not require the facets to deflect or the
creases to stretch in order to fold. In many origami-based engineering applications, the facet ma-
terial should not deflect, as this will introduce additional degrees of freedom to the mechanism.
Also, if a facet deforms during actuation it stores energy, complicating actuation and posing po-
tential problems when the energy is released.
The behavior of the barrier during deployment and use relies heavily on the origami crease
pattern utilized. A crease pattern fit for a deployable ballistic barrier would be rigid foldable to
allow stiff facets, flat-foldable to enable storage in a compact shape, have a large frontal area in the
deployed state, and have a low number of degrees of freedom for ease of actuation. The choice of
crease pattern also affects the final deployment ratio and the stability of the barrier during use.
Choosing an origami pattern to apply to the barrier is key to achieving these goals. While
there are many origami patterns that perform these functions, not all lend themselves well to thick
origami. Three thick-foldable and rigid-foldable patterns were explored in detail: the Miura-ori
[6, 7], square twist [8], and Yoshimura (or Diamond) [9, 10].
The Miura-ori was considered due to its ability to fold compactly and unfold to a large
sheet. However, the Miura-ori pattern is more difficult to achieve in thick materials as many varia-
tions of the pattern have four layers of material nesting inside one vertex. The square-twist pattern
deploys to a large flat shape, but it has a relatively small deployment ratio. Both the Miura-ori and
the square-twist deploy to near-flat states, which would require external supports to achieve stabil-
ity in the deployed state. While the Yoshimura pattern has the disadvantage of multiple degrees of
freedom, which makes it harder to control, it forms a curved shape while in its intermediate state.
This partially deployed shape offers more stability and could provide front and flank protection for
the user.
Variations of the Yoshimura pattern were then explored. To reduce the number of degrees of
freedom in the pattern and provide a more practical shape for manufacturing purposes, the barrier
pattern in Figure 2.1 was developed. It is composed of identical, horizontally mirrored degree-4
vertices with the diagonal creases set to 45 degrees to allow the sides to fold backward at a 90
degree angle in the collapsed state (see Fig. 2.4). This is similar to a pattern first analyzed by
Huffman [11], with his A/B and C/D lines made parallel, and similar to a pattern which Gattas et

6
Figure 2.1: The crease pattern selected for the barrier design.

al. called an extended Yoshimura [12]. A nearly identical pattern is found in a magician’s device
called a “Troublewit,” with the earliest known publication on the device seen in 1676 [13].
After pattern selection, the remaining independent variables in the design were the number
of rows and columns of facets in the pattern. For practical manufacturing, it is advantageous to have
fewer, taller rows. If the same total vertical dimension is kept but more rows are added, collapsed
height increases and collapsed base dimensions decrease. Storage efficiency also decreases because
the empty volume in the middle of the collapsed pattern increases (Fig 2.4). A barrier pattern with
6 rows and 3 columns was selected because it has small collapsed-base dimensions while keeping
the fabrication process simple.

2.2.3 Surrogate Folds

Because most materials do not crease like paper, achieving crease-like behavior is a chal-
lenge typical to most origami-based engineering applications. When paper is folded to create a
crease, the fold becomes a localized area of decreased stiffness. This decrease in stiffness is what
enables the hinge-like characteristic of paper folds. Many other materials, however, experience a
localized increase in stiffness when folded. For example, folding or bending metal does not en-
able repeated rotational motion at that location [14]. Therefore, creating crease-like behavior in
materials other than paper is a key component of origami-based design.
Surrogate folds allow for hinge-like motion in a non-paper material. Several methods for
creating surrogate folds have been developed to provide options besides mechanical hinges for
obtaining rotational motion [15, 16]. The living hinge method involves locally reducing stiffness

7
by reducing material thickness along the fold line. Another surrogate fold known as a Lamina-
Emergent Torsional (LET) joint, involves cutting material away in specific patterns to allow torsion
along the fold line, instead of purely bending, to create the folding motion [17]. In the membrane
fold technique [18], separate stiff panels are attached to a flexible membrane, which allows for
rotational motion at the breaks between panels, as seen in Figure 2.2.
In order to meet a NIJ level IIIa standard of ballistic resistance [19], twelve sheets of
ballistic-grade aramid fiber fabric are normally required. The membrane fold technique was cho-
sen to create surrogate folds in the barrier design because the twelve layers of aramid fiber fabric
could be used as a multi-layer membrane. The membrane fold technique usually involves fitting a
membrane onto rigid panels, but here the membrane was first chosen and its thickness identified,
followed by designing the rigid facets. The panels are designed to match the facets of the fold
pattern, with additional space between each panel to allow fold motion. Thickness accommodation
techniques were used to design the panel geometry and spacing.

2.2.4 Thickness Accommodation

The paper used in traditional origami is thin enough to assume a zero-thickness model,
where the thickness of the folding material is ignored. Most engineering applications require the
use of non-paper materials. Therefore, methods have been developed to accommodate thickness
in origami-based engineered systems [20]. Many of these methods focus on the kinematics of an
origami pattern and accommodate thickness through the movement of rotational hinges [21, 22] or
by changing the shape of the facets in the pattern [23, 24]. These methods can result in patterns
where holes are present in the hinges during some or all of the folding motion. Methods that use
compliant joints to accommodate thickness have also been developed [18].
The membrane fold technique used to create surrogate folds was adjusted to accommodate
folding of the thick facets, which are comprised of multiple aramid fabric layers and the rigid
panels. This application differs from the typical implementation of the membrane fold technique
in that here the panels are thinner than the membrane and the membrane is several fabric sheets
stacked together rather than a single flexible material. Figure 2.2 compares the membrane fold
technique (a) to two thick-membrane adaptations considered. The first approach, called a “turn of
cloth” (b), is common in sewing. This results in a pattern where the fabric in an unfolded state

8
Figure 2.2: Profile view of the membrane fold (a), the ”turn of cloth” fold (b), and the center-panel
gap fold (c).

bunches out from the fold line. The second approach, the center-fold gap approach (c), does not
use bunches to accommodate folding, but provides a larger gap between the two rigid panels.
The turn of cloth technique resulted in a hinge area that is less stiff than one where no
turn of cloth is performed. An early full-scale canvas prototype was constructed, which helped
to evaluate this technique. The flexibility of the hinges allowed the barrier to fold compactly, but
the approach added manufacturing complexity and made it more difficult to create a self-standing
barrier. The gap fold approach could not fold as compactly, but the added hinge stiffness aided
in maintaining the deployed position and simplified the manufacturing process. To accommodate
both mountain and valley folds, it was decided to place the rigid panels in the middle of the 12
aramid fabric layers. This also helps prevent shrapnel from the rigid panels reaching the user.
The spacing between rigid panels was designed by adapting a recently developed approach
[25] to the barrier. The radius of curvature, or width of the gap between panels, is defined based
on the thickness of the panels involved in the crease and the additional length needed due to the
membrane curving around the rigid panels rather than folding at a right angle. A wider gap is
needed for folds that involve nested panels, or panels that rest inside other panels in the folded
position. By adding the thickness of the aramid fabric membrane to the model, we defined the
minimum gap sizes allowable while still maintaining 180 degrees of fold motion. The layout of the
panels, including relative gap sizes, is shown in Fig.2.3.

9
Nesting Fold,
Gap Size 4.5 cm
Non-Nesting Fold
Gap Size 2.0 cm
Rigid Panels

Figure 2.3: Drawing of final barrier topology. Each larger nesting fold encloses the two adjacent
panels in the folded position. White sections are rigid panels, and the yellow area is the aramid
fabric covering both sides of the panels.

2.2.5 Stability

An effective ballistic barrier must be able to retain its deployed state under external forces.
A key issue with stability is the weight of the barrier compared to its hinge stiffness. The paper
used in traditional origami is light compared to the stiffness of the folds in the paper. As the fold
pattern is scaled up to thicker materials, the increase in the weight-to-fold-stiffness ratio causes the
pattern to collapse under its own weight. Therefore, methods were explored to stabilize the barrier
without hindering deployability.
One major aid in achieving stability is the chosen origami fold pattern. In an actuated
system, a lower number of degrees of freedom is beneficial to increase stability. The extended
Yoshimura pattern used is based on the classic Yoshimura pattern, but with specific vertices elon-
gated to create an additional horizontal fold line [12]. This results in hexagonal-shaped units com-
posed of two trapezoid panels. The vertex elongation effectively reduces the Yoshimura pattern to a
single-degree-of-freedom mechanism, which makes the barrier easier to stabilize and actuate with
external supports such as telescoping rods or gas springs. Also, in the deployed state, the barrier
pattern resembles a crescent shape from a top-down view. Figure 2.4 shows a top-down view of the
barrier deployment. The crescent shape increases the footprint of the barrier and shifts the center
of gravity toward the middle of the footprint, which stabilizes the barrier against external forces.
The number of rows in the overall fold pattern also affects the stability of the barrier through
the deployed footprint and diagonal sheering of the panels. Decreasing the height of each row in
the pattern, so a larger number of rows is included in the same vertical dimension, increases the

10
X

(a) (b) (c)


Figure 2.4: Top-down view of barrier pattern deployed states. (a) Stowed state (b) Deployed state,
’X’ marking center of gravity (c) Fabrication state (scaled)

footprint in the deployed state. Dividing the fold pattern into more rows also increases the distance
between the diagonal creases. This separation eliminates one degree of freedom in the final barrier
design. The closer diagonal fold lines are to being collinear, the more diagonal sheering, or folding
along those two hinges as if it were one continuous hinge, occurs.
A final source of stability is the stiffness of the fold lines or joints. For ease of folding, low
stiffness is preferred. For stability, it is beneficial to have a larger stiffness. If the barrier folds were
stiff enough to prevent the barrier from collapsing, it would be difficult to actuate. Therefore, an in-
termediate value of stiffness was sought that would facilitate stability in the deployed state without
compromising ease of actuation. Motion limiting features, analogous to the umbrella deployment
mechanism, were added to define the deployed state and prevent the barrier from collapsing back
to the stowed state.
The design of the folds is largely defined by the chosen surrogate fold method. The mem-
brane fold technique leaves only the spacing between the two rigid panels and the method of
combining the layers of aramid fabric as variables. Decreasing the space between rigid panels in-
creases the joint stiffness because the hinge length decreases but the required angular deflection
remains the same. Increasing the space between the rigid panels has the opposite effect but also
increases the amount of parasitic motion in the joint. After experimentation with several sample
panels, gap sizes of 4.5 cm for the nesting folds and 2.0 cm for the non-nesting folds were chosen
for the barrier, as seen in Figure 2.3.
Many options exist for combining multiple layers of fabric so they behave as one thicker
piece of fabric. Several potential paths are similar to manufacturing methods for creating carbon
fiber layups, including vacuum fusion to impregnate epoxy into the fabric in the the facet areas.

11
Vacuum epoxy infusion could eliminate the need for a separate rigid material inside the aramid
fabric, as the epoxy deposited in the appropriate areas, with epoxy-free gaps for fold lines, could
automatically create facets and folds. Other lamination methods include heat pressing the fabric
with sheets of glue between layers and applying spray adhesive between layers. The latter method
was used in manufacturing the ballistic barrier.

