Вы находитесь на странице: 1из 155

Folded Shell Structures

A thesis submitted for the degree of


Doctor of Philosophy
31 August 2011

Mark Schenk
Clare College
University of Cambridge

Supervisor: S.D. Guest

Declaration

I declare that, except for commonly understood and accepted ideas, or


where specific reference is made to the research of other authors, this
dissertation is the result of my own original work and includes nothing
which is the outcome of work done in collaboration. I further state
that this dissertation has not been previously or is currently being
submitted, either in part or as a whole, for any degree, diploma, or
other qualification at any other university. The thesis presented is 148
pages, containing approximately 32000 words and 91 figures, and does
thus not exceed the limit of length prescribed by the Degree Commitee
of the Faculty of Engineering.

Acknowledgements

I would like to extend a warm thank you to my supervisor Simon Guest


for his time, advice and insights throughout my PhD. A few further
words of thanks are due. To Keith Seffen for many interesting discussions on the mechanics of folded sheets, which helped crystallise several
ideas on unit cell kinematics. To Julian Allwood for his infectious enthusiasm when presented with the challenge of manufacturing a Miura
sheet from metal, which opened up a whole new direction of study.

ii

Abstract
A novel type of shell structure was analysed, folded shell structures. These shell
structures have a distinct structural hierarchy: globally they can be regarded as
thin-walled shells, but at a meso scale they consist of tessellated unit cells, which
in turn are composed of thin-walled shells joined at distinct fold lines. It is this
structural hierarchy that imbues the folded shell structures with their interesting mechanical properties. The global sheet deformations are a combination of
bending along the folds, and deformation of the interlying material. The former
is primarily a kinematic problem, with a parallel in the flexibility of hinged plate
structures. A review of the mathematics of rigid origami provides the necessary
background to develop non-trivial geometries of these folded shells that still exhibit
a soft deformation mode.
Two example folded shell structures are introduced, the Miura and Eggbox sheet.
Both consist of a tessellation of parallelogram facets; the first is developable, while
the other has points of positive and negative Gaussian curvature. The first property of interest is their increased in-plane flexibility, by virtue of the opening and
closing of folds. The Miura and Eggbox sheet respectively have an effective negative and positive in-plane Poissons ratio. Secondly, both sheets can modify their
global Gaussian curvature, with no stretching at the material level. Thirdly, both
sheets exhibit an oppositely signed Poissons ratio for in-plane and out-of-plane
deformations; e.g. when bending the Miura sheet it exhibits a negative Poissons
ratio behaviour and deforms anticlastically.
The salient global deformations of the sheets were analysed in terms of the kinematics of the constituent unit cells. The characteristic in-plane and out-of-plane
properties of the sheets followed directly from developable deformations of the
tessellated unit cells. A more holistic top-down numerical approach modelled the
sheets as an array of unit cells. The sheets were represented by a pin-jointed
bar framework, and additional planarity constraints between facets enabled the
inclusion of a bending stiffness for the facets and fold lines. A modal analysis of
the sheets stiffness matrix showed that the characteristic deformation modes are
among the dominant eigenmodes of the sheets for a wide range of geometries and
material properties.
Many folded shell structures can be folded from flat sheet material, with only
minimal material deformations. The manufacturing processes must overcome the
intrinsic kinematics of the sheets, whereby the sheet contracts in two directions
simultaneously. Existing methods were reviewed, and classified into synchronous
folding, gradual folding and pre-gathering techniques. A novel cold gas pressure
manufacturing method was introduced, and it was shown that a simple plastic
hinge model cannot yet fully account for the total required forming energy.

iii

Contents
1 Introduction
1.1

Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 Background and Concepts


2.1

2.2

2.3

Textured Shell Structures . . . . . . . . . . . . . . . . . . . . . . .

2.1.1

Texture Bending Stiffness . . . . . . . . . . . . . . . . . .

2.1.2

Texture Flexibility . . . . . . . . . . . . . . . . . . . . . .

12

2.1.3

Texture Impact Absorption . . . . . . . . . . . . . . . . .

15

2.1.4

Folded Shell Structures . . . . . . . . . . . . . . . . . . . .

15

Origami Folding and Foldability . . . . . . . . . . . . . . . . . . . .

16

2.2.1

Curvature and Creases . . . . . . . . . . . . . . . . . . . . .

16

2.2.2

Rigid Foldability . . . . . . . . . . . . . . . . . . . . . . . .

19

2.2.3

Fold Pattern Design . . . . . . . . . . . . . . . . . . . . . .

28

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

31

3 Folded Shell Structures


3.1

32

Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

32

3.1.1

Example Sheets . . . . . . . . . . . . . . . . . . . . . . . . .

32

3.2

Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . .

34

3.3

Structural Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

4 Kinematic Analysis
4.1

4.2

47

Planar Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . .

47

4.1.1

Miura Sheet . . . . . . . . . . . . . . . . . . . . . . . . . . .

48

4.1.2

Eggbox Sheet . . . . . . . . . . . . . . . . . . . . . . . . . .

54

Out-of-Plane Kinematics . . . . . . . . . . . . . . . . . . . . . . . .

58

4.2.1

Miura Sheet . . . . . . . . . . . . . . . . . . . . . . . . . . .

63

4.2.2

Eggbox Sheet . . . . . . . . . . . . . . . . . . . . . . . . . .

67

iv

CONTENTS

4.3

Conclusion & Discussion . . . . . . . . . . . . . . . . . . . . . . . .

5 Numerical Analysis
5.1

5.2

5.3

72
73

Mechanical Model: Bar Framework . . . . . . . . . . . . . . . . . .

73

5.1.1

Governing Equations . . . . . . . . . . . . . . . . . . . . . .

74

5.1.2

Kinematics: Compatibility

. . . . . . . . . . . . . . . . . .

74

5.1.3

Stiffness: Material Stiffness . . . . . . . . . . . . . . . . . .

75

5.1.4

Extensions to Framework Analysis . . . . . . . . . . . . . .

78

Modal Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

79

5.2.1

Symmetry Analysis . . . . . . . . . . . . . . . . . . . . . . .

87

5.2.2

Curve Veering & Imperfections . . . . . . . . . . . . . . . .

87

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

93

6 Manufacture of Folded Sheets

94

6.1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

94

6.2

Review of Manufacturing Methods . . . . . . . . . . . . . . . . . .

95

6.2.1

Synchrononous Folding Processes . . . . . . . . . . . . . . .

97

6.2.2

Gradual Folding Processes . . . . . . . . . . . . . . . . . . . 102

6.2.3

Pre-Gathering Processes . . . . . . . . . . . . . . . . . . . . 107

6.2.4

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.3

6.4

Cold Gas-Pressure Folding . . . . . . . . . . . . . . . . . . . . . . . 117


6.3.1

Process Description . . . . . . . . . . . . . . . . . . . . . . 117

6.3.2

Process Calculations . . . . . . . . . . . . . . . . . . . . . . 117

6.3.3

Manufacturing Trials . . . . . . . . . . . . . . . . . . . . . . 126

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

7 Conclusions & Future Work

133

7.1

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

7.2

Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

References

137

Chapter 1

Introduction
Traditionally, thin-walled shells have been designed with the purpose of carrying external loads efficiently and rigidly, by an appropriate design of their global
geometry. This thesis forms part of ongoing work to extend the capabilities of
conventional shell structures, and expand the applications beyond their traditional remit. For example, using a combination of geometry, material anisotropy
and pre-stress it is possible to design multi-stable shells. These shells can switch
repeatedly and reliably between multiple states, adapting to different functional
requirements. Examples include bistable deployable cylindrical booms (Guest &
Pellegrino, 2006), and a bistable helicopter blade that changes shape during hover
and vertical lift (Daynes et al., 2009). In both cases the bistability is introduced
by tailoring the fibre layup of a composite laminate shell.
Alternatively, the mechanical properties and the range of applications can be extended by means of textured shell structures. By introducing a local texture (such
as corrugations, dimples, folds, etc.) to otherwise isotropic thin-walled sheets, the
global mechanical properties of the sheets can be favourably modified. The local
texture has no clearly defined scale, but lies somewhere between the material and
the structural level. A classic example is a corrugated sheet, where the bending
stiffness is increased through the local geometry; a more advanced extension is a
pre-stressed bistable corrugated sheet, where the local texture locks in the stored
strain energy (Norman et al., 2008b).
The research into textured thin-walled shells branched into the development of
compliant shell mechanisms, where the texture provides a distinct kinematic layer
to the mechanical properties (Seffen, 2011); Figure 1.1. The folded shell structures
described in this thesis are examples of such sheets. Globally they can be regarded
as thin-walled shells, but at meso scale they consist of a tessellation of unit cells,
which in turn are composed of thin-walled facets joined at distinct fold lines. These
folds may be straight or curved, and the surface may have points or lines of nonzero Gaussian curvature where the facets meet. It is the hierarchical interaction
between the articulation at the fold lines and the deformation of the interlying
facets, which imbues these shells with unconventional mechanical properties.

Figure 1.1: The morphing curved corrugated shell investigated by Norman et al.
(2009) is an example of a compliant shell mechanism. Under uni-axial tension the
shell coils up tighter (top-right), and under bending it can assume global shapes
with both positive or negative Gaussian curvature (bottom).
The characteristic behaviour of folded shell structures straddles that of structures
and mechanisms. Firstly, they have the ability to undergo large displacements
within soft mechanistic deformation modes. Consequently, there is an anisotropy
in deformation modes, as the fold lines enable motion in some, and inhibit deformation in other modes. There may also be a non-trivial interaction between
different modes, such as in-plane and out-of-plane coupling. A common thread
further appears to be the ability of the shells to change their global Gaussian
curvature, i.e. the curvature of an equivalent mid-plane, with only developable deformations at material level. Our focus is on the mechanistic deformations of the
folded sheets, and we generally consider the deformation of the interlying material
as developable (i.e. no stretching of the material).
The presented study is a first analysis of folded shell structures and their mechanical properties. The aim is to describe the kinematic and structural properties,
and find suitable modelling methods to capture their salient behaviour. Low-level
effects such as material fatigue at the fold lines are not currently considered. A review of the mathematics of rigid origami provides a solid theoretical background for
the design of folded shell structures, which ensures the presence of a soft kinematic
mechanism. Furthermore, a potential impediment to their practical applications
is removed by investigating suitable manufacturing techniques. This study has
generally been agnostic towards any specific applications; instead, a broad foundation was established for the design, analysis and manufacture of these types of
structures. Nonetheless, potential applications of folded shell structures certainly
exist and will briefly be highlighted.

Morphing Skins An emerging field of research is concerned with the technical


challenges of designing morphing aircraft; the ultimate aim is to modify the wings
aerodynamic profile during flight, to adapt to changing conditions. An important
aspect is the development of a suitable morphing skin. The requirements are demanding and contradictory: the skin should be light-weight, have a low membrane
stiffness (small actuation forces), high flexural stiffness (to withstand aerodynamic
loads) and be able to attain large strain rates (Olympio, 2009). Thill et al. (2008)
provides an extensive review of proposals, but notes that the level of maturity
for morphing skins is low and the existing concepts are very early in development. Yokozeki et al. (2006) explored a promising solution using highly anistropic
corrugated skins; this provides the desired combination of mechanical properties,
but morphing is limited to one direction. Olympio (2009) explores the concept
of morphing honeycombs, whose Poissons ratio can be tailored to requirements.
Those honeycombs either support a closed elastic skin, or are filled with a low
modulus material to form a closed surface. Ramrakhyani et al. (2005) proposed
several skin concepts, including a folded skin mounted on a load bearing variable
geometry truss; however, this concept has not been further pursued. The folded
shell structures provide an unexplored area for morphing skins, with a promising
combination of mechanical properties. They can be designed to have low in-plane
moduli, whilst maintaining their bending stiffness; they can expand biaxially, and
attain doubly-curved configurations. Furthermore, some folded shells can be tailored for a desired coupling between extension and bending. Many issues, however,
are unresolved. For example, the maximally attainable in-plane strains and resulting fatigue at the fold lines are currently unknown, and the shells strongly
non-linear Poissons ratio may be undesirable.

Compliant Mechanisms Compliant mechanisms are designed to transmit both


motion and force mechanically, relying on elastic deformation of its elements rather
than conventional joints and hinges. Advantages include high displacement accuracy, reduced hysteris, zero backlash and wear, and ease of manufacture without
assembly. The design of such mechanisms treads a fine line between achieving
adequate stiffness in order to exert sufficient force at the end-effector, and yet be
flexible enough that the required motion due to applied loads is realised. The
folded shell structures may extend this concept beyond the current linkage-based
compliant mechanisms towards compliant surfaces; the shells stiff deformation
modes may then be employed to actively transmit forces, rather than merely passively resist external loads. A possible application may be found in supportive
exoskeletons. For example, the X-arm project by InteSpring (2011) investigates
the potential applications of a passive exoskeleton, to support a users upper body
weight and thereby reduce health problems for people in professions with heavy
lifting and long uncomfortable postures, such as healthcare workers and mechan-

1.1 Outline

ics. Compliant shell mechanisms may provide the capability to conform to the
users movements, but preserve the ability to transmit forces.

Architectural Fa
cades Much of recent theoretical work on rigid origami folding was motivated by the desire to design deployable architectural structures, such
as facades, roofs and shelters. The folded shell structures may serve a different
architectural purpose, providing a flexible building envelope that can adjust to
changes in shape, whether for structural, functional or visual purposes. Furthermore, the textured sheets may also be used as static cladding material, for their
ability to conform to doubly-curved surfaces.

1.1

Outline

Chapter 2 establishes the necessary context and concepts for the study of folded
shell structures. The state of the art in textured shell structures is described, and
examples from the literature were classified by their dominant structural purpose.
The mathematics of curvature and creases was summarised to provide essential
concepts for the deformations of these shells. The distinct fold lines in the shell
surface invite a strong link to origami folding, and the recent developments in
the mathematics of rigid origami are presented as a means to design folded shell
structures with soft internal deformation modes.
Chapter 3 introduces two representative example folded shell structures. Their
characteristic mechanical properties are described and illustrated with some demonstrative experiments, and potential modelling methods are discussed.
Chapter 4 describes the global deformations of folded shell structures, by analysing
the kinematics of its constituent unit cells. Their characteristic in-plane and outof-plane properties were described in terms of developable deformations of the unit
cells. This analysis provides the compatibility equations, as a first step towards
an equivalent continuum model. The numerical analysis in Chapter 5 takes a
more holistic approach to the global sheet deformations. The folded shells were
modelled as pin-jointed trusses with additional planarity constraints to provide
bending stiffness for the folds and facets. A modal analysis establishes the dominant deformation modes for a range of geometries and material properties.
Chapter 6 describes the manufacturing challenges of developable folded sheets.
Existing manufacturing processes are reviewed and classified. A novel cold gas
pressure manufacturing method is proposed, and the required forming pressures
are derived analytically and compared with experimental data.

Chapter 2

Background and Concepts


2.1

Textured Shell Structures

A versatile and promising means of extending the mechanical properties of thinwalled shells is to introduce a local texture pattern, with surface features at a scale
intermediate to the material and global structural scale. The shells base material
will generally be isotropic, and the change in mechanical properties is entirely due
to the local geometry. Paradoxically, the main structural applications for these
types of shell structures either exploit the texture for its increase in stiffness, or
its increase in flexibility.

2.1.1

Texture Bending Stiffness

The primary application of local texture in thin-walled shells has been to increase
the shells bending stiffness, by increasing the second moment of area. The classic
example is corrugated sheets. Invented in 1829 by Henry Palmer, they enabled
the construction of large-span enclosures such as the great railway stations of
the Victorian era (Mornement & Holloway, 2007), and have been ubiquitous ever
since. Exploiting their increased stiffness at minimal expense of weight, Junkers
(1929) used corrugated sheets as the outer skin for several aircraft designs. The
mechanical properties of corrugated sheets are now well established; Briassoulis
(1986) provided the equivalent shell properties of smoothly corrugated sheets,
which Samanta & Mukhopadhyay (1999) extended to folded plate corrugations.
Miura (1969) introduced a category of textured shell structures, Pseudo-Cylindrical
Concave Polyhedral (PCCP) shells, inspired by the stable inextensional postbuckling geometry of axially compressed cylinders, known as the Yoshimura pattern. Rather than considering the buckled sheets as failed, Miura (1969) exploited
the increased circumferential bending stiffness. These globally curved shells consist of planar facets joined at straight fold lines, and their local geometry increases
the shells bending stiffness, but at the expense of its in-plane stiffness. Proposed
applications included undersea pressure hulls due to the increased buckling resis-

2.1 Textured Shell Structures

tance (Knapp, 1977); a stress analysis of PCCP shells under both axial load and
hydrostatic pressure was therefore performed by Tanizawa & Miura (1975). A
commercial application was found in textured drinking cans, which enable the use
of thinner material (Ishinabe et al., 1992). Another proposed application of PCCP
shells is their ability to cover large unsupported spans, and is very closely related
to the use of folded plates in architecture (Salvadori & Heller, 1963; Engel, 1968;
Buri, 2010). For the textured shells, the features are generally less pronounced
and the appearance of a globally thin-walled shell is therefore stronger.
Many textured sheets aim to increase the second moment of area along any bending axis; see Figure 2.1. Ewald (1948) describes a sheet where the texture consists
of alternating protrusions, whose tessellation pattern is staggered to avoid any direct fold lines in the core. The dimpled sheet described by Pfistershammer (1952)
aims to minimise the variability in bending stiffness by a suitable choice of dimple pattern and their up/down orientation. Farmer & Spangler (1962) studied
the bending properties of such Collattan sheets, and observed remaining preferred
bending axes. The stiffness of sheets with interlocking embossed patterns was investigated by Tewes (1966) and measurements confirmed an increased, but highly
anistropic bending stiffness. A great variety of such interlocking embossing patterns is possible (de Swart, 1955), but no studies of preferred patterns have been
found. Mirtsch et al. (2006) describe a vault-structured material, where the material is locally buckled into a three-dimensional shape under hydrostatic pressure.
The described manufacturing process aims to minimise material deformation as
well as required forming forces (Behrens & Ellert, 2005).
An important application of textured sheets is as a stiffening element in a combined
panel. Rauscher et al. (2008) describe the manufacture of hump plates, whereby
two sheets are connected through humps which were formed into one or both of
the cover sheets. More commonly, face sheets are added on either side of the
textured sheet to form a sandwich panel; in contrast to conventional honeycomb
cores, these textured cores not only provide compressive and shear strength, but
also significantly contribute to the bending stiffness of the panel. Multiple sheets
may be stacked and joined together to form a core material (e.g. Pfistershammer,
1952). Countless variations of textured cores have been found in the literature,
and a detailed comparison is beyond the scope of this review. Two types of core are
noteworthy. Schott (1975) describes a core of a square array of alternating conical
protrusions, with an important feature that the core material can be deformed into
both single and compound curvatures with minimal loss of load carrying capacity
of the core; see Figure 2.2. Another important textured core material was first
introduced by Rapp (1960), based on a doubly-corrugated sheet which can be
folded from a flat sheet without stretching of the base material: a folded core.

2.1 Textured Shell Structures

(a) Ewald (1948).

(b) Pfistershammer (1952).

(c) de Swart (1955).

(d) Mirtsch et al. (2006).

Figure 2.1: Examples of textured sheets with an increased stiffness along any bending axis, with (a) a staggered tessellation of protrusions, (b) an array of dimples,
(c) interlocking embossing patterns, and (d) a locally bulged vault-structured
material.

2.1 Textured Shell Structures

(a) Schott (1975) describes a core material with conical protrusions,


which can be adjusted to fit curved surfaces.

(b) a core material that can be folded from flat sheet material, with only bending operations (Rapp, 1960).

Figure 2.2: Two types of textured sandwich panel cores, with (a) designed to
conform to double curvatures, and (b) the first known example of a folded sandwich
panel core.

2.1 Textured Shell Structures

(a) the Zeta core provides increased bonding contact area, at the expense of no longer
being developable (Miura, 1975b).

(b) first known example of curved folded


core material (Sehrndt, 1960).

Figure 2.3: Examples of fold core geometries.


Folded Core A similar core structure, termed Zeta core was analysed by Miura
(1972), who derived analytical predictions for the core shear strength, with an
optimal efficiency found for a fold angle of 55 . Among its advantages were listed
the high shear modulus and strength, an isotropic or controllable shear modulus,
and improved buckling resistance due to the ruled surface along the facets. By
providing bonding surfaces on the crests of the corrugations, the Zeta core sheet
is no longer developable, and therefore cannot be folded from sheet material.
This decision was further motivated by the absence of suitable mass-production
methods. The Zeta cores were instead manufactured by vacuum forming plastics,
or progressive press forming of aluminium sheets (Miura, 1975b).
In recent years there has been a revived interest in folded core material for
lightweight sandwich panels for aircraft fuselages (Heimbs et al., 2010). An important motivation is the fact that fold cores possess open ventilation channels,
solving the problem of moisture condensation in the core material. Furthermore,
high-performance materials such as resin-impregnated aramid paper (e.g., Nomex
or Kevlar) can be folded, and the unit cells can be designed with respect to specific
mechanical requirements, as they offer a large design space for tailored properties.
The application in the aerospace industry also spurred the development of continuous manufacturing techniques for fold cores; a review is given in Chapter 6.
Extensive work has taken place on experimental and numerical testing of the mechanical properties of fold cores, with a specific focus on impact protection. Heimbs
et al. (2010) provide a good overview of recent work, and identify two major issues
in the virtual testing simulations. The first point is a reasonable implementation
of imperfections in the numerical models, which are inevitably present in cellular
structures resulting from the manufacturing process, affecting the buckling and

2.1 Textured Shell Structures

strength properties of the foldcore. A range of methods to introduce appropriate imperfections to the finite element models is discussed by Heimbs (2009) and
Baranger et al. (2011b). The second point is a correct constitutive modelling of
the cell wall material; especially for the resin-impregnated aramid paper fold cores
this requires extensive correlation with experimental data (Fischer et al., 2009).
Lebee & Sab (2010) derived analytical and numerical upper and lower bounds for
the shear stiffness of the chevron folded core materials in a sandwich panel. For
some existing patterns, both the numerical and analytical bounds were found to
be too loose. While for honeycomb cores the discrepancy between the bounds was
determined to be a skin effect, in the case of the folded core the large discrepancy
still has no explanation and requires more refined models, to take into account
interaction between skin and core.
Klett & Drechsler (2009) identified three design scales for folded cores: macro-level,
concerned with the global geometry of the core material; meso-level, which tunes
the unit cell geometry to desired mechanical requirements, and micro-scale, at
the level of the sheet material. Especially the ability to attain globally non-planar
geometries without distortion of the core material is an important benefit of folded
cores; see Figure 2.4. The possibilities include single curvature panels (Sehrndt,
1960; Khaliulin & Desyatov, 1993), doubly-curved cores (Talakov, 2010), as well
as helical core material designed to wrap around a cylindrical fuselage (Akishev
et al., 2010). The development of computational design tools has further enabled
the design of more freeform geometries, with varying curvatures, densities and fold
depths (Klett et al., 2007).

Bending Anisotropy
The introduction of a texture to thin-walled shells will generally introduce an
anistropy in its bending stiffness. While often considered undesirable, this anisotropy
may be also be positively exploited, for example in bi- and multistable structures.
Composite shells with anistropic bending stiffness have been used to enable bistability in cylindrical shell structures (Guest & Pellegrino, 2006). A combination
of pre-stress and simple corrugations was used to design multi-stable corrugated
shell structures (Norman et al., 2008b), where the introduced geometry locks in
the stored strain energy. Greater control over the bistability can be obtained by
exploring texture patterns; for example, Norman et al. (2008a) describe a doublycorrugated pattern, which introduces the necessary anistropy for a bistable cylindrical shell. Another example of the combination of pre-stress and texturing was
found in a study into hierarchical multi-stable dimpled shells: an array of bistable
dimples provides local control over the release of pre-stress, and thereby determines
the global geometry of the sheet (Seffen, 2006). When all dimples are oriented the

10

2.1 Textured Shell Structures

(a) transversely curved core.

(b) longitudinally curved core.

(c) freeform core geometry, with varing density, fold depth and curvature; the fold pattern is found by reflection in the two prescribed bounding curves.

(d) helical fold core, enclosing a cylindrical


fuselage (Akishev et al., 2010).

Figure 2.4: The versatility of folded cores is illustrated by the ability to design
globally curved geometries, whilst folding from flat sheet material with minimal
material deformation. Images by Klett et al. (2007) and Akishev et al. (2010).

11

2.1 Textured Shell Structures

same way, there is a uniform state of self stress, but the sheet will have preferred
axes of cylindrical bending orthogonal to the directions of least packing of dimples;
Seffen (2007) provides a simplified homogenisation of the dimpled sheet by assuming cylindrical beam-like bending and integrating the second moment of area. An
alternative approach to controlling the bending anistropy of thin-walled shells, is
the use of perforation patterns (Duncan & Upfold, 1963; Fort, 1970).
While the ability to control the anistropy of laminated composite shells is well
established (e.g. Halpin, 1992; Nettles, 1994) the possibilities and limitations of
texture patterns remain largely unknown.

2.1.2

Texture Flexibility

Another application of textured sheets is for their in-plane flexibility, for example
to provide stress relief for applications with large changes in temperature, such as
linings for cryogenic storage vessels. Various patents have been found describing
an expandable sheet with a number of parallel corrugations set at an angle to each
other, thereby allowing expansion of the sheet in multiple directions; see Figure 2.5.
The challenge in those cases is finding a satisfactory solution for the point where
the corrugations intersect. Dunajeff (1939) used a perforation at the vertices of
crossing corrugations to avoid stress concentrations. Arne (1973) places expandible
conical frusta at the intersection, whereas Girot (1964) describes expansion joints
that can be folded from the sheet material along with the corrugations. Similarly,
French & Petty (1965) describe expansion joints which are formed by twist-folding
the material at the point where corrugations meet.
Whereas the corrugated sheets concentrate the expansion along specific locations
in the sheet, examples exist of sheets which distribute their in-plane flexibility
throughout the surface; see Figure 2.6. Dunajeff (1941) describes several concepts
for resilient sheets, including the tessellation of expansible hexagonal units, as
well as a zigzag corrugation that enables extension along two orthogonal axes.
Brunner (1968) describes expansible sheets, consisting of parallelograms joined at
their edges such that each parallelogram is inclined to the midplane of the sheet;
this includes both the Eggbox and Miura sheet described in this thesis. Lueke
(1994) provides a brief review of flexible linings for cryogenic storage vessels, and
suggests that strains of about 1-2% are necessary. Furthermore, sheets with a
negative Poissons ratio are considered desirable, in part because they are also
flexible at 45 to the primary axes. The proposed sheet is a smoothed version
of the doubly-corrugated fold pattern to avoid stress concentrations at the folds,
and the depth of the pattern is limited for weight efficiency. Lastly, Fritz et al.
(1996) describe a modification of an Eggbox pattern, with a corrugation within
the parallelogram facets, enabling expansion along all axes.

