Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/367167560

Statics of integrated origami and tensegrity systems

Preprint · January 2023

CITATIONS READS

0 220

4 authors, including:

Ma Shuo Muhao Chen


Zhejiang University of Technology Texas A&M University
18 PUBLICATIONS 126 CITATIONS 39 PUBLICATIONS 196 CITATIONS

SEE PROFILE SEE PROFILE

Robert E Skelton
University of California, San Diego
523 PUBLICATIONS 17,253 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Tensegrity Approaches to In-Space Construction of a 1-g Growable Habitat View project

Tensegrity construction of space structures View project

All content following this page was uploaded by Muhao Chen on 16 January 2023.

The user has requested enhancement of the downloaded file.


Statics of integrated origami and tensegrity systems

Shuo Maa , Muhao Chenb,∗, Hongying Zhangc , Robert E. Skeltonb


a Collegeof Civil Engineering, Zhejiang University of Technology, 310023, China
b Department of Aerospace Engineering, Texas A&M University, College Station, TX, 77840, USA
c Department of Mechanical Engineering, National University of Singapore, 119077, Singapore

Abstract
This study presents an explicit form of the static equilibrium equations of integrated origami and tensegrity systems. The analytical
approach allows one to model and analyze the isolated origami and tensegrity paradigms as a whole system. The tensegrity
and origami members are described by the nodal coordinates and hinge angels between the origami panels. The nonlinear static
equations of the integrated system are derived by the Lagrangian method. By Taylor’s expansion theory, we also presented its
linearized form. The developed approach is capable of conducting the following comprehensive statics studies for any integrated
origami and tensegrity systems: 1) Performing loading analysis, where bars and strings can have elastic or plastic deformations.
2) Conducting infinitesimal and large deformation analysis, which is helpful in understanding the stress in structural members and
actuation strategies. 3) Dealing with various kinds of boundary conditions, for example, fixing or applying static loads at any
nodes in any direction (i.e., gravitational force, some specified forces, or moments). 4) Conducting stiffness analysis, including
eigenvalues and their modes. Three examples, a Miura origami unit, an integrated tensegrity origami shelter, and a cable-driven
Kresling structure, are carefully selected and studied. This study provides a deep insight into structures, materials, as well as
performances. The integration idea also promotes our ability to design and build deployable structures at large.
Keywords: Origami, Tensegrity, Statics equilibrium, Deployable structures, Integrated origami and tensegrity structures

1. Introduction shapes [10, 11]. 4). Metamaterials. One can neat the bar-string
patterns to get materials with metamaterial properties [12, 13].
In the field of deployable structures and soft robots, there 5). Shape control. One can actuate the strings to achieve large
are two leading structure paradigms: tensegrity and origami, morphing objectives [14, 15]. The cons of tensegrity are 1).
that have attracted significant interest from researchers. Both There is no cover for the bar-string network. One can add mem-
of them have shown great potential in engineering future branes, shells, and sheets to the bar-string network. But related
lightweight systems. To help understand the fundamental idea research is still quite limited. 2). The joints are complicated to
of integrating the two paradigms, we first introduce a little bit make. Since one node may connect several bars, and each bar
of each paradigm and its advantages and limitations. can have a different degree of freedom requirements, the joints
The tensegrity art form was first created by Ioganson (1921) are usually non-standard types.
and Snelson (1948) [1], which denotes a network of compress-
ible (bars/struts) and tensile (strings/cables) members. Buck- The origami art form or commonly known as paper folding
minster Fuller coined the two words: tensile and integrity in is probably as old as the invention of paper itself by the ancient
1959 [2]. After decades of study, tensegrity has shown its Chinese, and the ancient Japanese combined the two words oru
great capacity in designing lightweight structures, deployable (means fold) and kami (namely paper) [16]. Origami was pre-
infrastructures, and robotics, such as space platforms [3], cable viously used as an art form with lucky meanings for celebrating
domes [4], space landers [5], etc. To summarize the strengths weddings and festivals in Asia. It was not until the 1950s that
and weaknesses of tensegrity, we know the pros of tensegrity origami gained popularity worldwide through the famous artist
systems are: 1). Mass efficiency. All one-dimensional struc- Akira Yoshizawa, who created more than 50,000 models and
tural members are axially loaded. One can place the struc- wrote 18 books on the art of origami [16]. After decades of re-
ture elements following the load path to achieve lightweight search, origami has been used to make space solar panels [17],
designs [6, 7]. 2). Actuate model. There is no bar bending, energy absorbers [18], robotics [19], etc. We also summarize
which brings accurate models. Accurate models can obtain the plus and minus points of origami. The pros of origami sys-
more precise control [8, 9]. 3). Tunable structure parameters. tems are: 1). Complicated shapes. The model is easy to achieve
One can adjust the structure parameters (i.e., prestress, length with complex folding shapes [20, 21]. 2). Compact size. The
of strings) to achieve various equilibrium states, stiffness, and 2D sheets can be easily stowed in small sizes and then deployed
to a shape with large volume [22, 23]. 3). Low-cost manufac-
∗ Correspondingauthor.+1 979-985-8285 turing. One can use 2D sheets to make 3D, or 4D structures
Email address: muhaochen@tamu.edu (Muhao Chen) [24, 25], which can also significantly simplify the assembly
Preprint submitted to International Journal of Solids and Structures January 16, 2023
process. 4). Metamaterial properties. Many origami structures
have tunable negative Poisson’s ratio, structural bistability, and
self-locking properties [26, 27]. The cons are 1). The origami
structure is usually of low stiffness. One can add actuators or
lock mechanisms to increase the stiffness, but it requires more
effort. 2). The transient dynamics during the shaping morphing
are challenging to control. Actuating the origami’s hinges is
possible to solve the problem partially, but it would complicate
the design.
It is natural to ask the question: tensegrity or origami, which
is a more efficient structure paradigm? On one side, people hold
the idea that tensegrity is more fundamental. Many nature sys-
tems implement this technology, i.e., cells [28], elbows [29],
spider fibers [30], etc. On the other side, some people believe
that origami is better. There are also nature pieces of evidence
Figure 1: An illustration of the bar-and-hinge model: (a) The bar-and-hinge
that bio-structures follow such mechanisms. i.e., flowers [31], model of a Miura-ori unit cell. It is composed of 5 nodes and 8 bars. It has four
horse-chestnut leaves [32], earwig wings [33], etc. Thus, inte- panels labeled as numbers with circles, and four hinges connect four panels.
grating the two paradigms may benefit each individual instead (b) a hinge connects two triangular panels. i, j, k, and l are the indices of the
of making the either-or decision. A little work has been con- nodes. m and n are the normal vectors of the two panels, (c) the cross-section
of the hinge p. The thickness and the length of the hinge p are labeled as th,p
ducted to integrate the two systems. To name a few, Rohmer and lh,p . The rotation angles and moment are denoted as θ p and M p , (d) the
et al. presented an experimental study of shape memory alloy- bending strain on the hinge p. We assume a linear curve about the neutral axis,
based tensegrity cylinder with an origami structure [34]. Mi- the strain ϵ = (θ p − θ0,p )x, where x here is the distance about the neutral axis.
randa et al. showed active solar facades using tensegrity D-
Bar to actuate origami sunscreens for energy-efficient build-
ings [35]. Park designed a geodesic dome for Martian agricul- method with nodal vector as the generalized coordinates. For
ture based on the combined tensegrity and origami units [36]. origami, recent analytical work by [38, 39] paved the way for
However, the analytical work of the integrated system is still modeling any origami structures. The critical assumption about
very limited. This paper bridges the two paradigms by de- the origami the papers made is the bar-and-hinge model, where
riving the governing equations of the integrated system. This the origami panels can be modeled as the bars and hinges. The
paper proposes an explicit form of the static equilibrium equa- idea is shown in Figure 1. Let us take a close look at this model.
tions approach that allows one to model and analyze the isolated When the rotational stiffness of the hinges is zero, one can view
origami and tensegrity paradigms as a whole system. this bar-and-hinge model as a truss. In other words, the bar-
The following paper is structured as follows. Section 2 gives and-hinge model is basically the combination of a truss model
the notations and assumptions. Section 3 derives the geome- with multiple hinges. Since the truss model is just a particu-
try and kinematics of the integrated origami and tensegrity sys- lar case of the tensegrity model, now one can see the connec-
tems. Section 4 provides the nonlinear and linearized static tion between the tensegrity and origami structures. Thus, based
equilibrium equations by the Lagrangian method. Section 5 on the tensegrity model, we added rotational angles to describe
demonstrates three examples to verify the proposed methods. the hinge kinematics to formulate the integrated origami and
Section 6 summarizes the conclusions. tensegrity systems. Moreover, in such a way, we can write an
explicit form of the integrated form to make the modeling and
computing much more effortless and benefit prestress design,
2. Notations and Assumptions stiffness design, and shape control.
2.1. Bar-and-hinge model As shown in Figure 1(a), the configuration of the bar-and-
hinge model is determined by the nodal coordinates of the end
In order to describe the equation of motion of the integrated nodes of bar elements. Let there be nn number of nodes, the X-,
origami and tensegrity system, we have to identify a proper way 3
to model the two paradigms, tensegrity and origami, as a whole h iT of the ith node ni ∈ R in the vector form
Y-, and Z-coordinates
is ni = xi yi zi . By stacking ni for i = 1, 2, · · · , nn , we
system. Let us first look at each of them. Recent theoretical
work by [9, 37] probably provided one of the most accessible can get the nodal vector n ∈ R3nn for the whole structure:
ways to explicitly model the tensegrity structures that allows h iT
one to 1) perform rigid body dynamics with acceptable errors, n = nT1 nT2 ··· nTnn . (1)
2) simulate FEM statics and dynamics accurately with elastic or
plastic materials, 3) deal with various kinds of boundary condi- Sometimes, the nodal coordinates of the structure are written in
tions (i.e., gravitational force, specified forces, or arbitrary vi- a useful matrix form, called nodal matrix N ∈ R3×nn :
brations), 4) conduct accurate modal analysis, 5) analyze struc-
tures with clustered strings, and 6) study tensegrity, truss, and h i
membranes. The approach is derived based on the Lagrangian N = n1 n2 ··· nnn . (2)

