Accepted Manuscript Reshaping of tensegrities using a geometrical variation approach K. Koohestani PII: DOI: Reference:
S0020-7683(15)00289-9 http://dx.doi.org/10.1016/j.ijsolstr.2015.06.025 SAS 8828
To appear in:
International Journal of Solids and Structures
Received Date: Revised Date:
29 October 2014 13 May 2015
Please cite this article as: Koohestani, K., Reshaping of tensegrities using a geometrical variation approach, International Journal of Solids and Structures (2015), doi: http://dx.doi.org/10.1016/j.ijsolstr.2015.06.025
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
1
Reshaping of tensegrities using a geometrical variation
2
approach
3
K. Koohestani *
4
Department of Structural Engineering, Faculty of Civil Engineering, University of Tabriz, Tabriz, Iran
5
Abstract
6
We propose a novel approach for the reshaping of tensegrities and pre-stressed pin-
7
jointed frameworks. Both free standing and restrained frameworks are studied. First, a
8
geometrical variation is introduced and considered as an initial geometrical change to a
9
self-equilibrated tensegrity. A new nonlinear set of equilibrium equations is then
10
established based on this variation and subsequently linearized using an expansion. The
11
linearization provides us with an underdetermined system of linear equations with a
12
rectangular Jacobian matrix which is derived in a very simple and compact matrix form.
13
The system is finally solved iteratively using a Newton-Raphson method employing a
14
Moore-Penrose pseudo inverse operation to establish a new equilibrium state. The
15
method promises to generate very accurate results with a remarkable convergence rate.
16
The outstanding features and performance of the proposed method are demonstrated
17
and studied comprehensively through three examples.
18
Keywords: Reshaping; Geometrical variation; linearization; Newton-Raphson method
19
* Corresponding author. Tel.: +98 411 33392391; fax: +98 411 33356024.
20
Email address:
[email protected];
[email protected]
21
1
22
1. Introduction
23
A tensegrity is a lightweight, pin-jointed framework with a rigidity which results from a
24
self-stressed equilibrium between cables and struts. In recent years, a wide range of
25
innovative applications have been developed for tensegrities due to their unique
26
characteristics. We briefly refer to applications including aerospace structures (Sultan,
27
2009; Tibert and Pellegrino, 2002), architectural and structural design (Motro, 2003;
28
Rhode-Barbarigos et al. 2010), molecules and cells (Stamenovic and Ingber, 2009),
29
biomechanics (Luo et al. 2008), smart systems (Ali and Smith, 2010; Moored et al. 2011)
30
and advanced engineering materials (Fraternali et al. 2012).
31
The analysis and design of tensegrity structures normally depend on two main variables
32
(considering that the connectivity is known) including geometry and the pre-stress level
33
of its elements. In the most general case, both the geometry and pre-stress level of the
34
elements are unknown. The solution process which simultaneously determines the self-
35
stressed equilibrium and the corresponding geometry is called form-finding. Form-
36
finding of tensegrities is usually performed through analytical and numerical methods.
37
The numerical methods are more practical for large and asymmetrical tensegrities
38
whilst the analytical methods are more appropriate for tensegrities with a high order of
39
symmetry. Form-finding of tensegrities has been widely studied using different
40
numerical methods. Motro (1984) employed the dynamic relaxation method and
41
Pellegrino (1986) suggested a nonlinear programming approach. Sultan et al. (1999)
42
proposed a reduced coordinate method and Masic et al. (2005) developed an algebraic
43
method based on invariant tensegrity transformations. Form-finding of tensegrity
44
structures using concepts of finite elements has been presented by Pagitz and Mirats Tur
45
(2009). Estrada et al. (2006), Tran and Lee (2010, 2013) and Zhang and Ohsaki (2006)
2
46
proposed numerical methods that employ either iterative eigenvalues or singular value
47
decompositions of the force density and equilibrium matrices. Koohestani and Guest
48
(2013) introduced a combined form of the equilibrium and geometrical compatibility
49
equations for analytical and numerical form-finding of tensegrity structures. Recently,
50
Zhang et al. (2014) developed an efficient stiffness matrix-based form-finding method
51
for tensegrities and form-finding of irregular tensegrities has been studied by Li et al.