12

2.1 Textured Shell Structures

(a) Arne (1973).

(b) Girot (1964).

(c) French & Petty (1965).

Figure 2.5: Examples of plane expansible surfaces with expansion joints at the
intersection of crossing corrugations.

13

2.1 Textured Shell Structures

(a) Dunajeff (1941).

(b) Brunner (1968).

(c) Fritz et al. (1996).

Figure 2.6: Examples of plane expansible sheets, including (a) the first known
example of the Miura sheet for engineering applications, (b) the first known example of a flexible Eggbox sheet, and (c) a modified Eggbox pattern that enables
in-plane expansion along all axes.

14

2.1 Textured Shell Structures

Most flexible textured materials only consider in-plane flexibility. Yokozeki et al.
(2006) explore the idea of ultra-anistropic corrugated surfaces of composite materials. For the intended application in morphing aircraft wings, it is desirable
for the skin to be stiff in span direction, but flexible in chord. Analytical expressions were found for the bending stiffness along and across the corrugations, and
the anisotropy in bending was further increased by inserting rigid rods into the
corrugations.

2.1.3

Texture Impact Absorption

A further application for textured sheets is for their ability to absorb impact
through plastic deformation of the texture pattern. Deshpande & Fleck (2003)
looked at the collapse mechanism of a doubly-corrugated Eggbox sheet, and describe feasible geometries that induce a travelling plastic knuckle, rather than
tearing of the material. Zupan et al. (2003) further detail the collapse mechanisms, and shows the dependency on the in-plane kinematic constraints. Basily &
Elsayed (2004) investigated the impact absorption of folded core material; compared to equivalent honeycombs the folded core provides a more consistent forcedisplacement profile during impact, and the folded sheets have the ability to absorb
energy in all directions of impact.

2.1.4

Folded Shell Structures

The folded shell structures described in this thesis are a type of textured shell structure. The shell consists of a tessellation of piecewise developable facets, bounded
by distinct fold lines. These fold lines may be curved or straight, and the sheets
may be developable or may have points and lines of non-zero Gaussian curvature.
The distinguishing feature of the folded shell structures are the distinct fold lines
which enable flexibility, but also provide a texture with an increased second moment of area. This combination creates an anistropy in the shells deformation
modes, which is more complex than for example for the highly orthotropic corrugated material of Yokozeki et al. (2006). Furthermore, as described by Seffen
(2011), these structures display unusual properties due to the hierarchical interaction of the deformation along the fold lines and the interlying material; these are
described in Chapter 3.
Several textured shells discussed previously fall within the definition of folded
shell structures. However, they were either exclusively employed for their bending
stiffness (e.g., Miura, 1969) or their planar expansibility (e.g., Brunner, 1968), and
where any coupling between in-plane and out-of-plane properties was observed this
was neglected (Lueke, 1994).

15

2.2 Origami Folding and Foldability

2.2

Origami Folding and Foldability

A key feature of the folded shell structures is the distinct fold lines in the surface,
which begs a strong link to origami folding. Many folded shells can also be folded
from flat sheet material, further solidifying this connection. The mathematics
of origami is a rich and thriving field (Demaine & ORourke, 2007); the topics
presented here form only a small subset, and are selected for their application
to the analysis and design of folded shell structures. We shall first discuss the
intrinsic geometry of surfaces and folds, before exploring the rigid foldability of
tessellated fold patterns, and resulting design methods.

2.2.1

Curvature and Creases

Any introduction to the mathematics of origami benefits from the discussion of


the curvature of surfaces, and specifically the concept of Gaussian curvature. Differential geometry describes the properties of surfaces; an excellent introduction is
given by Hilbert & Cohn-Vossen (1952), with more detailed information in Struik
(1961). For sufficiently smooth surfaces (C 2 , smooth in second derivative), the
principal curvatures 1 2 are defined for each interior point as the maximum
and minimum (signed) curvatures for the geodesics through the point. The Gaussian curvature K is then given as the product of the two principal curvatures:

K = 1 2

(2.1)

By virtue of Gauss Theorema Egregium 1 (Gauss, 1828) the Gaussian curvature


remains invariant under bending of the surface; a bending is any deformation
for which the arc lengths and angles of all curves drawn on the surface remain
invariant. The invariability means the Gaussian curvature is thus an intrinsic
property of the surface, independent of any external coordinate system. Moreover,
it enables a description of curvature at folds and vertices where the conventional
notion of the radius of curvature no longer holds.
An intrinsic description of Gaussian curvature considers a small area F on the
surface, and the solid angle G it subtends. The Gaussian curvature is then given
as
G
K = lim
(2.2)
F 0 F
which only involves information locally on the surface. In order to find the subtended angle G, we introduce the spherical representation of a surface; see Figure 2.7(a). Consider a closed contour k oriented counterclockwise on the surface
1 Latin:

Remarkable Theorem. A translation of Gauss treatise on geometry of surfaces is


published as Gauss (1902).

16

2.2 Origami Folding and Foldability

and enclosing a point on the surface; each point on the contour has an associated
unit vector normal to and oriented away from the surface. Now transfer this set
of normal vectors to the centre of a unit radius sphere, the Gaussian sphere. The
ratio K of the area G enclosed by the resulting closed contour k 0 (the trace of the
contour k) on the sphere, to the area F enclosed by the contour k has a definite
limit as k shrinks to a point, and is the Gaussian curvature. When k 0 is oriented
counterclockwise, the calculated area and therefore the curvature is positive (and
vice versa).
This approach may now be extended to folds and polyhedral vertices, as shown in
Figure 2.7(b). At a fold line the normal vector is not uniquely defined and the arc
on the trace covers all possible directions; the arc length therefore corresponds to
the dihedral fold angle between the two planes. Any closed curve around a point on
the fold line will map onto a single arc on the unit sphere, which therefore has zero
Gaussian curvature. Using results from spherical trigonometry it can be shown
that at the vertex the area G within the trace is equal to 2 minus the sum of the
sector angles of the corresponding surface (Huffman, 1976; Calladine, 1983); this
is called the spherical excess, or angular defect. It follows that the spherical excess,
and thereby the Gaussian curvature is independent of the fold angles beween the
plane sectors, and it provides an intrinsic measure of the curvature at the vertex.
When the sum of sector angles falls short of 2 the vertex has positive Gaussian
curvature, when it exceeds 2 it is negative, and when they add up to 2 the
vertex has zero Gaussian curvature.
Of interest to origami folding is the concept of developable surfaces. These surfaces
locally have Gaussian curvature everywhere zero, and are globally isometric to the
plane. A developable surface may alternatively be defined as a ruled surface (there
exists a line segment through any point on the surface) for which the tangent
plane is the same at any point along a line embedded in the surface. Under any
bending or folding deformation, the zero Gaussian curvature everywhere on the
sheet is preserved, which limits its smooth deformations to cylindrical, conical and
tangent surfaces.
The presence of creases (or folds) in the surface greatly affects the attainable
geometries. Demaine et al. (2009) characterise the facets of a crease pattern,
which are by definition regions folded without creases; the creases are defined
as discontinuities of the first derivative of the surface. One result was that for
polygonal facets (bounded by straight lines) of the crease pattern, the fold lines
fold to straight line segments in space, and the facets must therefore stay planar
(provided they have no boundary edges). The geometry of fold lines and vertices
was first investigated in a landmark paper by Huffman (1976); using the trace and
the condition that its net surface must be zero, an algebraic relationship between
the dihedral fold angles of degree-4 vertices was found. Miura (1989) showed that

17

2.2 Origami Folding and Foldability

(a) the spherical image of a closed contour around a point on a surface


is produced by tracing the contour and transferring the unit normals
to the centre of a unit sphere. The Gaussian curvature is defined as
the limit of the ratio of the area G of the trace over the area F of
the curve on the surface, as F shrinks to a point. If the contour k is
traversed counterclockwise, a clockwise enclosed area on the spherical
image is considered to be negative, as is the case for the saddle shape.

(b) spherical image for polyhedral vertex. The fold angle between
facets is reflected by the arc length on the spherical image. The
enclosed area on the trace is given as the sum of the sector angles
minus 2, and is independent of the fold angles.

Figure 2.7: The spherical representation provides an intrinsic view of the Gaussian
curvature, both for (a) smooth and (b) polyhedral surfaces. Images from Hilbert
& Cohn-Vossen (1952).

18

2.2 Origami Folding and Foldability

the simplest origami fold is a vertex with four folds, one valley and three mountain
folds. Huffman (1976) also considered curved folds; curved folding is beyond the
scope of this review, and the reader is referred to Duncan & Duncan (1982) and
Fuchs & Tabachnikov (1999) for the necessary differential geometry.

2.2.2

Rigid Foldability

Rigid foldability considers an origami crease pattern as freely hinged flat rigid
panels, and explores whether a continuous folding can take place without any
deformation of the facets. Unlike the study of flat-foldability, which is concerned
with the final state only, rigid foldability describes the continuous route from flat
to fully folded state.
The investigation of rigid foldability has largely been motivated by the design
of deployable structures, both developable and non-developable. For example,
Dureisseix et al. (2011) describe deployable structures for architectural applications, which maintain their flat-foldability for compact stowage, but are purposely
non-developable to improve the structures load-bearing capacity in its fully deployed state. For the design of folded shell structures, the mathematical concepts
of rigid foldability and the resulting design methods are also of great interest.
Namely, if the tessellated fold pattern is rigid-foldable there exists at least one
soft internal mechanism by virtue of the bending along all fold lines. Furthermore, rigid folding provides the ability to manufacture a textured surface from
flat sheet material with minimal material deformation.
First several modelling methods for rigid origami are described, followed by an
overview of recent developments in describing the rigid foldability of 1 DOF multivertex fold patterns. In recent years significant progress has been made in the
understanding of rigid foldable patterns, providing an unprecedented freedom of
design.
Origami Modelling
The deformation of developable surfaces can be represented in a variety of ways,
with a suitable choice depending on the application (Balkcom, 2004). The overview
given here is restricted to rigid origami, where the material does not stretch and
facets do not bend.
Nodal Coordinates A convenient choice of representing the folding process of
rigid origami, is by the position and motion of its vertices. The vertices are then
modelled as pin-joints and the fold lines as bars (Schenk & Guest, 2011). For
non-triangular facets, additional bars and planarity constraints must be added to

19

2.2 Origami Folding and Foldability

avoid deformation of the facets. This is the simplest mathematical representation


of origami folding, as the Jacobian of the bar length constraints is linear in the
nodal coordinates. Alternatively, the facets can be modelled by hinged plane stress
elements in a finite element analysis (Resch & Christiansen, 1971).

Fold Angles Representing the rigid folding process in terms of the fold angles
of the crease lines is not only more intuitive, but also fosters the development of
a more fundamental understanding of rigid foldability.
Belcastro & Hull (2002) describe the modelling of non-flat origami using piecewise affine (flat facets) and isometric (no stretching) transformations around each
vertex; this results in a series of rotation matrices that include both the sector
and fold angles around the vertex. The loop closure constraint (no cutting of
material) then results in the identity of those rotation matrices (Kawasaki, 1994).
This condition is a necessary, but not sufficient condition for rigid foldability as
it does not take into consideration any self-intersection (Belcastro & Hull, 2002).
For single vertex foldings Streinu & Whiteley (2005) further showed that the 3D
folded state can be attained by continuous motion without bending of the facets,
by analogy with spherical polygonal linkages. Wu & You (2010) also exploit the
analogy with spherical linkages, and describe the loop closure in terms of the rotations of normal vectors to the panels, using quaternions. Balkcom (2004) used
mechanism theory with forward/inverse kinematics, by virtual cutting along some
creases, and much of the geometric work by Stachel (2010a) relies on the coupling
between multiple spherical 4-bar linkages around facets.
For the numerical simulation of rigid origami, the non-linear loop closure equations
must be continuously satisfied. Tachi (2009c) describes such a numerical method,
based on a modified version of the equations of Belcastro & Hull (2002). If the
surface is a topological disk, the closure of any loop around vertices can be reduced
to the combination of local constraints around interior vertices. For each interior
vertex and its incident fold lines Li with fold angles 0 , . . . , n (see Figure 2.8) the
rotation matrix identity condition is written as
F (0 . . . n ) = 0 . . . n1 n = I

(2.3)

where i represents the rotation around each of the fold

1
0
0
cos i sin i 0

i = 0 cos i sin i sin i


cos i 0
0 sin i
cos i
0
0
1

lines

(2.4)

with i the sector angle between fold lines i and i + 1. When deriving the Jacobian
of the constraint equation it is useful to realise that the partial derivative with

20

2.2 Origami Folding and Foldability

L1

L2

1
0

2
3
3

L0

L3

Figure 2.8: At each vertex the incident fold lines Li have an associated fold angle
i , and sector angles i between Li and Li+1 .
respect to a fold angle i represents an instantaneous rotation around the direction
vector of its fold line; this is a skew symmetric matrix

0 a c
F

= a
(2.5)
0 b
i
c b
0
where a, b, c constitute the direction cosines for the fold line; the number of
independent constraints per vertex thereby reduces to three. The constraints for
every vertex in the sheet can subsequently be combined into a 3M N matrix

J1
1
0
. . .
. . .
J =
(2.6)
. . = .
JM
N
0
with N the number of creases and M the number of vertices. If there are holes
in the surface, additional loop constraints must be added (Tachi, 2010c), but
are here neglected for simplicity. The nullspace of the Jacobian J provides the
possible motions that satisfy the constraints. Next configurations are found using
an Euler integration and projecting each step onto the constraint space using the
Moore-Penrose pseudo-inverse, in combination with a Newton-Raphson iteration
to eliminate errors.
For a multi-vertex pattern the Jacobian J will generally be overconstrained with
3M > N . The nullity of the matrix, and thereby the DOF of the fold pattern,
must therefore follow from Ns redundant constraints: DOF = N 3M + Ns .
Finding multi-vertex crease patterns that supply that singular configuration is the
challenge of rigid foldability.

21

2.2 Origami Folding and Foldability

Multi-Vertex Rigid Foldability


Firstly, as the number of vertices, facets and edges are related by the EulerPoincare characteristic of the surface, the number of degrees of freedom for the
overall system is limited. Specifically, a crease pattern has at most N0 degrees of
freedom (assuming that all facets are triangulated), where N0 is the number of
vertices on the boundary of the surface (Tachi, 2010c).
In the case of quadrilateral mesh origami, with degree-4 vertices, the number of
fold lines is generally smaller than the number of constraints. Let us consider a
tessellation of degree-4 vertices, with an n n array of quadrilateral facets:
N
M
DOF

=
=

2n(n 1)
(n 1)

(2.7)

(2.8)
2

= N 3M = (n 2) + 1

(2.9)

where the DOF become negative for n > 2. A similar derivation was given by
Dureisseix (2011) using recursion of mobility equations. Alternatively, consider
that in a multi-vertex pattern with degree-4 vertices, an assigned set of fold angles
will generally conflict; see Figure 2.9. The flexibility of a quadrilateral mesh is
therefore due to redundancy of constraints, such as parallel fold lines or an intrinsic
symmetry at the vertices. The best known example of a rigid foldable degree-4
vertex pattern is the Miura-ori (Figure 2.10) and it will be used throughout this
thesis.
Watanabe & Kawaguchi (2009) described a test for infinitesimal rigid-foldability of
a crease pattern with mountain-valley assignment, using the Jacobian and Hessian
of the rotation matrix introduced by Belcastro & Hull (2002), but a more fundamental understanding of the necessary conditions for rigid foldability of multivertex patterns is desirable.
Rigid foldability of a general planar quadrilateral mesh is still an open question,
but the conditions for several classes of fold patterns are now understood. We
shall describe some recent developments in generalising rigid origami folding, and
point towards a direction by which these may be extended further.

Plane Tessellation A first type of rigid foldable mesh is described by Huffman


(1976) and Kokotsakis (1933, Fig. 15). It is the plane tessellation of a degree4 vertex whereby the vertex is rotated in order to avoid contradiction of the
fold angles; see Figure 2.11. This was considered self-evidently rigid-foldable by
Huffman (1976), but a more rigorous geometric proof was given by Stachel (2009).

22

2.2 Origami Folding and Foldability

Figure 2.9: Conflicting fold angle assignment for a mesh of general degree-4 vertices. When a single incident fold angle 0 is prescribed, all fold angles at the
vertex are subsequently defined. When looping around the central facet, this results in a conflict at the last vertex, as two incident fold angles are prescribed. The
pattern will therefore generally not be rigid-foldable. Figure after Tachi (2010a).

Figure 2.10: The classic Miura-ori pattern is both rigid-foldable and flat-foldable;
the redundancy of its constraints follows both from intrinsic symmetry at its vertices, as well as global symmetry with parallel fold lines. The Miura pattern forms
the basis for the generalisations of rigid-foldable developable origami.

23

2.2 Origami Folding and Foldability

Figure 2.11: Rigid foldable tessellated fold pattern described by Huffman (1976)
and Stachel (2009). The plane tessellation of any convex quadrilateral, by iterated
180 rotations about the midpoint of the sides, is rigid foldable as all fold angles
automatically connect without contradiction. Image by Stachel (2009).
Flat Foldable and Developable Tachi (2009a) described the first general
extension for rigid-foldable flat-foldable degree-4 vertices, to form a generalised
freeform Miura-ori surface. The consideration of flat-foldability came from a desire for compact stowage of deployable structures. The Kawasaki-Justin theorem
provides a necessary condition for local flat-foldability (Belcastro & Hull, 2002;
Demaine & ORourke, 2007). The theorem states that for a vertex where n is
even,
n
X
(1)i i {0, 2, 2}

(2.10)

i=1

holds for flat-foldable vertices. For the case of a developable degree-4 vertex this
reduces to
0 = 2

and 1 = 3

(2.11)

The relationship between the fold angles i and j incident to the vertex can be
found using spherical trigonometry, and is given as follows (Tachi, 2010b):
(

Aij tan 2j (i j = 1 or 3)
i
tan
=
(2.12)
j
2
(i j = 2)
tan 2
where the latter represents that pairs of opposing fold lines have an equal absolute folding angle, and Aij is the coefficient between these two equivalent pairs,
determined by 0 . . . 3 . In other words, it is an intrinsic measure of the crease

24

2.2 Origami Folding and Foldability

pattern. Specifically, if |0 | = |2 | > |1 | = |3 |, then


s
1 + cos(0 1 )
|A01 | =
1 + cos(0 + 1 )

(2.13)

This relationship is a special case of the general formula for an origami vertex by
Huffman (1976). The formulation by Huffman (1976) provides further insight into
the degree-4 vertex, including conditions for local self-intersection.
Tachi (2009a) uses a modified version of Equation 2.12 to map between the alternating pairs of fold angles, in terms of a conversion coefficient. Using a loop
argument around each facet he provides a necessary condition for the existence
of rigid-foldable flat-foldable quadrilateral mesh origami. More importantly, it is
shown that if there exists a partially folded state (where every fold line is semifolded, i.e. 6= 0, 6= , ), it is finitely rigid foldable. An alternative way of
deriving this result is to note that Equation
2.12 o
is linear in tan 2i . Therefore, if
n
0)
an arbitrary semi-folded configuration tan i (t
2
angles, there must exist a finite path:
 


i (t0 ) tan 2t
i (t)
= tan
tan
2
2
tan t20

is found that satisfies all fold

(2.14)

where t (0 t ) is the parameter that defines the amount of folding. This


directly leads to a numerical design method: starting from a partly folded configuration of a known rigid foldable pattern such as the Miura-ori, the nodes in
the crease pattern can be incrementally modified whilst satisfying a number of
constraints: developability, flat-foldability and planarity of facets. This provides
an effective means to design freeform rigid-foldable surfaces. A useful corollary
from Tachi (2009a) states that if the conversion coefficient of every inner vertex
is constant in each row or each column, the pattern is rigid foldable.
Flat Foldable and Non-Developable The extension to non-developable surfaces made use of recent results on the finite integrability (continuous motion) of
a discrete Voss surface (Schief et al., 2007). This is a planar quadrilateral mesh
surface composed of degree-4 vertices, each of which satisfies 1 = 3 and 0 = 2 ;
this includes the Eggbox pattern described in Chapter 3. By introducing the concept of complementary fold angles, 0i = i , Tachi (2010b) showed that the
flat-foldable Miura and Voss vertices share an intrinsic symmetry, and their kinematics are identical. The resulting hybrid Miura-Voss patterns are bi-directionally
flat-foldable (Tachi, 2010b). Dureisseix et al. (2011) also describe a generalisation
of Miura-ori into doubly-curved non-developable surfaces, but it retains a plane
of reflective symmetry. It is interesting to note that in its generalised form the
Miura vertex will always expand in all directions simultaneously, whereas the Voss
vertex expands in one and contracts in the orthogonal direction.

25

2.2 Origami Folding and Foldability

Kokotsakis Meshes Recent work has shown that both the generalised Miura
and Voss patterns are a subset of flexible Kokotsakis meshes (Kokotsakis, 1933;
Stachel, 2010a,b). This connection unifies and further extends the families of
possible rigid-foldable quadrilateral mesh patterns.
In discrete differential geometry there is an interest in polyhedral structures composed of quadrilaterals, i.e. quadrilateral surfaces. When all quadrilaterals are
planar, the edges are geodesics and form a discrete conjugate net. When each
quadrilateral is seen as a rigid body and only the dihedral angles can vary, the
question arises under which conditions such structures are flexible. Stachel (2010a)
states a theorem due to Schief et al. (2007), that a discrete conjugate net in general position is continuously flexible if and only if all its 3 3 complexes, i.e. all
included Kokotsakis meshes, are continuously flexible. A Kokotsakis mesh is a
polyhedral structure consisting of an n-sided central polygon P0 surrounded by a
belt of polygons; see Figure 2.12. Each vertex Vi is the meeting point of 4 facets;
each facet is a rigid body and only the dihedral angles can vary. The lengths
of the sides of the central polygon have no bearing on the connectivity, and its
kinematics are therefore represented by a coupling of spherical linkages.
Stachel (2010b) presents several known categories of continuously flexible Kokotsakis meshes (see Figure 2.12):
I plane-symmetric : reflection in the plane of symmetry of V1 and V4 maps
each horizontal fold onto itself while the vertical ones are exchanged.
II translational : there is a translation V1 7 V4 and V2 7 V3 mapping the
three faces on the right-hand side of the vertical fold through a2 onto the
triple on the left-hand side of the vertical fold through a4 .
III isogonal : at each vertex the fold angles are congruent.
IV orthogonal : here the horizontal folds are located in parallel (say: horizontal)
planes, and the vertical folds in vertical planes. P0 is a trapezoid.
V line-symmetric : a line-reflection maps the linkage at V1 7 V4 and V2 7 V3 .
Case V is newly derived by Stachel (2010a), and includes the earlier example of
the plane tessellated quadrilateral shown in Figure 2.11. It is emphasised that a
complete classification of continuously flexible Kokotsakis meshes has not yet been
achieved.
Stachel (2010b) rephrases case III, such that a Kokotsakis mesh is flexible if at
each vertex Vi opposite angles are either equal or complementary: i = i , i = i ,
or i = i , i = i . This describes the intrinsic symmetry that unites
the generalised Miura-ori pattern and Voss surface, and shows why both are

26

2.2 Origami Folding and Foldability

(a) Kokotsakis mesh with n = 4, with sector angles i , i , i , i at vertex Vi .


The polygons need not be planar, i.e. it can be non-developable.

(b) I - plane-symmetric.

(c) II - translational.

(d) IV - orthogonal.

(e) V - line-symmetric.

Figure 2.12: For a Kokotsakis mesh with n = 4 (a), there exist 5 known classes
that ensure continuous flexibility (b-e). Images from Stachel (2010b).

27

2.2 Origami Folding and Foldability

continuously flexible. It is noted that the traditional Miura-ori pattern is of types


II, III and IV simultaneously, providing its redundant flexibility.
Concluding from the analysis of continuously flexible Kokotsakis meshes, future
extensions of generalised rigid-foldable degree-4 meshes will have to consider conditions around facets, rather than on individual vertices to establish finite flexibility.
The restriction to flat-foldable meshes can then be relinquished, opening up the
possibility of new types of rigid-foldable origami patterns.

Topological Extensions
In the previous discussion of rigid foldability, it has been assumed that the quadrilateral mesh is homeomorphic to a disk. Several of the concepts, however, have
also been extended to cylindrical surfaces in order to design rigid-foldable tubes
and architectural coverings (Tachi, 2009b, 2010b).

2.2.3

Fold Pattern Design

Developments in the mathematics of rigid foldability, as well as the emergence of


computational design tools have enabled great freedom in the design of origami
patterns. For architectural applications the ability to design a complex freeform
surface geometry is desirable, but the increased complexity comes at the cost
of difficulty of manufacture. Most technical applications such as folded cores
therefore use relatively simple fold patterns, as the manufacturing processes rely
on the repetition of a small set of operations. Klett & Drechsler (2009) further
note that the relative simplicity of most technically relevant fold patterns is also a
consequence of optimised structural and functional efficiency. In general, multiple
overlapping folds and material layers do not add much performance to the resulting
folded core but result in higher mass and material consumption. We here briefly
describe freeform origami design methods, as well as tools to design row-tessellated
fold patterns for engineering applications.

Freeform Surfaces The minimal requirement for the design of origami fold
patterns is developability, i.e. the sector angles at each vertex add up to 2.
Tachi (2009a) describes a freeform origami design methodology, where a partlyfolded sheet can be interactively modified by moving individual vertices, whilst
preserving the necessary developability constraints. Further constraints may be
added, such as local flat foldability and global symmetries, as well as the ability for
quadrilaterals to collapse into triangular facets when nodes merge (Tachi, 2010a).
Kling (2007b) developed an Aspect-Shaping-Floating algorithm, where an initial

28

2.2 Origami Folding and Foldability

Figure 2.13: Overview of available modifications of a Miura-ori unit cell. Possibilities include transverse and longitudinal curvature (mod 1 and mod 6), increased
bonding area (mod 5) and an increase in local density (mod 4). Image from
Khaliulin (2005).
partly-folded surface is transformed to a desired global geometry, before solving
the developability constraint for each of the nodes by floating on the surface.