2
2.2. Truss elements ith column of Ca ) can be written as:
In the integrated system, we use truss elements to model the  
 j 
bars and strings from tensegrity structures and the bars from
[Ca ](:,i) =  k  . (6)
 
origami’s bar-and-hinge model. The strings in the tensegrity
m
 
part will benefit the stiffness and shape control of the whole
structure by tuning their prestress and length.
For example, the panels in Figure 1(a) corresponds to the fol-
lowing:
2.2.1. Connectivity matrix of truss
The truss topology can be given by the index information 
5 5 5 5

of all the nodes, labeled by a connectivity matrix C ∈ Rne ×nn , Ca = 2

1 3 4 .

(7)
where ne is the total number of line segments (bars and strings 
1 3 4 2

from node to node) in the structure. Suppose the ith segment
is from node j to node k ( j and k = 1, 2, · · · , nn ). Then, the
It should be noted that the order of the three nodes reflects
connectivity matrix C ([C]mi is the mth row and ith column of
the normal direction of the panel. For example, if the order
C) can be written as:
of nodes is clockwise, the normal direction is inward of the

−1, i = j panel. If it is counter-clockwise, the normal direction is out-

 ward. The quadrilateral structure can be represented by the
= 1, i = k .

[C]mi (3)


 N5B8 (five nodes and eight bars) model, and the cross-section
 0, i = else

area of the bars is given in [39] for similar mechanical prop-
erties in both in-plane stretching, shearing, and out-of-plane
2.2.2. Cross-section information of truss bending and folding.
Apart from the geometric information, we also assign ma-
terial properties to describe the physics by defining the cross-
sectional area, Young’s modulus, tangent Young’s Modulus, 2.4. Hinges
length, rest length, and force vector of the ith truss element as
Ai , Ei , Eti , li , l0i , and ti . The corresponding vector of the whole There are bending and folding hinges in the adjacent panels
structure is A, E, Et , l, l0 , and t ∈ Rne . The force in the ith [38]. The scalable stiffness for both types of hinges is stud-
member ti satisfies: ied by [39]. The formulation for bending and folding hinges
is identical, and the difference is the stiffness properties. To
Ei Ai simplify our formulation, we do not distinguish between the
ti = (li − l0i ), (4)
l0i bending and folding hinges.
As shown in Figure 1(b), a rotational spring is used to model
where Ei is the secant Young’s Modulus, which contains the the pth rotational hinge in the conjunction of two triangle pan-
nonlinear constitutive law of the material. Then, the force vec- els. The dihedral angle is θ p , and the rest dihedral angle is θ0,p ,
tor t can be written as: representing the angle with a moment of M p = 0. The length,
thickness, Young’s Modulus, and tangent Young’s Modulus of
t = Ê Â l̂0−1 (l − l0 ), (5)
the ith hinge is lh,p , th,p , Eh,p , and Eth,p , as shown in Figure 1(c).
The corresponding vector containing information of the whole
where v̂ transforms a vector v into a diagonal matrix. Note that
structure is lh , t h , Eh , Eth , θ, θ0 , and M ∈ Rnh . Let the number
we differentiate the tangent and secant Young’s Modulus here.
of hinges be nh . Suppose the hinge material deforms linearly,
The tangent Young’s Modulus is used in linearized equilibrium
and the hinge’s plane sections remain planar after bending. In
equations. And the secant Young’s Modulus is for calculating
other words, the strain in the neutral axis is zero, and the bend-
the material strain considering the nonlinear constitutive law
ing strain follows a linear rule, as shown in Figure 1(d). The
of the material. Thus, the secant modulus can represent linear
moment of the pth hinge M p (θ p ) can be written as:
elastic, multi-linear elastic, and plastic material properties.
M p (θ p ) = Eh,p I p (θ p − θ0,p ), (8)
2.3. Panels
2.3.1. Connectivity matrix of panels where Eh,p is the secant modulus of the hinge. And I p =
1 3
As shown in Figure 1(a), we use triangle panels (or some 2 lh,p th,p is the second moment of area of the hinge’s cross-
people call them plates) as the element in the origami structure. section. The rotational stiffness of the pth hinge is:
Each triangle contains three nodes and three edges. The rota-
tional springs connect the adjacent panels to model the hinges ∂M p 1
kp = = Eth,p I p = 3
Eth,p lh,p th,p . (9)
in between the panels. We use Ca ∈ R3×n p to represent the con- ∂θ p 12
nectivity of the panel, where n p is the number of the panel. Sup-
pose the ith panel is connected by node j, k, and m in a counter- The rotational stiffness of all hinges in the structure is kh =
clockwise sequence. The connectivity matrix C ([Ca ](:,i) is the [k1 , k2 , · · · , knh ]T ∈ Rnh .
3
3. Geometry and Kinematics 3.2.2. Hinge related nodes
3.1. Truss geometry and kinematic To write the equilibrium equation of the whole structure, we
This part calculates the length of the ith (i = 1, 2, · · · , ne ) need to connect the nodal coordinates of the four nodes (as
truss. The vector of the ith member hi ∈ R3 can be calculated shown in Figure 1(b)) related to a hinge to the nodal coordi-
as: nates of the whole structure. We use nep ∈ R12 to represent the
four nodal coordinates labeled i, j, k, and l in the two panels
hi = Ci ⊗ I3 n. (10) connected to the pth hinge, as shown in Figure 1(b):
And its length can be expressed as: h iT
1 nep = nepi T nep j T nepk T nepl T . (20)
li = ||hi || = [n T
(CiT Ci ) ⊗ I3 n] .
2 (11)
We give the two useful forms of the derivation of the ith truss The relation between nep and nodal coordinate of the whole
length with respect to the nodal coordinate in the denominator structure n is connected by Eh p ∈ R3nn ×12 :
and numerator layouts:
∂li ∂nep T
= li−1 (CiT Ci ) ⊗ I3 n, (12) Eh p = . (21)
∂n ∂n
∂li
= li−1 hTi Ci ⊗ I3 . (13) Similar to the definition in Section 3.2.1, the above equation
∂nT can be written as follows to get nep from n:
∂li
Stack the ∂n T (i = 1, 2, · · · , ne ) in a column, we have the com-