52
(2010) using the Monte Carlo method. Rieffel et al. (2009) proposed an evolutionary
53
form-finding method, Koohestani (2012), Paul et al. (2005) and Xu and Luo (2010) used
54
genetic algorithms and Chen et al. (2012a, b) used colony systems. Also, Koohestani
55
(2013) suggested an unconstrained minimization approach to the simultaneous form-
56
finding and design of tensegrities with geometrical and force constraints.
57
Kanno (2013a) proposed a mixed integer linear programming approach to the
58
formation of tensegrities with kinematical indeterminacy (see also Kanno, 2012, 2013b
59
for kinematically determinate tensegrities). The method needs no information about the
60
topology of a tensegrity but requires a ground structure with fixed nodes. As a result,
61
new and irregular tensegrities (from the topological point of view) can be formed.
62
Formation of irregular tensegrities has been studied by Zhang et al. (2006) using
63
dynamic relaxation method with kinetic damping. The method needs an initial geometry
64
but the configuration may be extended. Furthermore, during the form-finding process,
65
either the force or length of some elements can be fixed by an appropriate choice of
66
related stiffness.
67
A marching procedure for the form-finding of tensegrity structures has been presented
68
by Micheletti and Williams (2007a). They have also developed a method for the shape-
69
changing of tensegrities (Micheletti and Williams, 2007b). The method starts from an
70
initial known configuration and is based on the characterization of the rank-deficiency 3
71
manifold in the space of nodal coordinates. The change in shape is obtained using
72
element lengths as control parameters in compatibility equations. The compatibility
73
equations and orthogonality conditions form together a system of ordinary differential
74
equation and is solved efficiently with the Runge-Kutta integration method.
75
Symmetrical reconfiguration of tensegrity structures has been presented by Sultan et al.
76
(2002) employing nonlinear dynamics of tensegrities and symmetrical motions (see also
77
Sultan, 2014 for tensegrity deployment by imposing motions along infinitesimal
78
mechanisms). There are a considerable number of studies related to the control and
79
deployment of tensegrity structures. In the most of these methods either cables or struts
80
are served as actuators with adjustable pre-stress level and length i.e., shape changing is
81
preformed through changing the length of elements employing efficient control
82
algorithms. However note that these interesting research topics may have some
83
common points with those of our paper, but there are major differences in assumptions,
84
theory and objectives. We do not review these methods for the sake of brevity.
85 86
In this paper, we study the sensitivity of the equilibrium equations of a self-stressed
87
tensegrity (free standing or restrained) with respect to a geometrical variation. The
88
main aim of this paper is to establish a computational framework which enables us to
89
reshape a tensegrity based on an initial geometrical variation, while providing a new set
90
of force densities, nodal coordinates, and most importantly retains the self-stressed
91
state of the tensegrity. The main challenge is that we cannot necessarily form a self-
92
equilibrated tensegrity based on an arbitrarily selected geometrical variation. However,
93
it is possible to find an adequately close variation to the initially selected geometry
94
which corresponds to a self-stressed state. We address this challenge by proposing an
95
iterative approach which is based on a linearized form of the underdetermined 4
96
nonlinear system of equilibrium equations. A highly accurate solution with a very good
97
convergence rate is then obtained by using a Newton-Raphson method employing a
98
Moore-Penrose pseudo inverse operation. The method enables us to form a wide variety
99
of irregular tensegrities from those of regular or existing ones. Note that our method
100
only takes into consideration the change in the geometry of a tensegrity, therefore the
101
irregularity we pointed out is merely related to the geometry and subsequently in the
102
force density of elements but not topology of a tensegrity. Furthermore, this platform
103
may be viable for the design of tensegrities with special geometries or force
104
distributions. The performance of the method is comprehensively studied and tested
105
through three examples.