2.2.6 Ballistics Protection

A U.S. National Institute of Justice (NIJ) ballistic resistance rating of IIIa was chosen as
a goal for the barrier. Many bullet-resistant vests worn by law enforcement officers have a IIIa
rating. A IIIa rated ballistic panel should prevent penetration of bullets at or below a .44 magnum
caliber [19], which includes the popular .45 ACP round. 12 layers of ballistic-grade aramid fiber
fabric are usually required to reach a IIIa rating if aramids are the only ballistic material included.
The barrier was therefore designed with 12 layers of aramid fiber fabric.
To maintain a IIIa rating over the entire surface of the barrier, it was necessary to maintain
12 layers of fabric at every location, including the hinges of the fold pattern. This requirement led
to the application of the membrane-fold thickness-accommodation technique. For such a solution
to be viable, the membrane itself must meet the ballistic resistance requirements and be flexible
enough to permit folding motion. Ballistic-grade fabrics, such as aramids (Kevlar®) and ultra-
high-molecular-weight polyethylene (UHMWPE) fabric, satisfy these requirements when several
layers are combined. An aramid fabric was chosen for the barrier design due to the higher cost of
UHMWPE, but UHMWPE is a viable and lighter alternative.
To become rigidly foldable, aramid fabric needs to be stiffened in certain locations and
left relatively flexible in others to mimic the fold pattern of the desired origami shape. The panel
stiffening approach taken for the ballistic barrier application was to insert a flat, rigid material
that serves a structural stiffening purpose rather than a ballistic function. This solution provides
needed stiffness while ensuring continuous ballistic protection at the facets and folds. Alternatively,
tougher or thicker ballistic panels can be integrated to serve both ballistic and structural functions,
increasing the level of protection the barrier can achieve.

12
Figure 2.5: Prototype sequence, top left to bottom right: (1) Thick-folding prototype inspiring
barrier design, using 2.5 cm-thick foam board for panels and fabric tape for surrogate folds (2)
Initial paper prototype (3) Wood and felt (4) Canvas and fiberglass (5) Weighted paper model (6)
Full-scale model with ballistic-grade aramid fiber fabric

2.2.7 Prototyping

The general sequence of prototypes leading to the final barrier design is shown in Figure
2.5. The prototype inspiring the creation of the barrier was a thick-folding prototype originally
designed to demonstrate how the hinge-shift technique could be combined with tapered panel hard
stops [24] to create unique deployed geometry [26]. Simple paper models were created to quickly
evaluate various modifications of the elongated Yoshimura pattern. Wood and felt models assisted
in solidifying the correct thickness accommodation technique for the barrier. A full-scale canvas
and fiberglass prototype was created to explore possible actuation methods and to determine the
effect of scaling on the weight-to-stiffness ratio and how this would effect deployment motion and
stability. Weighted card stock models were used to quickly evaluate various actuation techniques.
A full-scale, ballistic-resistant prototype was then fabricated.

13
Figure 2.6: Police officer deploying final barrier prototype

2.3 Results

2.3.1 Full-Scale Ballistic Resistant Prototype

A full-scale ballistic barrier was constructed to demonstrate the utility of the origami-based
design developments discussed above. Figure 2.6 shows the final prototype, which is constructed
using an adapted membrane fold technique, where the thick aramid fabric membrane is on both
sides of the rigid panels. Table 2.1 shows the design requirements and the final measured values,
for those that were non-obvious and measured after the design was completed. This modified
surrogate-fold technique simplifies thickness accommodation, as both mountain and valley folds
have the same gap size between the rigid panels. The ballistic-grade fabric used is Kevlar® 29
Style 745. Rigid Omega-Bond® panels are inserted between layer 6 and 7 to provide the stiffening
necessary to separate facets from folds. The specific panel used is an Omega-Bond® Premium
sandwich composite panel, with two layers of aluminum sandwiching a polyethylene core. The
aluminum sheets are each 0.3 mm thick, and the whole panel is 3 mm thick.

14
Table 2.1: Table showing the original design requirement goals versus values measured from final
prototype testing.

Measured
Requirement Goal
Values
Ballistic Level 3a 3a
4.2 sec with case,
Deployment Time 5 sec
2.8 sec without
Storage Volume 4.74 ftˆ3 4.04 ftˆ3
Persons to operate 1 1
Stability Self-standing Self-standing
Frontal Area 48 in x 36 in 58 in x 40 in
Weight 50 lbs 53 lbs

2.3.2 Fabrication

Fabrication of the prototype barrier seen in Fig.2.6 was performed at Brigham Young Uni-
versity. First, the rigid OmegaBond® panels were cut to shape with a water-jet cutter. These were
placed and arranged on top of a single sheet of aramid fabric and adhered to it using 3M® Super
77 spray adhesive. Six layers of aramid fabric were then adhered on top of the panels with spray
adhesive applied between each layer, followed by a single layer of black nylon fabric. The compos-
ite was then turned over and five more layers of aramid fabric were adhered, finished with another
layer of ballistic nylon. Figure 2.7 shows the layers of the barrier. Immediately after the layers were
arranged, the barrier was pre-creased and folded to bias the folds in the correct orientation before
the adhesive set. The unfinished edges of the barrier were trimmed with a strip of ballistic nylon
and sewn to prevent the aramid fabric from fraying. The use of vacuum epoxy infusion to stiffen
the facet areas of the barrier is a viable solution for panel stiffening and manufacturing in greater
quantities.

Several additions to the construction serve to help with the functionality of the barrier.
Aramid fibers readily absorb water and degrade quickly when exposed to UV radiation. To prevent
both of these damaging elements from reaching the aramid fabric, both sides of the barrier were
finished with one layer of ballistic nylon, which is a nylon-based, water-proof, and UV resistant

15
Figure 2.7: All 15 layers in barrier construction, labeled

fabric commonly used in ballistic armor and consumer products such as backpacks and computer
cases.
Two hardware additions relate to the stability of the barrier. Because the fold pattern adapted
to the barrier creates a single-degree-of-freedom mechanism, the barrier can be actuated with
extension poles or gas springs placed in strategic locations. The barrier prototype includes gas
springs, similar to those seen in automotive trunks or hatches, to assist in manual actuation. Two
gas springs are used in the prototype, each with a spring force of 110 N. They are mounted horizon-
tally near the top of row 2 and the bottom of row 5, as seen in Figure 2.8. The mounting hardware
includes bolts and small plates on the front-facing edge of the barrier to spread the bolt force over
a wider area, which prevents the bolt head from being pulled back through the fabric.
The second hardware addition is the inclusion of over-center toggle supports, seen in Figure
2.8. The gas springs successfully open the barrier as the user lifts up on the attached handles, but
parasitic motion introduced from torsion in the fabric folds of the barrier causes the outer edge
of the bottom two layers to sag after deployment. As the barrier is being opened, the over-center
supports pass their toggle position and hit a hard stop that keeps the two legs of the supports at an
angle slightly greater than 180 degrees. This allows the over-center supports to hold a compressive
load, which keeps the bottom two vertices of the barrier open to the desired angle of 45 degrees.

2.3.3 Ballistics Demonstration

To demonstrate performance of the barrier, a ballistics test was performed with three goals
in mind. First, it was necessary to determine if the design and construction of the barrier was
resistant to handgun rounds. Testing of the full barrier was necessary to observe how the entire

16
Figure 2.8: Rear view of deployed ballistic barrier. Gas springs are placed horizontally to assist
in actuation. Over-center toggle supports, in lower left and right corners of the barrier, constrain
parasitic motion.

system performed, as the interaction of dozens of different panels in one system could not be
witnessed with individual sample panels. Second, a stability analysis was required to determine
if the unsupported barrier would remain stable when fired upon by various handguns. Third, we
desired to determine the effect that various design parameters have on ballistic resistance. This
was accomplished using sample panels with various spacing in the fold between the rigid panels,
different rigid panel materials, and flat panels fired at with various angles of attack.
Standards published by the National Institute of Justice require Type IIIa armor to be tested
with .357 magnum and .44 magnum handgun rounds [19]; therefore, these rounds were used for
testing. The same standard also specifies bullet velocities, angles the bullets should strike the test
panels, how many shots should be tested on each panel, and how close together the shots are
allowed to be. Each of these test requirements was followed on two test panels per test type,
although the NIJ specifies that 4 of each test type should take place on 4 different panels. 9mm
Luger rounds were also used, as 9mm caliber firearms are the most commonly confiscated firearms
in the Unites States. [27]
The sample panels and the full barrier were fired at from a distance of 4.3m, using the above
specified rounds. A Phantom V1610 high-speed camera recording at 30,000 frames per second

17
Figure 2.9: (a) 9mm round striking top center of barrier (b) Panel deflection 20 ms after impact (c)
Permanent panel deflection

was used to capture the bullets hitting the barrier and to determine the muzzle velocity of various
calibers. The .357 magnum, .44 magnum, and 9mm rounds reached a muzzle velocity of 556, 417,
and 425 m/s, respectively. All test specifications given by the NIJ were met or exceeded in our
testing, except the .44 magnum muzzle velocity, which was 19 m/s slower than required. Figure
2.9 shows a bullet striking the barrier and the corresponding panel deflection at two different times
after impact.
A total of 20 handgun rounds were fired at the full barrier, and 42 rounds were fired at 7
different sample panels. Of the 62 handgun rounds fired, three bullets penetrated; two bullets passed
through when they struck bolt heads, and one of the ten rounds fired at the fold lines in the test
panels penetrated. These results demonstrated that the barrier could pass NIJ 3a class certification
testing with slight modification. The hardware mounting area of the barrier and the valley fold,
where the panels are separated to allow fold motion and some fabric delamination occurs, are both
areas of lower ballistic resistance. This information was used to improve the next prototype, where
the hardware issue was addressed. To test the stability of the barrier, a .44 magnum round was fired
at the top center of the barrier while it was in a deployed state without any external supports. The
barrier did not tip, slide on gravel, or shake a significant amount.
The ballistics testing helped demonstrate that recent advances in rigid foldability, thickness
accommodation, and surrogate fold techniques can effectively be combined to design a deployable,
origami-based ballistic barrier.

18
Figure 2.10: Fold pattern of the final bullet-resistant barrier design.

Future work on the barrier should involve testing the fatigue limits of the barrier, as internal
friction in the hinges may cause wear as the barrier is cyclically actuated.

2.4 Final Barrier Design

Several months after the conclusion of the ballistic barrier project, the fold pattern was re-
visited in an attempt to make several improvements to the previous design. For this report, Version
1 of the barrier is considered to be the first bullet-resistant prototype that was tested at the firing
range. In Version 2, which is briefly mentioned at the end of Section 2.5, hardware improvements
were made the increase stability and bullet resistance. Version 3, whose fold pattern is seen in
Figure 2.10, has the largest frontal area (4.0 ft by 5.1 ft) and smallest storage volume (2.57 ftˆ3) of
the three designs. This design also removed several of the weaknesses inherent in the fold pattern
and design used by Versions 1 and 2. The fold pattern for Version 3 can be seen in Figure 2.10.
First, all the major fold lines are vertical, which makes this design much more stable against
gravity and allows the barrier to be self-standing, and hold additional weight, without any support-
ing or actuating hardware.
Second, all of the panels are accessible from the outside of the barrier. This means flaps
can be added to the edge of the barrier to access and remove the rigid interior panels, allowing for
the possibility of upgrading to rifle-round plating.

19
Figure 2.11: Left: Front-facing side of ballistic barrier V3. Right: Back/User-facing side of V3.
Note that no external hardware is needed for the barrier to be self-stable.

This feature also allows the barrier to be sewn instead of adhered with spray glue. Because
the rigid panels can be added after the layers are attached together, the entire barrier can be fabri-
cated in a flexible state, allowing the manufacturer to roll portions of the barrier and sew the layers
together. In Versions 1 and 2, the rigid panels were not accessible from the edge of the pattern, so
they were included in very beginning of the fabrication process. This prevented traditional sewing
machines from reaching the center of the barrier.
Lastly, while the final pattern is also a single degree-of-freedom pattern like Versions 1
and 2, the horizontal and vertical motions are almost separated. When unfolding the barrier, nearly
all of the vertical expansion occurs before the “wings” on the side begin to expand. This also
allows the user to collapse the barrier horizontally without losing much vertical height, allowing
unhindered travel through narrow doorways.
Figure 2.12 shows a to-scale representation of all three major barrier patterns, including
the original full-scale canvas model, the first two ballistic models (Version 1 and 2), and the final
ballistic model (Version 3). The patterns and fully-folded state representations are all drawn at the
same scale, allowing for a direct comparison between patterns to be made.