patibility matrix Bl ∈ Rne ×3nn : nep = EThp n. (22)


∂l
Bl = = l̂−1 b.d.(H)T (C ⊗ I3 ), (14) Stack the Eh p for p = 1, 2, · · · , nh , we have the transformation
∂nT
matrix of all the hinges-related nodes in the whole structure:
where H = NCT ∈ R3×ne is the element matrix, b.d.(H) means
the block diagonal matrix of H. One can also Eq. (14) as a h i
Eh = Eh1 Eh2 ··· Ehnh . (23)
standard compatibility equation:
Bl dn = dl. (15)
3.3. Hinge kinematics
Here, we should also give another useful form of the equi-
librium matrix A2 of the tensegrity system in [9] with member 3.3.1. First order kinematics
force as the variable, where the equilibrium equation is given as Based on [38], the Jacobian matrix of the pth hinge angle θ p
A2 t = w. The equilibrium matrix A2 is: to the related four nodal coordinates vector nep in Figure 1(a)
are:
∂lT
A2 = = C ⊗ I3 b.d.(H) l̂−1 . (16)
∂n ∂θ p ∥rk j ∥
e = m, (24)
One may observe that A2 is the transpose of compatibility ma- ∂npi ∥m∥2
trix A2 = BTl , which can be proved by the virtual work princi- ∂θ p ∥rk j ∥
ple, and the detailed derivation is referred to [9]. In summary, e =− n, (25)
∂npl ∥n∥2
the equilibrium matrix A2 and compatibility matrix Bl are the
Jacobian matrix of member length to nodal coordinate in the ∂θ p ri j · rk j ∂θ p rkl · rk j ∂θ p
e =( − 1) e − , (26)
numerator and denominator layouts. ∂np j ∥rk j ∥ 2 ∂npi ∥rk j ∥2 ∂nepl
∂θ p rkl · rk j ∂θ p ri j · rk j ∂θ p
e =( ,
3.2. Nodal Transformation Matrices − 1) e − (27)
3.2.1. Free nodes and constrained nodes ∂npk ∥rk j ∥2 ∂npl ∥rk j ∥2 ∂nepi
For a general model, we also need to deal with the nodal
constraints. Here, we use two index matrices Ena ∈ R3nn ×na and where
Enb ∈ R3nn ×nb to label the free nodal coordinate vector na ∈ Rna
rk j = nepk − nepk j , (28)
and fixed nodal vector nb ∈ Rnb in the total nodal vector n ∈
R3nn , where na and nb are the number of free and constrained
and the normal vectors m and n can be computed as:
nodal coordinates. We have the following equations to extract
na and nb from n:
m = ri j × rk j , (29)
na = ETna n, (17) n = rk j × rkl . (30)
nb = ETnb n. (18)
In the above derivation, we have the Jacobian matrix of one
We also have the following equation to combine na and nb into
hinge angle to the four connected nodes in the two connected
n:
panels. Extending the kinematics relation in one unit to the
n = Ena na + Enb nb . (19) whole structure, the Jacobian matrix of the hinge angle vector
4
with respect to the structure’s nodal coordinate vector is: Stack Gh,p for p = 1, 2, · · · , nh in a matrix, we have the matrix
Gh ∈ R3nn ×(3nn ×nh ) representing the second-order partial deriva-
∂θT X
nh
∂nep ∂θ p ∂θT T
= , (31) tive of the hinge angle to the nodal coordinate vector of the
∂n p=1
∂n ∂nep ∂θ p whole structure:
 ∂θ1   ∂θT
··· 0
 h i
  ∂n1
 e
  ∂θ1
   Gh = Gh,1 Gh,2 · · · Gh,nh (40)
 .. .. ..   ..
 
∂ne1 T ∂nenh T 
=  . . .   . (32) ∂ 2
θ
h 2 i
= ∂ θ1T ∂ θ2 ,

∂n ··· ∂n  2
···
nh
(41)
∂θnh    ∂θT  ∂n∂n ∂n∂n T T ∂n∂n
0 ···

∂nnh
e ∂θnh
The formulation of Gh will be used to derive the explicit tangent
∂θi
= Eh b.d.( e )Inh , (33) stiffness matrix of the whole structure in Section 4.4.
∂ni
∂θi
where b.d.( ∂ne) is the block diagonal matrix of 4. The Equilibrium Equation Formulation
i
∂θnh
 