Modular Unit Cells For many structural applications of folded sheets, modular unit cells are tessellated in orthogonal rows and columns. By tailoring the local
geometry of the unit cells, a wide range of mechanical properties can be achieved.
Khaliulin (2005) describes the synthesis of folded cores by modifying existing rowarranged patterns. The modifications are parametric (modifying sector angles at
the vertices), structural (adding additional facets and creases), or a hybrid of both;
see Figure 2.13. The available modifications provide the ability to introduce global
(double) curvature, local increase in core density, and increased contact area with
face sheets. This synthesis method offers a large scope for pattern design, without
straying from simple repetitive unit cells that can be manufactured continuously.
A computational design methodology for Doubly Periodic Folded (DPF) surfaces is
proposed by Kling (2005). The method relies on prescribing desired row/column
cross sections, with intersection points, by which the interlying surface can be
constructed; see Figure 2.14. The resulting DPF surfaces are rigid foldable, as a
result of the underlying mathematics described in Kling (1997). The local isometry
used to design the doubly periodic sheets can be regarded as reflection of the rows
in mirroring planes, which thereby implicity introduces the local flat-foldability
conditions around the degree-4 vertices.

29

2.2 Origami Folding and Foldability

Figure 2.14: A Doubly Periodic Folded surface can be constructed by specifying


a row cross section and a column reflection scheme. Image from Kling (2007b).

Figure 2.15: A modified Miura fold pattern which is not flat foldable, and thereby
produces a rigid cellular core material (McKay, 1984).
Self-Locking Patterns By purposely introducing non-flat-foldable vertices to
a fold pattern, a folded structure can be designed to lock into a prescribed configuration due to facet-to-facet contact. For example, Akishev et al. (2010) add
self-locking to the helical fold core pattern to fix the configuration and provide flat
bonding areas between adjacent helical curves; see Figure 2.4(d). McKay (1984)
describes folded core material which self-locks, and thereby introduces vertical
separating walls which provide additional compressive strength; see Figure 2.15.
Lastly, self-locking patterns are of great interest when considering self-assembly of
folded sheets, for example for MEMS applications.

30

2.3 Conclusions

2.3

Conclusions

In textured shell structures, a local texture pattern (e.g. corrugations, dimples,


folds) is introduced to a thin-walled structure, in order to modify its global mechanical properties. Existing applications either exploit the increased in-plane
flexibility (e.g. to accommodate large thermal strains) or the increased out-ofplane bending stiffness (e.g. corrugated sheets, or folded sandwich panel cores)
provided by the texture patterns.
The folded shell structures described in this thesis exploit the anisotropy of deformation modes introduced by the fold pattern. The distinct fold lines increase
the shells second moment of area, whilst simultaneously providing compliant deformation modes. The compliance is driven by the hierarchical interaction of
mechanistic articulation around the fold lines, in combination with the inextensional deformations of the thin-walled shells. The mechanics of these folded shell
structures has not previously been studied, and forms the topic of Chapters 4
and 5.
A review of origami mathematics provides a background for the study of folded
shell structures. In rigid origami a fold pattern is modelled as rigid facets connected by frictionless hinges. For tessellated fold patterns, such as folded shell
structures, the resulting mechanism rapidly becomes overconstrained and a folding motion is only possible under specific geometric conditions. Recent work in
origami mathematics has revealed several rigid-foldability conditions. In the case
of the folded shell structures, this enables the design of soft kinematic modes where
the shells deform through articulation around the fold lines. It also has important
consequences for the manufacture of folded sheets, as described in Chapter 6.

31

Chapter 3

Folded Shell Structures


3.1

Description

The distinguishing feature of the folded shell structures described in this thesis,
is the presence of distinct fold lines in the shell surface, which provides the sheet
with an inherent degree of flexibility. Globally the folded shell structures can be
regarded as thin-walled shells, built up of a tessellation of unit cells, which are in
turn composed of thin-walled facets joined at the distinct fold lines.
This description covers a wide range of possible configurations, and we shall limit
ourselves to a subset of the folded shell structures. The folded sheets need not
necessarily be developable, and may have points or lines of non-zero Gaussian
curvature. The fold lines, and thereby the facets, may in general be curved, but
here we only consider straight fold lines and planar facets. Furthermore, the
sheets considered here consist of regular tilings of degree-4 vertices, where four
fold lines meet. If constructed of rigid panels with hinges, these sheets would form
a mechanism with a single degree of freedom (DOF). Higher-order vertices would
provide additional flexibility in the sheet, but thereby also reduce its load-carrying
capacity. Lastly, although global curvatures can be introduced by modifying the
tessellation pattern, we only consider initially planar folded sheets.

3.1.1

Example Sheets

Two representative examples of folded shell structures will be used throughout


this thesis: the Miura and Eggbox sheet, shown in Figure 3.1. Both sheets consist of a regular tessellation of identical parallelogram facets, but the Miura sheet
is developable and can therefore be folded from flat sheet material, whereas the
Eggbox sheet consists of alternating apices and saddle points with equal and opposite angular defect; see Figure 3.2. These two patterns provide representative
examples of both developable and non-developable folded shell structures.
The Miura sheet is named after Koryo Miura who first introduced this fold pattern to engineering applications, and it has remained the most commonly studied

32

3.1 Description

(a) overview of Miura sheet.

(b) overview of Eggbox sheet.

(c) close-up of unit cells

Figure 3.1: photographs of (a) the Miura, and (b) the Eggbox sheet. The models
are made of standard printing paper, and the parallelograms in both sheets have
sides of 15mm and an acute angle of 60 . The Miura sheet is developable, whereas
the Eggbox sheet has an (equal and opposite) angular defect at its apices and
saddle points.

33

3.2 Mechanical Properties

Figure 3.2: The Miura sheet is folded from a single flat sheet of paper (left); in
contrast, the Eggbox sheet (right) is made by joining individual strips of paper,
which introduces the angular defects at the vertices.
fold pattern. It can be straightforwardly modified to create sheets with a global
(double) curvature, varying densities and tapering fold depths, and helical configurations (e.g., Klett & Drechsler, 2009; Akishev et al., 2010; Talakov, 2010); some
examples of modified Miura patterns are shown in Figure 3.3. In the literature
the pattern goes by many monikers, such as chevron or herringbone pattern,
Z-crimp and zigzag corrugation. The Eggbox pattern was simply named after
its resemblance to boxes used for the storage of eggs. The pattern is much less
common in engineering literature, although it has previously been proposed as an
expansible sheet material (Brunner, 1968), and a deployable structure for architectural applications (Tachi, 2010b).

3.2

Mechanical Properties

The first interesting property of the folded sheets is their ability to undergo relatively large deformations, by virtue of the folds opening and closing. Moreover,
the fold patterns enable the sheets to locally expand and contract and thereby
change their global Gaussian curvature without any stretching at material level.
Our interest lies with the macroscopic behaviour of the sheets, and we therefore
consider the global Gaussian curvature of an equivalent mid-surface of the folded
sheet. Both the Eggbox and Miura sheets are initially flat, and thus have a zero
global Gaussian curvature. Now, unlike conventional sheets, both folded sheets can
easily be twisted into a saddle-shaped configuration which has a globally negative
Gaussian curvature see Figure 3.4(a) and Figure 3.5(a).
The sheets most intriguing property, however, relates to their Poissons ratio.
Both sheets have a single in-plane mechanism whereby the facets do not bend and
the folds behave as hinges; by contrast, facet bending is necessary for the out-ofplane deformations. As shown in Figure 3.4(b) and Figure 3.5(b), the Eggbox and

34

3.2 Mechanical Properties

Figure 3.3: A selection of folded Miura geometries. Left to right: angled sheet;
continuously transversely curved; planar sheet; stacked set of sheets that expand
in a coordinated manner.
the Miura sheet respectively have a positive and a negative Poissons ratio in their
planar deformation mode. A negative Poissons ratio is fairly uncommon, but
can for instance be found in foams with a reentrant microstructure (Lakes, 1987),
chiral honeycomb lattices (Prall & Lakes, 1997) and materials with hinged rotating
units (Grima et al., 2005). Conventionally, materials with a positive Poissons ratio
will deform anticlastically under bending (i.e., into a saddle-shape) and materials
with a negative Poissons ratio will deform synclastically into a spherical shape. As
illustrated in Figure 3.4(c) and Figure 3.5(c), however, both folded textured sheets
behave exactly opposite to what is conventionally expected, and their Poissons
ratio is of opposite sign for in-plane stretching and out-of-plane bending. This
remarkable mechanical behaviour has only been described theoretically for auxetic
composite laminates (Lim, 2007) and specially machined chiral auxetics (Alderson
et al., 2010), but is here observed in folded sheets made of conventional materials.

35

3.2 Mechanical Properties

(a)

(b)

(c)

Figure 3.4: mechanical behaviour of the Miura sheet; it can be twisted into a
saddle-shaped configuration with a negative global Gaussian curvature (a). Secondly, the Miura sheet behaves as an auxetic material (negative Poissons ratio)
in planar deformation (b), but it assumes a saddle-shaped configuration under
bending (c), which is typical behaviour for materials with a positive Poissons
ratio.

36

3.2 Mechanical Properties

(a)

(b)

(c)

Figure 3.5: mechanical behaviour of the Eggbox sheet. Firstly, it can change
its global Gaussian curvature by twisting into a saddle-shaped configuration (a).
Secondly, the Eggbox sheet displays a positive Poissons ratio under extension (b),
but deforms either into a cylindrical or a spherical shape under bending (c). The
spherical shape is conventionally seen in materials with a negative Poissons ratio.

37

3.2 Mechanical Properties

Demonstrative Experiments
To demonstrate the unexpected and contrasting bending properties of the Miura
and Eggbox sheets, simple three-point bending tests were carried out; Figures 3.7
3.11. The Miura and Eggbox sheets consisted of 5 5 unit cells, composed of
parallelograms with sides of 15mm and an acute angle of 60 . The Miura sheet
was folded from standard 80gr printing paper; the Eggbox sheet was constructed
by glueing together two orthogonal sets of strips of parallelograms of the same
paper. The sheets were simply supported, and the top vertex of the central unit
cell was displaced downwards by 10mm from its rest configuration; see Figure 3.6.
The coordinates of the top vertices were subsequently measured using a threeaxis coordinate measuring machine, to within 0.5mm. In order to characterise the
out-of-plane bending, the curvature 1 along, and 2 transversely to, the bending
line were calculated by fitting a circle to the measured points. Of main interest
is the ratio 1 /2 , which describes the coupling between the curvatures along the
principal axes. The results are given in Figures 3.7(c)3.11(c).
The experiments affirmed the remarkable bending properties observed in the Miura
and Eggbox sheets, under minimal loading and boundary conditions. Several
observations are of interest. For the Miura sheet, the deformed configuration
differed when bending along or transversely to the corrugations, whereas one would
expect the ratio of the curvatures to be inverted for the two orientations. This
can be attributed to the difficulty of imposing appropriate boundary conditions,
and this simple experiment will therefore not excite a precise bending mode, but
rather a combination of bending and stretching. Secondly, the symmetry of the
Eggbox sheet would suggest a spherical deformation mode with equal principal
curvatures, when bending along the principal axes; this is not the case, and can
be attributed to an induced in-plane strain along the bending axis. Furthermore,
unlike the Miura sheet, when bending at 45 we do not obtain a twisting mode,
but rather a cylindrical deformation mode, as in this configuration uninterrupted
bending lines extend across the sheet.

38

3.2 Mechanical Properties

d = 10mm

Figure 3.6: Experimental set-up for the three-point bending of the folded sheet.
The sheet is simply supported on two blocks, and the top node of the central unit
cell is displaced downwards by 10mm from its rest configuration.

39

3.2 Mechanical Properties

(a) overview

(b) side views

(c) measured nodal coordinates, with circles of curvature.

Figure 3.7: Three-point bending experiment of the Miura sheet, longitudinally


along the corrugations. The ratio 1 /2 2.4, with therefore a relatively weak
coupling between the two bending axes.

40

3.2 Mechanical Properties

(a) overview

(b) side view

(c) measured nodal coordinates, with circles of curvature.

Figure 3.8: Three-point bending experiment of the Miura sheet, transversely


across the corrugations. The fitted circles of curvature provide the ratio 1 /2
1, indicating a strong coupling between the two axes.

41

3.2 Mechanical Properties

(a) overview

(b) side views

(c) measured nodal coordinates, with circles of curvature.

Figure 3.9: Three-point bending experiment of the Miura sheet, with bending axis
at 45 to the corrugations; this is effectively identical to the twisting mode. The
fitted circles of curvature showed that the ratio of 1 /2 1.3.

42

3.2 Mechanical Properties

(a) overview

(b) side views

(c) measured nodal coordinates, with circles of curvature.

Figure 3.10: Three-point bending experiment of the Eggbox sheet, into the spherical deformation mode. The ratio 1 /2 1.6, which can be attributed to compression along the bending line.

43

3.2 Mechanical Properties

(a) overview

(b) side views

(c) measured nodal coordinates, with circles of curvature.

Figure 3.11: Three-point bending experiment of the Eggbox sheet, along the
double-corrugation. The ratio of 1 /2 6.5, and the deformed configuration
is therefore almost cylindrical.

44

3.3 Structural Analysis

3.3

Structural Analysis

The folded shell structures can be regarded as a type of compliant structure, where
the dominant deformation modes can be considered as mechanisms with a nonzero stiffness. Their mechanical properties therefore straddle that of mechanisms
and structures, and a suitable modelling method must capture this behaviour.
For the numerical analysis of folded sandwich panel cores, non-linear Finite Element Analysis is used, as it provides the necessary detailed modelling of local
buckling and crushing (e.g., Heimbs, 2009; Baranger et al., 2011a). Resch &
Christiansen (1971) used a simple folded plate finite element model to analyse
both the kinematics and stiffness of a triangulated folded sheet. An analytical approach was employed by Norman (2009) for (curved) corrugated sheets, where the
global behaviour was described using an equivalent mid-surface. For the analysis
of compliant shell mechanisms which are closely related to the folded sheets
discussed in this thesis Seffen (2011) places a stronger emphasis on the hierarchical nature of these types of structures, and the proposed modelling method
describes the global geometry in terms of unit cell kinematics. This is also the
approach taken in this thesis, as it best elucidates the interrelationship between
the various mechanical scales of the folded shell structures. In our analysis the
sheet material of the folded shell structures is assumed to have zero thickness.
This implies that fold lines are infinitely sharp, whereas in reality there must be
a finite (but large) curvature. Furthermore, the stiffness analysis of the fold lines
must necessarily be simplified, as both the curvature and material thickness are
singular. These simplifications, however, remove many of the clouding details to
gain a conceptual understanding of the kinematics of the folded shell structures.
First the kinematics of the unit cells are described in Chapter 4, followed by a
more holistic stiffness matrix approach in Chapter 5 where the sheets are given a
simplified material model. Both approaches provide complementary insights into
the deformations of the sheets.

Moving Vertices Some deformations observed in experimental models challenge the underlying assumption for the mechanical models used thus far: the fold
pattern changes during the sheets deformation. Figure 3.12 shows a model of a
planar Miura sheet manufactured by vacuum forming High Impact Polystyrene
(HIPS) onto a mould. When bending the sheet, it can attain a single (transverse) curvature; this is made possible by virtue of the vertices moving plastically
through the material, thereby modifying the fold pattern. The shifting of fold vertices through the material has previously been employed for the energy dissipation
in automobile crash boxes (Ma & You, 2011), but not for its ability to introduce
additional flexibility to fold patterns.

45

3.3 Structural Analysis

Figure 3.12: A model of a planar Miura sheet [left], made by vacuum forming High
Impact Polystyrene (HIPS). When bending, the fold pattern changes slightly as
the vertices move plastically through the material, enabling the sheet to attain a
single curvature [right].

46

Chapter 4

Kinematic Analysis
4.1

Planar Kinematics

The planar kinematics of the folded textured sheets can be described algebraically,
by assuming that the facets bounded by the fold lines remain rigid and the folds
behave as frictionless hinges. This analysis provides interesting insight into the
values of the instantaneous Poissons ratio for planar deformations.

Mobility A single degree-4 vertex (whereby 4 fold lines meet), by analogy with a
spherical 4-bar linkage, has a single degree of freedom (Streinu & Whiteley, 2005).
The challenge lies in the finite flexibility of a tessellated pattern of degree-4 vertices, such as the Miura-ori and the Eggbox pattern. Using geometric arguments,
Stachel (2009) shows the 1 DOF flexibility of both these patterns. Stachel (2010b)
describes the continuous flexibility of more general plane tessellations of degree4 vertices, known as a Kokotsakis mesh, and Schief et al. (2007) introduce the
general case of flexibility of a planar quadrilateral mesh surface, a discrete Voss
surface, which includes the Eggbox sheet. As a result, for both example folded
textured sheets a single parameter suffices to describe its current configuration.

Poissons Ratio When studying the planar kinematics of the sheets, an important parameter is its Poissons ratio, or expansion coefficient. This describes the
amount the structure expands or contracts orthogonally to an applied uni-axial
strain. In classical continuum mechanics the Poissons ratio is a material property, and for isotropic materials it is identical regardless of the direction of the
applied strain. In our case, however, the Poissons ratio is primarily a kinematic
property and will vary considerably with the direction of the applied strain. Furthermore, due to the local structure of the sheets the Poissons ratio will also be
strain-dependent and non-linear, and we shall therefore make use of the tangential
Poissons ratio or the Poisson function (Smith et al., 1999), where the calculated

47

4.1 Planar Kinematics

strains are the instantaneous true (or Hencky) strains


i =

xi
xi

(4.1)

with xi the length in the i-direction. The instantaneous Poissons ratio is now
given by
1
ij = ji
=

j
i

(4.2)

where i is the applied strain, and j the resulting strain orthogonal to the applied
direction. This approach to the Poissons ratio is commonly used in literature on
auxetic materials (Grima et al., 2005) and cellular solids (Gibson & Ashby, 1999),
where the microstructure plays an important role in the internal kinematics. See
Smith et al. (1999) for an overview of possible definitions of the Poissons ratio,
and the difficulties in measuring its value in experiments.
Writing the width of the unit cell in the two orthogonal directions Wi (), Wj ()
as a function of a single path length parameter , we can write the strains as:
i =

Wi
1 dWi
=

Wi
Wi d

(4.3)

and therefore the tangential Poissons ratio, or expansion coefficient, as


ij =

j
Wi dWj
=
i
Wj dWi

(4.4)

The choice of a convenient set of orthogonal axes for defining the Poissons ratio
depends on the geometry of the unit cell. Calculating the Poissons ratio at an
angle to those chosen axes requires the effective Youngs modulus along the initial
axes (Grima et al., 2005; Gibson & Ashby, 1999) to enable the coordinate transformation of the resulting stiffness tensor (Nye, 1957). As our focus is on the unit
cell kinematics, this is currently beyond the scope of the analysis.

4.1.1

Miura Sheet

For the planar Miura sheet, the unit cell can be defined by the side lengths a and
b, and the acute angle of the parallelogram elements see Figure 4.1. The
dihedral fold angle between the facets and the mid-plane determines the current
configuration. The height H, the width 2S and length 2L of the unit cell are
calculated as follows:
H = a sin sin
S =b p
L=a

cos tan
1 + cos2 tan2

(4.5)
,

(4.6)

1 sin2 sin2

(4.7)

48

4.1 Planar Kinematics

a
H

2S

2L
x

Figure 4.1: Geometric parameters of a Miura unit cell. The values a, b and
describe the shape of the parallelogram, whereas the angle determines the
current configuration of the unit cell (=0, 55, 85 are shown). Alternatively, the
unit cell can be described by H, S, L and V .
Other parameterisations for the Miura pattern are common in literature (Klett
& Drechsler, 2010; Ichikawa, 1995) and may provide additional insights. When
specifying the geometry directly by the unit cell dimensions H, S, V , L, the final
folded sheet configuration can straightforwardly be converted to the fold pattern
parameters (see Figure 4.2):
p
L0 = H 2 + L2 = a
(4.8)
r
H 2V 2
S0 =
+ S 2 = b sin
(4.9)
H 2 + L2
LV
V0 =
= b cos
(4.10)
H 2 + L2
This enables the design of fold patterns, based on the desired mechanical properties
and geometry of the final folded sheet. Further useful parameters are
tan = cos tan

(4.11)

sin = sin sin

(4.12)

and the fold angle is given as


sin = sin / sin

(4.13)

49

4.1 Planar Kinematics

2S0

V0

2L0

Figure 4.2: Dimensions of the unfolded Miura unit cell. These can be backcalculated from the desired geometry of the unit cell.
Poissons Ratio Using Equation 4.4, with the width S and length L of the
Miura unit cell given by Equations 4.6 and 4.7, the Poissons ratio is given as
SL =

L
S dL
=
S
L dS

(4.14)

which eventually reduces to


SL = cos2 tan2

(4.15)

The results are plotted in Figure 4.3 and 4.4. Using Equation 4.11, the Poissons
ratio can be rewritten as
SL = tan2

(4.16)

which implies the Poissons ratio is only a function of the angle in the xy-plane.
This relationship has important implications for some of the manufacturing techniques discussed in Chapter 6, where flat sheet material is folded using an accompanying non-planar master sheet; at any given configuration both sheets will share
the same , S and L, and therefore share the same kinematics. Another interesting application may be found in deployable structures; for example, a deployable
sandwich structure, where two face sheets are placed on either side of the folding
core, as shown in Figure 4.5.

50

4.1 Planar Kinematics

[deg]
90
0

80

70

60

50

40

30

20

10

80

90

1
2

Poissons ratio SL []

3
4
5
y
6
7

9
10

(a) Poissons ratio versus angle .


0
0.5
45

50

Poissons ratio SL []

1.5

55

60
2.5

65
70

75

3.5
4
4.5
5

10

20

30

40

[deg]

50

60

70

(b) Poissons ratio versus for various values of ; the curve for =
60 is emboldened. The arrows indicate the primary strain direction
S.

Figure 4.3: Analysis of the instantaneous Poissons ratio SL of a Miura sheet,


which is given as SL = cos2 tan2 = tan2 . For the geometry used in our
calculations, = 60 , this means that the sheet has a Poissons ratio of 1 at
55 .

51

4.1 Planar Kinematics

0.6

0.4
=90

L []

0.2

=55
1

0.8

0.6

0.4

0.2

0.2

0.4

0.2

0.4

=0

0.6

S []

Figure 4.4: The relationship of the engineering strain of a Miura sheet with =
60 , with respect to the reference configuration = 55 . The slope of the curve
provides an alternative formulation of the Poissons ratio SL ; the primary strain
direction S is indicated by the arrows.

52

4.1 Planar Kinematics

Figure 4.5: Deployable sandwich panel based on the planar Miura pattern. The
core and the face sheets share the same planar kinematics, and the combination
is therefore deployable.

Figure 4.6: The individual Miura sheets [bottom] can be stacked, using an intermediate folded layer with modified geometry [middle], whilst maintaining their
in-plane kinematics. Multiple layers of the Miura sheets can be stacked to create
a micro-structured material with highly anisotropic properties [top].

53

4.1 Planar Kinematics

y
H

2W

2B

Figure 4.7: Geometric parameters of an Eggbox unit cell. The values a, b and
describe the shape of the parallelogram, whereas the angle 0 determines
the current configuration of the unit cell (=10, 30, 45, 50, 58 are shown).

4.1.2

Eggbox Sheet

The Eggbox sheets consists of a repeated pattern of identical parallelograms, which


meet at points of positive and negative Gaussian curvature (respectively where
four acute, and four obtuse angles of the parallelogram meet). A unit cell is
shown in Figure 4.7. The unit cell geometry is described by the acute angle
of the parallelograms, and their side lengths a and b; the current configuration
is given by the angle at the apex. The angle 2 lies between the two edges of
length a and the angle 2 lies between the two edges of length b. It holds that
0 , . The three angles are related as follows:
cos cos = cos

(4.17)

54

4.1 Planar Kinematics

The height H, width W and depth B of the Eggbox unit cell are given by
H = a cos + b cos

(4.18)

W = a sin

(4.19)

B = b sin

(4.20)

Another aspect of interest is the dihedral fold angles of the sheet. The dihedral
fold angles and , respectively along edges with length b and a, are given by


sin
(4.21)
= 2 arcsin
sin


sin
= 2 arcsin
(4.22)
sin
Poissons Ratio

The instantaneous strain in W -direction can be described by

W
1 dW
=

W
W d
and analogously for B . The tangential Poissons ratio is now given as:
W =

W B =

B
W dB
=
w
B dW

(4.23)

(4.24)

Performing the derivation, the following expression is obtained


W B =

cos2 tan2
cos2 cos2

(4.25)

Note that the expansion coefficient is only a function of and , and not of a or
b. The results for the Eggbox sheet with = 60 is shown in Figure 4.8.
Neutral Configuration The paper models of the Eggbox sheet described in
Chapter 3 are manufactured by glueing together two orthogonal sets of strips of
parallelograms; alternatively it could be made by weaving the strips. As a result
of this manufacturing process every facet has double paper thickness, and along
every fold line one layer of paper is bent. This means the final Eggbox sheet will
be prestressed, and there is therefore a neutral configuration which minimises the
total strain energy. Thus every Eggbox pattern has a default Poissons ratio. The
folds are modelled as simple torsion springs, with torsional stiffness K per unit
length. The total fold energy can then be written as

1
(4.26)
U = K a 2 + b 2
2
and subsequently rewritten in non-dimensional form as
= 2U = 2 + 1 2
U
(4.27)
Ka
r
with r = a/b. Now, for every configuration and r, the minimum fold energy and
corresponding minimizing can be found; see Figure 4.9.

55

4.1 Planar Kinematics

5
4.5
4

Poissons ratio WB []

3.5
3

45
50
55

2.5

60

65

1.5

70

75

0.5
0

10

20

30

[deg]

40

50

60

(a) instantaneous Poissons ratio W B for various values of ; the


curve for = 60 is emboldened.
==60

0.8

0.6

B []

0.4

0.2

0
1

0.8

0.6

0.4

=45

0.2

0.2

0.4

0.2

=0

0.4

w []

(b) relationship of engineering strain with respect to reference configuration of = 45 . The slope provides an alternative formulation of
the sheets Poissons ratio.