2.5 Discussion and Conclusion

Designing the origami ballistic barrier has revealed that surrogate fold and thickness ac-
commodation techniques can be used to adapt origami design and patterns to previously unrelated

20
21
Figure 2.12: Diagram comparing the three different patterns from the full-scale barrier prototypes. Each line drawing is to scale with the
other five line drawings.
fields, such as ballistic shielding. The membrane fold technique can also be modified to include a
membrane much thicker than the rigid panels as long as the gaps between panels are carefully de-
signed. It was also reinforced that the benefits of origami, such as customizable deployment shapes
and motions and large deployment ratios, can transfer to various fields of study with proper appli-
cation of origami adaptation and accommodation techniques. In the final design and prototype,
all of the original design requirements were met or exceeded, with the exception of the operating
temperature range, which was not directly tested.
Several weaknesses to the current design were identified and remedied. Parasitic motion
in the folds allowed the outer edges of the barrier to sag, which led to the addition of over-center
hinges to support the lower four horizontal vertices. Also, penetrating the aramid fabric layers
with hardware or allowing the layers to delaminate is detrimental to the ballistic resistance of the
barrier. A second prototype, with the same geometry and materials as the first, was created with
improvements to alleviate these issues. After experimentation with scaled models, the gas springs
were mounted in different locations, which allows the barrier to be self-standing without the use
of over-center hinges. The springs were also mounted to plastic plates, which were then sewn into
the rear 6 layers of the barrier. This allowed mounting of the springs without penetrating the fabric
with additional hardware.

22
CHAPTER 3. CYLINDRICAL DEVELOPABLE MECHANISMS FOR MINIMALLY
INVASIVE SURGICAL INSTRUMENTS

3.1 Introduction

The objective of this work was to demonstrate practical applications of novel mechanisms
that collapse and expand and research techniques for creating hinge motion in these mechanisms.
This paper presents an analysis of the properties of cylindrical developable mechanisms, a newly
developed deploying mechanism on cylinders, and discusses the application to minimally invasive
surgical devices. It should be noted that this work focuses on demonstrating design feasibility
rather than reviewing specific engineering designs. Emphasis was placed on achieving the desired
motion with the novel developable mechanisms rather than completing an in-depth device design.
Medical devices have been produced and developed for millennia, and improvements con-
tinue to be made as new technologies are adapted into the field. Developable mechanisms, which
are mechanisms that conform to or emerge from certain curved surfaces, were recently introduced
as a new mechanism class [28]. These mechanisms have unique behaviors that are achieved via
simple motions and actuation methods. These properties may enable developable mechanisms to
create novel hyper-compact medical devices. In this chapter I discuss the behaviors, characteristics,
and available functions of cylindrical developable mechanisms, review a design process for adapt-
ing these mechanisms to medical devices, and review two example minimally invasive surgical
device prototypes that demonstrate design feasibility.

3.2 Background

Foundational work leading to this research followed a relatively linear path: first compliant,
or flexible, mechanisms were developed [29], then they were created out of flat surfaces with
motion that emerges from the plane (these mechanisms are called lamina emergent mechanisms, or
LEMs [30, 31]. Arrays of compliant lamina emergent joints were then used to create developable

23
surfaces from rigid materials, including curved-fold origami models [32, 33]. This inspired the
adaptation of compliant and traditional rigid mechanisms to developable surfaces to create a new
mechanism class [28]. In this work, these new developable mechanisms are adapted to medical
devices.

3.2.1 Compliant and Lamina Emergent Mechanisms

A compliant mechanism is a mechanism that transfers motion, force, or energy, and gains
its mobility through the flexibility of its members [29]. They can reduce cost due to their low part
count and offer increased performance, including increased precision and reliability and decreased
wear, weight, and maintenance. The development of analysis techniques such as the pseudo-rigid-
body model [29, 34, 35] and topology optimization [36, 37] have simplified analysis and allowed
compliant mechanisms to become prevalent in many industries.
Lamina Emergent Mechanisms (LEMs) are mechanisms fabricated from sheet material that
emerge out of plane as they actuate. LEMs can also be made fully compliant, so they can share
the same advantages as compliant mechanisms. Thanks to stereolithography fabrication and the
scalability of LEMs, they have enabled mechanisms and motions at the micro- and nano-scale.
Examples include an on-chip nanoinjector [38] and a demonstrative belltower [39]. The initial flat
state of LEMS introduces the limitation that only mechanisms with a change point can be used
in their design [30]. A noteworthy LEM is the Lamina Emergent Torsional joint [40, 41], or LET
joint, which provides sheet material greater flexibility in a desired location and along a desired
axis.

3.2.2 Developable Surfaces

A developable surface can be flattened into a plane without stretching or compressing the
surface. One example is ship hulls, where bending the plate material into a developable surface
requires much less energy than traditional forming methods, which can require the sheet material
to stretch or compress. Ushakov [42] stated the following characteristics of developable surfaces:

1. They are ruled surfaces


2. They have zero gaussian curvature everywhere

24
Figure 3.1: Three types of developable surface classes. Generalized cylinder, generalized cone, and
tangent developable.

3. They are isometric (do not stretch or bend when folded flat)
4. Each is an envelope of a 1-parameter family of planes

A ruling line is a straight line in space, and a ruled surface is created by sweeping a ruling
line through space. In other words, through every point on a developable surface there must be a
direction where a straight line can be fit to the surface. A developable surface also has zero gaussian
curvature everywhere, or the product of the maximum possible and minimum possible curvatures
at any point is zero. This means that the surface needs to be curved in only one direction at every
point, like a cylinder. Doubly-curved surfaces have a positive or negative Gaussisan curvature, such
as the sphere with positive Gaussian curvature, or the saddle with negative Gaussian curvature.
Developable surfaces are also isometric, which here means that the path length between two
points on the curved surface is identical to the path length between those points when the surface
is flattened. Finally, a developable surface is an envelope of a one-parameter family of planes.
Figure 3.1 shows the three types of developable surfaces that exist besides planes. These are
generalized cylinders, generalized cones, and tangent developable surfaces. Developable surfaces
can be defined by their ruling lines. In a generalized cylinder, all of the ruling lines are parallel.
Generalized cones are formed when all of the ruling lines meet at a point. A tangent developable
surface is formed by keeping the ruling lines tangent to any three dimensional curve [43].

3.2.3 Developable Mechanisms

Developable Mechanisms are a new classification of mechanism currently being researched


[28]. A developable mechanism is a mechanism that conforms to or emerges from a developable
surface. Lamina Emergent Mechanisms emerge from a flat plane, but they cannot be formed onto

25
Figure 3.2: Three developable mechanisms (DMs). From left to right, a cylindrical DM with a
planar four-bar linkage, a conical DM with a spherical four-bar linkage, and a tangent DM with a
Bennett four-bar linkage.

a curved surface without distortion. This can cause the mechanism to bind or change its motion
and behavior. Developable mechanisms satisfy the need for mechanisms that can fit to a curved
surface.
Three example developable mechanisms are illustrated in Figure 3.2. Developable mech-
anisms can be created from any developable surface by aligning the hinge axes of an applicable
kinematic mechanism with the ruling lines of the surface. For example, four-bar mechanisms can
be fit to a plane surface, generalized cylinder, or generalized cone [28]. Planar mechanisms map
well to generalized cylinders, and spherical mechanisms map to generalized cones. The Bennett
four-bar linkage and all 7R linkages, or seven-link mechanisms with revolute joints, can be fit to
tangent developable surfaces. The large number of links in 7R tangent developable mechanisms
creates many change points, which can make these mechanisms difficult to implement in practice.
Developable mechanisms can solve a variety of problems and be fit to a wide number of applica-
tions, such as doors on a train that emerge from the curved body, wheels that have an embedded
mechanism to create a walking motion, or casts and splints that quickly attach to a limb [28].

3.3 Properties of Cylindrical Developable Mechanisms

Here we will outline the behaviors and characteristics of cylindrical developable mecha-
nisms and possible functions of developable medical devices to provide a guide for how to utilize
these mechanisms in device creation. The Behaviors column of Figure 3.3 lists behaviors from

26
classical mechanisms that are possible in cylindrical developable mechanisms. The Characteristics
column lists general characteristics of developable mechanisms, and the last column shows both
general functions and functions that can be derived from the specific mechanisms discussed in this
paper.

3.3.1 Behaviors of Developable Mechanisms

Because developable mechanisms share similarities to classical kinematic mechanisms,


many behaviors will also be shared with classical linkages. The major novel behavior appears
when we consider the adaption of classical linkages to developable surfaces. This behavior, termed
Emergent Motion, describes how the mechanisms can be designed from and move out of the ma-
terial that makes up a developable surface and re-conform to that surface after actuation. The
prototype in Figure 3.4 demonstrates how the wall of a cylinder can be cut in certain areas to allow
a 4-bar mechanism to actuate, using the cylinder wall itself as linkages, but still resemble a cylin-
der in the closed configuration. These behaviors also exist in lamina emergent mechanisms, but
important differences will be manifest as we consider the characteristics of and functions available
to developable mechanisms.
Other behaviors of developable mechanisms include predictable motion (comprising the
various motions described by Joskowicz [44]), energy storage, and force transfer. The behavior
of energy storage exists when compliant joints are used, which also allows for the behavior of
reaching multiple stable points to exist.

3.3.2 Characteristics of Developable Mechanisms

Nelson et al. [28] stated the following as the required criteria for a mechanism to belong to
the developable mechanism class:
A developable mechanism must

1. Conform to or lie within a developable surface when both are modeled with zero thickness
2. Preserve the shape of the developable surface throughout its motion
3. Have mobility

27
28
Figure 3.3: Behaviors and characteristics of cylindrical DMs and possible functions of developable medical devices. The functions list
is not comprehensive but is included to show potential tasks. Functions marked with a (p) can be considered processes.
Figure 3.4: A prototype demonstrating emergent motion behavior. Embedded mechanisms can
cleanly emerge from and conform to surfaces to preserve surface function and appearance, poten-
tially including the water-tight behavior of solid cylinders.

Two of the characteristics listed in Figure 3.3 for developable mechanisms are comprised
of the first two requirements above. These requirements also highlight the need for the axes of the
mechanism joints to align with ruling lines in the developable surface, which Nelson [28] terms the
joint-axis ruling condition. As a subset of criteria 1, a developable mechanism has a characteristic
that allows it to be manufactured in a flat state and then curved to a developable surface state. The
requirement of mobility was discussed earlier as a behavior.
When considering the connection to kinematics, developable mechanisms follow the rules
and classifications of classical kinematics, which is another defining characteristic. Also, if each
joint of the developable mechanism is fully compliant, it can be characterized as a monolithic
mechanism.
If both criteria 1 and the joint-axis ruling condition are met, a visual or geometric charac-
teristic emerges. These two characteristics will force the links of the mechanism to contain one or
more curved surfaces. Figure 3.5 shows a straight-line mechanism with curved links that conform
to and can be concealed within the cylindrical surface and a coupler link that partially conforms.

3.3.3 Functions

The collection of functions of developable mechanisms discussed here does not represent
a comprehensive list, but is provided to explore the breadth of tasks these mechanisms could com-
plete. The successful designer or engineer will use the behaviors and characteristics of developable
mechanisms to create new features and find new applications in their work. Functions labeled with
a “p” can be considered a processes, or a combination of functions, possibly from other non-

29
Figure 3.5: A Roberts’ straight line mechanism as a developable mechanism on a cylinder, demon-
strating the conforming characteristic. The first and third links are conforming and the coupler link
is partially conforming.

Figure 3.6: A prototype demonstrating the anchoring function. Feet on the coupler link could
provide stability or prevent a cylinder from rolling.

developable surgical tools, used to perform a more detailed task. The “anchor” function can be
seen in Figure 3.6. The mechanism shown may prevent a cylinder from rolling or may provide
support for a cylindrical vehicle.