∂θ1 ∂θ2
∂ne1 ∂ne2 ···. Similar to the truss kinematics,
∂nenh 4.1. The Lagrangian method
we name the above Jacobian matrix as the equilibrium matrix
of hinge rotation Ah . The compatibility matrix of hinge We use the Lagrangian method to derive the governing equa-
rotation Bh is also the transpose of the equilibrium matrix: tions of the equilibrium equation. All the formulation here is
consistent with the principle of stationary total potential en-
∂θT ∂θi ergy and the principle of virtual work. However, using the La-
Ah = = Eh b.d.( e )Inh , (34)
∂n ∂ni grangian method to derive the equilibrium equation will make
∂θ ∂θi it easy to extend to the future study of the dynamics.
Bh = ATh = T = Inh b.d.( e )T ETh . (35) The general form of the Lagrangian equation is:
∂n ∂ni
We will explain in Section 4.3 that equilibrium matrix Ah d ∂L ∂L
− = fnpa , (42)
gives the equilibrium relation between hinge moment and ex- dt ∂ ṅa ∂na
ternal force.
where L = T − V is the Lagrangian function, T and V are the ki-
3.3.2. Second order kinematics netic energy and potential energy of the system, na is the gener-
The Hessian matrix for the pth hinge has size 12×12 and con- alized coordinates of the system, fnpa ∈ Rna is the non-potential
sists of 4×4 sub-matrices of size 3×3 [40]. Since the Hessian force vector exerted on the free nodal coordinate. The relation-
matrix is symmetric, ten independent sub-matrices are provided ship of fnpa with the non-potential force fnp exerted on all the
in the Appendix of [38]. The first 3×3 block of the Hessian ma- nodes is:
trix of θ p to nepi is given here as an example:
fnpa = ETa fnp . (43)
∂ θp
2
∥rk j ∥
=− (m ⊗ (rk j × m) + (rk j × m) ⊗ m). (36) For the statics problem, the kinetic energy T is zero, and Eq.
∂nepi ∂nepi T ∥m∥4
(42) degenerates to:
The above Hessian matrix for the pth hinge can be then written
in a compact form as: ∂V
= fnpa . (44)
 ∂2 θ p ∂2 θ p ∂2 θ p ∂2 θ p 
 ∂na
 e e T
 ∂npi2∂npi ∂npi ∂np j
e e T
∂npi ∂npk
e e T
∂npi ∂npl 
e e T 

 ∂ θ p ∂2 θ p ∂2 θ p ∂2 θ p   4.2. Energy Functions
∂2 θ p  ∂nep ∂nep T ∂nep ∂nep T ∂nep ∂nep T ∂nep ∂nep T 
=  2 j i j j j k j l  . (37) The potential energy of the whole system contains three
∂nep ∂nep T  e∂ θ pe T ∂2 θ p ∂2 θ p ∂2 θ p  
parts: the strain potential energy in the truss, the strain potential
 ∂npk ∂npi ∂npk ∂np j
e e T
∂npk ∂npk
e e T
∂npk ∂npl 
e e T  
 ∂2 θ p ∂2 θ p ∂2 θ p ∂2 θ p   energy in the hinges, and the gravitational potential energy of
∂npl ∂npi
e e T e
∂npl ∂np j
e T e e T e
∂npl ∂npk
e T
∂npl ∂npl all the elements.
In the above derivation, we have the Hessian matrix of one
hinge angle with respect to the four connected nodes in a basic 4.2.1. Strain energy in truss
element with two connected panels. The Hessian matrix for the The strain potential energy of all the truss elements is:
pth hinge angle with the whole structure nodal coordinate is ne Z li
Gh,p ∈ R3nn ×3nn :
X
Vt = ti du. (45)
l0 i
∂ θp2 ∂nep T ∂ θ p ∂nep
2 i=1
Gh,p = = (38)
∂n∂nT ∂n ∂nep ∂nep T ∂nT This is a general form of potential energy since it can accom-
modate any material’s properties, i.e., linear elastic and plastic.
∂ θp
2
= Eh p EThp . (39) In many papers, one can see this equation is in the form of 12 k∆l,
∂nep ∂nep T which only assumes linear elastic materials.
5
The partial derivative of Vt to na is: where m pi is the mass of the ith node, and mp ∈ Rnn is the point
mass vector.
∂Vt Then, the partial derivative of Vg to na is:
= ETa Kn = ETa A2 t, (46)
∂na
∂Vg ∂Vge ∂Vgp
where K is the nonlinear stiffness matrix (since K is a function = + (54)
∂na ∂na ∂na
of n) of the whole system. The detailed derivation is referred to
1 h iT
[9, 37]: = ETa ( |C|T me + mp ) ⊗ a x ay az (55)
2
K = (CT l̂−1 tC) ⊗ I3 . (47) = Ea g,
T
(56)

where g ∈ R3nn is the gravitational force vector.


4.2.2. Strain energy in the hinges
Similarly, we can write the strain potential energy of the 4.3. Nonlinear equilibrium equations
hinges:
The potential energy of the system is a summation of three
nh Z
X θi parts:
Vh = Mi (φ)dφi . (48)
i=1 θ0i V = Vt + Vh + Vg . (57)
The partial derivative of Vh to na is: Substitute Eqs. (46) and (50) into Eq. (44), and we have:

∂nT ∂θiT ∂Vhi


nh
∂Vh X ∂V
" #
t
= (49) = Ea Ao
T
= fnpa − ETa g, (58)
∂na ∂na ∂n ∂θi ∂na M
i=1
nh
X ∂θiT where Ao is the equilibrium matrix for integrated origami and
= ETa Mi (θi ) = ETa Ah M, (50)
i=1
∂n tensegrity systems:
h i
h iT Ao = A2 Ah . (59)
where M = M1 M2 · · · Mnh ∈ Rnh is the moment vector
of all the hinges in the structure. Note that the partial derivative The first part of Eq. (59) is the equilibrium matrix correspond-
of truss strain energy Vt and hinges strain energy Vh to na in ing to truss elements in Eq. (16), and the second is the equi-
Eqs. (46) and (50) has a similar form. librium matrix corresponding to hinge elements " #as in Eq. (34).
t
The Eq. (59) is a linear equation in terms of , which makes
4.2.3. Gravitational potential energy M
The gravitational potential energy is relative to any member the force design of complex structures possible by force density
that has mass. In the system, all the axial members and point methods [41, 42, 43]. To perform structural analysis subject to
mass will contribute to gravitational potential energy: external forces, Eq. (59) becomes a nonlinear equation in terms
of free nodal coordinate dna because Ao , t, and M are all func-
Vg = Vge + Vgp , (51) tions of dna . The linearization of the equilibrium equation will
be discussed in the next section.
where Vge and Vgp are the potential energies of the axial mem- The compatibility equation of the structure is:
bers and point mass. " #
The gravitational potential energy corresponding to the axial dl
Bo dna = . (60)
elements Vge [37] can be written as: dθ
X mei h i Based on the principle of virtual work, the compatibility matrix
Vge = ne a x ay az |Ci | ⊗ I3 n
i=1
2 is:
1 h i " T# " #
A B
= (mTe |C|) ⊗ a x ay az n, (52) Bo = ATo = 2T = l . (61)
2 Ah Bh
where mei is the mass of the ith axial element, and me ∈ Rne is It should be noted that the equilibrium Eq. (58) and the com-
the mass vector of all axial elements, a x , ay , az are the gravita- patibility Eq. (60) can describe the equilibrium and compatibil-
tional acceleration values in the X, Y, and Z-direction. ity properties of the structure information considering the axial
The gravitational potential energy of the point mass Vgp is: force members and hinges simultaneously. They degenerate to
  pure tensegrity structures considering only axial deformation of
nn
X h i  xi  bar and strings [37], which are:
Vgp = m pi ⊗ a x ay az yi 
zi
 
i=1
h i A2 t = fnpa − ETa g, (62)
= mTp ⊗ ax ay az n, (53) Bl dna = dl. (63)

6
The degenerated equilibrium equation Eq. (62) can also be in-
terpreted as the external force on the right-hand side is in the left
null space of Ah . In other words, the external forces only work
in the axial members without generating a moment in hinges.
This scenario happens when all external forces are in the space
spanned by the connected axial members. Eqs. (58) and (60)
can also degenerate to rigid origami structures considering only
hinge rotational deformation:

Ah M = fnpa − ETa g, (64)


Bh dna = dθ. (65) Figure 2: Configuration of the illustrative example, modified from Li [40]. On
the left-hand side, the four rectangular sheets, given by solid lines, are modeled
Eq. (64) can be seen as the external force is in the left null space by eight panels (separated by four dotted lines). The right-hand side is the
corresponding origami model. Nodes 1, 2, and 5 are pinned to the ground.
of A2 , which means the external forces only yield a moment in
the hinges without generating force in the axial members. This
scenario happens when all external force is in the null space It should be noted that the material stiffness of trusses Km1
spanned by the connected axial members. and hinges Km2 are in a very similar form, since A2 and Ah are
the equilibrium matrix for truss and hinges and Ê Â l̂−1 and k̂h
4.4. Linear equilibrium equations
are the stiffness of truss and hinges. The geometry stiffness of
Using Taylor’s expansion to Eq. (58) about an equilibrium trusses Kg1 and hinges Kg2 are also similar if we write the Kg1
configuration ni in the ith iteration step, we have the linearized in an equivalent form as:
equilibrium equation:
" # Kg1 = Gl t ⊗ I3nn , (75)
t
T
Ea Ao ni + ETa KT dni = fnpa − ETa g, (66)
M ni where the Gl ∈ R3nn ×(ne ×3nn ) is the in the similar form with Eq.
(41):
where KT ∈ R3nn ×3nn" is #the tangent stiffness matrix of the whole
t
h i
structure, Ao ni and are the equilibrium matrix and inner Gl = Gl,1 Gl,2 · · · Gl,nh (76)
M ni
∂ ln
2
h 2 i
force in ni configuration. Substituting Eq. (19) into Eq. (66), = ∂ l1T ∂2 l2
· · · ∂n∂nhT . (77)
∂n∂n ∂n∂nT
we have:
" # Comparing Eqs. (74) and (75), we can observe they are sim-
t ilar. But we prefer to write Kg1 in the form of Eq. (72) because
T
Ea Ao ni + KT aa dnia = fnpa − ETa g − KT ab dnib , (67)
M ni the formulation is explicit and simple [37]. It should be noted
that the geometry stiffness related to hinges Kg2 in Eq. (74) is
where
difficult to be written into the form as Kg1 in of Eq. (72) because
KT aa = ETa KT Ea , KT ab = ETa KT Eb . (68) the nonlinear properties in the hinge kinematics.

Substitute Eqs. (34) and (41) into the definition of tangent


stiffness matrix, we have: 5. Numerical examples

∂(Kn + Ah M) 5.1. A Miura origami unit cell


KT = (69)
∂nT 5.1.1. Basic information
= A2 Ê Â l̂−1 AT2 + (CT l̂−1 tC) ⊗ I3 + Ah k̂h ATh + Gh M ⊗ I3nn . An illustrative example of a Miura origami unit cell is shown
(70) in Figure 2. In this example, we try to be as detailed as
In Eq. (70), the first and second parts are the material stiffness possible to help readers to understand the whole formulation.
and geometry stiffness of all truss elements, which is identical The Miura-ori unit cell is composed of 5 nodes and 8 bars.
to tensegrity formulation [9]: The quadrilateral model is simplified as triangle panels, which
means only folding hinges are needed in the unit cell. The inner
Km1 = A2 Ê Â l̂−1 AT2 , (71) angle of the unit cell is β = 75◦ . The nodal coordinate matrix
of the muri-ori unit cell is:
Kg1 = (CT l̂−1 tC) ⊗ I3 . (72)
 
−1.4142 −0.3660 −0.3660 1.4142 0
The third and fourth parts are the material stiffness and geome-
N =  0 1.3660 −1.3660 0 0 . (78)
 
try stiffness related to the hinges:
0 0 0 0 0
 

Km2 = Ah k̂h ATh , (73)


Nodes 1, 2, and 5 are pinned to the ground. The cross-sectional
Kg2 = Gh M ⊗ I3nn . (74) area of all the bars is set to be A = 1×10−4 m2 . The connections
7
and rest length of bars are:
1 2 1.7218
   
2 4
 2.2439
   
4 3 2.2439
  
3 1 1.7218
Cin =   , l0 =  . (79)
1 5 1.4142
2 5
 1.4142
 
  1.4142
4 5  
3 5 1.4142
Figure 3: The eigenvalues and modes of the illustrative example: (a) eigenval-
The panels have a thickness of t = 6 × 10−4 m and Young’s ues vs. order of the modes and (b) their corresponding modes.
modulus of E = 1 × 106 Pa.The connectivity of panels to nodes
are:
 
5 5 5 5
Ca = 2 1 3 4 . (80)
 
1 3 4 2
 

The rest angle and rotational stiffness of the folding hinges


are:
 ◦
180  2.546
 
180◦  2.546
θ =  ◦  , Kh = 10−4 ×   N·m/rad. (81)
180  2.546

180 2.546
The moments in hinges of the flat Miura origami are zero.
The transformation matrix of the first hinge relates the node i, j,
k, and l in the local element to the node numbers 2, 1, 5, and 3 in
the global nodal coordinate. The corresponding transformation
for the four hinges are:
Figure 4: Configuration of the mode shape of the Miura–ori unit cell: (a) the
0 1 0 0 0 0 0 0
   
initial and final shapes, (b) stiffness of the structure, (c) stiffness contribution of
1 0 0 0 0 1 0 0
origami, bars, and hinges, and (d) normal of angle difference at different angles
Eh1 = 0 0 0 1 ⊗ I3 , Eh2 = 0 0 0 0 ⊗ I3 , (82)
   
of α.
0 0 0 0 1 0 0 0
   
0 0 1 0 0 0 1 0
5.1.3. Infinitesimal deformation analysis
0 0 0 0 1 0 0 0
   
0 0 0 1 0 0 0 0 Figure 4(a) shows that the Miura origami unit cell is 80%
Eh3 = 1 0 0 0 ⊗ I3 , Eh4 = 0 1 0 0 ⊗ I3 . (83) folded. The nodes 1, 2, and 5 are fixed, and a unit external load
   
0 1 0 0 0 0 0 1 is applied at node 4. The external force is in the Y and Z planes,
and the angle between the force and the Y axis is θ. We perform
   