Figure 4.8: Analysis of the in-plane properties of an Eggbox sheet with = 60 ,


using (a) the true strain and (b) engineering strain. The arrows indicate the
primary strain direction W .

56

4.1 Planar Kinematics

90
80
70

[deg]

60
50
40
30
20
10
0

10

20

30

40

50

[deg]

60

70

80

90

Figure 4.9: The fold energy contours are plotted for Eggbox unit cells, with the
black line indicating the least energy configurations for a given geometry of r =
a/b = 1 and varying facet angle ; the shaded area indicates an infeasible geometry,
where > . For the example Eggbox sheets which have r = 1 and = 60 , the
minimum energy confirms the observed default configuration of = 45 (and
therefore = 45 ).

57

4.2 Out-of-Plane Kinematics

4.2

Out-of-Plane Kinematics

A crucial point of consideration for the out-of-plane kinematics of the folded textured sheets is that the quadrilateral facets of the unit cells do not remain planar
during the global deformation. In fact, the assumed global shapes are generally
attained through a combination of bending of the facets and along the fold lines of
the unit cells. A careful consideration of the deformation of the unit cells during
the observed global deformation modes revealed a number of subtleties; these will
be outlined before describing the unit cell deformations of the Miura and Eggbox
sheets, under small-displacement and developability assumptions.

Developability Considerations
At unit cell level the folded sheets can be regarded as thin-walled shells, joined
along distinct fold lines. The deformation of thin shells has been extensively
studied in the past (e.g., Timoshenko, 1964; Calladine, 1983). In general, the
deformation of thin-walled shells is composed of bending and stretching components, and their interrelationship is described by the Von Karman equations.
These are non-linear, coupled fourth-order differential equations, and are usually
only solved under simplifying assumptions (Mansfield, 1989; Witten, 2007). For
very thin shells, the bending behaviour dominates and in the asymptotic limit
the deformations can be described as developable (i.e. only bending). Crucially,
developable deformations preserve the Gaussian curvature at every point on the
surface (Hilbert & Cohn-Vossen, 1952).
For the Miura and Eggbox sheet, the facets of the unit cell are initially planar
and thus have zero Gaussian curvature. Under developable deformations this
Gaussian curvature is preserved, and consequently at any point on the deformed
facet surface at least one straight line must extend to the boundary. It is therefore
the boundary conditions that determine the feasible deformed geometries. In the
folded sheets the quadrilateral facets of the unit cells are bounded by straight
fold lines, and the facets must thus remain piecewise planar and no cylindrical or
conical deformations are possible. Demaine et al. (2009) confirmed that straight
fold lines in an origami pattern indeed remain (piecewise) straight after folding.
The assumed developable deformation behaviour at unit cell level therefore implies
that the parallelogram facets of the Miura and Eggbox sheet bend along a new
diagonal crease line; see Figure 4.10. The orientation of the triangulation is irrelevant when only considering first-order displacements (either provides out-of-plane
deformation of the facet), but for large displacements the choice of triangulation
determines the possible kinematics. Visual inspection of the deformed Eggbox
and Miura sheets suggests that the facets bend along the minor axis, and this will

58

4.2 Out-of-Plane Kinematics

O(2)

O(2)

(a)

(b)

Figure 4.10: (a) the Eggbox and Miura unit cells are composed of parallelogram
facets, which must bend to attain the global out-of-plane kinematics of the sheet.
When the bending deformation of the parallelogram facets is developable, it will
bend around a new diagonal crease line. (b) To first order approximation the
orientation of the bending line is irrelevant, as all other nodal displacements are
of higher order. However, for large displacements and stiffness analysis, the choice
of the major [top] or minor [bottom] bending axis is of importance.
be presumed throughout the kinematic modelling. As a result of the additional
fold lines the number of folds meeting at the vertices will increase, and it is these
higher order vertices that provide the additional flexibility beyond the 1 DOF
planar mode.

Observed Deviations When studying the deformation of the example folded


textured sheets and similar hierarchical shell structures with developable surfaces
joined at distinct fold lines (e.g., Norman, 2009; Seffen, 2011), several small deviations from the described developable deformations can be observed.
Firstly, upon close inspection the facets of the deformed Miura and Eggbox sheets
appear to have a subtle double curvature along the introduced bending line of the
facet. This behaviour was described by Lobkovsky et al. (1995) for a kite deformation (see Figure 4.11). Although the exact shape of the deformed configuration
was not explicity derived, a simplified calculation showed that the sagging along
the ridge line minimises the combined bending and stretching energy, which introduces a slight negative curvature along the ridge. When assuming the boundary
conditions (the bounding fold lines) do not change, the Gauss-Bonnet theorem
further implies that the deformed sheet contains areas of positive curvature, with

59

4.2 Out-of-Plane Kinematics

d
R

Figure 4.11: For thin sheets, the bending deformation of the facets will not exactly
conform to pure developability. This kite deformation illustrates the doublecurvature introduced due to sagging along the ridge, as observed in the Eggbox
and Miura models; figure after Witten (2007).
the integral summing to zero (Struik, 1961; Witten, 2007). When the boundary
conditions are more flexible, the ridge line may possibly shorten, but the effect
is certainly of higher order than the vertex displacements due to the bending deformations. Secondly, an important consideration is whether straight fold lines
truly remain straight in the deformed configurations. One example of a folded
sheet which challenges this assumption is the folded hyperbolic paraboloid (hypar), where the fold lines appear to twist (Seffen, 2011); see Figure 4.12. Demaine
et al. (2009) proved that a straightforward triangulation of the pattern would make
the pattern foldable, yet inspection of folded models of the hypar reveals indistinct
bending of the facets, and rather prominent twisting of the facets and associated
helical deformation of the fold lines. The subtleties of this near-developable behaviour are currently not yet fully understood.

Large-Displacement Considerations
The relationship between in-plane strain and change in Gaussian curvature is
given, for small displacements from a flat configuration, by Calladine (1983):
K=

2 xy
2 y
2 x

xy
x2
y 2

(4.28)

More generally these relations can be expressed in terms of curvilinear coordinates to enable the description of the deformation of non-planar surfaces. Crucially, Equation 4.28 shows that for shapes with a constant Gaussian curvature
(e.g. a sphere), the strain would have to grow quadratically with the distance from
the centrepoint. This means that for the observed global deformation modes of
the Miura and Eggbox sheets, despite their apparently straightforward geometry,
the deformation of the unit cells must vary across the surface; see Figure 4.13.

60

4.2 Out-of-Plane Kinematics

Figure 4.12: The folded hyperbolic paraboloid (Demaine et al., 2009) shows clear
twisting of the facets, and corresponding helical deformation of the fold lines. The
facet deformation is therefore not purely developable, and some stretching of the
material is required.

Figure 4.13: When a flat disc is deformed into a shallow spherical shell, the introduced constant positive Gaussian will induce an in-plane strain, which is quadratic
with the distance from the centre. This is here graphically illustrated by drawing geodesics (great arcs) on the spherical surface. As the unit cells tessellate
along the highlighted slices, they must contract transversely (and thereby expand
longitudinally) in-plane. The induced in-plane strain thus requires that, modulo
obvious symmetry considerations, the deformation of the Eggbox unit cell will
have to vary across the spherical surface.

61

4.2 Out-of-Plane Kinematics

Nonetheless, we aim to capture the global sheet geometry in terms of the kinematics of a single unit cell which can be tessellated. In effect, this provides a
first-order approximation of the global deformation modes.

Kinematic Analysis
For the kinematic analysis, the unit cells of the folded sheets are modelled as
a pin-jointed truss with inextensional members; for details see Chapter 5. The
parallelogram facets are triangulated, to enable bending along their minor axis.
The tessellation boundary conditions were prescribed by equating the (change in)
angle between the facets on opposing sides of the unit cell. By capturing the unit
cell deformation within bounding planes, the local deformations can be translated
to the next scale in the structural hierarchy (Seffen, 2011).
The tessellation boundary conditions were naturally added to the mechanical description of the unit cell. The Jacobian of the tessellation constraints was added to
the compatibility matrix of the unit cell. The nullspace of this combined compatibility matrix provides all deformations (i.e. change in angle of the fold lines and
bending of facets) that satisfy both the bar length constraints (i.e. no stretching)
and the tessellation boundary conditions. Any kinematic deformations will be a
linear combination of the vectors in this nullspace. In order to gain insight into this
deformation space, ad hoc boundary conditions were added. For example, adding
a symmetry condition would yield only symmetric deformation modes. By then
orthogonalising the original nullspace with respect to the symmetric deformation
mode using a QR-decomposition (Strang, 2006), the anti-symmetric deformation
modes could be separated. In this manner the nature of the deformation modes
could be studied.
It is important to emphasise that the unit cell deformations described here are not
exhaustive. The range of deformations was restricted by taking into account the
tessellation boundary conditions. In reality, many more unit cell configurations
are observed; in fact, it is those deformations that enable the large displacement
configurations with strain varying across the surface.

62

4.2 Out-of-Plane Kinematics

Figure 4.14: An undeformed Miura unit cell, with its bounding planes [left]. The
tessellation boundary conditions are specified by equating the (change in) angles
between the emboldened lines on opposing sides of the unit cell; symmetry in the
yz-plane [right] helps impose additional boundary conditions.

4.2.1

Miura Sheet

The Miura unit cell is shown in Figure 4.14. The fold lines introduced to enable
bending of the facets create a degree-6 vertex at the centre of each unit cell, and
the bending of facts and folds lines will thus be coupled. The nullspace of the
compatibility matrix contained three independent deformation modes, which were
separated into the following modes:
Planar When imposing reflective symmetry in the yz-plane, the deformation
space reduced to two independent modes. The linear combination that
constrains the bending of the facets produces the planar mode described
analytically in Section 4.1.
Saddle By orthogonalising the subspace of symmetric deformation modes to the
planar mode using a QR-decomposition, the global saddle mode was found;
see Figure 4.15. This deformation mode is obtained by all facets bending equally, and the fold lines adjusting accordingly. The coupling ratio
yy /xx is plotted in Figure 4.16. This saddle-shaped bending mode of
the Miura sheet was previously described by Zakirov et al. (2004), by mapping the unit cell to a hyperbolic surface. The actual deformations of the
unit cell, however, were not elucidated.
Twisting The twisting mode was found by orthogonalising the deformation space
to the planar and saddle deformation modes; see Figure 4.17. This deformation mode is anti-symmetric in the yz-plane, whereby both facets bend
into mountain or valley folds, and the sawtooth folds remain unchanged; the
zigzag folds adjust accordingly.

63

4.2 Out-of-Plane Kinematics

(a) first-order saddle mode. As can be seen in the side views [right], each pair of bounding planes
tilts towards or away from each other, but does not twist. This produces an oppositely signed
curvature for the two orthogonal axes.

+
-

+
-

(b) facet and fold deformations of the saddle mode. The facets either bend into a mountain (M)
or valley (V) fold, and the existing folds either fold further (+) or unfold (-).

Figure 4.15: Saddle deformation mode of the Miura sheet, with (a) the unit cell
with its bounding planes, and (b) the constituent facet and fold deformations.

64

4.2 Out-of-Plane Kinematics

dyy
dxx /2
2L

2S

(a) the curvature along the principal axes is given by xx =


yy =

dyy
2L

dxx
2S

and

, with S and L the dimensions of the undeformed unit cell.

0
0.5
1
1.5

yy /xx

2
2.5
3
3.5
4
4.5
5

10

20

30

40
50
fold angle [deg]

60

70

80

90

(b) the coupling ratio yy /xx plotted with respect to the fold
angle , for a Miura sheet with = 60 , a/b = 1.

Figure 4.16: The ratio of principal curvatures yy /xx for the saddle-shaped
deformation mode of the Miura-ori can be calculated by finding the change in
angle dxx and dyy of the bounding planes. Plotting this coupling ratio with
respect to the fold angle revealed the remarkable result of it being identical to
the in-plane Poissons ratio shown in Figure 4.3(b).

65

4.2 Out-of-Plane Kinematics

(a) first-order twisting mode. As can be seen in the side views [right], each pair of bounding planes
remains parallel (and therefore no bending is involved), but rotates with respect to each other,
producing a pure twist deformation.

=
-

=
+

(b) facet and fold deformations of the twisting mode. The facets either bend into a mountain
(M) or valley (V) fold, and the existing folds either fold further (+), unfold (-) or remain unchanged (=).

Figure 4.17: Twisting deformation mode of the Miura sheet, with (a) the unit cell
with its bounding planes, and (b) the constituent facet and fold deformations.

66

4.2 Out-of-Plane Kinematics

Figure 4.18: An undeformed Eggbox unit cell, with its bounding planes [left]. The
tessellation boundary conditions are specified by equating the (change in) angles
between the emboldened lines on opposing sides of the unit cell.

4.2.2

Eggbox Sheet

For the Eggbox sheet, the introduced facet bending lines create an alternating
array of degree-4 and degree-8 vertices. This means that bending of the facets
and opening/closing of the fold lines is decoupled. The nullspace of the compatibility matrix again contained three independent deformation modes, which were
separated into the following modes:
Planar The combination of tessellation boundary conditions and symmetry in
the yz-plane reduced the configuration space to two independent vectors; the
linear combination that constrained facet deformations provided the planar
mode. This is the only mode that involves bending of the fold lines.
Spherical Orthogonalising the symmetric deformation modes to the planar mechanism using a QR-decomposition, revealed the spherical deformation. Here
all facets deform equally, producing an equally oriented curvature along both
principal axes; see Figure 4.19. The ratio of principal curvatures yy /xx
for a range of unit cell configurations is plotted in Figure 4.20. Seffen (2011)
describes the spherical deformation mode of the Eggbox sheet in terms of
the deformations of its constituent unit cells. There it is suggested that the
assumed unit cell deformations require in-plane deformations of the facets,
in contrast to the results in this thesis.
Twisting The facets deformations of the twisting mode are anti-symmetric in
both the xz- and the yz-plane, and therefore alternatingly form mountain
and valley folds; see Figure 4.21.
Cylindrical A linear combination of the spherical and twisting mode, which fixes
the bending of one diagonally opposing pair of facets, provides the cylindrical
mode; see Figure 4.22.

67

4.2 Out-of-Plane Kinematics

(a) first-order spherical mode. Both pairs of bounding planes tilt either away or towards each
other, and thereby produce an equally signed curvature along both principal axes.

=
=

=
=

=
=

M
=

=
=

(b) facet and fold deformations of the spherical mode. All facets bend into a mountain (M) fold,
and the existing folds remain unchanged (=).

Figure 4.19: Spherical deformation mode of Eggbox sheet, with (a) the unit cell
with its bounding planes, and (b) the constituent facet deformations.

68

4.2 Out-of-Plane Kinematics

5
4.5
4

yy / xx

3.5
3
2.5
2
1.5
1
0.5
0

10

20

30

[deg]

40

50

60

Figure 4.20: Plotting the coupling ratio yy /xx of the Eggbox sheet ( = 60 ,
a/b=1) with respect to the configuration angle revealed the same relationship
as for the in-plane Poissons ratio shown in Figure 4.8.

69

4.2 Out-of-Plane Kinematics

(a) first-order twisting mode. As can be seen in the plan view [top-right], the bounding pairs
remain parallel (and therefore no bending is involved), but as seen in the side views [bottom], they
rotate with respect to each other, producing a pure twist deformation.

=
=

=
=

=
=

V
=

=
=

(b) facet and fold deformations of the twisting mode. The facets alternatingly bend into a mountain
(M) or valley (V) fold, and the existing folds remain unchanged (=).

Figure 4.21: Twisting deformation mode of Eggbox sheet, with (a) the unit cell
with its bounding planes, and (b) the constituent facet deformations.

70

4.2 Out-of-Plane Kinematics

(a) first-order cylindrical mode. When viewing the unit cell at 45 to the reference axes [right], it
can be seen that there is curvature along one axis and none along the other, thereby producing a
global cylindrical mode.

=
=

=
=

=
=

V
=

=
=

(b) facet and fold deformations of the twisting mode. Two facets bend into a valley (V) fold, and
the other facets and folds remain unchanged (=).

Figure 4.22: Cylindrical deformation mode of Eggbox sheet, with (a) the unit cell,
and (b) its constituent facet deformations. The cylindrical mode is constructed by
a linear combination of the twist and spherical mode, whereby one pair of facets
does not bend.

71

4.3 Conclusion & Discussion

4.3

Conclusion & Discussion

The kinematic analysis of the example folded shell structures highlights the hierarchy of their deformation mechanics. The feasible global deformation modes of the
folded shell structures are prescribed by the deformations of their unit cells, which
in turn rely on the constrained deformation of their constituent facets. Understanding the kinematics of the unit cells therefore provides insight into the global
sheet geometries that can be attained.
For both the Eggbox and Miura sheet, the observed global deformation modes were
successfully described in terms of the unit cells, and the developable deformation
of its facets. Constraining the bending of the unit cell facets produces a single
planar mechanism, whereas any out-of-plane deformation mode generally requires
a combination of facet and fold bending.

Planar Deformation The planar kinematics of the Eggbox and Miura sheet
were described algebraically, and their in-plane Poissons ratios were found. A
remaining challenge is to find the Poissons ratio at any angle to the selected axes;
this cannot be described within a purely kinematic analysis, and requires strain
energy considerations. This is illustrated by comparing the Eggbox and Miura
sheets under in-plane loading at 45 to the chosen axes: the first is inextensible,
whereas the second freely expands.

Out-of-Plane Deformation An important feature of the folded shell structures is their ability to easily change their global Gaussian curvature. These
global deformation modes were described to first-order in terms of their unit cell
kinematics, by prescribing tessellation boundary conditions. It was shown that the
globally doubly-curved configurations can be attained through purely developable
deformations at the material level of the sheets. The counterintuitive oppositely
signed Poissons ratio for bending and stretching directly followed from the unit
cell deformations. Intriguingly, for the unit cell geometries studied, the out-ofplane coupling ratio yy /xx is equal and opposite to the in-plane Poissons
ratio yx .
It is important to note that for large global bending deformations, the induced
change in Gaussian curvature necessitates in-plane strain. Therefore, even for
seemingly simple globally doubly-curved configurations, the unit cells deform differently throughout the sheet. The modes described here are limited to first-order,
due to the repetitive boundary conditions.

72

Chapter 5

Numerical Analysis

5.1

Mechanical Model: Bar Framework

The salient behaviour of the folded sheets straddles kinematics and stiffness: there
are dominant mechanisms, but they have a non-zero stiffness. To gain insight into
their mechanical behaviour, a suitable modelling method therefore needs to cover
this behaviour. In our analysis we are not interested in the details of the stress
distributions (Tanizawa & Miura, 1975), but rather the effect of the introduced
geometry on the global properties of the sheet. Furthermore, the out-of-plane
kinematics of the sheets involve bending of the facets, so straightforward rigid
origami modelling of the fold pattern is insufficient (Tachi, 2009c; Balkcom, 2004).
Our approach is based on modelling the partially folded state of a folded pattern
as a pin-jointed truss framework. Each vertex in the folded sheet is represented
by a pin-joint, and every fold line by a bar element. Additionally, the facets are
triangulated to avoid internal mechanisms (in effect, in-plane strain), as well as
provide a first-order approximation to bending of the facets see Figure 5.1.
Although the use of a pin-jointed bar framework to represent origami folding has
been hinted at on several occasions (Tachi, 2009c; Balkcom, 2004), it has not been
fully introduced into the origami literature. The method provides useful insights
into the mechanical properties of a folded textured sheet, and has the benefit of an
established and rich background literature. A similar methodology for calculating
the kinematics and elastic properties of a developable triangulated folded plate
system, the Resch pattern, was described by Resch & Christiansen (1971). In
their finite element formulation, each facet is represented as a triangular plane
stress element, and for the stiffness analysis the fold lines are given an additional
bending stiffness.
An underlying assumption of the mechanical model is that straight fold lines
remain straight in their folded (and deformed) state. In reality some twisting
and other deformations of the fold lines can be observed in folded structures,
as discussed in Section 4.2, but the approximation suffices for our analysis. In

73

5.1 Mechanical Model: Bar Framework

combination with the assumption of developability, the deformation of the facets


will be piecewise planar and therefore the triangulation approximation is a valid
one. For first-order approximation of the out-of-plane displacements of the nodes,
the direction of triangulation of the facets is not relevant; the triangulation in
our analysis is chosen to represent the softest deformation mode, i.e. the shortest
fold line (see Figure 5.1). Lastly, in our analysis the facets are considered to be
thin-walled shells, with the thickness of the sheet material an order of magnitude
smaller than the dimensions of the sheets unit cells.

5.1.1

Governing Equations

The analysis of pin-jointed frameworks is well-established in structural mechanics.


Its mechanical properties can be described by three linearised matrix equations:
equilibrium, compatibility and material properties (Guest, 2006).
At = f

(5.1)

Cd = e

(5.2)

Ge = t

(5.3)

where A is the equilibrium matrix, which relates the internal bar tensions t to the
applied nodal forces f ; the compatibility matrix C relates the nodal displacements
d to the bar extensions e and the material equation introduces the axial bar
stiffnesses along the diagonal of G. It can be shown through a straightforward
virtual work argument that C = AT , the static-kinematic duality.

5.1.2

Kinematics: Compatibility

The linear-elastic behaviour of the truss framework can now be described by


analysing the vector subspaces of the equilibrium and compatiblity matrices (Pellegrino & Calladine, 1986). Of main interest in our case is the nullspace of the
compatibility matrix, Cd = 0, as it provides nodal displacements that to first
order have no bar elongations: internal mechanisms. These mechanisms may
either be finite or infinitesimal, but in general the information from the nullspace
analysis alone does not suffice to establish the difference. First-order infinitesimal
mechanisms can be stabilised by states of self-stress; higher-order compatibility
equations may shed further light on finite mechanisms (Kumar & Pellegrino, 2000),
or a full tangent stiffness matrix can be formulated to take into account any geometric stiffness resulting from reorientation of the members (Guest, 2006).
In the case of the folded textured sheets, the nullspace of the conventional compatibility matrix does not provide much useful information: the facets can easily

74

5.1 Mechanical Model: Bar Framework

bend due to the triangulating fold lines, which is reflected by an equivalent number of internal mechanisms. The solution is to introduce additional contraints.
One interpretation of the compatibility matrix is to consider it as the Jacobian
of the bar length constraints, with respect to the nodal coordinates. This parallel
can be used to introduce additional equality constraints to the bar framework. In
our case we add a constraint on the dihedral angle between two adjoining facets.
The angular constraint F is set up in terms of the dihedral fold angle between
two facets. Using vector analysis, the angle between two facets can be described
in terms of cross and inner products of the nodal coordinates p of the two facets,
F = sin () = sin ( (p)) = . . .

(5.4)

as described in Figure 5.2. The Jacobian then becomes


1 X F
J=
dpi = d
cos ()
pi

(5.5)

The Jacobian of additional constraints J can now be concatenated with the existing
compatibility matrix
"
#
"
#
C
e
d=
(5.6)
J
d
and the nullspace of this set of equations produces the nodal displacements d that
do not extend the bars, as well as not violate the angular constraints. In effect,
we have formulated a rigid origami simulator no bending or stretching of the
facets is allowed. In order to track the motion of the folded sheet, one iteratively
follows the infinitesimal mechanisms whilst correcting for the errors using the
Moore-Penrose pseudo-inverse (Strang, 2006; Tachi, 2009c). Our interest, however,
remains with the first-order infinitesimal displacements.
In the case of the two example textured sheets, the kinematic analysis provides a
single degree of freedom planar mechanism that was previously described analytically in Chapter 4; see Figure 5.3. In this mechanism the facets neither stretch
nor bend, and it is the mechanism a rigid origami simulator would find.

5.1.3

Stiffness: Material Stiffness

A kinematic analysis of a framework, even with additional constraints, can clearly


only provide so much information. The next step is to move from a purely kinematic to a stiffness formulation. Equations 5.15.3 can be combined into a single
equation, relating external applied forces f to nodal displacements d by means of
the material stiffness matrix K.
Kd = f

(5.7)

K = AGC = CT GC

(5.8)

75

5.1 Mechanical Model: Bar Framework

Kfold
Kfacet

Figure 5.1: Unit cell of the Eggbox sheet, illustrating the pin-jointed bar framework model used to model the folded textured sheets. The facets have been
triangulated, to avoid trivial mechanisms and provide a first-order approximation
for the bending of the facets. Bending stiffness has been added to the facets and
fold lines, Kfacet and Kfold respectively.

Figure 5.2: The dihedral fold angle can be expressed in terms of the nodal
coordinates of the two adjoining facets. Using the vectors a, b and c, the following
1
expression holds: sin () = sin()1sin() |a|3 |b||c|
(a (c a)) (a b). Here is the
angle between a and b, and the angle between a and c.

76

5.1 Mechanical Model: Bar Framework

(a)

(b)

Figure 5.3: The Eggbox (a) and Miura (b) sheet both exhibit a single planar
mechanism when the facets are not allowed to bend. The reference configuration
is indicated as dashed lines.
In fact this approach can straightforwardly be extended to other sets of constraints
by extending the compatibility matrix.
"
K=

C
J

#T "

G
0

0
GJ

#"

C
J

#
(5.9)

Depending on the constraint and the resulting error that its Jacobian constitutes,
either a physical stiffness value can be attributed in GJ or a weighted stiffness
indicating the relative importance of the constraint. In our case, the error is the
change in the dihedral angle between adjacent facets. In effect, we introduce a
bending stiffness along the fold line (Kfold ) and across the facets (Kfacet ) see
Figure 5.1. As a result, we obtain a material stiffness matrix that incorporates the
stiffness of the bars, as well as the bending stiffness of the facets and along the fold
lines. A more rigorous substantiation for the ability to extend the tangent stiffness
matrix of trusses with additional constraints is provided by Kovacs (2011), using
the variational approach to mechanics (Lanczos, 1986).
Plotting the mode shapes for the lowest eigenvalues of the material stiffness matrix
K provides insight into the deformation kinematics of the sheets. Of main interest
are the deformation modes that involve no bar elongations (i.e., no stretching of
the material), but only bending of the facets and along fold lines. These modes
are numerically separated by choosing the axial members stiffness of the bars to
be several orders of magnitude larger (say, 106 ) than the bending stiffness for the
facets and folds. In our analysis only first-order infinitesimal modes within K are
considered.