3.4 Adaptation to Surgical Devices in Shafts

Minimally Invasive Surgery is rising in popularity because it reduces trauma on the pa-
tient compared to open surgery, which leads to less pain, decreased recovery times, and improved
cosmesis [45, 46]. The tools involved in MIS are often cylindrical in shape, and the general trend
is toward smaller diameter devices to further reduce pain and trauma. As the scale of these devices
continues to shrink, they become more difficult to design and manufacture. Some devices have
parts in the tenth of a millimeter scale. Parts in the meso-scale range are too large to be created via
micro-manufacturing processes and too small for conventional machining, like milling and CNC
machines. Therefore, a low part count and low design complexity are important to maintain feasi-

30
bility in manufacturing. Current MIS shafts often allow only one tool to operate at the distal end,
especially when the shaft diameter is small.
Developable mechanisms can address some of these issues and therefore positively impact
the cost, speed, and recovery times associated with MIS. These mechanisms can create relatively
simple devices with low part counts, yet provide a net increase in functionality to existing tools.
The increase in functionality occurs because of their conforming nature: developable mechanisms
provide the opportunity to combine devices (and therefore functions) into a single MIS tool with
only a small increase in complexity. A developable instrument could be included within the walls
of a cylindrical shaft and enter a workspace through a single entrance in combination with another
instrument on the end of the shaft. This combination has the potential to:

1. Lower the time required to perform a task in a confined/remote workspace by reducing the
number of tooling changes required.
2. Reduce the trauma/damage to the boundary of the workspace by reducing the number of
entrance holes/points required.
3. Reduce the complexity of the control system used in conjunction with the tooling setup,
because fewer shafts would be required to enter the space.
4. Reduce the cost of the procedure.

For example, an existing MIS tool could operate at the end of a shaft in conjunction with
a developable mechanism tool that actuates outward from the cylinder (extramobile mechanism),
or intermittently with a developable tool that actuates inside the cylinder (intramobile mechanism)
[47].

3.4.1 Modeling

Each of the three developable mechanism property categories (behaviors, characteristics,


and functions) can be associated with specific design processes.

1. Behaviors - Energy and Motion Modeling


2. Characteristics - Geometry and Manufacturing Options
3. Functions - Link Shape

31
Various modeling techniques exist that enable creation of the behaviors required to achieve
a certain design or function. If bistable behavior is desired, an energy analysis [48] or an estab-
lished bistable design process [49] could be utilized to create a bistable compliant mechanism.
If a specific motion behavior is desired, such as translation or cylindrical motion, then kinematic
position, motion, or path generation modeling could be used to create it.
The characteristics category aids in the construction and geometric design of the mod-
eled mechanism. During analytic modeling, the first three characteristics must be complied with
to ensure that the finished mechanism will be developable when it is constructed. The optional
characteristics, monolithic design and sheet material construction, inform the designer of different
geometric design options available. For example, if a specific four-bar has been specified using
path generation, the characteristics section informs us that the device can be made monolithic or
be constructed from sheet material. If proper compliant mechanism design is used, the adaption to
a monolithic structure or sheet material construction will not alter the path generation analysis.
As discussed in [30], lamina emergent mechanisms must be change point mechanisms to
lie in a plane at some point in their motion. However, this restriction does not apply to devel-
opable mechanisms initially constructed from sheet material. As the sheet is rolled or shaped into
a developable surface, the distance between links changes, potentially changing the mechanism
classification. Therefore, sheet material can be used to create any developable mechanism without
the change point mechanism restriction.
The function category of developable mechanism properties is where the shape of the links
can be altered to accomplish the specific task desired. Because link shape in a kinematic mecha-
nism is independent of the mechanism’s motion, we can alter the shape of the links in a developable
mechanism throughout the design process. For example, extending a link beyond the connecting
joint and thinning it down to a point can enable that link to act as a needle to perform the function
of “injecting.”
The chart in Figure 3.7 shows an example of using these properties to assist in the creation
of a desired device. An engineer may wish to create a device from a sheet material that can wrap
or retro-fit around an existing cylinder and provide anchoring or a hoist point. Most of the charac-
teristics presented are inherent to cylindrical developable mechanisms, but this design relies on the
concealing and shape preservation characteristics. The monolithic and sheet material characteris-

32
Required Main Desired
Behaviors Characteristics Function
1) Predictable Motion - 1) Conceal to/be made from A flexible sheet with an
relative motion the surface they exist on/in embedded mechanism, when
2) Energy Storage - 3) Preserve shape of base wrapped around a tube, can
bistability developable surface deploy stablizing legs.
3) Emergent Motion throughout motion Key features:
6) Can be monolithic Wrap/Retrofit
7) Can originate from sheet Sheet
material
Stabilize (link shape)

Figure 3.7: A chart showing the functions, characteristics, and behaviors needed to complete the
design of an example device.

tics would also be beneficial in this application to simplify the design, increase packing density,
and reduce costs. A monolithic design would also allow for compliant joints that can enable a
bistable behavior so the device can be stable in the fully deployed and fully conformed state. The
relative motion and emergent motion behaviors are also required so the mechanism can actuate off
the cylinder wall and reach a certain position with respect to the cylinder.

3.5 Examples

During the course of this research, multiple minimally invasive surgical devices were de-
veloped. We will explore two exemplary prototypes and discuss the path through the design guide
that can be traced to create these devices.

3.5.1 Internal Scissors

The first design presented is an internal scissor mechanism, seen in Figure 3.8. A triple
rocker four-bar mechanism is the basis for this device, with link ratios of 1:0.7368:1.289:0.5263,
where the first link is ground. The curved blades of the scissors are shaped from the links, with the
shortest link having a large extension beyond the hinge to create one of the blades (see Fig. 3.9.)
A kinematic model of this device is shown in Figure 3.9. Note that the ground link from
the kinematic model is not physically present in the prototypes. The scissor mechanism is actuated
by rotating an internal cylinder (blue) with respect to an outer cylinder (dashed line, or black

33
Figure 3.8: Rendering of an Internal Scissors design embedded in a shaft.

cylinder in the prototype). Rotating the inner blue cylinder is equivalent to actuating a virtual link
that is pinned at the center of the concentric cylinders. The ground link is therefore a virtual link
connecting this center-pinned link and the orange link.
An example of where this device could be used is in a biopsy procedure. This tool, with
a forceps tool inserted in the shaft, could enter into a body cavity, grip a piece of tissue, retract it
partially into the tube, and then actuate the developable tool to cut off and contain a biopsy sample
of the tissue. Figure 3.10 illustrates this configuration.
During the design process, multiple developable mechanism properties from the design
guide were used to create the internal scissors mechanism. For the initial concept to be properly
designed, the relative motion and convergent motion behaviors need to be incorporated. Relative
motion between two adjacent links, afterwards shaped into blades, would allow scissor-like motion.
Convergent motion, or the ability for the mechanism to form back to the surface after actuation,
would allow the center of the cylindrical shaft to remain completely open after the scissors were
used. If the mechanism conformed to the rules of a developable mechanism, namely Characteristic
#1, then the scissors would conform to the cylinder walls and allow another device to pass through
the center. After modeling these behaviors and characteristics, shaping the links 2 and 3 of the
mechanism into blades enables the cutting function initially desired.

3.5.2 Multiplying Cylinder

The second design presented here is called the “multiplying cylinder” (Figure 3.11). The
mechanism is a triple rocker with link ratios 1:0.7368:1.289:0.5263, where the first link is ground,
or the cylinder wall. The design shown in Fig. 3.12 has links shaped to hold LEDs. In the latest

34
Figure 3.9: Kinematic skeleton diagram model of the scissors mechanism, the developable mech-
anism version, and a 3D-printed prototype with altered link shapes. Note: The ground link is not
physically present in the prototype, and blue link is the inner cylinder, actuated by rotating with
respect to the outer cylinder.

Figure 3.10: A 3D-printed internal scissors prototype with a forceps tool inserted.

35
Figure 3.11: Kinematic skeleton diagram model of the multiplying cylinder mechanism, the devel-
opable mechanism version, and a 3D-printed prototype with altered link shapes.

prototypes, bistability was achieved using the design technique presented by Jensen and Howell
[49], resulting in the inclusion of a compliant joint between links 2 and 3.
A kinematic model of a single multiplying cylinder mechanism is shown in Figure 3.11.
The orange link spans 180 degrees of the cylinder to potentially allow for different cylindrical,
disk-shaped, or spherical tools to be affixed to or snapped into this link and be deployed outside
the cylindrical shaft as the mechanism actuates. Again, only two of the four modeled links are
physically present in the prototype, because the curved slider and the ground link are virtually
pinned at the center of the cylinder.
The potential applications of this device are numerous. The ability to conceal and deploy
multiple lines or tools from the same shaft opens new possibilities for simplifying surgical proce-
dures. As shown in Figure 3.12, placing two of these mechanisms on a shaft can allow for two light

36
Figure 3.12: Multiplying Cylinders mechanism with ISI forceps inserted. LED lights included to
demonstrate potential application as a dual-source flashlight or a vehicle for stereoscopic vision
cameras.

Figure 3.13: Rendering of multiplying cylinder superimposed on open source image from MIS
appendectomy procedure.

or imaging devices to accompany a conventional MIS tool acting through the cylindrical shaft. The
use of two cameras could enable stereoscopic vision with a larger depth of field than conventional
3D laparoscopic cameras due to the additional space between lenses. A light, camera, and for-
ceps could be placed on one shaft, allowing the appendectomy procedure rendered in Figure 3.13
to be performed with two tools instead of three, while moving the image and light source to a
point-of-view position on the forceps.
The multiplying cylinder development process also involved drawing several properties
from the developable mechanism design guide. The initial idea behind the concept was to create

37
Figure 3.14: Secondary channel created when an elongated version of the multiplying cylinder
mechanism is opened.

secondary working channels, so the emergent motion behavior would need to be maximized to
create the largest possible working channel. It was also initially desired for the mechanism to be
bistable, with stable equilibrium positions at the deployed and stowed positions. This required the
energy storage behavior, resulting in the included compliant member.
A monolithic design would be desirable for a multiplying cylinder placed at the end of
a working shaft. This would aid in making the device disposable or interchangeable if it holds or
channels one-time-use tools. The amount of functions available with the multiplying cylinder high-
lights its adaptability. While this is not a comprehensive list, this device was designed to perform
expanding, observing (with the use of embedded cameras), snapping (bistable), guiding, protect-
ing, and duplicating or multiplying functions. The lengthened multiplying shafts in Figure 3.14
could be used to perform the multiplying, guiding, and protecting functions.

3.6 Conclusion

Developable Mechanisms, or mechanisms fit to developable surfaces, are useful because


they increase the functionality of already existing surfaces by embedding deployable, concealable
mechanisms on or beneath the surface. Here we have outlined these mechanisms’ behaviors, char-
acteristics, and potential functions, which can assist engineers in adapting developable mechanisms
into their design work. Two cylindrical developable mechanisms, aimed at providing functionality

38
in minimally invasive surgical procedures, have been discussed. These include the internal scissors
mechanism, which could cut or grip objects inside a minimally invasive tube, and the multiplying
cylinders mechanism. The multiplying cylinders was demonstrated in two configurations, one that
deploys two additional stationary tools from one minimally invasive shaft, and a second that turns
one cylindrical shaft into two functional working channels.