0 0 1 0 0 0 1 0
linear elastic material with small displacement static analyses of
5.1.2. Tangent stiffness analysis the structure. Figure 4(a) shows the displaced shapes obtained
The spectral decomposition of the tangent stiffness matrix by the static analysis as θ = 45◦ , and the deformation of node
Eq. (70) reveals the stiffness properties by eigenvalue λi and its 4 is in towards the Z-direction, which is identical to the fold-
corresponding deformation mode vi of free nodal coordinate: ing trajectory in the rigid folding assumption. The stiffness of
the structure is defined by the reciprocal of compliance, which
KT aa vi = λi vi . (84)
equals the work done by the unit of external force. The semi-
Figure 3(a) is the eigenvalue of tangent stiffness with differ- log plot of stiffness to orientation angle θ is shown in Figure
ent orders. Figure 3(b) shows the mode shapes corresponding 4(b). The result shows that the structure gets stiffer as θ is close
to each eigenvalue. We can observe that the first and second to 180◦ and 360◦ . And the stiffness is the lowest as θ is close
modes are plane deformation, in which nodes 3 and 4 moves to 90◦ and 270◦ , which is in the direction close to the mech-
to the opposite and the same direction in the Z-coordinate. anism trajectory. Figure 4(c) gives the percentage of stiffness
The first two modes mainly involve the folding hinges rotation contributed by bars and hinges. We can observe that bars and
deformation, which requires less energy, so the eigenvalue is hinges contribute around 2/3 and 1/3 of the total stiffness, re-
much lower than the other deformation modes. Modes 3 to 6 spectively. However, when α is around 180◦ and 360◦ , there is
mainly contain the bar length deformation in the plane, which a huge reduction of hinge stiffness. This is because the displace-
needs higher energy and, thus, the higher eigenvalue of the tan- ment is in the null space of the compatibility matrix Bh , which
gent stiffness matrix. means the nodal displacement caused by the external force will
8
Figure 6: Configuration of an integrated tensegrity origami shelter. The left-
hand side is the 2D view. On the right-hand side, the shelter is a combination
of origami sheets and a tensegrity frame.

changing the configuration. The maximum stiffness in X- and


Y-directions is obtained when the structure is flat. The stiffness
in the Y-direction is generally higher than in the X-direction be-
cause the in-plane stiffness of panels strengthens the Y-direction
stiffness. In the Z-direction, the stiffness is maximum when bar
7 is close to vertical.

5.2. Integrated tensegrity origami shelter


5.2.1. Basic information
Figure 5: Folding information of the Miura–ori unit cell: (a) the folding pro- Accordion shelters are origami-like deployable shelters de-
cess, (b) stiffness of the structure, (c) angles of the panels, and (d) moment in
the hinges at different load factors. veloped by the U.S. military [45]. To increase the stiffness and
controllability of the deployable shelter, an integrated tensegrity
origami shelter is proposed, as shown in Figure 6. The shelter
not change the angle of hinges Bh dna = 0. Figure 4(d) gives the comprises two planar D-bar tensegrity structures [46] braced
norm of angle difference in external force, and we can see that with out-of-plane truss and a six-fold origami panel modeled
at around 180◦ and 360◦ , the norm of angle difference is zero. by N5B8 unit [39]. The number of D-bar in planar tensegrity
This means the material stiffness of rotational hinges does not structures is defined as the complexity p of the structure, and
contribute to the structural stiffness. we choose p = 3 for the structure, as given in Figure 6. The
length of bars in the D-bar and the width of the structure in the
5.1.4. Large deformation analysis X-direction is 1.4140 m. The angle of the bars in the D-bar
The integrated origami and tensegrity formulation can also in the Y-direction is 40◦ . The cross-sectional area of bars and
capture large deformations of the origami structures. For ex- strings is 10−4 m2 , the area of the diagonal bars in N5B8 unit is
ample, when actuating a Miura-ori unit cell by changing the 10−7 m2 , and the area of vertical strings is reduced to 10−7 m2
rest angle of folding hinges, the Miura-ori unit cell will de- to allow large deformations in the folding process. The Young’s
form from a flat configuration to a folded one, as shown in Modulus of bars and strings is set as E = 2 × 1011 Pa for steel
Figure 5(a). Note that the rest angle of one mountain fold and materials. The rotational stiffness of the rotational and bending
three valley folds increases and decreases respectively by 180◦ hinges are 2.546 × 10−4 N·m/◦ and 1.8 × 10−2 N·m/◦ . Note that
in the folding process, as shown in Figure 5(c). The Newton- the rotational hinge stiffness is decreased for the consideration
Raphson method is used as the static solver in large displace- of the reduction of Modulus due to the plastic deformation.
ment analysis since the singularity of the stiffness matrix does
not occur in this problem. Note that if the structures are sub- 5.2.2. Tangent stiffness analysis
ject to highly geometric nonlinearity, the modified generalized Figure 7(a) gives the first 15 eigenvalues of the tangent stiff-
displacement control method (MGDCM) [44] can be used as ness. The 10th to 15th modes correspond to the in-plane de-
the solver. Figure 5(d) shows the moment in the four folding formations of the center nodes in the N5B8 unit of the origami
hinges, in which the relation between the folding angles and the part. The 7th to the 9th modes correspond to the lateral defor-
moments is calculated by Eq. (8). The moment in the mountain mations of the origami panels in the X-direction. Note that the
fold is negative, while the moments in the two symmetric val- deformation happens in the upper part instead of the bottom part
ley folds are the same. The characteristic stiffness for the X-, because the brace truss in the bottom is stiffer than the diagonal
Y-, and Z-directions is calculated in the folding process. Figure bars in the origami panels. The 1st to the 6th modes corre-
5(b) shows that the stiffness of the Miura-ori can be tuned by spond to the out-of-plane deformations of the 6 origami panels,
9
Figure 7: The eigenvalues and modes of the illustrative example: (a) eigenval-
ues vs. order of the modes, (b) mode 15, (c) mode 8, and (d) mode 5.

which is reasonable because the origami panels are compara-


tively weaker in the out-of-plane direction than the tensegrity
part.

5.2.3. Large deformation analysis


The integrated origami and tensegrity shelter can be folded
or deployed by changing the rest length of horizontal strings
actively, as shown in Figure 8(a). The rest length of the hori-
zontal strings is reduced by 70% in the folding process, while
the rest length of other members is constant. The length of ver-
tical strings is elongated to allow deformation of the D-bar, as
shown in Figure 8(c). The force of bars and vertical strings
increases, and the force of horizontal strings increases in the
first half part and decreases in the second half part, as shown in
Figure 8(d).
The characteristic stiffness for the external force exerted on Figure 8: Folding information of the shelter: (a) the folding process, (b) stiff-
ness of the structure, (c) length of structural elements, and (d) force in the struc-
the two end nodes in the X-, Y-, and Z-directions are calculated tural elements at different load factors.
in the folding process. Figure 8(b) shows that the stiffness of
the shelter can be tuned by changing its configuration. The stiff-
ness in Y- and Z-directions increases as the structure is folded,
which is reasonable because the cantilever length decreases.
The stiffness in the Z-direction is generally higher than in the
Y-direction because the height of the structure is higher than in
the Z-direction. The stiffness in the X-direction decreases as
the structure is folded because the increment in the angle of the
D-bar with the X-axis reduces the X-direction stiffness.

5.3. Cable-driven Kresling structure


5.3.1. Basic information
The cable-driven Kresling structure is a combination of a
Kresling pattern origami [38, 47] and strings, as shown in Fig-
ure 9(a). The cable-driven Kresling structure is composed of
vertical bars, horizontal bars, diagonal bars, and strings, and
the connectivity is shown in Figure 9(b). The complexity of
the structure in horizontal and vertical directions is p and q, Figure 9: The geometric configuration of the cable-driven Kresling pattern: (a)
axonometric view, (b) expansion view of the cylinder, (c) side view of one layer
which are the number of sides of the polygon and the layers of
(there are q layers in total), and (d) top view.
the structure. The angles between the diagonal bar, the hori-
zontal bar, and the vertical bar are α and β. For Kresling pat-
tern origami under compressive load [47] when α + β < 90◦ ,
the strain in each member remains in tension or compression
throughout the deployment and collapse processes. The strain
10
Figure 10: Stiffness of the cable-driven Kresling structure in X-, Y-, Z-, and
rotational directions.

may vary between tension and compression when α + β > 90◦ .