77

5.1 Mechanical Model: Bar Framework

di

6
Facets

Sawtooth

ZigZag

Figure 5.4: The softest eigenmode of a Miura sheet with 9 9 unit cells, is the
global twisting mode. The change in fold angle di of respectively the facets,
the sawtooth and zigzag folds is plotted. As expected, the unit cell deformations
vary across the sheet, but a trend is visible. The central unit cell is plotted
in its deformed configuration, with its change in fold angles highlighted. The
global twisting mode maps to the following unit cell deformations: two convex
and two concave facets, a respective opening and closing of the zigzag folds, and
the sawtooth folds remaining effectively unchanged.

5.1.4

Extensions to Framework Analysis

Further extensions to the bar framework approach provide additional tools to


obtain insight into the sheets kinematics. They are here briefly described.

Coordinate Transformation Currently all properties of the folded sheet are


expressed in terms of the displacements of the nodal coordinates. The change in
fold angles is more convenient when studying the underlying unit cell kinematics;
see Figure 5.4. A transformation matrix T converts nodal displacements d to
changes in angle d:
d = Td

(5.10)

where T is here identical to the Jacobian in Equation 5.5.

Folding Kinematics For rigidly foldable patterns, the presented model can also
be used to calculate the folding motion of the sheet, by following the nullspace
of its Jacobian J, which maintains both the length and planarity constraints. A

78

5.2 Modal Analysis

first displacement step d0 can be projected onto the constraint space using the
Moore-Penrose pseudo-inverse of the Jacobian, indicated with J+ :
d1 = [I J+ J]d0

(5.11)

As the deformation continues, future steps will be selected from the nullspace of J,
or, equivalently the row vector with smallest singular value. After each step, the
error is reduced by means of the Newton-Raphson iteration (Press et al., 2007):
di+1 = J+ r

(5.12)

with the error vector r derived from the original constraint functions F (p) = r. In
effect this produces a rigid origami simulator, although one less computationally
efficient than presented by Tachi (2009c), which has a minimal coordinate representation with fewer explicit constraints and therefore the ability to calculate the
pseudo-inverse without full SVD decomposition.
Large Displacements In the modal analysis, only first-order infinitesimal mechanisms are considered. Of interest in practical applications, is the compliance of
the sheet in some of its deformation modes. To solve the large-displacement,
small-strain behaviour of the sheets, the deformations are found using a numerical
integration scheme in combination with a Newton-Raphson iteration to correct for
errors. It must be noted, however, that the model used in this chapter purposely
provides a significant simplification of the actual deformation mechanics of the
sheets. For detailed analysis of the fold lines and facet bending, a finely meshed
finite element analysis is required.

5.2

Modal Analysis

In the modal analysis of the Eggbox and Miura sheets, the eigenvalues can be
plotted with respect to any combination of the parameters that define the sheets
geometry and configuration. In this analysis we have focused on the stiffness
ratio and the sheet configuration, whilst keeping the geometric parameters of the
unit cells constant ( = 60 , r = a/b = 1); the resulting eigenvalue loci plots
are effectively two-dimenstional slices across a multi-dimensional parameter space
(see Figure 5.9 and 5.10). For the calculations a 4 4 grid of unit cells was used,
as a compromise between large (infinite) sheets, and a single unit cell; we did not
wish to impose a priori repetitive boundary conditions.
Stiffness Ratio
An important parameter in the folded textured sheets turns out to be Kratio =
Kfacet /Kfold . This is a dimensionless parameter that represents the material prop-

79

5.2 Modal Analysis

erties of the sheet. When Kratio we approach a situation where rigid panels
are connected by frictionless hinges; values of Kratio 1 reflect folded sheets
manufactured from sheet materials such as metal, plastic and paper; and when
Kratio < 1 the fold lines are stiffer than the panels, which is the case for workhardened metals or situations where separate panels are joined together, for example by means of welding or architectural fittings.
Figure 5.5 and 5.6 show a log-log plot of the eigenvalues versus the stiffness ratio
Kfacet /Kfold . It can be seen that for both sheets the salient kinematics (the softest
eigenmodes) remain dominant over a large range of the stiffness ratio; this indicates
that the dominant behaviour primarily depends on the geometry, rather than the
exact material properties. Note that for the Eggbox sheet, the spherical, twisting
and cylindrical modes observed in the paper models (see Chapter 3) have a slope
of approximately one on the log-log plot; the eigenvalues are thus proportional
to the Kfacet and no bending of the folds takes place. The stiffness of the planar
mode, on the other hand, is proportional to Kfold as the facets do not bend. For
the Miura sheet, the stiffness of the twisting and saddle modes are not directly
proportional to either Kfacet or Kfold over a long range of the Kratio , and are
therefore a combination of facet and fold bending. As Kratio those modes,
however, change to become approximately proportional to Kfacet .
Sheet Configuration
Figure 5.7 and 5.8 show the normalised eigenvalues of the Miura and Eggbox
sheets with respect to their respective configuration parameter and , at a
fixed Kratio = 2. In effect these plots show the change of the softest out-ofplane deformation modes, with respect to the sheets in-plane deformation. For
the Miura sheet, the saddle and twisting mode remain dominant up to a certain
fold depth, after which a planar squeezing mode takes over. For the Eggbox
sheet, the spheroidal mode remains dominant over the entire range of geometry,
whereas the twisting mode becomes stiffer for low values of and . Due to the
sheets symmetry, at = 45 the eigenvalue loci of the two cylindrical modes
cross. At first glance, the cylindrical modes appear to be a higher mode (cf.
buckling modes). However, in this configuration there is a line of inflection along
the longest, and therefore stiffest, fold line along the diagonal of the Eggbox sheet;
this seemingly higher mode therefore actually minimises its stiffness.

80

5.2 Modal Analysis

Stiffness of eigenmode / Kfold

100

10

0. 1

0. 01
0. 01

0. 1

10

Kfacet/Kfold

100

saddle

twisting

planar

Figure 5.5: This figure shows the relative stiffness of the six softest eigenmodes
of the Miura sheet, with = 60 , a/b = 1 and initial configuration = 55 . The
twisting deformation mode remains the softest eigenmode over a large range of
Kratio , while the saddle-shaped mode is also dominant. As Kratio the planar
mechanism identified in the kinematic analysis becomes the softest eigenmode.

81

5.2 Modal Analysis

Stiffness of eigenmode / K fold

100

10

0.1

0.01

0.1

0.01

Kfacet/Kfold

10

100

cylindrical (2x)

spherical

planar

twisting

Figure 5.6: Here is plotted the relative stiffness of the nine softest eigenmodes of
the Eggbox sheet, with = 60 , a/b = 1 and = 45 . It can be seen that the
twisting deformation mode remains the softest eigenmode over a large range of
Kratio . The spherical and cylindrical deformation modes observed in the models
are also dominant. As Kratio the planar mechanism becomes the softest
eigenmode; this corresponds with the result from the kinematic analysis.

82

5.2 Modal Analysis

Stiffness of eigenmode / Kfold

10

0.1

0. 01

10

20

30

40

50

60

70

80

90

Fold Angle [deg]

saddle
squeeze

twisting

Figure 5.7: The relative stiffness of the six softest eigenmodes of a Miura sheet are
plotted with respect to the initial fold angle , for = 60 , a/b = 1 and Kratio = 2.
The twisting and saddle deformation mode remain dominant over a large range of
the fold angle, but eventually a planar squeezing mode assumes the lowest modal
stiffness.

83

5.2 Modal Analysis

Stiffness of eigenmode / Kfold

10

Fold Angle [deg]

60

59

58

10

20

55

49

45

39

30

40

45

50

60

0.1

0. 01

Fold Angle [deg]

spheroidal
cylindrical (2x)
twisting

Figure 5.8: The nine softest eigenmodes of the Eggbox sheet are here plotted,
versus its unit cell geometry parameter , for = 60 , a/b = 1 and Kratio = 2.
Note that the plot is symmetric with respect to = 45 , as the geometry and
the eigenmodes are rotated by 90 .

84

5.2 Modal Analysis

Stiffness of Eigenmode / K fold

100

10

0.1
100
0.01
0

10
20

1
40

Fold Angle [deg]

Kratio

0.1

60
80

0.01

(a)

Stiffness of Eigenmode / K fold

100

10

0.1
100
0.01
0

10
20

1
40

Fold Angle [deg]

Kratio

0.1

60
80

0.01

(b)

Figure 5.9: Eigenvalue surface of Miura sheet (a), showing the 6 softest eigenmodes with respect to the initial folded configuration and stiffness ratio Kratio =
Kfacet /Kfold . A slice can be made across the surface, for example as shown in (b)
at Kratio = 2.

85

5.2 Modal Analysis

Stiffness of Eigenmode / K fold

100
10
1
0.1
0.01
3

10

100
4

10

10

0
1

20
Fold Angle [deg]

Kratio

0.1

40
60

0.01

(a)

Stiffness of Eigenmode / K fold

100

10

0.1
100
0.01
0

10
1

20
Fold Angle [deg]

Kratio

0.1

40
60

0.01

(b)

Figure 5.10: Eigenvalue surface of the Eggbox sheet (a), showing its 9 softest eigenmodes with respect to its configuration and stiffness ratio Kratio = Kfacet /Kfold .
A slice can be made across the surface, for example as shown in (b) at = 45 .

86

5.2 Modal Analysis

5.2.1

Symmetry Analysis

The symmetry of the Eggbox structure enables us to perform a symmetry analysis


(Kangwai et al., 1999). The sheet belongs to the C4v symmetry group, which has
6 symmetry subgroups: A1 , A2 , B1 , B2 , and the two-dimensional group E. Using standardized symmetry tables, a symmetry adapted coordinate system Vsym
can be set up for the structure, which spans the same coordinate space but is
subdivided into the various symmetry groups. This is effectively a coordinate
transformation, which means that each eigenmode found in the stiffness matrix
resides in a specific symmetry subgroup. This is nicely demonstrated in FigT
ure 5.11, which shows the sparsity pattern of the matrix Vsym
U where U contains
all selected eigenmodes as column vectors. A dot on position {i, j} indicates that
the i-th column of the symmetry space Vsym , and the j-th column of the eigenvectors U are linearly dependent and that the eigenvector therefore lies in that
specific symmetry subgroup.

A1

A2

Components in Symmetry Subspace

50

B1
100

B2

150

E11

200
E22

10
15
Selected Eigenmode

20

25

Figure 5.11: The sparsity pattern shows in which subspace of the C4v symmetry
group each eigenmode of the Eggbox sheet has a component. As can be seen, each
eigenmode belongs to a specific symmetry subgroup.

5.2.2

Curve Veering & Imperfections

During the analysis of the eigenvalues and eigenmodes of the stiffness matrix,
the curve veering phenomenon was encountered: at a point where two eigenvalue

87

5.2 Modal Analysis

loci appear to cross, they instead approach closely and then abruptly veer away,
each taking the path and mode shape of the other. In the veering region the mode
shapes undergo rapid (but continuous) changes, and the system is therefore highly
sensitive to the system parameter. Curve veering is observed in a wide range of
fields where eigenvalue problems exist.
As shown by Perkins & Mote (1986), curve veering is linked to a coupling between the eigenmodes. A weak coupling results in very abrupt veerings with high
changes in curvature of the locus, whereas for strong coupling, the modes are often
sufficiently spaced apart to make the veering difficult to spot. In the case of zero
modal coupling the curves are free to cross; this represents a situation where the
systems eigenvalues are mathematically degenerate. For non-self-adjoint eigenproblems for example in non-conservative dynamic problems one may also
find cases of eigenvalue coalescence (implying eigenvector degeneracy) and veering with in the eigenvalue curves (Perkins & Mote, 1986). Eigenvalue coalesence
is encountered in practical problems, for example when studying the flutter of
rotating blades and brake squeal. Afolabi & Mehmed (1994) cite both mathematical (based on catastrophe theory) and engineering reasons why curve crossings
will very seldom be encountered, and when observed in theoretical studies, may
well have been introduced by simplifications whereby modal coupling is ignored.
Balm`es (1993) identifies only three types of conservative structures as allowing
truly multiple modes: 1) symmetric structures with multiple modes within a symmetry group, 2) multi-dimensional substructures for which motions in different
dimensions uncouple, such as plates having a bending and a torsional mode at the
same frequency, and 3) structures with fully uncoupled substructures. Therefore,
in general, curve veering is far more likely to occur than curve crossing. In the
case of the folded sheets, curve crossings occur relatively frequently as a result of
the symmetry of the sheets; the introduction of geometric imperfections to the
folded sheets will break their symmetry and will therefore increase the number of
curve veerings.

Rotation of Eigenvectors As emphasized by du Bois et al. (2009) it is important to recognize the presence of curve veering in numerical models, and take
it into account for methodologies such as system identification, and modal parameter extraction. They also suggest the use of the Modal Assurance Criterion
(MAC) as a quantative measure of mode shape transformations (Allemang, 2003).
In the case of the folded textured sheets the eigenvectors will be orthogonal in the
physical coordinate system due to the implied unity mass matrix; in this case the
modal rotation is given by arccos of the inner product of the normalised eigenvectors. The rotation is not necessarily over 90 since the modes may change shape
as the system parameter varies; see Figure 5.12.

88

Eigenvalue

5.2 Modal Analysis

10

0.3

10

0.1

10

11

10

11

12

13

14

15

12

13

14

15

Rotation [deg]

90
45
0

Kratio

Figure 5.12: For the Miura sheet with = 60 , a/b=1, and = 55 the second
and third eigenmodes veer at Kratio 11.5. The figure shows the eigenvalues and
eigenmodes of the two eigenmodes, as well as the rotation of the eigenvectors with
respect to the stiffest eigenvector at Kratio = 8. Within the veering region the
eigenmodes rapidly, but continuously, rotate and eventually assume each others
shape; the rotation does not fully reach 90 as the mode shapes change slightly
with respect to the system parameter.

89

5.2 Modal Analysis

Geometric Imperfections Pierre (1988) showed that imperfections in repetitive systems may introduce curve veering and mode localizations, which are shown
to be two manifestations of the same phenomenon. The effect of geometric imperfections on the eigenvalue loci of the Eggbox and Miura sheet was investigated; at
a given configuration, the nodal x, y, z coordinates were randomly perturbed, with
a standard deviation specified as a percentage of the bar length. The eigenvalue
curves produced in this section are merely intended as indicative of general trends,
because the eigenvalue curves will differ for every randomly perturbed configuration. To enable comparison, the graphs of each of the sheets are all created using
the same random draw from a normal distribution, but with a different standard
deviation.
The results of the imperfection analysis are shown in Figure 5.13 and 5.14; as
expected, the curve crossings are replaced by curve veerings due to the break in
symmetry and introduction of coupling between modes. Importantly, for both
sheets the salient behaviour does not change materially as a result of the imperfections, and especially for lower values of Kratio the mode shapes remain very
similar. For both sheets a planar mode will dominate as Kratio , but for the
imperfect configurations the stiffness of the planar mode is no longer proportional
to Kfold as it now also involves bending of facets as well as the folds. What is more,
the planar mode emerges as softest mode at a higher value of Kratio compared to
the unperturbed configuration, as the imperfections must now be overcome.

90

10

10

10

10

10

10

Stiffness of eigenmode / K fold

Stiffness of eigenmode / K fold

5.2 Modal Analysis

10

10
Kratio

10

10

10

10

10

10

10

10

10

10

10

10

10

10

10

10
Kratio

10

10
Kratio

10

10

10

(b) = 2.5% of bar length.

Stiffness of eigenmode / K fold

Stiffness of eigenmode / K fold

(a) no imperfections; = 0.

10

10

(c) = 5% of bar length.

10

10

10

10

10

10

10

10
Kratio

10

(d) = 10% of bar length.

Figure 5.13: Effect of geometric imperfections in Miura sheet, on its eigenvalue


loci. The sheet is at = 55 , and the x, y, z coordinates are perturbed with the
same normal distribution, but with differing standard deviation .

91

10

10

10

10

10

10

Stiffness of eigenmode / K fold

Stiffness of eigenmode / K fold

5.2 Modal Analysis

10

10
Kratio

10

10

10

10

10

10

10

10

10

10

10

10

10

10

10

10
Kratio

10

10
Kratio

10

10

10

(b) = 2.5% of bar length.

Stiffness of eigenmode / K fold

Stiffness of eigenmode / K fold

(a) no imperfections; = 0.

10

10

(c) = 5% of bar length.

10

10

10

10

10

10

10

10
Kratio

10

(d) = 10% of bar length.

Figure 5.14: Effect of geometric imperfections in Eggbox sheet, on its eigenvalue


loci. The sheet is at = 45 , and the x, y, z coordinates are perturbed with the
same normal distribution, but with differing standard deviation .

92

5.3 Conclusions

5.3

Conclusions

The simple mechanical model presented in this chapter, which represents the
partly folded patterns as pin-jointed bar frameworks with additional planarity
constraints, successfully captures the key mechanical behaviour observed in Chapter 3. The counterintuitive out-of-plane modes were are among the softest, and
therefore most dominant, eigenmodes of the stiffness matrix. The dominant kinematics were also shown to be primarily a result of the geometry, rather than the
sheets material properties (represented by Kratio ). Furthermore, analysis of the
curve veering behaviour of the eigenvalue loci revealed the relative insensivity of
the dominant eigenmodes with respect to geometric imperfections. This result
supports observations by Norman (2009) regarding the behaviour of continuously
doubly-corrugated sheets. The modal analysis further enables the selection of unit
cell configurations where certain modes are dominant over others, or of equivalent
stiffness; both cases may be of interest in the design of morphing structures.
The analysis was purposely limited to non-extensible, and therefore developable,
deformations of the folded sheets. Bending of the facets was modelled by straightforward triangulation of the quadrilaterals; for first-order analysis the direction of
triangulation is not relevant, but for larger displacements the assumed kinematics
may no longer be valid.

93

Chapter 6

Manufacture of Folded Sheets


6.1

Introduction

Using conventional forming methods such as stamping for folded sheets would lead
to thinning and possibly fracture of the facets due to the pronounced folds. An important feature of sheets such as the Miura-ori is that they are developable and can
thus be folded from a flat sheet of material, without any stretching of the material
between the fold lines; they can therefore be formed using only low-energy bending operations. Many patterns are also rigidly foldable, and can be folded even
without any bending of the facets. However, the folded sheets also present unique
manufacturing challenges. As the fold lines do not extend across the sheet, there
is a strong coupling between the folds and many of the patterns can effectively
be described as 1 DOF mechanisms. This means it is difficult to simultaneously
have both folded and unfolded regions in the sheet material. Furthermore, during
the folding process the sheets significantly contract both longitudinally and transversely, while also expanding in thickness. Specialised forming techniques have
been developed to manufacture folded sheets continuously and reliably.
In recent years the development of manufacturing methods for folded sheets has
been spurred by a revived interest in folded sandwich panel cores for aerospace
applications. Geometric imperfections introduced during the manufacturing process were found to have a very significant detrimental effect on the load carrying
capacity of those folded cores, even within the linear-elastic range (Baranger et al.,
2011a). To enable accurate numerical simulations of the folded core material, the
folded sheet geometries are scanned after manufacture (Heimbs et al., 2010) or
simplified folding models are used to emulate the effect of imperfections in the
fold pattern (Baranger et al., 2011a). In essence, the geometric imperfections in
the folded sheets are currently modelled without any reference to the underlying
manufacturing method. Knowledge of the available manufacturing methods, the
associated material deformations and how they affect the accuracy of the folded
sheets, is therefore important for more widespread use of folded sheets in engineering applications.

94

6.2 Review of Manufacturing Methods

This chapter starts with a literature review of existing manufacturing methods


for folded sheets, followed by the description of a novel forming process using
gas pressure at room temperature that is particularly suited for rapid prototyping of different geometries. The required operating pressures are predicted using
analytical calculations, and these results are compared with manufacturing trials.

6.2

Review of Manufacturing Methods

In this review, we focus on manufacturing methods where the sheet material is only
(or substantially) subjected to bending operations, and material strains are minimal. This clearly excludes forming methods such as deep-drawing or conventional
vacuum forming operations. A range of specialized manufacturing techniques for
folded sheets has been developed, and a limited review was given by Khaliulin &
Batrakov (2005). Our aim is to consolidate the literature on manufacturing methods for folded sheets, provide a consistent classification of forming techniques and
identify the material deformations involved in the forming methods.

Classification Scheme
In this review we shall distinguish between synchronous processes, where the
folding takes place along all fold lines simultaneously, gradual folding processes,
whereby the folded sheet gradually transitions from a flat to its fully folded state,
and lastly, pre-gathering processes, which first form the sheet material into
singly corrugated sheets before creating the final double corrugation. Within these
categories, further refinements will be given with respect to the pattern orientation,
and whether it concerns a batch, cyclic or continuous forming process.

Fold Patterns
The material deformations not only depend on the choice of manufacturing process, but also on the desired fold pattern. As highlighted by Klett & Drechsler
(2010), the use of isometric and rigid origami patterns has the benefit of reducing deformation of the sheet material. However, even for rigidly foldable patterns
several undesirable deformations may take place during manufacturing of folded
sheets, in addition to bending along the intended fold lines. As will be shown,
facets of the fold pattern may bend, fold lines may move through the material,
or sign reversal of the fold lines can take place. By far the most studied folding
pattern is the Miura-ori, and it will be used as the example sheet throughout
the review. Most manufacturing techniques also extend to more general folding
patterns, but some make explicit use of the Miura patterns properties.

95

6.2 Review of Manufacturing Methods

(a) transverse pattern orientation.

(b) longitudinal pattern orientation.

Figure 6.1: The Miura pattern can be orientated with its corrugations extending
(a) transversely or (b) longitudinally to the direction of travel of the sheet material.
Miura - Pattern Orientation For many of the manufacturing processes the
orientation of the fold pattern with respect to the sheet material has a
strong influence on the forming operations and associated material deformation. We shall distinguish between whether the zigzag corrugations extend
transversely or longitudinally across the sheet material; see Figure 6.1.
Miura - Geometric Parameters The unit cell geometry of the Miura pattern
(see Section 4.1.1) not only affects the mechanical properties of the folded
sheet, but also its manufacturability. Important factors are the longitudinal
and lateral contraction ratios, respectively given by ix = 1 sin /sin
and iy = 1 cos , which prescribe the required amount of slip and/or
pre-gathering during the manufacturing process. The ratio V /L describes
the patterns deviation from a single corrugation, and may influence the
choice for a longitudinal or transverse pattern orientation. The only known
parametric study of the manufacturability of different Miura-ori geometries
is provided by Ichikawa (1995), and is limited to a specific manufacturing
process.
Globally curved sheets can be obtained through straightforward modifications of the Miura pattern; given their application in sandwich panel cores,
it is important to consider the suitability of the manufacturing methods for
this purpose.

96

6.2 Review of Manufacturing Methods

6.2.1

Synchrononous Folding Processes

In the synchronous folding methods folding takes places along all fold lines simultaneously, closely approximating the theoretical rigid folding process. The main
challenge is to coordinate the simultaneous folding of all fold lines. This is generally achieved by supporting the sheet during the folding process along the facets
or (some of) the fold lines using a transformable die.
Vacuum Actuation Gewiss (1968) describes a range of manufacturing methods, using vacuum actuation of the folding process. A distinction is made between
a moulding structure i.e. a master-sheet that is in contact with the facets
of the fold pattern, and a motive structure which coordinates the folding motion of the sheet. In its basic embodiment, the flat sheet material is sandwiched
between two flattened master-sheets, and then placed between two partly-folded
motive structures, whose initial volume enables actuation using a vacuum bag; see
Figure 6.2. The forming method makes use of the property of the Miura sheet,
whereby both the sheet material and the motive structure share the same kinematics; see Section 4.1.1. Accurate alignment of the moulding and motive structure
on either side of the sheet material is an important issue. One proffered solution
is to place the sheet material and moulding structures inside a separate vacuum
bag, which ensures complete flattening of the master sheets. Alternatively, the
moulding and motive structure are joined along their mating fold lines to form
a combined forming structure, or the moulding structure is removed entirely. In
another described embodiment a flat sheet is placed on either side of the motive
structure and using the vacuum bag both sheets are folded simultaneously. As
well as increasing the external pressure, it is suggested that additional forming
forces can be applied by placing the forming structure between a press. To enable
the motive structure to transfer the additional forces, vertical strengthening strips
are placed along the ridges, which are loaded in compression; see Figure 6.2(b).
Khaliulin & Dvoeglazov (2001) describe a similar vacuum actuated manufacturing
process, which makes use of a transformable matrix composed of hinged rigid
panels. The transformable matrix can be constructed either with or without a
master sheet that is in continuous contact with the sheet material during folding.
The two opposing transformable matrices are supported via orthogonal guides,
which enable the biaxial contraction of the sheets during the folding process, as well
as provide a means of controlling the fold depth; see Figure 6.3. Alternatively, as
described by Akishev et al. (2009), one side may be replaced by a vacuum bag that
presses the sheet against the matrix, solving the problem of accurate alignment
of the two matrices. The use of vacuum actuated transformable matrices was
extended to globally transversely curved Miura sheets by Khaliulin et al. (2002),
but here some slip between the sheet material and the die may take place.

97

6.2 Review of Manufacturing Methods

(a) combined moulding and motive structure, joined at the mating fold lines. The sheet
material is sandwiched between the two structures, before being folded as the two motive
structures contract under vacuum.

(b) the sheet material and forming structure


are placed in a vacuum bag, optionally with
increased external pressure.

(c) when extra forming forces are applied by placing the forming structure
between a press, the motive structures
can be strengthened with undulating
strips along the ridge lines.

Figure 6.2: Gewiss (1968) describes a series of embodiments for vacuum actuation
of the folding mechanism, comprising a moulding structure that is in contact with
the blank sheet material, and a partially-folded motive structure that coordinates
the folding process and whose initial volume enables vacuum actuation.

98

6.2 Review of Manufacturing Methods

(a) two orthogonal guides ensure alignment of the transformable matrices, as well as guide and control the folding
process.

(b) the sheet can either be placed directly between two transformable matrices (top), or together
with a master sheet that also makes contact along the facets (bottom).

Figure 6.3: A batch manufacturing process using a pair of opposing transformable


matrices, actuated using a vacuum bag (Khaliulin & Dvoeglazov, 1998, 2001).