106
107
2. Self-equilibrium equations
108
This section describes two counterpart formulations of self-equilibrium equations for a
109
general three-dimensional tensegrity structure with n nodes and m elements. Consider
110
a typical node i which is adjacent (connected by an element) to all nodes belong to the
111
set Ω . The equilibrium of internal and external forces at node i in the x-, y- and z-
112
directions can be written as:
113
∑ f ik ( x i − x k ) / Lik = pix
(1)
k∈Ω
114
∑ f ik ( y i − y k ) / Lik = piy
(2)
k∈Ω
115
∑ f ik (z i − z k ) / Lik = piz
(3)
k∈Ω
5
116
where fik and Lik are the internal force and length of element ( i , k ) respectively and
117
pix , piy and piz are the external forces at node i in the x-, y- and z-directions. These
118
external forces are generally set to zero since tensegrities are self-equilibrated
119
structures. To form the equilibrium equations for a three-dimensional tensegrity, the
120
above equations must be defined and integrated for all nodes. This gives the force
121
density formulation of the equilibrium equations (Schek, 1974) as:
122
G[x , y , z] = [0, 0, 0] , G = BQBt
123
Here, G ∈R n×n is the force density matrix and Q = diag(q) , q = [q1 , q2 , K , qm ] t is the
124
diagonal matrix of force densities. Note that force density of an element, labeled by a
125
single number j , is defined as f j / L j . In Eq. (4), x = [x 1 , x 2 ,K , x n ] t , y = [ y1 , y2 ,K , yn ] t and
126
z = [z1 , z 2 ,K , z n ] t are vectors of the Cartesian coordinates of nodes. B = [bij ]n×m , known as
127
node-member incidence matrix, is defined as:
128
− 1 if i is the start node of element j bij = 1 if i is the end node of element j 0 otherwise
129
Equation (4) plays a crucial role in the form-finding of tensegrity structures. In general,
130
for a d-dimensional tensegrity in a state of self-stress, Eq. (4) must have at least d
131
independent solutions. However, due to the intrinsic rank deficiency of G , there is
132
always a trivial solution which is a vector of ones. Thus the minimum rank deficiency of
133
G for a d-dimensional tensegrity is d + 1 . This also means that G must have at least d + 1
134
zero eigenvalues. The stability of tensegrities may also be studied using the spectral
135
characteristics of the force density matrix, i.e. G (Zhang and Ohsaki, 2007). However,
136
this form of the equilibrium equations is appropriate only when finding a set of nodal
(4)
(5)
6
137
coordinates corresponding to a set of known force densities of the elements. For cases in
138
which the geometry of a tensegrity is known, whilst the vector of force densities is
139
unknown, another form of the equilibrium equations is used. Equation (6) gives the
140
relationships between the elements projected lengths and nodal coordinates as:
141
d x = B t x , d y = B t y , d z = Bt z
142
where d x , d y and d z are the vectors of the elements projected lengths in the x-, y- and z-
143
directions respectively. Substitution of the relationships given in Eq. (6) into Eq. (4)
144
yields three sets of equilibrium equations:
145
BQd x = B diag(d x )q = 0
(7-a)
146
BQd y = B diag(d y )q = 0
(7-b)
147
BQd z = B diag (d z )q = 0
(7-c)
148
By combining Eqs. (7a-c), an alternative set of equilibrium equations is expressed as:
149
B diag(d x ) Aq = 0 , A = B diag(d y ) B diag(d z )
150
where A ∈R dn×m (d is 2 or 3 for two or three-dimensional tensegrities) is the rectangular
151
equilibrium matrix. For cases where the geometry of a tensegrity is known, the
152
coefficient matrix in Eq. (8), A is known and a single state or multiple states of self-
153
stressed equilibrium can be identified according to the dimension of its null space. Note
154
that different bases can be calculated for the null space of a matrix (Stewart, 1973)
155
including triangular bases (normally obtained by Gauss elimination or LU factorization
156
methods) and orthogonal bases (normally obtained by singular value decomposition,
157
SVD).