The front and top views of a one-layer cable-driven Kresling
structure are shown in Figure 9(c). The height of one layer
Kresling structure is h. The twist angle between the bot-
tom and top polygon is φ. The cross-sectional area of all the
bars and strings is 10−4 m2 . The panels have a thickness of
t = 6 × 10−4 m, and Young’s modulus of E = 1 × 106 Pa. For
the cable-driven Kresling structure in Figure 9(a), the geometry
parameters are p = 5, q = 3, r = h = 5 m, and φ = 15◦ .

5.3.2. Stiffness analysis


The configurations of the cable-driven Kresling structure Figure 11: Folding information of the cable-driven Kresling structure: (a) the
folding process, (b) stiffness of the structure, (c) length of structural elements,
with different h ∈ [5, 15] and φ ∈ [−45◦ , 45◦ ] are shown in and (d) force in the structural elements at different load factors.
Figure 10, given complexity p = 5, q = 3 and radius r = 5 m.
The normalized force vector is exerted on the top nodes in X-
, Z-, and rotational directions. As shown in Figure 10(a) and The characteristic stiffness for X-, Y-, Z-, and rotational di-
(b), the maximum stiffness in X- and Z-direction is obtained rections is calculated in the folding process. Figure 11(b) shows
when h = 10 m and φ = −45◦ . The minimum stiffness in the that the stiffness of the cable-driven Kresling structure can be
X-direction is at φ = 45◦ , while the minimum stiffness in the tuned by changing its configurations. The stiffness in X- and Y-
Z-direction is at h = 6 m or 14 m and φ = −45◦ . The string ac- directions increases as the structure folds, which is reasonable
tuation stiffness is defined as the sensitivity of force increment because the cantilever length decreases. The stiffness in the Z-
to string length decrement. As shown in Figure 10(c), for string direction is generally higher than in X-and Y-directions because
actuation in the deployment, h = 5 m and φ = −45◦ is the most the Z-direction forces compress the structure while the X- and
difficult configuration for actuation. At the same time, it would Y- direction forces bend the structure. The stiffness in the Z-
be much easier to actuate as the height increases. As shown in direction decreases as the structure folds due to the decrease of
Figure 10(d), the rotational stiffness at φ = −45◦ is more than angles between bars and the X- and Y-planes. The rotational
four times higher than at φ = 45◦ , and the maximum stiffness stiffness reaches the highest point in the middle of the folding
in the rotation is obtained at h = 6 m or 14 m and φ = −45◦ , process.
which is the minimum stiffness in the Z-direction.