99

6.2 Review of Manufacturing Methods

Figure 6.4: The blank is sucked onto the forming mandrel using an internal vacuum
bag, followed by the folding process using vacuumisation of the bag surrounding
the auxiliary dies (Akishev et al., 2009).
Khaliulin et al. (2002) also provide a (incomplete) force analysis of the transformable matrix under vacuum actuation, but no relationship between the applied pressure and the rigidity of the sheet material was established. Shabalin
et al. (2010) numerically simulated the folding process with a transformable matrix, and studied the complex deformation of the sheet material at the vertices.
Contact between the dies and the sheet material was shown to be initially along
the entire length of the fold lines, but gradually concentrated at points near the
vertices. This means that the folds and the area around the vertices are formed
without any direct contact with the dies. No account of required forces or total
forming energy was given.
An important feature of the vacuum actuated synchronous methods with transformable dies is the uniform application of the forming forces, and the minimal
material deformation during the folding process; any significant strains will be
limited to the areas around the vertices. The batch methods also provide flexibility to create globally-curved panels, fold combined patterns and add facilities for
curing or other post-treatment procedures. Disadvantages include unsuitability
for large sheets due to the costly increase in components, limitations on sheet material rigidity and the large number of manual steps involved. These processes are
therefore generally recommended for small scale prototyping purposes (Khaliulin
& Batrakov, 2005).

100

6.2 Review of Manufacturing Methods

Mechanical Actuation The synchronous folding process may also be actuated


mechanically, for example by means of a pantographic linkage (Paterson, 1999),
or by axially compressing the two opposing transformable matrices (Khaliulin
& Desyatov, 1991). Another general approach is to sufficiently bias the folds
mechanically, before exerting a compressive force around the perimeter to fold the
pattern to its desired depth (Khaliulin et al., 2005b; Akishev et al., 2007). This
is often used as a final calibration step in continuous manufacturing processes.

Successive Stamping Although strictly speaking not a folding process, successive stamping has been successfully used to create metal sandwich panel cores.
As the number of successive moulds is increased, the closer the folding process is
approximated and the more the slip between material and mould, and subsequent
stretching of the material, is minimized. The process benefits from simplicity
of equipment design, and is suitable for relatively ductile materials, with limited
pattern density (Miura, 1975a,b).

Self-Assembly Self-assembly of fold patterns has been considered for MEMS


devices, where strain mismatches across the fold lines would enable the selfassembly of the folded sheet (Bassik et al., 2009). The final configuration can
further be controlled by introducing areas in the folded structures, where the pattern will lock into place due to facet-to-facet contact (Kling, 2007b).
Another interesting approach to self-assembly is to utilize the fact the Miura-ori
pattern emerges as a post-buckling solution for thin flat sheets (Tanizawa & Miura,
1978). Mahadevan & Rica (2005) decribe how the biaxial compression of a thin,
stiff elastic film supported on a thick, soft substrate yields into a Miura-ori pattern
without any external guidance other than induced by the isotropic compressive
strains due to the relative expansion and contraction of the film and substrate.
Initially, straight primary buckles are formed; non-linear deformations of these primary buckles through global compression lead to instabilities wherein the buckles
collectively deform through soft modes, which are energetically cheaper than the
local extension or compression of the individual buckles. The disparity between
the energetics of bending and stretching a thin sheet favours the concentration of
curvature, which is energetically inexpensive.
Further work on self-folding origami membranes was performed by Pickett (2007),
who described a method relying on geometry to force a controlled collapse by
manipulating self-interactions and gentle external influences. The pre-creased thin
membranes were modelled as freely-hinged triangular facets, with a mild uniform
biaxial compression. The collapse was achieved through a Brownian dynamic
simulation; a myriad of collapsed states occur when the membranes collapse is
guided by internal, symmetric interactions, but a small initial global curvature

101

6.2 Review of Manufacturing Methods

proved to be sufficient to guide the system to a unique pre-determined state. Thus


the linear response of the system to an initial condition is enough to guide the
entire future behaviour of the system. This process effectively gives local control
over the mechanical properties of the collapsed sheet, by imposing a priori global
curvatures. This was shown for the waterbomb tessellation used by Kuribayashi
et al. (2006), as well as the Miura-ori. An interesting parallel is the post-buckling
pattern found in axially compressed thin-walled cylinders, the Yoshimura pattern,
which is consistently inwardly oriented (Yoshimura, 1955; Miura, 2002).

6.2.2

Gradual Folding Processes

In gradual folding processes, the sheet material transitions continuously from a


flat to its fully folded state, thereby gradually deepening and reducing its dimensions in-plane. As many fold patterns are rigidly foldable with a single kinematic
degree of freedom, this process necessarily requires some deformation of the facets
bounded by the fold lines. Interestingly, for the Miura pattern this gradual transition can be achieved with only (minimal) bending and no stretching of the facets.
Gewiss (1959) described a machine to fold a constructional sheet material from
flat sheets of metal, primarily intended for heat exchanger surfaces; see Figure 6.5.
Two sets of dies are arranged on either side of the sheet material; the individual
dies of each set are arranged to move relative to each other, both longitudinally
and laterally, using a set of push-rods and linear guides. During each cycle, the
sets of dies are moved apart and are extended such that the dies are above the
row formed in the previous cycle. As the dies are closed and the folds deepened,
the sheet material is automatically fed through in preparation of the next cycle.
The gradual folding can also be achieved in a continuous manner, for example by
means of a series of patterned rollers, each incrementally deepening the pattern
to its final fold depth (Kling, 2007a). A downside of this process is the costly
manufacture of the complex rollers, and the difficulty in tuning their relative rotational speeds. Another continuous gradual folding method features a looped
master sheet which continuously folds and unfolds, thereby shaping the sheet material (Paterson, 1999; Zakirov et al., 2009). As the sheet material is fed through,
it is brought in contact with the master sheet (e.g. using air pressure or bristle
brush rollers) and is subsequently gradually deepened as the master sheet folds
up. When the desired fold depth is reached, the sheet material is released, and
the master sheet is gradually unfolded ready for the next cycle; see Figure 6.6.

102

6.2 Review of Manufacturing Methods

Figure 6.5: Gewiss (1959) describes a folding method using a gradually deepening
set of dies. In each forming cycle the dies move apart, before adding a new row
and deepening the existing folds.

Figure 6.6: The advancing sheet material is pre-patterned before being pressed
against a master sheet using a bristle belt, and folded gradually to the desired
depth. At this point the material is released and the master sheet unfolds to
restart the cycle (Zakirov et al., 2009).

103

6.2 Review of Manufacturing Methods

Patterning and Gathering An important category within the gradual folding processes is the patterning and gathering method, which in recent years has
been successfully used for the continuous production of (curved) folded cores for
sandwich panels. First, the desired fold pattern is embossed onto the flat sheet
material using patterned rollers. The resulting residual stresses initiate the folding process, which is then progressively continued to the desired fold depth. An
important feature of this process is that the sheet material is not, or only intermittently, accompanied along the fold edges or facet surfaces during its folding
process. Instead, the process largely relies on the embossed sheet material folding, or rather buckling and collapsing, along the desired fold lines under minimal
external guidance.
Drechsler & Kehrle (2004) and Kehrle (2005) describe a one-step process for the
continuous manufacture of folded sandwich panel cores for aerospace applications;
see Figure 6.7(a). In the pre-processing step, the fold lines are embossed, impressed, scored or creased into the material using a set of rollers. Either one roller
is structured and the other of smooth, resilient material, or there are recesses
corresponding to the protrusions of the other. The folding process is initiated
using air nozzles above and below the material, or a mechanical arrangement that
contacts the material pointwise or along the folds. Next, the folding process is progressively continued. For example, using bristle brush rollers the forward travel is
decelerated to accommodate, or cause, the sheets longitudinal contraction. There
may also be guides between the fold initiation and the bristle rolls to constrain
and guide the vertical expansion, or transversely tapering channels that further
develop the folding. Finally, the folded sheet is passed through post-treatment
rollers which further emboss, and thus sharpen or emphasise, the folds. Zakirov
et al. (2006) describe a similar crease and bend manufacturing process where the
folding process is again initiated by creasing the sheet material along the fold lines,
using pairs of rollers with male knife dies and a female elastic back die. Locally the
procedure consists of complex shaping operations, with sheet material bending on
the elastic back die, as well as local material stretching. It is suggested this creasing process results in a groove relief of up to 20% of the required height. Pairs
of gathering rollers with radially oriented rods further deepen the fold pattern,
before the folded sheet is passed through tapering guides to achieve its final fold
depth; see Figure 6.7(b) and 6.8.
An important parameter in this process is the configuration of the patterned
rollers. These can either be configured as a single pair of mating rollers, or as
pairs of rollers creasing a single side at a time. The first option ensures good
alignment of the fold patterns on both sides of the rollers, but introduces problems at the vertices where mountain and valley folds meet and the material may
tear. Having more than one roller pair reduces their individual complexity, but

104

6.2 Review of Manufacturing Methods

(a) one-step patterning and gathering method (Kehrle, 2005).

(b) a crease and bend machine (Zakirov et al., 2008).

Figure 6.7: In the patterning and gathering methods the fold pattern is embossed or creased into the sheet material using patterned rollers, which initiates
the folding process. In (a) the longitudinally oriented Miura pattern is further
deepened using air nozzles, before the sheet is gathered using bristle rollers and
passed through a textured roller that imparts the final shape. In (b) the transverse Miura pattern is directly gathered using porcupine rollers, and subsequently
narrowed to the desired fold depth using tapering guides.

105

6.2 Review of Manufacturing Methods

(a) patterning rollers emboss the fold pattern on the sheet material;
this may be done with multiple one-sided rollers, or a pair of rollers
imparting the pattern on both sides simultaneously.

(b) gathering rollers with radially oriented rods deepen the fold pattern.

(c) photos of a demonstrator model (Zakirov et al., 2010).

Figure 6.8: Zakirov et al. (2006, 2010) describe a creasing and bending technique,
where the fold pattern is embossed onto the sheet material using patterned rollers,
before being gathered using a set of porcupine rollers.

106

6.2 Review of Manufacturing Methods

introduces the problem of poor alignment with the already partially folded sheet
material. Zakirov et al. (2006) note that position inaccuracy of the sawtooth lines
has less influence on the final fold height, and these might therefore be added with
a separate roller. Zakirov & Alekseev (2007) discuss the design parameters of the
creasing rollers, converting the planar geometry of the transverse Miura pattern
to cylindrical coordinates on the patterned rollers.
No conclusive preference is provided in the literature for either the longitudinal
or transverse pattern orientation, and examples of both Miura patterns have been
found. The choice of pattern orientation appears to be driven by the application
of the core material, rather than any obvious limitations of the manufacturing
process. For example, Alekseev et al. (2011) show a folded sandwich panel core
with tapering thickness, which can only be achieved with a transverse pattern orientation. For curved folded sheets a preferred solution would be to align the global
curvature longitudinally with the advancing sheet material. In combination with
the intended application, this will determine the preferred pattern orientation.
A major advantage of the gradual folding techniques, and especially the patterning
and gathering method, is the possibility of continuous production of the folded
sheets with minimal material deformation. It must be noted that the Miura-ori has
a very convenient gradual deformation which only involves bending of the facets,
but this may not necessarily extend to other fold patterns. An underexplored
aspect of the patterning and gathering methods is the limits on the sheet thickness
and material stiffness that can be processed. The folding forces are not exerted
directly onto the sheet material, and the forming process is an interplay between
the desired local in-plane buckling (into the fold pattern) and the global out-ofplane buckling of the sheet.

6.2.3

Pre-Gathering Processes

The third manufacturing category aims to overcome the coupled longitudinal and
transverse contraction by first pre-gathering the sheet material into a singly corrugated sheet, to a width substantially corresponding to the final width of the
folded sheet, before forming the double corrugation.

Cyclic Forming For the cyclic forming processes the double-corrugation is


added row-by-row, by feeding the pre-gathered material through to the shaping zone in a cyclic manner. For the transverse pattern orientation the process
is characterised by bending of the material along the zigzag lines, accompanied
by a sign reversal of the dihedral fold angles along the longitudinal corrugations
bounded by these lines (Hochfeld, 1959; Gewiss, 1976; Khaliulin & Skripkin, 1997);
see Figure 6.9. During the forming process the unconstrained facets must undergo

107

6.2 Review of Manufacturing Methods

(a) the sheets double corrugation is formed row by row, by inverting


the fold lines between the dashed lines.

(b) cyclic folding process; left to right: clamp, fold, release, feed.

(c) cyclic forming machine by Hochfeld (1959), which continuously


gathers the sheet before it enters the shaping zone.

Figure 6.9: For the transverse pattern orientation the double corrugation is formed
by cyclically clamping a pre-corrugated sheet using mating jaws, and inverting the
longitudinal corrugations (Hochfeld, 1959; Khaliulin & Skripkin, 1997).
significant bending, and less ductile materials may fracture when the fold angles
are inverted. Also note that the folded sheet leaves the folding area at an angle to the plane of the incoming sheet material. For the longitudinal pattern
orientation, at each forming cycle three shaping zones are placed across the pregathered sheet material, and the middle one is moved transversely to create the
double corrugation; see Figure 6.10. This folding operation may be performed
on a row-by-row basis (Khaliulin et al., 2007) or multiple rows simultaneously.
Gewiss (1960) describes parallel sets of mating blades which are placed across the
corrugations, and whose transverse movement is coordinated using a link-chain
running along a tapering guide. Khaliulin & Batrakov (2005) comment that the
transition from linear to zigzag corrugation is accompanied by large (elastic) bending deformation of the facets, which increases for smaller parallelogram angles
of the Miura pattern. For the longitudinal orientation the fold lines either have

108

6.2 Review of Manufacturing Methods

(a) during the forming of a longitudinal Miura pattern from precorrugated material, the fold lines necessarily migrate through the
material (Khaliulin et al., 2007).

(b) in a process described by Gewiss (1960), an array of blades is


placed into the pre-corrugated sheets, and their sideways movement
is coordinated using a link-chain and a tapering guide.

Figure 6.10: Cyclic manufacturing of a longitudinally oriented Miura pattern from


pre-gathered sheet material (Gewiss, 1960; Khaliulin et al., 2007).
to migrate through the material or the facets must undergo a shear deformation.
The difficulty of controlling these deformations is considered a disadvantage of
these techniques. Among the advantages listed by Khaliulin & Batrakov (2005)
of the cyclic pre-gathering forming methods, is their relatively simple and reliable
equipment assemblies, as well as the integrated calibration steps. This can be done
by either pressing the folded sheet material into a dense packet, or stretching of
the facets at the final stage of the shaping cycle (respectively for the longitudinal
and transverse orientation).

Bunch & Crunch Basily et al. (2006) described a manufacturing method where
the corrugated material is passed through a set of mating textured rollers that directly impart the final shape. This modifies the biaxial contraction of the sheet
to successive transverse and longitudinal contraction, thereby significantly reducing the amount of slip between the sheet material and the textured rollers. The
pre-gathering is achieved by passing the flat sheet material through a series of cir-

109

6.2 Review of Manufacturing Methods

(a) schematic overview of the gradual pre-gathering of the sheet material before passing the corrugated sheet through textured rollers
(not shown).

(b) photos of an experimental manufacturing setup, with textured rollers producing doublycorrugated folded sheets from pre-gathered sheet material. Rollers for both the transverse and
longitudinal pattern orientation are shown.

Figure 6.11: Elsayed & Basily (2004, 2010) describe a manufacturing process
whereby the sheet material is first gradually pre-gathered using grooved rollers,
before the final double-corrugation is imparted using textured rollers.
cumferentially grooved rollers; by increasing the number of grooves sequentially,
the chance of the material being locked between two grooves is reduced and material shredding is eliminated. It was noted that the method had successfully folded
thin sheets of paper, copper, aluminium and stainless steel. An estimate of the required forming energy was provided by Elsayed & Basily (2004), but no account of
actual deformations during the folding process was given. Elsayed & Basily (2010)
suggest the direction of the engraved folding pattern on the last set of rollers can
be made longitudinal or perpendicular to the roller axis, or any desirable angle in
between; the latter can be achieved by mirroring the pattern along the centre of
the sheet, which counters any imbalance in forming forces.
Kling (2005) describes the same two-phase process, and comments that the fact
that the conversion from fluted material to faceted material using the patterned
rollers takes place without significant stretching of the material inside the rollers
is completely non-obvious. Kling (2007a) notes that in the bunch and crunch
method the material undergoes its longitudinal contraction in the transition zone
from just before it is fed into the rollers to approximately the midpoint between
the roller axes. During this process the crease lines must migrate through the
sheet as the fold depth increases. It is suggested that in many cases it is actually

110

6.2 Review of Manufacturing Methods

impossible for the sheet to slide over the successive teeth to reach the desired
folded pattern without in-plane distortion, due to the friction with the rollers and
the difficulty in having creases migrate through the material. Consequently, when
manufacturing metal sheet, the depth of the pattern needs to be severely limited
to avoid punctures or tears. Kling (2007a) suggests manufacturing the rollers from
hard materials with a low friction coefficient, as it improves slippage between the
sheet material and the roller. Furthermore, the longitudinal pattern orientation is
preferred over the transverse shown in Basily et al. (2006), as it allows for deeper
fold patterns. Kling (2007b) qualitatively discusses design parameters for the
textured rollers. At the contact region several teeth of the roller will engage with
the sheet simultaneously. The further the teeth go into the sheet material, the
greater the longitudinal contraction. However, tangentially, the teeth spacing in
the longitudinal direction remains approximately constant. Thus there is a relative
velocity between the roller and the sheet. The larger the roller circumference to
the period length, the more the teeth will have to slide over the material due to
this relative velocity, and the greater the resulting material strains. The use of
preferably less than 20 fold edges going around the roller is suggested. Very slender
rollers, however, will bend and deflect due to the contact forces and may need to
be supported by backing rollers. To reduce the amount of slip between rollers and
the sheet material, Kling (2007a) further suggests passing the corrugated sheet
through two or more pairs of textured rollers that gradually achieve the desired
fold depth.
Kehrle (2005) refers to a two-step process where the initial embossing precedes
a transverse contraction and thickness expansion, followed by out-of-plane folding and longitudinal contraction in the second phase (either through compressed
air or a suitable mechanical arrangement). Drechsler & Kehrle (2004) mention
the procedure works for tough materials like nomex paper, but is not suited for
materials such as aluminium sheets. Fischer et al. (2008) show a prototype of a
manufacturing machine for this process, folding a longitudinally oriented pattern.
Ichikawa (1995) describes a transversely oriented bunch and crunch method,
where it is noteworthy that the wavelength of the pre-corrugated material does
not appear to match that of the folded sheet produced by the textured rollers,
although the contraction ratio is substantially the same. Ichikawa (1995) is the
only known publication to provide a combined analysis of the desired mechanical
properties of a folded core and the limits imposed on the geometry by the chosen
manufacturing method; see Figure 6.12. Using a series of experiments, parametric
relationships are established for shear strength (parallel shear of face sheets) and
bending rigidity, respectively providing desired geometries of 0.4 < H/L < 1.4
and D/L 0.5. The manufacturing limit is determined by the pre-narrowing
ratio, with an experimentally determined maximum of iy = 10%; exceeding this

111

6.2 Review of Manufacturing Methods

(a) pre-gathered sheet material is passed


through textured rollers.

(b) geometry parameterisation used by


Ichikawa (1995).

(c) manufacturability diagram, with the shaded area indicating a combination


of desirable and feasible geometries.

Figure 6.12: Ichikawa (1995) describes a manufacturing method where pregathered sheet material is passed through textured rollers (a). A chart with manufacturing limits (c) was developed, using a combination of mechanical properties
and manufacturability requirements. The desired range of H/L followed from
shear tests, whereas the allowable width narrowing ratio iy < 10% was determined through manufacturing trials. A combination of these parameters uniquely
defines D/N , and its upper limit.

112

6.2 Review of Manufacturing Methods

ratio creates too much local wrinkling and buckling in the crunching process. This
provides a feasible design space in terms of amplitude ratio H/L and meandering
ratio D/N ; the parameter D/L remains free to choose.
The material deformations for the cyclic and bunch and crunch methods are comparable, but are strongly affected by the choice of pattern orientation. For the
longitudinal pattern orientation an important issue is the shifting of fold lines
through the material during the sheets transition from a single to a double corrugation; this will be especially relevant as the V /L ratio increases, and the hitherto
parallel corrugations increasingly overlap. For the transverse pattern orientation
an important factor is the inversion of fold lines, which limits its application to
ductile materials. The cyclic forming processes will benefit from lack of slip between the mould and sheet material, which is a crucial consideration when using
textured rollers. Also, the transversely folded sheet will wish to leave the forming
zone at an angle to the plane of the incoming sheet material, which will impart
some bending stresses to the folded sheet as it exits the textured rollers. The manufacturing limits and accuracies for either orientation are still largely unknown,
and no conclusive preference can thus be given. Although not explicity mentioned
in the literature, the bunch and crunch methods would be suited for globally
curved panels if the curvature aligns with the direction of travel of the sheet material. The pattern orientation would then be prescribed by the application of the
core material.
An important advantage of both these types of pre-gathering methods is that
relatively stiff materials can be formed as the folding forces are exerted directly
on the sheet material. A disadvantage of the textured rollers is their complexity
and thus costly manufacture, as well as the complex material deformations due to
the required slip between the sheet and the textured rollers.

Oscillating Discs Kling (2007b) describes an intriguing method to create sinusoidally doubly corrugated sheet, by passing a corrugated sheet through an array
of articulated rollers. As these discs oscillate, they impart an approximately sinusoidal corrugation to the sheet; see Figure 6.13. This roughed out version of the
sheet is then passed through one or more sets of secondary rollers that impress
the polygonal folding pattern. In effect, the oscillating discs accomplish most
of the longitudinal contraction prior to the textured rollers, thereby significantly
reducing slip in the final forming stage.

Continuous Inversion Described by Gewiss (1964), this method was aimed


at creating large numbers of small herringbone corrugations of pliable materials,
specifically cardboard. The material is first longitudinally corrugated, before being
fed onto a continuously moving belt which imparts the desired folded configura-

113

6.2 Review of Manufacturing Methods

Figure 6.13: Corrugated sheet material is passed through a set of oscillating


discs that impart a sinusoidal corrugation pattern. The roughed-out sheet is
subsequently passed through textured rollers or stamping dies to obtain its final
shape (Kling, 2007a,b).
tion; see Figure 6.14. As the corrugated material contacts the patterned belt, the
corrugation is either continously inverted, or (on the opposite side) it is immediately mated to the master pattern. Both the contact between and the eventual
separation of the sheet material and the belt is achieved using air pressure. By
design this method only applies to transversely oriented patterns, and the material
should be pliable enough to allow fold lines to move through the material.

Hybrid Folding The hybrid folding technique introduced by Khaliulin et al.


(2005a) features both single longitudinal corrugations and the Miura patterns
double corrugations within the same folded sheet. During each shaping cycle the
sheet material is folded flat, before a new double corrugation is added and the
hybrid pattern is folded simultaneously along all fold lines to a three-dimensional
state; see Figure 6.15. An important issue is the repeated folding and unfolding
of the sheet, which prohibits the methods use for materials sensitive to fatigue.
Both transverse and longitudinal patterns can be folded, but for the latter orientation the material must undergo a substantial shear deformation when the linear
corrugation is converted into a zigzag one. Khaliulin & Batrakov (2005) suggest
that these methods are convenient for fold patterns with a relatively large V0 /S0
ratio. Otherwise the transverse scheme would involve too large dihedral folds, and
for the longitudinal scheme the position of the deforming forces with respect to
the blank becomes unfavourable.

114

6.2 Review of Manufacturing Methods

Figure 6.14: In the continuous inversion method described by Gewiss (1964) the
pre-corrugated sheet material is continuously fed onto a moving belt, which imparts the desired doubly-corrugated pattern.

(a) longitudinal hybrid folding.

(b) transverse hybrid folding.

Figure 6.15: The hybrid folding process was described by Khaliulin et al. (2005a).
During each cycle the folded sheet is flattened, before being fed into the shaping
zones between A and C. Shaping takes place by simultaneously folding along
zigzag, sawtooth and straight lines; this forms a hybrid structure with the doublecorrugation joined with the triangular corrugation along the sawtooth lines.

115

6.2 Review of Manufacturing Methods

6.2.4

Conclusions

A wide range of manufacturing methods for folded sheets has been found in the
literature, but almost all most omit a quantative account of the material deformations during the folding process. Any complete overview of the suitability of
the various methods for given materials and fold patterns is therefore impossible
to obtain merely through a literature review. Especially the manufacturing accuracies for the final folded sheets is still an open question. Nevertheless, several
conclusions can be drawn.
The synchronous folding methods closely approximate the theoretical rigid folding kinematics of the sheets, and thus introduce the least amount of undesired
deformations in the sheet material. Making use of the kinematic properties of the
Miura-ori sheets, the vacuum-actuated methods provide a very suitable means to
produce sheets of a wide range of materials and geometries. However, other than
for the production of small numbers and small surface area, the methods will be
uneconomical.
The gradual folding methods find a middle ground between minimal material deformation and the possibility of continuous production. Especially the patterning
and gathering methods have been successfully shown to produce folded (curved)
sandwich panel cores in a wide range of materials. However, no studies have been
found into the patterns and materials that can be folded, and the amount of embossing or pre-weakening of the fold lines that is required to enable the sheet to
collapse into the desired fold pattern.
Several interesting pre-gathering methods have been found. Many, however, appear to require substantial deformation of the sheet material to attain the final
folded configuration. These deformations range from bending of the facets, to
shifting or inversion of fold lines, and shear deformation of the facets. Nevertheless, especially the bunch and crunch methods are interesting for mass production
of folded textured sheets, as relatively thick and stiff materials can be formed.
Finally, it should be noted that even for folding processes that significantly deform
the sheet material, the ultimately formed sheets will still undergo small material
strains compared to alternative conventional forming methods.

116

6.3 Cold Gas-Pressure Folding

6.3

Cold Gas-Pressure Folding

This section describes a novel forming process to fold a Miura pattern from metal
sheets, using a cold gas-pressure technique1 (Schenk et al., 2011). Although the
Miura-ori sheet is most commonly used, the manufacturing process clearly extends
to other folding patterns.