(6)
(8)
7
158
159
3. Stability
160
In general, statically and kinematically indeterminate pin-jointed structures exhibit
161
different kinds of stability. However, in this paper the stability of a tensegrity is checked
162
after reshaping on the base of the pre-stress stability and super-stability. In general, a d-
163
dimensional tensegrity structure needs to satisfy the following three conditions to be
164
the super-stable (Zhang and Ohsaki, 2007).
165
(a) The force density matrix G has the minimum rank deficiency d + 1.
166
(b) G is positive semi-definite.
167
(c) The rank of the geometry matrix is d (d + 1)/2 or, equivalently, the member
168
directions do not lie on the same conic at infinity (Connelly, 1982).
169 170
It is useful to note that the first two conditions above play a crucial role in the study of
171
the stability of tensegrity structures since the third condition is usually satisfied (Zhang
172
and Ohsaki, 2012).
173
A tensegrity that does not meet above conditions is not super-stable but may still be
174
stable. However, for these cases, the stability can be investigated based on a pre-
175
stress/stiffness ratio of elements and spectral characteristics of the tangent stiffness
176
matrix (sum of the linear elastic stiffness and geometrical stiffness matrices). The reader
177
may refer to Ohsaki and Zhang (2006) for the necessary and sufficient conditions for the
178
stability of pin-jointed structures, including tensegrities.
179
A kinematically indeterminate tensegrity is pre-stress stable if the state of self-stress
180
stiffens all the infinitesimal mechanisms (excluding rigid body motions). The left null
181
space of the equilibrium matrix A forms a basis for the infinitesimal mechanisms of a
8
182
tensegrity. This basis can effectively be calculated employing the singular value
183
decomposition (SVD) as follows.
184
[U, V , W ] = SVD ( A )
185
where the vectors corresponding to infinitesimal mechanics (displacement modes that
186
do not elongate elements) are those vectors in U , entirely denoted by matrix U m , that
187
satisfies A t U m = 0 .
188
The stability of a tensegrity can be assessed by checking the positive definiteness of its
189
tangent stiffness matrix. However if an infinitesimal mechanism is applied to this matrix,
190
the terms involving equilibrium matrix are vanished (see Estrada et al. 2007, and notes
191
about second order rigidity in Connelly and Whiteley, 1996) therefore the only term
192
which may impart stiffness is a section of geometric stiffness, i.e. I d ⊗ G . The pre-stress
193
stability of a tensegrity is then checked by evaluating eigenvalues of the following
194
quadratic form.
195
Λ = U tm (I d ⊗ G)U tm
196
Note that this is the case for a tensegrity with a single state of self-stress. For multiple
197
states of self-stress the process is more complicated (Calladine and Pellegrino, 1991;
198
Pellegrino and Calladine, 1986).
199
In order to ensure that self-stress state stiffens infinitesimal mechanisms and exerts
200
positive work, all eigenvalues of the above-mentioned quadratic form must be positive
201
excluding 3 or 6 zero eigenvalues corresponding to rigid body motions in 2 or 3-
202
dimensional spaces.
203
The method proposed in this paper does not check the stability of a tensegrity during
204
the reshaping process, however, since we normally work based on small geometric
205
variations, the method normally preserves the stability. In section 7 we demonstrate this
(9)
(10)
9
206
feature through numerical examples but formation of an unstable tensegrity after
207
reshaping implies that the geometrical variation considered is not appropriate and must
208
be modified.
209
210
4. Definition of a geometrical variation
211
This section describes geometrical variations and how these variations affect the
212
equilibrium equations. Consider we have a tensegrity in a state of self-stress where its
213
nodal coordinates and force density of elements are denoted by vectors x , y , z and q
214
respectively. The equilibrium equations for this tensegrity can be written based on Eq.