39
CHAPTER 4. HINGES AND CURVED LAMINA EMERGENT TORSIONAL JOINTS
IN CYLINDRICAL DEVELOPABLE MECHANISMS

4.1 Introduction

The objective of this work was to demonstrate practical applications of novel mechanisms
that collapse and expand and research techniques for creating hinge motion in these mechanisms.
This paper presents various hinge options available for small-scale cylindrical developable mecha-
nisms, including a study validating the assumption that lamina emergent torsional (LET) joints in
thin-walled cylinders can be modeled using the planar equations developed previously [50].
The work presented focuses on devices and hinges that can enable motion on cylinders
and mechanisms in scales commonly seen in the medical field, specifically in laparoscopic surgi-
cal tools. These tools virtually all exist in a cylindrical form, and are being scaled down to reach
smaller and more restrictive spaces in the body [51, 52]. Therefore, the simplistic design and con-
forming nature of cylindrical developable mechanisms (DMs) show large potential for impact in
industries such as the medical field where small-scale tooling is used. At the laparoscopic surgi-
cal scale, where cylindrical shafts can be as small as 3mm in diameter, design becomes difficult
due to the nature of manufacturing, part assembly, low gripping forces, and relatively high friction
forces [53, 54]. These difficulties can be addressed through compliant designs, which can have
much lower part counts and are more scalable than traditional pin and link mechanisms. The traits
of DMs could facilitate small-scale, compliant, and potentially monolithic cylindrical mechanisms.
The background given in Chapter 3 is suitable to prepare the reader for the following information.

4.2 Joint Requirements

In many kinematic models, it is assumed that a revolute joint:

1. Has no friction
2. Has no backlash

40
3. Enables one rotational degree of freedom
4. Maintains a constant center of rotation
5. Does not store energy.

It is also assumed in kinematic models that rigid link lengths do not change throughout
actuation, or the distances between joint axes of rotation do not change. While the properties of an
ideal joint depend on the application, friction and backlash are generally undesirable, and a single
degree of freedom per joint simplifies actuation of the mechanism. Energy storage can be desirable
in a mechanism, especially when attempting to create bistability through compliance [55, 56]. For
the purposes of this research, the ideal hinge for a cylindrical DM has no friction, no backlash, a
single degree of freedom in rotation, and a constant center of rotation.

4.3 Evaluation of Various Hinge Candidates

Previous work on cylindrical DMs [47, 57] incorporated traditional pin joints during the
prototyping phase. These classical methods of creating revolute joints will first be briefly discussed
and their limitations given. Additional hinge and modeling options will then be discussed, includ-
ing small-length flexural pivots (SLFP), initially curved cantilever beams, the membrane fold tech-
nique, and curved LET joints. Section 4.4 will then show a detailed analysis of thin-walled curved
LET joints, including the adaptation of planar LET joint models to curved LET joints.

4.3.1 Classical Pin Joints

Pin hinges have been in use for many years due to their simplicity and ease of manufactur-
ing. They allow infinite rotation, disassembly of parts, and different material options between the
pin and the links. However, the bearing surfaces often require lubrication, friction is inevitable, and
there are issues when scaling to small sizes. Generally, the diameter of the pin needs to be less than
the thickness of the link material. Therefore at small scales, such as those of laparoscopic surgi-
cal devices, direct shearing of the pin becomes a major concern. For example, many laparoscopic
devices are 5 mm in diameter [58] with wall thicknesses around 0.5 mm or less [59]. A pin joint
aligned with the ruling lines of the cylinder needs to have a diameter near 0.3 mm, or an area of
0.283 mm2 . At this dimension, a cylindrical pin made of 316 stainless steel would yield in direct

41
shear with a load of only 87 N, or 19.6 lbs. Friction in these joints is also relatively high compared
to the actuation force required, resulting in high energy losses between actuation and output force.
Other kinematic joints, such as prismatic, spherical, and cylindrical joints could also be
used to create cylindrical DMs, but they would share many of the same disadvantages as revolute
or pin joints.

4.3.2 Small-Length Flexural Pivots and Initially Curved Cantilever Beams

Small-Length flexural pivots (SLFP) are areas of reduced stiffness in an otherwise rigid
link that enable joint-like motion between rigid members [29]. If the reduced stiffness portion is
significantly shorter and more flexible than the longer, more rigid portion, then a characteristic
pivot (a location where a pin joint with a torsional spring can be modeled to take place) is located
in the middle of the shorter, less-rigid portion. This modeling technique is termed the pseudo-rigid-
body model (PRBM) and is diagrammed in Figure 4.1. With the PRBM, relatively simple equations
can be used to model the rotation and end coordinates of the rigid portion given an input moment or
forces. A closed-form elliptic integral solution for modeling initially curved small-length flexural
pivots was previously developed [60]. Modeling the flexible portion of the beam as an initially
curved cantilever beam, using the PRBM for that case [61], is a possible method that avoids using
elliptic integrals.
However, there are three conditions present in a cylindrical DM with initially curved beam
flexures that make modeling difficult or make the PRBM invalid for this case. First, when using the
PRBM to model an initially curved cantilever beam, the characteristic pivot lies near the first 15%
of the beam length, and is closer to the free end for higher curvature beams, as seen in Figure 4.1.
The pivot can also lie in space off the reference surface, or the developable surface used to create
the developable mechanism [47]. While this may enable new behaviors and features in DMs, it
can also make kinematic modeling more difficult. If the characteristic pivot lies off the reference
surface, the actual link length of the mechanism is different than the links of a DM with revolute
joints that lie on the reference surface. Small changes in link lengths could alter the behavior and
classification of the mechanism.
Second, for initially curved beams that decrease their curvature during a mechanism’s mo-
tion, the PRBM is only valid until the flexure reaches zero curvature, or a straight orientation [61].

42
F

M
F
l/2
(EI)L >> (EI)l M γl
(1-γ)l

l L

Figure 4.1: Left: The pseudo-rigid-body model (PRBM) demonstrated for a small-length flexural
pivot. Right: The PRBM applied on an initially-curved beam with a curved rigid link attached on
the end.

Figure 4.2: A 3D printed prototype of the expanding cylinder mechanism with one initially curved
flexure, highlighted in white. The path the right end of the flexure is forced through creates a
deflected shape that suggests end forces and moments, similar to a traditional fixed-guided beam.

Therefore, we must consider the deflected state of the mechanism and whether each initially curved
flexure deflects in a direction that increases or decreases its curvature. This limits the use of this
model to mechanisms that travel through relatively small deflections. Figure 4.2 shows the mech-
anism that will be discussed in Section 4.5 next to the same mechanism in the most deployed state
that can be analyzed with the PRBM.
The third difficulty arises when we consider the flexure’s end conditions and loads. In the
mechanism shown in Figure 4.2, the right end of the beam is guided through a non-conventional
path, resulting in an end condition consisting of a force vector with changing magnitude and di-

43
Figure 4.3: Top: The wall thickness reduction method (reduces h in beam) for creating SLFPs
or pre-curved flexures. Bottom: The width reduction method (reduces w in beam), which is only
viable when the wall thickness is less than the beam width.

rection and a variable moment load. To the authors’ knowledge, there is no existing PRBM for an
initially curved beam with end forces and moments. It can be seen in Figure 4.2 that the pre-curved
beam does not reach a flat state, which does occur with pre-curved beams with a pure moment on
the end, but a shape more resembling a beam under fixed-guided end conditions.
In regards to manufacturing, there are two simple methods for reducing stiffness to create
a SLFP or an initially curved cantilever beam to act as a joint - reduce either the width of the beam
or the cylinder wall thickness in select locations (Fig. 4.3). Reducing the width of the beam can be
done with many rotary manufacturing processes, such as a CNC machine or laser cutter. However,
this method is only beneficial in thin-walled cylinders where the wall thickness is less than the
width of the cut beam. It also does not aid in decreasing stress in the flexed member, as the depth
of the beam in the direction of bending is not reduced. Decreasing cylinder wall thickness has
opposing attributes - it is a more difficult manufacturing process, as small amounts of the cylinder
wall need to be removed from inside or outside the cylinder, but it does aid in decreasing stress on
the flexed member.

4.3.3 Membrane Technique

The membrane technique is one method for creating simple revolute motion and accom-
modating for thickness in origami designs [62–64], and for reducing parasitic motion in LET

44
Figure 4.4: Parallel guiding cylindrical developable mechanism with membrane joints. The cylin-
der is 38mm diameter polycarbonate tubing, cut on an Epilog M2 CO2 laser cutter, and the mem-
brane is red acrylic tape.

joints [65]. In this technique, a thin, flexible membrane is used to join the edges of rigid pan-
els to create a relative folding motion. If the membrane holds the panels together while leaving a
gap between panel edges, the folding and sandwiching of thick materials can be accomplished. The
membrane may also exist on both sides of the panels and be significantly thicker than the panels
themselves [66].
The membrane technique can be used to create hinge motion in cylindrical DMs, with some
consideration. With very thin-walled cylinders, the location of the membrane (whether is it attached
to the inside or outside of the cylinder walls) does little to affect the mechanism motion. However,
this location is important to consider when using cylinders with considerable wall thickness, as
the joints may then self-interfere and prevent the desired motion. Thickness accommodation tech-
niques from origami design could be further adapted to DMs, but this will not be undertaken in this
work. However, prototyping efforts have revealed that for initially extramobile links, or links that
move outward from the cylinder wall [47], the membrane should be applied to the outside of the
cylinder, and vice versa for intramobile links. A prototype cylindrical parallel guiding mechanism
with red membrane joints can be seen in Figure 4.4.
A second limitation or adaptation to consider is the shifting of the hinge position when
using a membrane adhered to the inside or outside of the cylinder wall. This effectively shifts the
joint locations to the cylinder wall, which can change the reference surface and the link lengths. In
the example shown in Figure 4.5, links 2-4 are held at the same lengths, but the ground link has to
change length if the joints are forced to the outside or inside of the cylinder wall. This may change

45
90˚ 104˚

Figure 4.5: Left: Skeleton diagram showing a parallel guiding mechanism aligned to a reference
surface (dotted line) in the middle of the cylinder walls. Right: The mechanism on the left, when
joints are pushed to the location of the membrane hinges, is forced to change its ground link length.

the mechanism’s classification and motion. In this example, the skeleton diagram of the linkage is
no longer a cyclic quadrilateral and the reference surface becomes a general cylinder.
Using the membrane technique on cylindrical DMs may enable small-scale devices, espe-
cially in the medical device industry. Some flexible endoscopes and catheters are finished with a
clear flexible coating, which coating could be utilized for the membrane technique with proper
application and selective cutting.

4.3.4 Lamina Emergent Torsional (LET) Joints

The LET joint is created by removing specific sections from a planar material to enable a
reduction of stiffness along a desired axis [50]. Many types of LET joints have been proposed and
analyzed [67–77]. These joints can allow for much greater deflections than simple single-beam
compliant flexures of similar thickness due to the combination of torsion and bending flexures in
series and parallel.
LET joints are promising for creating joint motion in cylindrical DMs because they allow
large deflections and can be created from a plain cylinder by removing material. This can enable
monolithic DMs. The adaptation of LET joints to curved surfaces was studied previously [78].
Equations were developed to determine the stiffness of a given curved LET joint and the stresses
seen during deflection. One example LET joint was analyzed analytically and compared to FEA
results, with good agreement, at up to 20° of deflection. However, the joints analyzed were made
from relatively thick walled material with large curvature.

46
Two of the disadvantages mentioned in regards to small-length flexural pivots and the
pseudo-rigid-body model in Section 4.3.2 also apply to LET joints. LET joints have a shifting
center of rotation off of the joint body, similarly resulting in the kinematic joint axes residing off
of the reference surface. This results in a combined loading condition similar to the initially curved
beams discussed above. However, the existing models assume either a pure moment load or rota-
tion displacement with no tensile or compressive forces. To complete the analysis desired in this
work, the assumption will again be made that the LET joints are being loaded with a pure moment.
Relatively thin-walled curved LET joints are analyzed here. It is expected that for a certain
range of thickness-to-radius ratios, the equations used to analyze planar LET joints in [40] will also
accurately predict joint stiffness in curved LET joints for a range of geometries. This assumption
would simplify the analysis of curved LET joints in relatively thin-walled cylindrical DMs.