6.3.1

Process Description

First the fold pattern is transferred onto the metal sheet, and the material is
weakened along the fold lines. This can be achieved through a range of techniques,
such as embossing, etching, melting, etc. In our initial implementation a simple
perforation pattern is created along the fold lines. Next, two such sheets are
packed into a sandwich structure, with the sheets separated by a series of simple
spacers that are placed into slots along the mountain ranges of the sheet; see
Figure 6.16. The combination is packed into an air-tight bag, which is evacuated
to near vacuum, combined with an increase in external pressure. This pressure
difference forces the sheets to bend locally and simultaneously along all fold lines;
see Figure 6.17. The biaxial contraction during the folding process is enabled
by the freely moving spacers. The fold depth is limited by the eventual contact
between the two sheets, which is determined by the height of the spacers.
The forming process introduced here differs from existing synchronous manufacturing methods for folded sheets in a number of ways. Firstly, it is does not
require expensive tooling such as folding matrices, and is therefore well suited to
create prototypes of different geometries. Also, the spacer plates provide minimal
guidance to the folding of the sheet, and the process therefore largely relies on
fold lines forming accurately and the sheet contracting uniformly, without external guidance. In contrast, a similar method described by Gewiss (1968) separates
two sheets by a more complex motive structure that controls the folding motion.
Lastly, relatively thick sheet materials can be formed using this method by increasing the external pressure, with an upper bound given by the buckling load of
the spacer plates.

6.3.2

Process Calculations

The aim of the following calculations is to predict the required forming pressure to
fold a Miura-ori sheet from flat sheet material. For reasons of simplicity, we shall
ignore many of the subtleties involved in sheet metal bending, such as shifting of
the neutral axis, the exact curvature of the fold line and subsequent springback
1 the

manufacturing method was developed in collaboration with Dr Julian Allwood, CUED.

117

6.3 Cold Gas-Pressure Folding

Figure 6.16: Manufacturing process. From right to left: aluminium sheet perforated along the fold lines; sheet with spacer plates slotted in place before the
second perforated sheet is placed on top to form a sandwich structure; a folded
sheet after the folding process.

Figure 6.17: The sandwich of patterned sheets is placed into an autoclave; after
creating a sufficient pressure difference the sheet is folded and the vacuum bag
opened to reveal the final product.

118

6.3 Cold Gas-Pressure Folding

2S

2L
x

Figure 6.18: Geometric parameters of a Miura unit cell. The values a, b and
describe the shape of the parallelogram, whereas the fold angle determines the
current configuration of the unit cell (=0,55,85 are shown).
(Marciniak & Duncan, 1992). Instead, we assume that during the folding process
a plastic hinge has formed along all fold lines (Johnson & Yu, 1980). By equating
the work done by the applied pressure and the energy dissipated plastically in the
hinges, an analytical formulation for the required folding pressure can be found.
Unit Cell Kinematics
For the calculations an analytical description of the kinematics of the folding
process of the Miura pattern is required. The unit cell can be defined by the
side lengths a and b, and the acute angle of the parallelogram elements see
Figure 6.18. The dihedral fold angle between the facets and the mid-plane
determines the current configuration. The height H, the width 2S and length 2L
of the unit cell are calculated as follows
H = a sin sin
S =b p
L=a

(6.1)

cos tan

(6.2)

1 + cos2 tan2

1 sin2 sin2

(6.3)

The projected area of the unit cell is then Aproj = 4SL. The fold angle is
described by
!
r
cos2
2
= arcsin 1/
+ sin
(6.4)
cos2

119

6.3 Cold Gas-Pressure Folding

Energy Balance
The forming pressure difference p can be calculated by equating the variation in
external work done by the applied pressure and the internal work required to
plastically deform the sheet material, as follows
dWext = dWint
X
pdV =
Mpi di

(6.5)
(6.6)

where di is the change in fold angle i, and Mpi the corresponding plastic moment.
For the Miura sheet we shall assume the compatibility equations of the folding
process to be described by the fold angle . The energy balance can then be
rewritten as
X
V
i
d =
Mpi
d
(6.7)
p

which must hold for any d for the sheet to be in equilibrium. The required
pressure difference p can subsequently be found for any fold depth.
External Work The external work consists of two components: the work exerted by the pressure on the top surface of the sheet and the work it exerts around
the perimeter, as the sheet folds see Figure 6.19. The two components of the
total displaced volume V are given as
A0
Aproj
(H h) +
h
2
2
For the virtual work equation, we find dV as


Aproj (H h) Aproj H
+
dV =
d

2
2
V = Vtop + Vside =

(6.8)

(6.9)

Note that increasing the height h of the spacers lowers the pressure required to
fold the sheet; however, in all calculations presented here, we shall assume that
the spacer height is double the desired fold depth.
It is further interesting to note that the component Vtop is by itself not sufficient
to fold the pattern to any desired depth. This is due to the fact that the displaced
volume Vtop reaches a maximum at = 45 see Figure 6.20. Thereafter dVtop
becomes negative and the external pressure provides negative work; i.e. it will
work to unfold the sheet. In order to fold the sheet further, the effect of the
pressure exerting a compressive force along the perimeter of the sheet takes over.
Internal Work Assuming a plastic hinge has formed along all fold lines, the
internal work done to deform a unit cell from its initial flat state to a given
configuration, is described by:



(6.10)
Wint = 2t2 y aa ( ) + bb
2

120

6.3 Cold Gas-Pressure Folding

(a) unit cell in a partially folded state. The volume under the unit
cell is given by Vtop = H
A
, with H the current fold depth of the
2 proj
pattern.

(b) partially folded sheet with spacers. The displaced volume is indi,
cated by the shaded areas, and is calculated as Vside = (A0 Aproj ) h
2
with h the spacer height.

Figure 6.19: The total displaced volume V during folding of the sheet consists of
the volume Vtop under each unit cell (a), and the volume Vside displaced by the
in-plane contraction of the sheet (b).
where is the dihedral angle between the vertical plane and the facets (see Figure 6.18), the degree of pre-weakening along the fold line and a, b the length of
the fold lines. The change in internal work dWint is then



dWint = 2t2 y bb aa
d
(6.11)

No folding will take place along the perimeter of the sheet, which could be taken
into account in the internal work equation. Assuming a sufficiently large sheet, the
effect along the perimeter becomes negligible and the folding process of an entire
sheet can be described using a single unit cell. Any additional plastic deformation
at the vertices, where several fold lines interact, has hitherto not been taken into
account. Within the assumed deformation kinematics, i.e. plastic hinges along the
length of the fold lines, the additional energy dissipated at the vertex is (t/2)3 ;
this component can therefore be neglected when t  a, b.

121

6.3 Cold Gas-Pressure Folding

1
total
side
top

0.9
0.8

2V / ( h A 0 ) []

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

10

20

30

40
50
[deg]

60

70

80

90

Figure 6.20: Normalized displaced volume, V = 2V / (hA0 ), as a function of the


fold angle; here h = 2hmax = 2a sin(). As can be seen, Vtop reaches a maximum
at = 45 (which is independent of geometry), after which dVtop < 0 and the
pressure on the top surface will therefore work to unfold the sheet. Klett &
Drechsler (2010) found the same result, whilst investigating the relative density of
a folded sandwich panel core.

122

6.3 Cold Gas-Pressure Folding

Maximum Pressure
Using Equations 6.7, 6.9 and 6.11, the required forming pressure at every stage
of the folding process can be calculated. Of greater practical interest, however,
is the maximum pressure needed to fold the sheet to a desired fold angle target .
In order to calculate the required pmax for Miura patterns of any geometry, the
variables governing the forming process have been normalized as follows:
p =

pmax A0
y t2eq

r=

a
b

target

where a, b, describe the geometry of the unit cell, A0 is the initial area of the unit

cell, target the desired fold angle, y the materials yield stress, and teq = t the
equivalent material thickness along the fold line. The resulting plot for r = a/b = 1
is shown in Figure 6.21 and enables quick lookup of the pressure required to fold
different patterns.

Coupling of Fold Angles


The relationship between the two fold angles and in the Miura sheet will
depend on the unit cell geometry and configuration see Figure 6.22. Values of
= 0 and = 90 provide edge cases where the two fold lines either coincide or
are orthogonal, and the fold angles are therefore respectively equal or completely
decoupled. For the geometry used in our manufacturing trials, = 60 , the
relationship is strongly non-linear. At the initial stage of folding, the ratio d(/2
)/d 0.5 and a plastic hinge will thus not yet have formed along both fold lines
simultaneously.

123

6.3 Cold Gas-Pressure Folding

(a)
70

16
15.5
15

65

14.5

6.5

14
7

60

p 9.3

15
14

13.5

13

7.5
8

55

16

12

[deg]

8.5
9

11

9.5

50

10

10.5

11

45

10
11.5

12

12.5 13

13.5
14

40

14.5

15

35

10

20

30

40

50

target [deg]

60

70

80

90

(b)

Figure 6.21: The normalized maximum pressure p = pmax A0 /(y t2eq ) required
during the folding process to attain target for varying , with the ratio r = a/b = 1.
By means of example, in (a) the surface is sliced at = 60 and target = 55 ,
yielding a maximum normalized forming pressure of p = 9.3. Alternatively, the
contour plot (b) enables straightforward lookup of the required pressure.

124

6.3 Cold Gas-Pressure Folding

90
80
70

90 [deg]

60

=0

50

=60

40
30
20

=88
10
0

10

20

30

40
50
[deg]

60

70

80

90

Figure 6.22: Coupling between the two fold angles of a Miura sheet, and ,
for varying geometry . At the intersection of the curves with the diagonal, the
coupling d(/2 )/d = 1.

125

6.3 Cold Gas-Pressure Folding

Figure 6.23: Modified three-point bending test to determine the yield stress in a
plastic hinge. A narrow wedge is moved downward, bending the sheet along the
perforated fold line. The measured displacement d and load F can be converted
into a plastic moment versus fold angle (= 2) relationship.

6.3.3

Manufacturing Trials

A series of manufacturing trials was performed to assess the novel manufacturing


method and validate the analytical calculations. The dimensions for the Miuraori pattern were chosen as a = b = 25mm, = 60 and m = n = 4, with
a desired fold angle of = 55 , which corresponds to a depth of 17mm. The
desired perforation pattern, with a = b = 0.4, was made using a CNC milling
machine, and the spacer plates were manufactured from 22 gauge stainless steel
with a waterjet cutter. Aluminium sheets of Al 5252-H22 alloy of two different
thicknesses 22 gauge (0.71mm) and 24 gauge (0.56mm) were used, and the
predicted maximum forming pressures were respectively 2.03 and 1.03 bar.

Plastic Hinge Measurements


In order to gain insight into possible strain-hardening behaviour and find a suitable
value for the materials yield stress in the plastic hinge, a series of modified threepoint bending tests was performed; see Figure 6.23. The tests were done on an
Instron 3367 testing machine. The measured load F and displacement d are
converted into a moment-angle relationship as follows:


F
t
Fw
Mp = tan ( arctan()) d
+
(6.12)
2
2
4


d
The fold angle, 2 2 arctan w/2
, follows from the geometry of the setup,
with w the width of the support and the coefficient of friction between the test
specimen and the support.

126

6.3 Cold Gas-Pressure Folding

40

35

Moment / Fold Length [N]

30

25

20

15
24 gauge; = 0.40; y = 221 N/mm2
24 gauge; = 0.95; y = 227 N/mm2

10

22 gauge; = 0.40; y = 270 N/mm

22 gauge; = 0.63; y = 275 N/mm2

22 gauge; = 0.79; y = 280 N/mm2


0

10

20

30

40
50
60
Fold angle [deg]

70

80

90

100

Figure 6.24: The plastic moment per unit length of the fold line, for the 22 and 24
gauge aluminium sheets, calculated from the measurements using a friction factor
of = 0.3. Each line is the average of four test specimens with equal perforation
factor . The equivalent y is calculated as the average over fold angles of 20100 .
The test results are shown in Figure 6.24 for the Al 5251-H22 alloy sheets used in
the manufacturing trials. For the purpose of our calculations, the measurements
sufficiently support the approximation of a rigid perfectly-plastic material model.
The equivalent yield stress y is taken as the average over fold angles of 20 100 .
Using a friction coeffient of = 0.3 this resulted in values of y = 270MPa for the
22 gauge (0.71mm) sheet and y = 220MPa for the 24 gauge (0.56mm) sheet.
A few remarks are in order, regarding the results. When assuming a lower value for
the friction coefficient , the bending moment at the fold line will increase and the
material will show a stronger tendency for strain hardening; see Figure 6.25. For
the 22 gauge sheets, the graph clearly shows the effect of some strain hardening,
before the bending moment drops off as the material cracks at the outer edge
of the bend. To gain more detailed insight into this issue, a more suitable test
setup is required. For example, a test used by Marciniak & Duncan (1992, pp.
90) applies a uniform bending moment, whereas Baranger et al. (2011a, Fig. 5)
use a Shanley column, where the specimen is attached to two rigid parts and the
combination is loaded in compression beyond its buckling load. Either approach
would eliminate the problem of friction, and provide a more accurate measurement

127

6.3 Cold Gas-Pressure Folding

45
0

40
0.1

35

0.2

Moment / Fold Length [N]

0.3
0.4

30

0.5

25
20
15

=0
= 0.1
= 0.2
= 0.3
= 0.4
= 0.5

10
5
0

10

20

30

40
50
60
Fold angle [deg]

70

80

90

100

Figure 6.25: In the conversion of the measured force-displacement values to the


desired moment-angle relationship (Equation 6.12), the assumed friction coefficient
has an important effect. Lower assumed values for the friction coefficient result
in greater strain-hardening behaviour as well as greater plastic moment.
of the plastic moment. Lastly, the sheets material properties will depend on the
fold lines orientation within the sheet due to the anisotropy introduced by the
sheet rolling process.

Trial Results
The novel forming process was successfully used to fold the Miura-ori pattern
from aluminium sheets; an example sheet manufactured of 22 gauge aluminium is
shown in Figure 6.26. For two sheets the applied pressure and acquired fold depth
were measured, and the experimental results are compared with the predicted
values in Figure 6.27. The fold depth was measured after pressure was released
(i.e. after springback) and the value was converted to fold angle , assuming the
theoretical folding kinematics of the sheet. As can be seen, the predicted values
lie below the measured pressures required to fold the sheet to a desired depth.
This suggests that even in this relatively pure form of folding, e.g. no moving fold
lines and minimal bending of the facets during folding, we cannot yet account for
the full forming energy required to fold the Miura pattern. The large disparity
between theory and measurements for low values of the fold angle is due to the
material elasticity, rather than the rigid perfectly-plastic material model assumed

128

6.3 Cold Gas-Pressure Folding

Figure 6.26: A Miura-ori sheet folded from a 22 gauge aluminium sheet, using the
novel folding process.
for the calculations. This should be resolved with a more advanced model, such
as Finite Element Analysis.
Looking at Figure 6.28 the vertices appear to undergo more complex and substantial plastic deformation than assumed in our simple kinematic model; this is also
seen in simulations by Shabalin et al. (2010). The holes drilled at the intended
vertices of the folded sheet are consistently not centred on the actual vertex in
the folded state. It is postulated that the curved fold line follows a lower-energy
route, and thereby avoids the intended vertex. As a result, the intended fold line
with the perforation pattern no longer lies along the actual fold line. The more
complex deformation kinematics at the vertices also resulted in small membrane
strains in the corners of the parallelogram elements. A closer look at the material
along the fold lines in Figure 6.28 also reveal cracks along the outer radius of the
fold lines, due to the tight fold angle.
Some tests were performed to establish the influence of the weakening factor
along the fold lines. The sheets that were pre-weakened less than = 0.4 did not
fold as desired and ended up effectively singly-corrugated, with the sheet forming
an almost straight corrugation along the unsupported ridges. There was also a
global instability as the two sheets bulged and buckled, damaging the spacer plates
and voiding their reuse.

129

6.3 Cold Gas-Pressure Folding

12

10

p A0 /(y t2) []

calculation (4x4)
calculation (unit cell)
22 gauge
24 gauge

10

15

20

25
30
[deg]

35

40

45

50

55

Figure 6.27: Comparison of the theoretical forming pressure with the measured
values; both are normalized. The measured values were normalized using the
experimentally determined yield stress. Given the relatively small dimensions of
the Miura sheet (4 4 unit cells), the absence of folding along the perimeter was
taken into account for the calculations.

130

6.3 Cold Gas-Pressure Folding

Figure 6.28: Close-ups of the fold lines of a Miura sheet folded from 22 gauge
aluminium sheet. Top shows a ridge supported by the spacer plates, and bottom
one that is freely formed. Note how the fold pattern avoids the intended vertex,
seeking a lower-energy solution.

131

6.4 Conclusions

6.4

Conclusions

A comprehensive review of (patent) literature revealed a range of specialised manufacturing methods for folded textured sheets specifically the Miura pattern.
These methods were categorised into synchronous, gradual and pre-gathering folding techniques. Spurred by investigations into folded sandwich panel cores, recent
years have seen the development of promising methods for the continuous manufacture of folded textured sheets. However, most publications describe the manufacturing process in qualitative terms, and minimal information is available regarding
the precise forming process and resulting deformations of the sheet material, or
the restrictions on the fold patterns and materials that can be folded. An important open area of research is the accuracy of the various manufacturing methods,
as this greatly affects the mechanical properties of the folded core material.
Furthermore, a novel manufacturing method was introduced, which uses cold gaspressure to fold the sheet material. The method is aimed at the production of
small-scale prototypes of relatively thick sheet material, with minimal initial tooling. The analytical calculations were shown to underestimate the pressure required
during the manufacturing trials; a possible explanation can be sought in the more
complex deformations at the vertices of the fold pattern. Further experimental
and numerical studies are required to more accurately determine the folding energy and forces, as well as establish the desired pre-weakening along the fold lines
to ensure accurate folding without compromising the structural properties of the
folded sheet.

132

Chapter 7

Conclusions & Future Work


7.1

Conclusions

A novel type of shell structure was introduced and analysed, folded shell structures. These shell structures have a distinct structural hierarchy: globally they
can be regarded as thin-walled shells, but at a meso scale they are constituted
of tessellated unit cells, which in turn are composed of thin-walled shells joined
at distinct fold lines. It is this structural hierarchy that imbues the folded shell
structures with their interesting mechanical properties. Of specific interest is the
sheets ability to change their global Gaussian curvature, with purely developable
deformations at material level (i.e. no stretching). Furthermore, for the two example sheets considered throughout this thesis, the effective Poissons ratio was
oppositely signed for stretching and bending.
In the presented structural analysis, an important consideration was the multiscale modelling of the sheets. Global deformations of the folded shell structures
were mapped to deformations of the unit cells and its constituent thin-walled
facets; conversely, the possible deformations of the unit cells prescribe the global
sheet mechanics. For example, the oppositely signed Poissons ratios of the folded
sheets were shown to follow directly from the unit cell kinematics.
The manufacture of developable folded shell structures, i.e. those that can be
folded from flat sheet material, presents unique challenges. An extensive literature
review established the available manufacturing methods for these types of folded
sheets. A novel manufacturing method was presented, as well as the first known
analytical and experimental comparison of the required forming energy.
In summary, the following conclusions can be drawn:
folded shell structures are a promising approach for creating compliant surfaces, as they can undergo large strains and change their global Gaussian
curvature with only low-energy bending deformations at the material scale.
two example folded shell structures, the Miura and Eggbox sheet, were shown

133

7.1 Conclusions

to have an oppositely signed effective Poissons ratio for their in-plane and
out-of-plane deformation modes.
a crucial aspect of the folded shell structures is their structural hierarchy; the
global surface consists of tessellated unit cells, which in turn are composed
of thin-walled facets joined at distinct fold lines.
the observed global modes were described in terms of unit cell deformations;
the sheets planar kinematics involve no bending of the unit cell facets, only
opening and closing of the folds, whereas any out-of-plane deformations of
the sheets require bending of the facets.
the sheets change in global Gaussian curvature can be attained through
only developable deformations of the unit cells, and the oppositely signed
Poissons ratio of the sheets follows directly from the unit cell kinematics.
due to the constrained kinematics of the unit cells it is difficult to describe
the unit cell deformations at axes other than those chosen for convenience.
the mechanics of the folded shells were modelled using a simplified representation as a pin-jointed truss, with additional bending stiffness for the facets
and fold lines; these additional constraints were systematically introduced
into the classical pin-jointed truss formulations.
a modal analysis of the folded sheets showed that the salient deformation
modes (those that change the sheets global Gaussian curvature) are among
the softest, and therefore most dominant, eigenmodes.
the dominant mechanical behaviour of the folded shell structures was shown
to be primarily a result of the geometry, rather than material properties,
and insensitive to imperfections in the fold pattern.
the manufacture of developable folded sheets requires specialised manufacturing techniques, to enable folding with minimal material deformations. A
literature review revealed a range of manufacturing processes that have been
successfully demonstrated in prototypes as well as industrial settings.
three distinct categories of manufacturing methods were identified, synchronous folding (where folding takes place along all fold lines simultaneously), gradual folding (where the sheet gradually transitions from flat to
fully folded state) and pre-gathering methods (where the material is precorrugated transversely to the desired width before the final fold pattern is
imparted).
an important factor in the manufacturing techniques is the material deformation during the forming process and the accuracy of the final product;

134

7.2 Future Work

only qualitative information was found and the operating limits of the manufacturing processes therefore remain largely unknown.
a new manufacturing method was developed, using cold gas pressure to fold
relatively thick metal sheets into the Miura pattern with minimal initial
tooling.
the first known comparison of the analytically derived forming energy with
experimental values showed that, even for a simple folding process, the total
forming energy of a Miura sheet is not completely accounted for by a simple
hinge plasticity model. A likely explanation is found in the more complex
plastic deformations at the vertices, where various fold lines interact.

7.2

Future Work

The work described in this thesis represents a first analysis of folded shell structures, and it has opened a range of possible future research avenues.
Structural Analysis The next step in the structural analysis of folded shell
structures would be to combine the compatibility equations (i.e. the kinematic analysis) with a material model, and using Castiglianos theorem to
express the resulting stiffness components in conventional continuum mechanics formulations. Capturing the large displacement out-of-plane behaviour will be the predominant challenge.
The folded sheets can undergo significant planar strains due to the opening and closing of the fold lines, but this has hitherto not been sufficiently
quantified. Taking into account the detailed deformation at the fold lines, as
well as material fatigue, is necessary to establish the practically achievable
in-plane strains. An initial starting point would be singly-corrugated sheets
with distinct fold lines, before extending the analysis to more complex folded
shell structures.
In this thesis developable deformation of the unit cell facets has been presumed, and shown to capture the observed global sheet behaviour. However,
in order to fully describe the mechanics of folded shell structures and other
compliant shell mechanisms, more knowledge of near-developable deformations is required.
Manufacture The proposed cold gas pressure forming technique warrants further study; more experimental data is needed for a better estimate of the
discrepancy between the analytically derived and experimentally determined
forming pressures. Numerical simulation of the deformation of the vertices,

135

7.2 Future Work

where several fold lines intersect and interact, will provide the necessary insight into the additional plastic work. Also, springback of the sheet material
has so far not been taken into account.
In some of the continuous gradual folding manufacturing methods, the sheet
material buckles and collapses along pre-weakened fold lines into the desired
fold pattern. The degree of pre-weakening or embossing required to ensure
accurate and reliable folding is not yet established.
Applications One area of application of folded sheets is as impact absorbing
materials. The work in this thesis has led to the development of a novel
meso-structured material based on the stacking of Miura sheets. Its inherent
flexibility in some modes, whilst maintaining stiffness in others, makes it very
promising as core material for blast-resistant sandwich panels.
The ability to control the Kfacet /Kfold ratio for a folded sheet would enable
tailoring of the dominant deformation modes for different applications. One
approach is to locally modify the facet bending stiffness, either increasing it
by introducing a local texture, or decreasing it with a perforation pattern.
Scaling down the fold patterns for use in MEMS devices would enable selfassembly of the patterns using established micro-manufacturing techniques
and create a new range of micro-materials.
Fold Patterns In this thesis two example patterns have been used to illustrate
the mechanical properties of folded shell structures. Tachi (2010b) showed
that the Eggbox and Miura patterns can be combined into single DOF deployable mechanisms, which provides a means to tailor the mechanical behaviour across the sheets.
Different patterns may provide additional and more desirable mechanical
properties. For example, the waterbomb tessellations (Halloran, 2009) provide an inherently greater flexibility, by virtue of the tessellation of degree-6
vertices.
In all current studies the fold pattern remains unchanged during folding and
under mechanical loads, although additional fold lines may be introduced.
It was found that in some plastic models, the vertices would move plastically
through the sheet material, adding a hitherto unexplored degree of freedom.

136

References
Afolabi, D. & Mehmed, O. (1994). On Curve Veering and Flutter of Rotating
Blades. Journal of Engineering for Gas Turbines and Power , 116, 702709.
Akishev, N.I., Zakirov, I.M. & Nikitin, A.V. (2007). Device for sheet material corrugation. US Patent Application 20070098835.
Akishev, N.I., Zakirov, I.M. & Nikitin, A.V. (2009). Device for sheet material corrugation. US Patent 7487658.
Akishev, N.I., Alekseev, K.A., Zakirov, I.M. & Nikitin, A.V. (2010).
Panel of curvilinear shape and method of its manufacturing. Russian Patent
RU2381955.
Alderson, A., Alderson, K., Chirima, G., Ravirala, N. & Zied, K.
(2010). The in-plane linear elastic constants and out-of-plane bending of 3coordinated ligament and cylinder-ligament honeycombs. Composites Science
and Technology.
Alekseev, K.A., Zakirov, I.M. & Karimova, G.G. (2011). Geometrical
model of creasing roll for manufacturing line of the wedge-shaped folded cores
production. Russian Aeronautics (Iz VUZ), 54, 104107.
Allemang, R.J. (2003). The Modal Assurance Criterion Twenty Years of Use
and Abuse. Sound & Vibration, 1421.
Arne, C. (1973). Stress oriented corrugations. US Patent RE28534.
Balkcom, D.J. (2004). Robotic Origami Folding. Ph.D. thesis, Carnegie Mellon
University.
`s, E. (1993). High Modal Density, Curve Veering, Localization: A Different
Balme
Perspective On The Structural Response. Journal of Sound and Vibration, 161,
358363.
Baranger, E., Cluzel, C. & Guidault, P.A. (2011a). Modelling of the Behaviour of Aramid Folded Cores Up to Global Crushing. Strain, 47.
Baranger, E., Guidault, P.A. & Cluzel, C. (2011b). Numerical modeling of
the geometrical defects of an origami-like sandwich core. Composite Structures,
93, 2504 2510.