6. Conclusions
5.3.3. Large deformation analysis
The cable-driven Kresling structure can be easily folded or A static analysis approach to nonlinear integrated tenseg-
deployed by increasing or decreasing the rest length of strings, rity and origami systems based on the Lagrangian method with
as shown in Figure 11(a). Note that the rest length of the hor- nodal coordinates and angles between the origami panels is
izontal strings is reduced by 80% in the folding process, while given in this paper. This approach allows one to conduct com-
the rest length of other members is constant, as shown in Fig- prehensive studies on any integrated origami and tensegrity sys-
ure 11(c). The difference between the length and rest length for tems with any node constraints and various load conditions (i.e.,
bars is much higher than that of strings, as in Figure 11(c). As gravitational force, specified forces, and moments). Results
a result, the force in the bars is much higher than in strings, as show that this method can 1) perform loading analysis of the
in Figure 11(d). integrated system with infinitesimal or large deformations. For
11
example, from the three examples, we can get information on [14] T. Rhodes, C. Gotberg, V. Vikas, Compact shape morphing tensegrity
stiffness, deformation, forces, and moments of all the structural robots capable of locomotion, Frontiers in Robotics and AI 6 (2019) 111.
[15] H. Zhou, A. R. Plummer, D. Cleaver, Distributed actuation and control of
elements. 2) help understand the stiffness contributions of the a tensegrity-based morphing wing, IEEE/ASME Transactions on Mecha-
bars and hinges. For example, in the Miura origami unit ex- tronics 27 (2021) 34–45.
ample, the bars and hinges contribute around 2/3 and 1/3 of [16] M. Meloni, J. Cai, Q. Zhang, D. Sang-Hoon Lee, M. Li, R. Ma, T. E.
the total stiffness. 3) study the deployment process or actu- Parashkevov, J. Feng, Engineering origami: A comprehensive review
of recent applications, design methods, and tools, Advanced Science 8
ation strategies of the cables. For example, in the integrated (2021) 2000636.
tensegrity origami shelter example, from the eigenvalue study, [17] T. Chen, O. R. Bilal, R. Lang, C. Daraio, K. Shea, Autonomous deploy-
we can see that the deformation happens in the upper part in- ment of a solar panel using elastic origami and distributed shape-memory-
stead of the bottom because the first six modes correspond to polymer actuators, Physical Review Applied 11 (2019) 064069.
[18] Y. Li, Z. You, Origami concave tubes for energy absorption, International
the out-of-plane deformations are comparatively weaker in the Journal of Solids and Structures 169 (2019) 21–40.
out-of-plane direction. 4) One can change the stiffness of the in- [19] Q. Ze, S. Wu, J. Nishikawa, J. Dai, Y. Sun, S. Leanza, C. Zemelka, L. S.
tegrated structure by changing its configurations with analytical Novelino, G. H. Paulino, R. R. Zhao, Soft robotic origami crawler, Sci-
guidance. For example, in the cable-driven Kresling structure, ence advances 8 (2022) eabm7834.
[20] E. P. Hernandez, D. J. Hartl, D. C. Lagoudas, Active origami, Active
by tuning the length of cables, we get the contour of stiffness Origami (2019).
with respect to the heights of the cylinder. Overall, this study [21] R. J. Lang, Origami design secrets: mathematical methods for an ancient
paves a way to help people design deployable structures from art, CRC Press, 2012.
[22] K. Huang, H. Elsayed, G. Franchin, P. Colombo, Complex sioc ceramics
the benefits of both tensegrity and origami systems, as well as a from 2d structures by 3d printing and origami, Additive Manufacturing
comprehensive understanding of the performance of both struc- 33 (2020) 101144.
tures and materials. [23] Y. Zhang, C. Wang, Y. Dong, D. Wang, T. Cao, S. Wang, D. Liu, Fold 2d
woven dna origami to origami+ structures, Advanced Functional Materi-
als 29 (2019) 1809097.
Acknowledgment [24] Q. Ge, C. K. Dunn, H. J. Qi, M. L. Dunn, Active origami by 4d printing,
Smart materials and structures 23 (2014) 094007.
The research was supported by the National Natural Sci- [25] Z. Zhao, X. Kuang, J. Wu, Q. Zhang, G. H. Paulino, H. J. Qi, D. Fang, 3d
ence Foundation of China (Grant No. 52208218) and the printing of complex origami assemblages for reconfigurable structures,
Soft Matter 14 (2018) 8051–8059.
Natural Science Foundation of Zhejiang Province (Grant No. [26] S. Kamrava, D. Mousanezhad, H. Ebrahimi, R. Ghosh, A. Vaziri,
LQ23E080021). Origami-based cellular metamaterial with auxetic, bistable, and self-
locking properties, Scientific reports 7 (2017) 1–9.
[27] H. Yasuda, J. Yang, Reentrant origami-based metamaterials with negative
References poisson’s ratio and bistability, Physical review letters 114 (2015) 185502.
[28] N. Wang, K. Naruse, D. Stamenović, J. J. Fredberg, S. M. Mijailovich,
[1] K. D. Snelson, Continuous tension, discontinuous compression struc- I. M. Tolić-Nørrelykke, T. Polte, R. Mannix, D. E. Ingber, Mechanical
tures, 1965. US Patent 3,169,611. behavior in living cells consistent with the tensegrity model, Proceedings
[2] R. B. Fuller, Tensile-integrity structures, Patente US3063521, concedida of the National Academy of Sciences 98 (2001) 7765–7770.
(1959). [29] G. Scarr, A consideration of the elbow as a tensegrity structure, Interna-
[3] K. M. Roffman, G. A. Lesieutre, Morphing tensegrity space platforms, tional Journal of Osteopathic Medicine 15 (2012) 53–65.
in: AIAA Scitech 2021 Forum, 2021, p. 0428. [30] F. Fraternali, N. Stehling, A. Amendola, B. A. Tiban Anrango, C. Hol-
[4] S. Ma, M. Chen, R. E. Skelton, Design of a new tensegrity cantilever land, C. Rodenburg, Tensegrity modelling and the high toughness of spi-
structure, Composite Structures 243 (2020) 112188. der dragline silk, Nanomaterials 10 (2020) 1510.
[5] K. Kim, A. K. Agogino, A. M. Agogino, Rolling locomotion of cable- [31] V. Caratelli, G. Fegatelli, D. Moscone, F. Arduini, A paper-based elec-
driven soft spherical tensegrity robots, Soft robotics 7 (2020) 346–361. trochemical device for the detection of pesticides in aerosol phase in-
[6] M. Chen, X. Bai, R. E. Skelton, Minimal mass design of clustered tenseg- spired by nature: A flower-like origami biosensor for precision agricul-
rity structures, Computer Methods in Applied Mechanics and Engineer- ture, Biosensors and Bioelectronics 205 (2022) 114119.
ing 404 (2023) 115832. [32] B. Kresling, Origami-structures in nature: lessons in designing “smart”
[7] A. Fraddosio, G. Pavone, M. D. Piccioni, Minimal mass and self-stress materials, MRS Online Proceedings Library (OPL) 1420 (2012).
analysis for innovative v-expander tensegrity cells, Composite Structures [33] J. A. Faber, A. F. Arrieta, A. R. Studart, Bioinspired spring origami,
209 (2019) 754–774. Science 359 (2018) 1386–1391.
[8] Z. Kan, F. Li, N. Song, H. Peng, Novel nonlinear complementarity func- [34] J. L. Rohmer, E. A. Peraza Hernandez, R. E. Skelton, D. J. Hartl, D. C.
tion approach for mechanical analysis of tensegrity structures, AIAA Lagoudas, An experimental and numerical study of shape memory alloy-
Journal 59 (2021) 1483–1495. based tensegrity/origami structures, in: ASME International Mechanical
[9] S. Ma, M. Chen, R. E. Skelton, Dynamics and control of clustered tenseg- Engineering Congress and Exposition, volume 57526, American Society
rity systems, Engineering Structures 264 (2022) 114391. of Mechanical Engineers, 2015, p. V009T12A064.
[10] X. Feng, W. Zhang, Y. Luo, S. Zlotnik, Optimal prestress investigation [35] R. Miranda, E. Babilio, N. Singh, F. Santos, F. Fraternali, Mechanics
on tensegrity structures using artificial fish swarm algorithm, Advances of smart origami sunscreens with energy harvesting ability, Mechanics
in Civil Engineering 2020 (2020). Research Communications 105 (2020) 103503.
[11] D. T. Trinh, S. Lee, J. Kang, J. Lee, Force density-informed neural net- [36] J. Hong Park, Tensegami: Design principle of combining tensegrity and
work for prestress design of tensegrity structures with multiple self-stress origami to make geodesic dome structure for martian agriculture, in:
modes, European Journal of Mechanics-A/Solids 94 (2022) 104584. Earth and Space 2021, 2021, pp. 978–984.
[12] A. Amendola, A. Krushynska, R. Miranda, F. Fraternali, Optimal pre- [37] S. Ma, M. Chen, R. E. Skelton, Tensegrity system dynamics based on
stress design of the band gap dynamics in tensegrity metamaterials, in: finite element method, Composite Structures 280 (2022) 114838.
Advances in Engineering Materials, Structures and Systems: Innovations, [38] K. Liu, G. Paulino, Nonlinear mechanics of non-rigid origami: an effi-
Mechanics and Applications, CRC Press, 2019, pp. 995–999. cient computational approach, Proceedings of the Royal Society A: Math-
[13] J. Bauer, J. A. Kraus, C. Crook, J. J. Rimoli, L. Valdevit, Tensegrity ematical, Physical and Engineering Sciences 473 (2017) 20170348.
metamaterials: Toward failure-resistant engineering systems through de- [39] E. Filipov, K. Liu, T. Tachi, M. Schenk, G. H. Paulino, Bar and hinge
localized deformation, Advanced Materials 33 (2021) 2005647.

12
models for scalable analysis of origami, International Journal of Solids
and Structures 124 (2017) 26–45.
[40] Y. Li, S. Pellegrino, A theory for the design of multi-stable morphing
structures, Journal of the Mechanics and Physics of Solids 136 (2020)
103772.
[41] J. Zhang, M. Ohsaki, F. Tsuura, Self-equilibrium and super-stability of
truncated regular hexahedral and octahedral tensegrity structures, Inter-
national Journal of Solids and Structures 161 (2019) 182–192.
[42] Y. Chen, Q. Sun, J. Feng, Improved form-finding of tensegrity structures
using blocks of symmetry-adapted force density matrix, Journal of Struc-
tural Engineering 144 (2018) 04018174.
[43] Y. Wang, X. Xu, Y. Luo, Form-finding of tensegrity structures via rank
minimization of force density matrix, Engineering Structures 227 (2021)
111419.
[44] S. E. Leon, E. N. Lages, C. N. De Araújo, G. H. Paulino, On the effect
of constraint parameters on the generalized displacement control method,
Mechanics Research Communications 56 (2014) 123–129.
[45] A. Thrall, C. Quaglia, Accordion shelters: A historical review of origami-
like deployable shelters developed by the us military, Engineering struc-
tures 59 (2014) 686–692.
[46] R. E. Skelton, M. C. De Oliveira, Tensegrity systems, volume 1, Springer,
2009.
[47] Z. Zhai, Y. Wang, H. Jiang, Origami-inspired, on-demand deployable and
collapsible mechanical metamaterials with tunable stiffness, Proceedings
of the National Academy of Sciences 115 (2018) 2032–2037.

13

View publication stats

You might also like