4.4 Modeling Curved LET Joints With a Planar Approximation

Zimmerman et. al [78] used continuum mechanics equations [79] to model the annular
torsion sections created when cutting LET joints into a curved surface. The equation used to deter-
mine stiffness of the annular torsion beam contains a summation term. When using parameters for
a LET joint in relatively thin-walled cylinders, such as given in Table 4.1, the summation reaches
infinity after a few iterations. This limitation motivated the adaptation of planar LET joint models
to curved joints.

4.4.1 Evaluating the Planar Approximation

To validate further analysis, the torsional constants of annular torsion beams at a specified
range of thickness-to-radius ratios (t/R) were compared to the same property from square torsion
beams with the same cross-sectional area. The annular cross section constants were obtained in
ANSYS Mechanical. All other values were computed analytically, with square cross section con-
stants being calculated using K = 2.25a4 [80], where a is half the length of one side of the square.
Only the torsional constants were compared because the majority of deflection in a LET
joint is due to twisting of the torsion beams. A visual comparison of the cross section with the
largest t/R analyzed to a square beam of equal area is shown in Figure 4.6, overlaid on a plot

47
Error in K for Annular vs Square X-Sec Beams

0.8

Absolute Value of Error (%)


0.6

0.4

0.2
Error values
Linear Fit

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
t/R
Figure 4.6: Plot showing relative error between torsional constants of annular cross section and
rectangular cross section beams. The finite mesh in cross sections input to ANSYS may account
for the non-smooth pattern from low to high t/R ratio.

k5 k6
k6
k5 k1 k2
k2
k1
k9 k9
k4 k3 k4
k3 k8
k7 k7 k8

Figure 4.7: Spring diagram for an inside LET joint.

showing the error between the square and annular cross section torsion constants. The low error
values demonstrated that further modeling may be fruitful.
The force deflection characteristic of the inside LET joint has been modeled by deriving
the equivalent spring stiffness, keq [40], such that:

T = keq θ (4.1)

48
Figure 4.8: Schematic describing the parameters used in the MATLAB analysis and elsewhere in
this chapter.

where T is the total torque applied to the joint and θ is the joint angular deflection. Considering
the analogous spring system depicted in Figure 4.7, keq may be expressed as:

1 1 1 1
= k1 k5 k2 k6
+ + k3 k7
(4.2)
keq
k1 +k5 + k2 +k6
k9
k3 +k7 + kk44+k
k8
8

where k1 , k2 , k3 and k4 represent the stiffness of the segments in torsion (i.e. green members in
Figure 4.8) , and k5 , k6 , k7 , k8 and k9 are related to the segments in bending (i.e. blue members in
Figure 4.8). Each stiffness constant can be obtained by means of the following relations:

Ci G
ki = , i = 1, .., 4 (4.3)
Li
EIi
ki = , i = 5, .., 9 (4.4)
Li

E
where E is the modulus of elasticity, G = 2(1+ν) is the shear modulus, Li is the member’s length
(i.e. Lt for segments in torsion, Lb for segments in bending and Lc for the central member, as
shown in Figure 4.8). Ci , a parameter analogous to the polar second moment of area for circular
cross-sections, and Ii , the cross section’s second moment of area, are defined as:

49
Table 4.1: LET parameters tested in the framework.

Parameter Value
R 5mm
t [0.05, 0.25]mm
wt 0.3mm
wb 0.3mm
wc 0.8mm
Lt 1.25mm
Lb 0.5mm
Lc 0.5mm

t4
  
3 1 64 t
Ci = wt t 1− , i = 1, .., 4 (4.5)
3 π 5 wt 12wt4
wbt 3
Ii = , i = 5, .., 8 (4.6)
12
wct 3
Ii = , i=9 (4.7)
12

where wt , wb and wc are the cross section’s width and t is the LET thickness.
A fast numerical routine aimed at validating the proposed theoretical models on the curved
inside LET joint has been set up in a Matlab environment. Thanks to the ANSYS APDL interfacing
capabilities, an integrated design framework in which Matlab manages the parametric study, sim-
ulations, and the data exchange operations, has been implemented, as shown in Figure 4.9. In this
routine, ANSYS APDL provides the torque-deflection characteristic of each candidate via batch
mode. The framework outputs the comparisons between theoretical and numerical results. Given
wt , wb , wc Lt , Lb , Lc , the behaviour of the inside LET joint is investigated for several t/R values.
Numerical values of each geometric parameter are presented in Table 4.1. To perform the study,
the following operations are addressed through a sequence of Matlab functions:

1. Creation of the design space and definition of N candidates.

2. FEA batch simulations.

For i = 1 to N

50
Figure 4.9: Schematic of the software framework

• Update the ANSYS Design parameter file (Input par.txt in Figure 4.9) with the i-th
value (t = t(i));
• Batch ANSYS Execution;
• Import of the i-th results set (Results.txt in Figure 4.9);
End
3. Perform the theoretical calculations.
4. FEA data fitting and results comparison.

Regarding the FE models, a mapped mesh of Shell 181 elements has been defined for all the
candidates. For boundary conditions, the LET joint is fully constrained at one extremity and guided
in a pure rotation (35◦ ) at the other. The employed material is Aluminum alloy 7075 (heat treated).
The Youngs modulus and Poissons ratio adopted are E = 71.0GPa and ν = 0.334, respectively.
The achieved results, in terms of stiffness (i.e. keq ), are reported in Figure 4.10 for different
values of t/R. From a direct comparison, the results obtained via FEA simulations show a good
agreement with the behavior predicted by the theoretical formulations. As a consequence, planar
equations can be used to design with acceptable accuracy curved LET joints for limited values of
t/R.

4.4.2 Comparison of Planar LET Joint Adaptation Prediction with Physical Prototype

To validate the analytical and FEA work performed, a small study was performed on the
stiffness of 3D printed curved LET joints. To obtain an accurate bending modulus of elasticity of

51
Curved-LET Joint Stiffness
14

12

Stiffness (Nmm/rad)
10

FEA
2 Fitted Curve
Planar Equations
0
0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
t/R
Figure 4.10: Plot showing a comparison between curved LET joint stiffness in FEA and planar
assumption models.

Figure 4.11: Experimental test setup. Torque gauge is connected to white wire, laser encoder is
connected to colored wires.

the printed material, sample flexion beams were printed on a Prusa i3 Mk2 in PETG with 100%
infill and according to ASTM D790-03. These were tested in a 3-point bending setup using an
Instron 3345 testing system. 3D printed PETG curved LET joints, with varying t/r ratios, were then
deflected using a guided rotation. The deflection path was obtained by exporting node locations

52
from a LET joint in ANSYS FEA with a pure moment applied. However, this path is not circular
due to the joint’s shifting axis of rotation. Due to the difficulty of applying a pure moment on a
compliant joint with its axis shift, the path was approximated as circular for the medium angle of
rotation being tested.
The node locations from FEA were plotted in CAD. Under the assumption that the path
was circular, a fixed point of rotation was found and a mounting system (Figure 4.12) was de-
signed to guide the LET joint through this approximated path. The deflection angle and torque
applied was measured using a US Digital E2-5000-375 laser encoder (resolution of +/-0.07°) and
the Omega TQ103-50 torque gauge, respectively. The test setup is shown in Figure 4.11. Due to the
slow manual actuation of the LET joint and the unidirectional testing to obtain the stiffness of the
PETG LET joints, hysteresis in the torque gauge was not seen in the test results. These results are
plotted in Figure 4.13, where the 3D printed joint stiffness is compared to the stiffness obtained by
both planar LET joint equations and the FEA model. The physical data agrees well with the FEA
results, despite the difficulties of maintaining consistent material properties with fused deposition
3D printing.

4.4.3 Discussion on LET Joints

LET joints show promise for application in cylindrical DMs due to their ability to enable
relatively large hinge motion using only the material in the cylinder itself. The analysis performed
validated the use of planar LET joint equations for LET joints in thin-walled cylinders, which could
improve efficiency on curved LET joint research. While LET joints enable larger deflections than
other compliant joints, a quick analysis using the MATLAB routine in Section 4.4 showed that a
single LET joint is incapable of producing the large deflections required in some cylindrical DMs
when cut into in a metal (steel, aluminum, or titanium) thin-walled cylinder. It is also important to
reiterate that the models analyzed in this work assume a pure moment or rotation is applied to the
LET joint, yet compliant joints in kinematic mechanisms are subject to combined loads. Further
research could focus on the forces and moments applied to LET joints in kinematic mechanisms
and the effects of the shifting center of rotation.

53
FEA node path
Approximated
path
Control arm

Control arm
line of rotation

Figure 4.12: Schematic showing the deflection path of the end of the LET joint arm (in red) and
the corresponding estimated circular axis of rotation (dashed line) used to design the experimental
setup.

PETG Curved-LET Samples in Pure Rotation


150

125
Stiffness (Nmm/rad)

100

75

50

FEA
25 Fitted Curve
Planar Equations
Experimental

0.03 0.04 0.05 0.06 0.07 0.08 0.09


t/R
Figure 4.13: Comparison of experimentally derived 3D printed curved LET joint stiffness to FEA
and planar LET models.

54
4.5 At-Scale Laparoscopic Surgical Device in Nitinol

A cylindrical DM was discussed [57] and designed for use as an end effector in minimally
invasive or laparoscopic surgery. End effectors are specialized tools at the end of laparoscopic
shafts designed to manipulate body tissue in a specific way. It was desired to analyze the same
device, termed the multiplying cylinder, with LET joints at a scale comparable to laparoscopic
surgical devices on the market. This would mean embedding LET joints into a cylinder that is 5
mm in diameter with a wall thickness of less than 0.5 mm. Using LET joints to achieve the large
deflections required in the multiplying cylinder device at this scale may require the use of exotic
materials. However, the routine discussed in Section 4.4.1 is valid for linear deflections only.
To demonstrate that LET joints can be used at this scale, the inside LET joint was modeled
in nitinol in ANSYS APDL. In the multiplying cylinder device, an outer cylinder with two joints
fits over an inner cylinder that attaches to the end of one of the links. The inner cylinder is then
rotated with respect to the outer to actuate the device. Between the stowed and deployed state, one
of the joints in this device travels through 90° of deflection. Preliminary analysis using the routine
from Section 4.4 showed that a single LET joint, in conventional metals and at the desired scale,
could not accomplish this deflection. It was therefore decided to model the device with superelastic
nitinol. Nitinol has a unique, non-linear stress-strain curve. Therefore, the at-scale model could not
be analyzed in Nitinol using the linear analysis method from Section 4.4.
To accomplish this task, a parametric CAD model of the Expanding Cylinder was created.
33 different Expanding Cylinder models were analyzed in ANSYS FEA with a multi-linear stress-
strain curve to match the behavior of superelastic nitinol. Solid187 elements were used with a
tetrahedral mesh that was refined at each LET joint. The rectangular sections were modeled as
perfectly rigid and used to replicate the cylinder-in-cylinder actuation of the original device. Each
device was actuated to full deployment, which involves rotating the ”inner cylinder” or rigid beams
180°, resulting in 90° of deflection in one of the LET joints. The model shown in Figure 4.14
performed best, with a maximum VonMises stress of 449 MPa and maximum strain of 0.0483.
While exact material properties for superelastic nitinol are difficult to obtain, the strain and stress
values are unlikely to cause yielding [81]. This model validates that LET joints cut into a nitinol
cylinder could enable cylindrical DMs to exist in the laparoscopic surgery market. Future work

55
First Load Step Intermediate Load Step

Undeformed
Fully constrained model
Fully actuated
model Guided
rotation,
180o in
Z-axis

Von-Mises
Stress (MPa)
449
399
y 349
299
249
200
x 150
99.8
z 50.0
.100

Figure 4.14: FEA results of the Expanding Cylinder mechanisms with LET joints in nitinol. The
rigid beams, pinned to the cylinder, were rotated about the center of the cylinder to mimic the
circular displacement of the rotating inner cylinder actuator.

should be done to integrate unique stress strain behaviors into analytical routines, such as the one
presented here.