137

Basily, B. & Elsayed, E. (2004). Dynamic axial crushing of multilayer core


structures of folded Chevron patterns. International Journal of Materials and
Product Technology, 21, 169185.
Basily, B., Elsayed, A. & Kling, D. (2006). Technology for continuous folding
of sheet materials. US Patent 7115089.
Bassik, N., Stern, G.M. & Gracias, D.H. (2009). Microassembly based on
hands free origami with bidirectional curvature. Applied Physics Letters, 95.
Behrens, A.W. & Ellert, J. (2005). Buckling Texturing Technology for Increase in Stability of Thin Sheet Metal Structures Simulation and Application.
Advanced Materials Research, 68, 623630.
Belcastro, S. & Hull, T.C. (2002). Modelling the folding of paper into three
dimensions using affine transformations. Linear Algebra and its Applications,
348, 273282.
Briassoulis, D. (1986). Equivalent orthotropic properties of corrugated sheets.
Computers & Structures, 23, 129138.
Brunner, A. (1968). Expansible surface structure. US Patent 3362118.
Buri, H.U. (2010). Origami - Folded Plate Structures. Ph.D. thesis, lUniversite
de Lausanne.
Calladine, C.R. (1983). Theory of Shell Structures. Cambridge University Press.
Daynes, S., Weaver, P.M. & Potter, K.D. (2009). Aeroelastic Study of
Bistable Composite Airfoils. Journal of Aircraft, 46, 21692174.
de Swart, J. (1955). Panel Material. US Patent 2858247.
Demaine, E.D. & ORourke, J. (2007). Geometric folding algorithms: linkages,
origami, polyhedra. Cambridge University Press.
Demaine, E.D., Demaine, M.L., Hart, V., Price, G.N. & Tachi, T. (2009).
(Non)existence of Pleated Folds: How Paper Folds Between Creases. In Abstracts from the 7th Japan Conference on Computational Geometry and Graphs
(JCCGG 2009), Kanazawa, Ishikawa, Japan.
Deshpande, V. & Fleck, N. (2003). Energy absorption of an egg-box material.
Journal of the Mechanics and Physics of Solids, 51, 187208.
Drechsler, K. & Kehrle, R. (2004). Manufacturing of folded core-structures
for technical applications. In Proceedings of the 25th International SAMPE Europe Conference, 508513, Paris.

138

du Bois, J.L., Adhikari, S. & Lieven, N.A.J. (2009). Eigenvalue curve veering
in stressed structures: An experimental study. Journal of Sound and Vibration,
322, 45.
Dunajeff, L.A. (1939). Resilient Sheet. US Patent 2158929.
Dunajeff, L.A. (1941). Resilient Sheet. US Patent 2233592.
Duncan, J.P. & Duncan, J.L. (1982). Folded Developables. Proceedings of the
Royal Society of London. Series A, Mathematical and Physical Sciences, 383,
191205.
Duncan, J.P. & Upfold, R.W. (1963). Equivalent elastic properties of perforated bars and plates. Journal of Mechanical Engineering Science, 5, 5365.
Dureisseix, D. (2011). An overview of patterns and mechanisms with origami,
accepted for publication in the International Journal of Space Structures.
Dureisseix, D., Gioia, F., Motro, R. & Maurin, B. (2011). Conception
dune Enveloppe Plisse Pliable-Dpliable. In 10e Colloque National en Calcul
des Structures (CSMA2011), Giens, France.
Elsayed, A. & Basily, B. (2010). Technology for continuous folding of sheet
materials into a honeycomb-like configuration. US Patent 7758487.
Elsayed, E.A. & Basily, B. (2004). A continuous folding process for sheet
materials. International Journal of Materials and Product Technology, 21, 217
238.
Engel, H. (1968). Structure Systems. Praeger, foreword by Ralph Rapson.
Ewald, W. (1948). Reinforced Structural Sheet. US Patent 2441476.
Farmer, G.D. & Spangler, W.B. (1962). Investigation of Calottan Sheet Stiffening Process. Tech. rep., Army Engineer Research and Development Labs Fort
Belvoir.
Fischer, S., Drechsler, K., Kilchert, S. & Johnson, A. (2008). Mechanical
Tests for Foldcore Base Material Properties. In CompTest 2008 - Composite
Testing and Model Identification, Dayton, Ohio, USA.
Fischer, S., Drechsler, K., Kilchert, S. & Johnson, A. (2009). Mechanical
tests for foldcore base material properties. Composites Part A: Applied Science
and Manufacturing, 40, 19411952.
Fort, P.L. (1970). Bending of perforated plates with square penetration patterns.
Nuclear Engineering and Design, 12, 122134.

139

French, M.J. & Petty, J.W.L. (1965). Extensible metal sheets. US Patent
3184094.
Fritz, P.J., Tice, D.A. & Rabb, L.R. (1996). Metal liner for a fiber-reinforced
plastic tank. US Patent 5535912.
Fuchs, D. & Tabachnikov, S. (1999). More on Paperfolding. The American
Mathematical Monthly, 106, 2735.
Gauss, C.F. (1828). Disquisitiones Generales Circa Superficies Curvas.
Gauss, C.F. (1902). General investigations of curved surfaces of 1827 and 1825..
Translation with notes and a bibliography by J. C. Morehead and A. M. Hiltebeitel.
Gewiss, L.V. (1959). Constructional sheet material and a machine for producing
the same. GB Patent GB815523.
Gewiss, L.V. (1960). Arrangement for the mechanical and continuous production
of developable herring-bone structures. US Patent 2950656.
Gewiss, L.V. (1964). Process and apparatus for forming transverse corrugations
of all forms in a sheet or band of malleable material. US Patent 3150576.
Gewiss, L.V. (1968). Process and devices for chevroning pliable sheet material.
US Patent 3397261.
Gewiss, L.V. (1976). Sheet with alternate protrusions and recesses. US Patent
3992162.
Gibson, L.J. & Ashby, M.F. (1999). Cellular Solids, Structure and Properties.
Cambridge University Press, 2nd edn.
Girot, P. (1964). Connecting element for expansion joints. US Patent 3118523.
Grima, J.N., Alderson, A. & Evans, K.E. (2005). Auxetic behaviour from
rotating rigid units. Physica Status Solidi (b) Solid State Physics, 242, 561
575.
Guest, S. & Pellegrino, S. (2006). Analytical models for bistable cylindrical
shells. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 462, 839854.
Guest, S.D. (2006). The stiffness of prestressed frameworks: A unifying approach. International Journal of Solids and Structures, 43, 842854.
Halloran, E. (2009). Concepts and Modeling of a Tessellated Molecule Surface.
In R.J. Lang, ed., Origami 4, Fourth International Conference on Origami in
Science, Mathematics, and Education (4OSME), 305314, A K Peters.

140

Halpin, J.C. (1992). Primer on Composite Material Analysis - Second Edition,


Revised . Technomic Publishing Co., Inc., 2nd edn.
Heimbs, S. (2009). Virtual testing of sandwich core structures using dynamic
finite element simulations. Computational Materials Science.
Heimbs, S., Cichosz, J., Klaus, M., Kilchert, S. & Johnson, A. (2010).
Sandwich structures with textile-reinforced composite foldcores under impact
loads. Composite Structures, 92, 14851497.
Hilbert, D. & Cohn-Vossen, S. (1952). Geometry and the Imagination. AMS
Chelsea Publishing, 2nd edn., translated by P.Nemenyi.
Hochfeld, H. (1959). Process and Machine for Pleating Pliable Structures. US
Patent 2901951.
Huffman, D.A. (1976). Curvatures and Creases: A Primer on Paper. IEEE
Transactions on Computers, C-25, 10101019.
Ichikawa, H. (1995). Method of production of reinforced composite corrugated
body and method of formation of corrugating rollers for use therein. US Patent
5443779.
InteSpring (2011). Company website. Last retrieved on 30 August 2011,
http://www.intespring.nl/.
Ishinabe, M., Nishimura, Y., Imazu, K., Kobayashi, S. & Matsubayashi,
H. (1992). Packing can. US Patent 5100017.
Johnson, W. & Yu, T.X. (1980). The angle of fold and the plastic work done
in the folding of developable flat sheets of metal. Journal of Mechanical Engineering Science, 22, 233241.
Junkers, H. (1929). Corrugated sheet-metal shape. US Patent US1704326.
Kangwai, R.D., Guest, S.D. & Pellegrino, S. (1999). An introduction to
the analysis of symmetric structures. Computers & Structures, 71, 671688.
Kawasaki, T. (1994). R()=I. In K. Miura, ed., Origami Science and Art: Proceedings of the Second International Meeting of Origami Science and Scientific
Origami .
Kehrle, R. (2005). Method and apparatus for producing a composite structural
panel with a folded material core. US Patent 6913570.
Khaliulin, V., Dvoeglazov, I. & Inkin, V. (2002). Isometric forming of
relief plates using vacuum energy. Kuznechno-Shtampovochnoe Proizvodstvo
(Obrabotka Metallov Davleniem), 1724.

141

Khaliulin, V.I. (2005). A Technique for Synthesizing the Structures of Folded


Cores of Sandwich Panels. Russian Aeronautics (Iz VUZ), 48, 712.
Khaliulin, V.I. & Batrakov, V.V. (2005). Technological Schemes of Zigzag
Crimp Shaping. Russian Aeronautics (Iz VUZ), 48, 6873.
Khaliulin, V.I. & Desyatov, V.E. (1991). Device for corrugation of sheet
material. Soviet Union Patent SU1690903.
Khaliulin, V.I. & Desyatov, V.E. (1993). Method of making a multilayer
panel of curvilinear shape. Soviet Union Patent SU1830326.
Khaliulin, V.I. & Dvoeglazov, I.V. (1998). Apparatus for corrugating sheet
material. Russian Patent RU2118217.
Khaliulin, V.I. & Dvoeglazov, I.V. (2001). On technological problems of fabrication of relief designs by isometric transformation of thin sheet. Transactions
of Nanjing University of Aeronautics & Astronautics, 18, 1116.
Khaliulin, V.I. & Skripkin, E.A. (1997). Method for making zigzag corrugation (variants) and apparatus for performing the same. Russian Patent
RU2100120.
Khaliulin, V.I., Batravkov, V.V., Dvoeglazov, I.V. & Menjashkin,
D.G. (2005a). Method of manufacture of a relief detail with a zigzag-shaped
channeled structure (versions). Russian Patent RU2264918.
Khaliulin, V.I., Dvoeglazov, I.V., Menjashkin, D.G. & Batrakov, V.V.
(2005b). Method of production of a filler with a zigzag corrugated structure.
Russian Patent RU2256556.
Khaliulin, V.I., Batrakov, V.V., Dvoeglazov, I. & Menjashkin, D.G.
(2007). Method for making parts with zigzag corrugated structure. Russian
Patent RU2303501.
Klett, Y. & Drechsler, K. (2009). Cutting Edge Cores: Multifunctional Core
Structures. In 2009 DGLR Luft- und Raumfahrtkongress.
Klett, Y. & Drechsler, K. (2010). Designing Technical Tessellations. In
P. Wang-Iverson, R.J. Lang & M. YIM, eds., Origami 5: Fifth International
Meeting of Origami Science, Mathematics, and Education (5OSME), CRC
Press.
Klett, Y., Drechsler, K., Kolax, M., Wentzel, H. & Kehrle, R. (2007).
Design of multifunctional folded core structures for aerospace sandwich applications. In Proceedings of 1st CEAS European air and space conference, 903908,
Berlin.

142

Kling, D.H. (1997). Doubly Periodic Flat Surfaces in Three-Space. Ph.D. thesis,
Rutgers, The State University of New Jersey.
Kling, D.H. (2005). Patterning technology for folded sheet structures. US Patent
6935997.
Kling, D.H. (2007a). Folding Method and Apparatus. US Patent Application
20070273077.
Kling, D.H. (2007b). Folding methods, structures and apparatuses. US Patent
Application 20070023987.
Knapp, R.H. (1977). Pseudo-Cylindrical Shells: a New Concept for Undersea
Structures. Transactions of the ASME Journal of Engineering for Industry, 99,
485492.

Kokotsakis, A. (1933). Uber


bewegliche Polyeder. Mathematische Annalen,
107, 627647.
Kovacs, F. (2011). Extended Truss Theory with Simplex Constraints. International Journal of Solids and Structures, 48, 472482.
Kumar, P. & Pellegrino, S. (2000). Computation of kinematic paths and
bifurcation points. International Journal of Solids and Structures, 37, 7003
7027.
Kuribayashi, K., Tsuchiya, K., You, Z., Tomus, D., Umemoto, M., Ito,
T. & Sasaki, M. (2006). Self-deployable origami stent grafts as a biomedical application of Ni-rich TiNi shape memory alloy foil. Materials Science and
Engineering: A, 419, 131137.
Lakes, R. (1987). Foam Structures with a Negative Poissons Ratio. Science,
235, 1038 1040.
Lanczos, C. (1986). The Variational Principles of Mechanics. Dover Books on
Physics & Chemistry, Dover Publications.
e, A. & Sab, K. (2010). Transverse shear stiffness of a chevron folded core
Lebe
used in sandwich construction. International Journal of Solids and Structures,
47, 2620 2629.
Lim, T.C. (2007). On simultaneous positive and negative Poissons ratio laminates. Physica Status Solidi (b) Solid State Physics, 244, 910 918.
Lobkovsky, A., Gentges, S., Li, H., Morse, D. & Witten, T.A. (1995).
Scaling Properties of Stretching Ridges in a Crumpled Elastic Sheet. Science,
270, 14821485.

143

Lueke, J.C. (1994). Tension and compression extensible liner for a primary vessel.
US Patent 5292027.
Ma, J. & You, Z. (2011). The Origami Crash Box. In P. Wang-Iverson, R.J. Lang
& M. YIM, eds., Origami 5: Fifth International Meeting of Origami Science,
Mathematics, and Education (5OSME), CRC Press.
Mahadevan, L. & Rica, S. (2005). Self-Organized Origami. Science, 307, 1740.
Mansfield, E.H. (1989). The Bending and Stretching of Plates. Cambridge University Press, 2nd edn.
Marciniak, Z. & Duncan, J.L. (1992). The Mechanics of Sheet Metal Forming.
Edward Arnold.
McKay, D.M. (1984). Structural systems for panels, boards, shelves, and laminates. UK Patent GB2123874.
Mirtsch, F., Weinert, N., Pech, M. & Seliger, G. (2006). Vault Structures
Enabling Sustainable Products. In 13th CIRP International Conference on Life
Cycle Engineering (LCE2006), 629633, Leuven, Belgium.
Miura, K. (1969). Proposition of Pseudo-Cylindrical Concave Polyhedral Shells.
ISAS Report, 34, 141163, institute of Space and Aeronautical Science, University of Tokyo report, no 442.
Miura, K. (1972). Zeta-Core Sandwich-Its Concept and Realization. ISAS Report, 37, 137164, institute of Space and Aeronautical Science, University of
Tokyo report, no 480.
Miura, K. (1975a). Cores of sandwich construction. GB Patent GB1390132.
Miura, K. (1975b). New Structural Form of Sandwich Core. Journal of Aircraft,
12, 437441.
Miura, K. (1989). A Note on Intrinsic Geometry of Origami. In 1st International
Conference of Origami Science and Technology, Ferrara, Italy.
Miura, K. (2002). PCCP Shells. In H.R. Drew & S. Pellegrino, eds., New Approaches to Structural Mechanics, Shells and Biological Structures, 329339,
Kluwer Academic Publishers.
Mornement, A. & Holloway, S. (2007). Corrugated Iron: Building on the
Frontier . Frances Lincoln ltd.
Nettles, A.T. (1994). Basic Mechanics of Laminated Composite Plates. NASA
Reference Publication 1351, NASA.

144

Norman, A.D. (2009). Multistable and morphing corrugated shell structures.


Ph.D. thesis, Cambridge University.
Norman, A.D., Golabchi, M.R., Seffen, K.A. & Guest, S.D. (2008a).
Multistable Textured Shell Structures, vol. 54 of Advances in Science and Technology, 168173. SciencePress.com, Acireale, Sicily, Italy.
Norman, A.D., Seffen, K.A. & Guest, S.D. (2008b). Multistable Corrugated
Shells. Proceedings of the Royal Society: Mathematical, Physical & Engineering
Sciences, 464, 16531672.
Norman, A.D., Seffen, K.A. & Guest, S.D. (2009). Morphing of curved
corrugated shells. International Journal of Solids and Structures, 46, 1624
1633.
Nye, J.F. (1957). Physical Properties of Crystals. Oxford University Press, 1st
edn.
Olympio, K.R. (2009). Compliant Load-Bearing Skins and Structures for Morphing Aircraft Applications. Ph.D. thesis, Pennsylvania State University.
Paterson, J.G.T. (1999). Method and apparatus for folding sheet materials with
tessellated patterns. US Patent 5947885.
Pellegrino, S. & Calladine, C.R. (1986). Matrix analysis of statically and
kinematically indeterminate frameworks. International Journal of Solids and
Structures, 22, 409428.
Perkins, N. & Mote, C. (1986). Comments on curve veering in eigenvalue
problems. Journal of Sound and Vibration, 106, 451463.
Pfistershammer, J. (1952). Honeycomb-type structural materials and method
of making same. US Patent 2738297.
Pickett, G.T. (2007). Self-folding origami membranes. Europhysics Letters, 78,
48003.
Pierre, C. (1988). Mode localization and eigenvalue loci veering phenomena in
disordered structures. Journal of Sound and Vibration, 126, 485502.
Prall, D. & Lakes, R. (1997). Properties of a chiral honeycomb with a poissons
ratio of 1. International Journal of Mechanical Sciences, 39, 305314.
Press, W.H., Teukolsky, S.A., Vetterling, W.T. & Flannery, B.P.
(2007). Numerical Recipes : The Art of Scientific Computing. Cambridge University Press, 3rd edn.

145

Ramrakhyani, D., Lesieutre, G., Frecker, M. & Bharti, S. (2005). Aircraft Structural Morphing using Tendon-Actuated Compliant Cellular Trusses.
Journal of Aircraft, 42, 16141620.
Rapp, E.G. (1960). Sandwich-type structural element. US Patent 2963128.
Rauscher, B., Karbasian, H., Tekkaya, A.E. & Homberghachen, W.
(2008). Manufacturing of hump plate walls by innovative forming techniques
and fields of application in commercial lightweight utility vehicles. In SCT 2008
- Steels in cars and trucks, Wiesbaden, Germany.
Resch, R. & Christiansen, H.N. (1971). Kinematic Folded Plate System. In
R. Krapfenbauer, ed., Proceedings of IASS Symposium on Folded Plates and
Prismatic Structures, IASS, Vienna, Austria.
Salvadori, M. & Heller, R. (1963). Structure in Architecture. Prentice-Hall.
Samanta, A. & Mukhopadhyay, M. (1999). Finite element static and dynamic
analyses of folded plates. Engineering Structures, 21, 277287.
Schenk, M. & Guest, S.D. (2011). Origami Folding: A Structural Engineering Approach. In P. Wang-Iverson, R.J. Lang & M. YIM, eds., Origami 5:
Fifth International Meeting of Origami Science, Mathematics, and Education
(5OSME), 293305, CRC Press.
Schenk, M., Allwood, J.M. & Guest, S.D. (2011). Cold Gas-Pressure Folding
of Miura-ori Sheets, accepted for presentation at the International Conference on
Technology of Plasticity (ICTP 2011); selected for publication in Steel Research
International.
Schief, W.K., Bobenko, A.I. & Hoffman, T. (2007). On the Integrability
of Infinitesimal and Finite Deformations of Polyhedral Surfaces. In Discrete
Differential Geometry, Oberwolf Seminars, vol. 38, 6793, Birkhauser Verlag.
Schott, L.A. (1975). Reinforced cellular panel construction. US Patent 3876492.
Seffen, K.A. (2006). Mechanical memory metal: a novel material for developing
morphing engineering structures. Scripta Materialia, 55, 411414.
Seffen, K.A. (2007). Hierarchical multi-stable shapes in mechanical memory
metal. Scripta Materialia, 56, 417420.
Seffen, K.A. (2011). Compliant Shell Mechanisms, accepted for Philosophical
Transactions of the Royal Society - Special Theme issue on the Geometry and
Mechanics of Layered Structures and Materials.
Sehrndt, G.A. (1960). Verbundplatte und Verfahren zu ihrer Herstellung. DD
Patent 1434112.

146

Shabalin, L., Sidorov, I. & Khaliulin, V. (2010). Simulation of Z-crimp


shaping with the use of the ansys finite element software. Russian Aeronautics
(Iz VUZ), 53, 339344.
Smith, C.W., Wootton, R.J. & Evans, K.E. (1999). Interpretation of experimental data for Poissons ratio of highly nonlinear materials. Experimental
Mechanics, 39, 356362.
Stachel, H. (2009). Remarks on Miura-Ori, a Japanese Folding Method. In
ICEGD 2009, International Conference on Engineering Graphics and Design.
Stachel, H. (2010a). A kinematic approach to Kokotsakis meshes. Computer
Aided Geometric Design, 27, 428 437.
Stachel, H. (2010b). On the Flexibility of Kokotsakis Meshes. In XVI Geometrical Seminar , Vrnjacka Banja/Serbia.
Strang, G. (2006). Linear Algebra and its Applications. Thomson, 4th edn.
Streinu, I. & Whiteley, W. (2005). Single-Vertex Origami and Spherical Expansive Motions. Discrete and Computational Geometry, 3742/2005, 161173.
Struik, D.J. (1961). Lectures on Classical Differential Geometry. Dover Publications, 2nd edn.
Tachi, T. (2009a). Generalization of Rigid Foldable Quadrilateral Mesh Origami.
Journal of the International Association for Shell and Spatial Structures, 50,
173179.
Tachi, T. (2009b). One-DOF Cylindrical Deployable Structures with Rigid
Quadrilateral Panels. In A. Domingo & C. Lazaro, eds., Proceedings of the
International Association for Shell and Spatial Structures (IASS) Symposium
2009 , 22952305, Universidad Politecnica de Valencia, Spain.
Tachi, T. (2009c). Simulation of Rigid Origami. In R.J. Lang, ed., Origami 4:
Fourth International Meeting of Origami Science, Mathematics, and Education
(4OSME), 175188, A K Peters.
Tachi, T. (2010a). Architectural Origami - Architectural Form Design Systems
based on Computational Origami. Lecture notes for MIT 6.849, Fall 2010.
Tachi, T. (2010b). Freeform Rigid-Foldable Structure using Bidirectionally FlatFoldable Planar Quadrilateral Mesh. In C. Ceccato, L. Hesselgren, M. Pauly,
H. Pottmann & J. Wallner, eds., Advances in Architectural Geometry 2010 ,
87102, Springer Vienna.

147

Tachi, T. (2010c). Geometric Considerations for the Design of Rigid Origami


Structures. In Proceedings of the International Association for Shell and Spatial
Structures (IASS) Symposium 2010 , Shanghai, China.
Talakov, M. (2010). Investigation of folded structure geometry with double curvature. Russian Aeronautics (Iz VUZ), 53, 334338.
Tanizawa, K. & Miura, K. (1975). Stress Analysis of a Concave Polyhedral
Shell. ISAS report, 40, 3960, institute of Space and Aeronautical Science,
University of Tokyo report, no 523.
Tanizawa, K. & Miura, K. (1978). Large Displacement Configurations of BiAxially Compressed Infinite Plate. Transactions of the Japan Society for Aeronautical and Space Sciences, 20, 177187.
Tewes, W.A. (1966). Structural Characteristics of Sheet Metal Embossed with
an Irregular Design. Tech. Rep. MEL-111/6, Navy Marine Engineering Lab
Annapolis.
Thill, C., Etches, J., Bond, I., Potter, K. & Weaver, P. (2008). Morphing
skins - A review. The Aeronautical Journal , 112, 117138.
Timoshenko, S. (1964). The Theory of Plates and Shells. McGraw-Hill.
Watanabe, N. & Kawaguchi, K. (2009). The Method for Judging Rigid Foldability. In R.J. Lang, ed., Origami 4: Fourth International Meeting of Origami
Science, Mathematics, and Education (4OSME), 165174, A K Peters.
Witten, T.A. (2007). Stress focusing in elastic sheets. Reviews of Modern
Physics, 79, 643675.
Wu, W. & You, Z. (2010). Modelling rigid origami with quaternions and dual
quaternions. Proceedings of the Royal Society A, 466, 21552174.
Yokozeki, T., ichi Takeda, S., Ogasawara, T. & Ishikawa, T. (2006). Mechanical properties of corrugated composites for candidate materials of flexible
wing structures. Composites Part A: Applied Science and Manufacturing, 37,
15781586.
Yoshimura, Y. (1955). On the mechanism of buckling of a circular cylindrical
shell under axial compression. NACA Technical Memorandum 1390 NACA-TM1390, NACA.
Zakirov, I., Katayev, Y. & Nikitin, A. (2004). Folded-Plate Structures Alteration When Bent. Sci Tech J Polyot, 5053.

148

Zakirov, I., Nikitin, A., Alekseev, K.A. & Mudra, C. (2006). Folded Structures: Performance, Technology and Production. In Latest Advancements of
Applied Composite Technologies Proceedings of the 27th International SAMPE
Europe Conference 2006 , 234239, Society for the Advancement of Materials
and Process Engineering (SAMPE), Paris.
Zakirov, I., Paimushin, V.N. & Alekseev, K. (2010). Composite Structures
from Polymer Matrix Materials for the Aerospace Sector. In Proceedings of the
3rd Congress for the Polymer Composite Materials Technologies, Modena, Italy.
Zakirov, I.M. & Alekseev, K.A. (2007). Parameters of a creasing-bending
machine as applied to the scheme of transverse rotary shaping of chevron structures. Russian Aeronautics (Iz VUZ), 50, 186192.
Zakirov, I.M., Nikitin, A.V., Akishev, N.I. & Karimullin, N.V. (2008).
Method for production of chevron filler and device for its realisation. Russian
Patent RU2341347.
Zakirov, I.M., Nikitin, A.V., Akishev, N.I. & Movchan, G.V. (2009).
Device for continuous corrugating of rolled material. Russian Patent RU2357828.
Zupan, M., Chen, C. & Fleck, N.A. (2003). The plastic collapse and energy
absorption capacity of egg-box panels. International Journal of Mechanical Sciences, 45, 851871.

149

Вам также может понравиться