4.6 Conclusion

In this work, possible options for hinges in cylindrical developable mechanisms were re-
viewed and analyzed. Classical methods such as pin hinges are not viable due to manufacturing and

56
loading concerns when reaching small scale parts. Simple compliant joints, such as small-length
flexural pivots, can be useful for large curvature joints but suffer from modeling issues due to the
combined loadings present in compliant joints in kinematic mechanisms. Membrane joints show
promise for small-scale mechanisms due to their simplicity, but the joint axis must shift to the out-
side or inside of the cylinder wall, which changes the initial or closed position of the mechanism
if all link lengths are held the same.
Lamina emergent torsional joints may be advantageous in cylindrical DMs due to the large
deflections often reached. A Matlab and ANSYS FEA integrated routine was developed to compare
planar LET joint stiffness models developed previously to finite element curved LET joint models.
For curved LET joints cut from thin-walled cylinders with a thickness to radius ratio of 0.032 to
0.08, there is very good agreement between the planar models and the curved FEA models. This
suggests the previously developed linear planar models are sufficient for curved LET joints in this
t/R range. The two analytical models were also compared to 3D printed curved LET joints, and
these results also showed good agreement. Finally, a cylinder-in-cylinder actuated cylindrical DM
was modeled in ANSYS FEA in nitinol, with dimensions similar to those seen in laparoscopic
surgical tools. This model showed that LET joints may be suitable for medical device applications
where exotic materials such as nitinol are more commonly found.

57
CHAPTER 5. CONCLUSION

In Chapter 2 it was reinforced that the customizable deployed shapes, unique motions, and
large deployment ratios of origami can be transferred to interesting fields, such as a deployable bal-
listic barrier, with the proper application of adaptation and thickness accommodation techniques.
A modified membrane fold technique was created that enables the use of membranes thicker than
the rigid facets, in contrast to the thick panel and thin membrane techniques studied previously.
Chapter 3 and 4 presented research related to cylindrical developable mechanisms. In Chap-
ter 3 the behaviors, characteristics, and properties of cylindrical DMs were outlined to assist de-
signers in using cylindrical DMs to accomplish novel tasks. Two cylindrical DM devices were
presented, both of which aim at increasing the functionality of laparoscopic surgical devices while
retaining the same dimensions. These devices included the internal scissor mechanism, which al-
lows a snipping or gripping operation in the center of the cylindrical shaft, and a multiplying
cylinder mechanism, where a secondary cylinder or channel emerges from the wall of the original
shaft.
Chapter 4 reviewed existing hinge options adaptable to cylindrical DMs and discussed their
strengths, weaknesses, and novel behaviors they may allow. LET joints on cylinders were discussed
in detail, and a planar modeling assumption was validated both analytically and empirically. This
assumption allows LET joints cut into thin-walled cylinders to be modeled with the simple planar
LET joint equations developed previously. The potential of LET joints in cylinders with exotic
materials was demonstrated by modeling a laparoscopic-scale multiplying cylinder device with
nitinol LET joints in ANSYS APDL.
In conclusion of this work, proper joint design is key to realizing origami and developable
mechanism designs in real-world engineering solutions. With traditional engineering designs, fail-
ures often occur at part interfaces. In origami and developable mechanism design, the part or link
interfaces are often the portion of the design preventing them from becoming a practical design.

58
While these mechanisms are still far from wide-spread real world application, this work has shown
advancements to the field through three different design stages. It was shown that LET joints in
thin-walled cylinders can be approximated as planar for a range of thickness-to-radius ratios, which
advanced the modeling of joints in cylindrical developable mechanisms. The laparoscopic device
prototypes shown in Chapter 3 and the feedback received from various laparoscopic and general
surgeons revealed that these mechanisms have promising applications in the medical field. Close
to the finished end of the design spectrum lies the origami-inspired ballistic barrier, which taught
that existing design techniques can be successfully modified to allow origami to exist in unique
applications. The final design of the barrier, Version 3, also taught that sometimes revising the
basics of a design, or altering the base mechanism or machine chosen for a task, can have highly
favorable outcomes. An engineer or designer should never assume that a portion of the design is
perfect because it has been completed - sometimes going back to the drawing board is better than
trying to fix a fundamentally flawed design.
From this work, a significant number of recommendations for future research were formed.
The origami ballistic barrier suffered from a high weight-to-stiffness ratio, requiring multiple
pieces of support and actuation hardware to be used. Research into various deployment and stiff-
ening methods specific to dense origami-based structures would aid in the design of full-scale
origami mechanisms.
Several recommendations for future work emerged during the work presented in Chapter 4.
Two recommendations stem from Section 4.3.2. First, a study should be done on how the behavior
of developable mechanisms with compliant joints changes when the pseudo-link does not lie on
the reference surface or within the joint material. Second, analyzing the path that various com-
pliant joints in developable mechanisms are driven through would aid in designing these joints.
In Section 4.3.3, thickness accommodation techniques from origami design were briefly used to
create a developable mechanism with membrane hinges. The adaptation of other thickness accom-
modation techniques to DMs would be a beneficial topic to research. The last recommendation is
regarding the analytical routine developed in Section 4.4. The routine currently relies on the as-
sumption of linear material behavior. Expanding this routine to handle non-linear and multi-linear
analysis would allow the study of yielded materials and more exotic materials such as nitinol in
their application to curved LET joints.

59
Due to the recent creation of developable mechanisms, this work is one of only a few
studies into their properties and applications. A large amount of research into DM behaviors, ap-
plications, and manufacturing methods could still be performed. This small amount of preliminary
work has demonstrated the potential for application to the medical field, and many more industries
and applications have yet to be discovered. As of the time of this work, it appears the largest barrier
to cylindrical DMs reaching industrial applications is manufacturing methods. It is therefore rec-
ommended that time be spent researching LET joints and membrane joints in DMs, as they appear
the most promising for enabling the desired motions on medium and small scales.

60
REFERENCES

[1] Peraza-Hernandez, E. A., Hartl, D. J., Malak, R. J., and Lagoudas, D. C., 2014. “Origami-
inspired active structures: a synthesis and review.” Smart Materials and Structures, 23(9).
[2] Rogers, J., Huang, Y., Schmidt, O. G., and Gracias, D. H., 2016. “Origami mems and nems.”
Materials Research Society, 41(2).
[3] Johnson, M., Chen, Y., Hovet, S., Xu, S., Wood, B., Ren, H., Tokuda, J., and Ho-Tse, Z. T.,
2017. “Fabricating biomedical origami: a state-of-the-art review.” International Journal of
Computer Assisted Radiology and Surgery, 12, pp. 2023–2032.
[4] Morris, E., McAdams, D. A., and Malak, R., 2016. “The state of the art of origami-inspired
products: A review.” In Proceedings of the ASME 2016 International Design Engineering
Technical Conferences, ASME.
[5] Maanasa, V., and Sri, R. L. R., 2014. “Origami-innovative structural forms and applications
in disaster management.” International Journal of Current Engineering and Technology, 4(5).
[6] Miura, K., 1991. “A note on intrinsic geometry of origami.” In 1st International Conference
of Origami Science and Technology, pp. 239–249.
[7] Schenk, M., and Guest, S. D., 2013. “Geometry of miura-folded metamaterials.” In Proceed-
ings of the National Academy of Sciences, pp. 3276–3281.
[8] Evans, T., Lang, R., Magleby, S., and Howell, L., 2015. “Rigidly foldable origami twists.” In
Origami 6, CRC Press, pp. 119–130.
[9] Yoshimura, Y., 1955. “On the mechanism of buckling of a circular cylindrical shell under
axial compression.” National Advisory Committee for Aeronautics.
[10] Stavric, M., and Wiltsche, A., 2014. “Quadrilateral patterns for rigid folding structures.”
International Journal of Architectural Computing, 12(1).
[11] Huffman, D. A., 1976. “Curvature and creases: A primer on paper.” IEEE Transactions on
Computers, C-25(10).
[12] Gattas, J., Wu, W., and You, Z., 2013. “Miura-base rigid origami: parametrizations of first-
level derivative and piecewise geometries.” Journal of Mechanical Design.
[13] J.M., 1676. Sports and pastimes, or, Sport for the city and pastime for the country with
a touch of hocus pocus. H.B. for John Clark, London seen in the Oxford Text Archive,
http://ota.ox.ac.uk/tcp/headers/A51/A51802.html.

61
[14] Francis, K., Rupert, L., Lang, R., Morgan, D., Magleby, S., and Howell, L., 2014. “From
crease pattern to product: Considerations to engineering origami-adapted designs.” In
Proceedings of the ASME 2014 International Design Engineering Technical Conferences,
ASME.
[15] Nelson, T., Lang, R., Pehrson, N., Magleby, S., and Howell, L., 2016. “Facilitating deployable
mechanisms and structures via developable lamina emergent arrays.” Journal of Mechanisms
and Robotics, 8(031006).
[16] Delimont, I., Magleby, S., and Howell, L., 2015. “A family of dual-segment compliant joints
suitable for use as surrogate folds.” Journal of Mechanical Design, 137(092302).
[17] Pehrson, N., Magleby, S., Lang, R., and Howell, L., 2016. “Introduction of monolithic
origami with thick-sheet materials.” In Spatial Structures in the 21st Century, IASS Annual
Symposium.
[18] Zirbel, S., Lang, R., Magleby, S., Thomson, M., Sigel, S., Walkmeyer, P., Trease, B., and
Howell, L., 2013. “Accommodating thickness in origami-based deployable arrays.” Journal
of Mechanical Design, 135(111005).
[19] Hagy, D., Morgan, J., Caplan, M., and Stoe, D., 2008. Ballistic Resistance of Body Armor
NIJ Standard-0101.06. National Institute of Justice.
[20] Lang, R., Tolman, K., Crampton, E., Magleby, S., and Howell, L., 2018. “Accommo-
dating thickness in origami-inspired engineered systems.” Applied Mechanics Reviews,
vol(number), p. pgs.
[21] Chen, Y., Peng, R., and You, Z., 2015. “Origami of thick panels.” Science, 349(6246),
pp. 396–400.
[22] Ku, J., and Demaine, E., 2016. “Folding flat crease patterns with thick materials.” Journal of
Mechanisms and Robotics, 8(3).
[23] Edmonson, B., Lang, R., Morgan, M., Magleby, S., and Howell, L., 2016. “Thick rigidly
foldable structures realized by an offset panel technique.” In Origami6, CRC Press, pp. 149–
161.
[24] Tachi, T., 2011. “Rigid-foldable thick origami.” In Origami 5, CRC Press, pp. 253–264.
[25] Pehrson, N., and Banik, J., 2018. “Folding approaches for tensioned precision planar shell
structures.” In Proceedings of the AIAA Spacecraft Structures Conference, American Institute
of Aeronautics and Astronautics.
[26] Tolman, K., 2017. “Developing hybrid thickness-accommodation techniques for new
origami-inspired engineered systems.” Master’s thesis, Brigham Young University.
[27] Brandon, T. E., 2016. Top Calibers Recovered and Traced in the United States and Territories.
Bureau of Alcohol, Tobacco, Firearms, and Explosives in the U.S. Department of Justice,
atf.gov/resource-center/firearms-trace-data-2016.

62
[28] Nelson, T., Zimmerman, T., Magleby, S., and Howell, L., 2019. “Developable mechanisms
on developable surfaces.” Science Robotics.
[29] Howell, L. L., and Midha, A., 1994. “A method for the design of compliant mechanisms with
small-length flexural pivots.” J. Mechanical Design, 116(1), pp. 280–290.
[30] Jacobsen, J. O., Winder, B. G., Howell, L. L., and Magleby, S. P., 2010. “Lamina emergent
mechanisms and their basic elements.” J. Mechanisms Robotics, 2(1).
[31] Aflattani, R., and Lusk, C., 2018. “A lamina-emergent frustrum using a bistable collapsible
compliant mechanism.” Journal of Mechanism Design, 140(12).
[32] Nelson, T. G., Lang, R. J., Pehrson, N. A., Magleby, S. P., and Howell, L. L., 2016. “Facilitat-
ing deployable mechanisms and structures via developable lamina emergent arrays.” Journal
of Mechanisms and Robotics, 8.
[33] Rabinovich, M., Hoffmann, T., and Sorkine-Hornung, O., 2018. “Discrete geodesic nets for
modeling developable surfaces.” ACM Transactions on Graphics, 37(2).
[34] Megaro, V., Zehnder, J., Bächer, M., Coros, S., Gross, M. H., and Thomaszewski, B., 2017.
“A computational design tool for compliant mechanisms.” ACM Transactions on Graphics,
36, pp. 82:1–82:12.
[35] Su, H., 2009. “A pseudorigid-body 3r model for determining large deflection of cantilever
beams subject to tip loads.” J. Mechanisms Robotics, 1(2).
[36] Pedersen, C., Buhl, T., and Sigmund, O., 2001. “Topology synthesis of large-displacement
compliant mechanisms.” International Journal for Numerical Methods in Engineering,
50(12), pp. 2683–2705.
[37] Sigmund, O., 1997. “On the design of compliant mechanisms using topology optimization.”
Mechanics of Structures and Machines, 25(4), pp. 493–524.
[38] Aten, Q. T., Jensen, B. D., Tamowski, S., Wilson, A. M., Howell, L. L., and Burnett, S. H.,
2012. “Nanoinjection: pronuclear dna delivery using a charged lance.” Transgenic Research,
21(6), pp. 1279–1290.
[39] Last, M., Subramaniam, V., and Pister, K., 2012. “Out-of-plane motion of assembled mi-
crostructures using a singlemask soi process.” In The 13th International Conference on
Solid-State Sensors, Actuators, and Microsystems, Vol. 2, IEEE.
[40] Jacobsen, J., Chen, G., Howell, L., and Magleby, S., 2012. “Lamina emergent torsional (let)
joint.” Mechanism and Machine Theory, 56, pp. 1–15.
[41] Xie, Z., Qiu, L., and Yang, D., 2017. “Design and analysis of outside-deployed lamina emer-
gent joint (od-lej).” Mechanism and Machine Theory, 114, pp. 111–124.
[42] Ushakov, V., 1999. “Developable surfaces in euclidean space.” J. of the Australian Mathe-
matical Society (Series A), 66(3), pp. 431–437.

63
[43] Cleave, J. P., 1980. “The form of the tangent-developable at points of zero torsion on
space curves.” Mathematical Proceedings of the Cambridge Philosophical Society, 88(3),
p. 403407.
[44] Joskowicz, L., 1990. “Simplification and abstraction of kinematic behavior.” Readings in
qualitative reasoning about physical systems, pp. 597–602.
[45] Jaffray, B., 2005. “Minimally invasive surgery.” Archives of disease in childhood, 90(5),
pp. 537–542.
[46] Robinson, T., and Stiegmann, G., 2004. “Minimally invasive surgery.” Endoscopy, 36(01),
pp. 48–51.
[47] Greenwood, J., Magleby, S. P., and Howell, L. L., 2019. An approach for designing devel-
opable mechanisms on regular cylindrical surfaces In review for IDETC2019.
[48] Hanna, B., Lund, J., Lang, R., Magleby, S. P., and Howell, L. L., 2014. “Waterbomb base:
a symmetric single-vertex bistable origami mechanism.” Smart Materials and Structures,
23(9).
[49] Jensen, B. D., and Howell, L. L., 2003. “Identification of compliant pseudo-rigid-body four-
link mechanism configurations resulting in bistableor behav.” Journal of Mechanical Design,
125(4), pp. 701–708.
[50] DeFigueiredo, B. P., Zimmerman, T. K., Russell, B. D., and Howell, L. L., 2018. “Regional
stiffness reduction using lamina emergent torsional joints for flexible printed circuit board
design.” Journal of Electronic Packaging, 140(4), May.
[51] Chang, J., Boules, M., Rodriguez, J., and Kroh, M., 2016. “Minilaparoscopy with inter-
changeable, full 5-mm end effectors: First human use of a new minimally invasive operating
platform.” J Laparoendosc Adv Surg Tech A, 26(1), pp. 1–5.
[52] Zoppi, M., Sieklicki, W., and Molfino, R., 2008. “Design of a microrobotic wrist for needle
laparoscopic surgery.” Journal of Mechanical Design, 130(10).
[53] J Kuchenbecker, K., Gewirtz, J., Mcmahan, W., Standish, D., J Mendoza, P., and Lee, D.,
2010. “Verrotouch: High-frequency acceleration feedback for telerobotic surgery.”.
[54] Gafford, J., Ding, Y., Harris, A., McKenna, T., Polygerinos, P., Holland, D., Walsh, C., and
Moser, A., 2015. “Shape deposition manufacturing of a soft, atraumatic, and deployable
surgical grasper.” Journal of Mechanims and Robotics, 7(2).
[55] Hwang, I.-H., Shim, Y.-S., and Lee, J.-H., 2003. “Modeling and experimental characteriza-
tion of the chevron-type bi-stable microactuator.” Journal of Micromechanics and Micro-
engineering, 13(6), aug, pp. 948–954.
[56] Opdahl, P. G., Jensen, B. D., and Howell, L. L., 1998. “An investigation into compliant
bistable mechanisms.” In DETC International Design Engineering Technical Conferences,
ASME.

64
[57] Seymour, K., 2019. Cylindrical developable mechanisms for minimally invasive surgical
instruments Unpublished, Accepted for DETC-CIE International Design Engineering Tech-
nical Conferences and Computers and Information in Engineering Conference, September.
[58] Ferreira, H., 2015. Equipment in Laparoscopic Surgery. 01, pp. 3–12.
[59] VitaNeedle, 2019. Vita needle master tubing gauge chart Company Website
https://www.vitaneedle.com/assets/files/ Vita Needle Master Tubing Gauge Chart.pdf.
[60] Midha, A., and Kuber, R., 2014. “Closed-form elliptic integral solution of initially-straight
and initially-curved small-length flexural pivots.” In ASME DETC-CIE International Design
Engineering Technical Conferences and Computers and Information in Engineering Confer-
ence, Vol. 5A.
[61] Howell, L. L., Midha, A., and Norton, T., 1996. “Evaluation of equivalent spring stiffness for
use in a pseudo-rigid-body model of large-deflection compliant mechanisms.” J. Mechanical
Design, 118(1), pp. 126–131.
[62] Lang, R. J., Tolman, K. A., Crampton, E. B., Magleby, S. P., and Howell, L. L., 2018. “A
review of thickness-accommodation techniques in origami-inspired engineering.” Applied
Mechanics Reviews, 70(1), Feb.
[63] Zirbel, S. A., Lang, R. J., Thomson, M. W., Sigel, D. A., Walkemeyer, P. E., Trease, B. P.,
Magleby, S. P., and Howell, L. L., 2013. “Accommodating thickness in origami-based de-
ployable arrays.” Journal of Mechanical Design, 135(11), Oct.
[64] Zirbel, S., Wilson, M. E., Magleby, S. P., and Howell, L. L., 2013. “An origami-inspired self-
deployable array.” In ASME SMASIS Conference on Smart Materials, Adaptive Structures
and Intelligent Systems, ASME.
[65] Chen, G., Magleby, S. P., and Howell, L. L., 2018. “Membrane-enhanced lamina emergent
torsional joints for surrogate folds.” Journal of Mechanical Design, 140(6), Mar.
[66] Seymour, K., Burrow, D., Avila, A., Bateman, T., Morgan, D. C., Magleby, S. P., and Howell,
L. L., 2018. “Origami-based deployable ballistic barrier.” In Proceedings of 7OSME, pp. 763–
777.
[67] Yanlin Li, Lifang Qiu, J. W., and Xie, Z., 2018. “Design and comprehensive performance
analysis of is-lej hinges.” In ASME DETC-CIE International Design Engineering Technical
Conferences and Computers and Information in Engineering Conference, Vol. 5A, ASME.
[68] Xie, Z., Qiu, L., and Yang, D., 2018. “Using the parts used to be removed to improve com-
pliant joints performance.” In ASME 2018 International Design Engineering Technical Con-
ferences and Computers and Information in Engineering Conference, American Society of
Mechanical Engineers, pp. V05AT07A006–V05AT07A006.
[69] Li, Y., Qiu, L., Wang, J., and Xie, Z., 2018. “Design and comprehensive performance analysis
of is-lej hinges.” In ASME 2018 International Design Engineering Technical Conferences and
Computers and Information in Engineering Conference, American Society of Mechanical
Engineers, pp. V05AT07A008–V05AT07A008.

65
[70] Klett, Y., 2018. “Paleo: Plastically annealed lamina emergent origami.” In ASME 2018
International Design Engineering Technical Conferences and Computers and Information in
Engineering Conference, American Society of Mechanical Engineers, pp. V05BT07A062–
V05BT07A062.
[71] Cao, Y., Wang, Q., Chen, G.-l., and Qin, Y.-l., 2015. “Structural synthesis of lems based on
planar kinematic chains.” In 2015 IEEE International Conference on Cyber Technology in
Automation, Control, and Intelligent Systems (CYBER), IEEE, pp. 618–623.
[72] Liu, K., Cao, Y., Ge, S.-y., and Ding, R., 2016. “Modeling and optimization of planar 2-
dof compliant rotational hinge.” In 2016 IEEE International Conference on Robotics and
Biomimetics (ROBIO), IEEE, pp. 2105–2110.
[73] Alfattani, R., and Lusk, C., 2016. “A lamina-emergent frustum using a bistable collapsible
compliant mechanism (bccm).” In ASME 2016 International Design Engineering Technical
Conferences and Computers and Information in Engineering Conference, American Society
of Mechanical Engineers, pp. V05AT07A014–V05AT07A014.
[74] Xie, Z., Qiu, L., and Yang, D., 2017. “Design and analysis of outside-deployed lamina emer-
gent joint (od-lej).” Mechanism and Machine Theory, 114, pp. 111–124.
[75] Xie, Z., Qiu, L., and Yang, D., 2018. “Design and analysis of a variable stiffness inside-
deployed lamina emergent joint.” Mechanism and Machine Theory, 120, pp. 166–177.
[76] Mutlu, R., Alici, G., and Li, W., 2015. “A soft mechatronic microstage mechanism based
on electroactive polymer actuators.” IEEE/ASME Transactions on Mechatronics, 21(3),
pp. 1467–1478.
[77] Mutlu, R., Alici, G., and Li, W., 2016. “Three-dimensional kinematic modeling of helix-
forming lamina-emergent soft smart actuators based on electroactive polymers.” IEEE Trans-
actions on Systems, Man, and Cybernetics: Systems, 47(9), pp. 2562–2573.
[78] Zimmerman, T., Butler, J., Frandsen, D., Burrow, D., Fullwood, D., Magleby, S., and How-
ell, L., 2018. “Modified material properties in curved panels through lamina emergent tor-
sional joints.” In 2018 ReMAR International Conference on Reconfigurable Mechanisms and
Robots, pp. 1–9.
[79] Jog, C., 2015. Continuum Mechanics, Foundations and Applications of Mechanics., 3rd ed.,
Vol. 1 Cambridge University Press.
[80] Roark, R. J., and Young, W. C., 1975. Formulas for stress and strain. McGraw-Hill.
[81] Qian, H., Li, H., Song, G., and Guo, W., 2013. “Recentering shape memory alloy passive
damper for structural vibration control.” Mathematical Problems in Engineering, 2013, 11,
pp. 1–13.

66

You might also like