215
(4) as follows:
216
(B diag(q)B t )[x , y , z] = [0, 0, 0]
217
Now let a variation in geometry be imposed on the tensegrity by applying changes
218
(denoted by ∆x , ∆y , ∆z ) to the nodal coordinates. The new set of equilibrium equations
219
can be written as:
220
(B diag(q)Bt )[x + ∆x , y + ∆y , z + ∆z] ≠ [0, 0, 0]
221
Clearly, we generally lose the equilibrium at all nodes, i.e. there will be out-of-balance
222
forces at all nodes. It is not possible to find a new set of force densities ( q + ∆q ) as given
223
in Eq. (13) since there is no one-to-one correspondence between an arbitrarily selected
224
geometry and a set of self-equilibrated force densities in tensegrities.
225
(B diag(q + ∆q)Bt )[x + ∆x , y + ∆y , z + ∆z] ≠ [0, 0, 0]
(11)
(12)
10
(13)
226
Only special geometries with their corresponding force densities provide a state of self-
227
stressed equilibrium. In this paper we aim to find a close variation to the predefined
228
geometry in such way that we could find a corresponding set of force densities ( q' ) and
229
retain the self-stressed equilibrium. This process is mathematically represented as
230
follows:
231
(B diag(q' )B t )[x + ∆x + ∆x' , y + ∆y + ∆y' , z + ∆z + ∆z' ] = [0, 0, 0]
232
In Eq. (14) [x + ∆x + ∆x' , y + ∆y + ∆y' , z + ∆z + ∆z' ] represents a new geometry close to that
233
generated by the initial variation in the nodal coordinates.
(14)
234
235
5. Proposed computational framework (free standing tensegrities)
236
The equilibrium equations given in Eq. (11) can be expanded by considering that the
237
three vectors of the nodal coordinates are joined together to make a single vector of
238
3n × 1 as follows:
239
B diag(q )B t f (q , x , y , z ) =
B diag(q )B
t
x 0 y = 0 B diag(q )B t z 0
(15)
240
In this new form of equilibrium equations f (q , x , y , z ) is a vector-valued function
241
consisting of equilibrium equations of all nodes in the x-, y- and z- directions as given in
242
Eq. (16).
11
243
f 1 x 0 M M f nx 0 f 1 y 0 f (q , x , y , z ) = M = M f ny 0 f 0 1z M M f nz 0
244
For example, f 1 x is the equilibrium equation in node 1 and in the x- direction. This vector
245
is clearly a function of the force density of all members and nodal coordinates of the
246
framework. Since we are investigating the case of a self-stressed equilibrium, the right
247
hand vector must be a zero vector.
248
We partition the node-member incidence matrix ( B ) into n rows as follows:
249
b1 b B = 2 M b n
250
where, b i is the ith row of B . According to the above definitions, the equilibrium
251
equation of node i in x-, y-, and z- directions can be expressed as:
(16)
(17)
252
f ix = b i diag(q )B t x
(18-a)
253
f iy = b i diag(q)B t y
(18-b)
254
f iz = bi diag(q )Bt z
(18-c)
255
The equilibrium equations defined in Eq. (15) are an underdetermined nonlinear system
256
of equations. The system is underdetermined since there are 3n equations with 3n + m
12
257
variables. We linearize this system of equations using the expansion process given in Eq.
258
(19).
259
f (q + ∆q , x + ∆x , y + ∆y , z + ∆z) ≈ f (q , x , y , z ) + J (q , x , y , z ) Ψ
260
In Eq. (19) J is the matrix of all first-order partial derivatives of the vector valued
261
function f (q , x , y , z ) , known as the Jacobian matrix, and Ψ is a vector of small variations
262
in the variables.
263
(19)
∆q ∆x Ψ= ∆y ∆z
(20)
264
The Jacobian matrix can be calculated using:
265
∂f J= ∂q
266
where its expanded form is as follows:
267
∂f 1 x ∂q 1 M ∂f nx ∂q 1 ∂f 1 y ∂q 1 J= M ∂f ny ∂q 1 ∂f 1z ∂q 1 M ∂f nz ∂q 1
∂f ∂x
∂f ∂y
∂f 1 x ∂qm M ∂f nx L ∂qm ∂f 1 y L ∂qm M ∂f ny L ∂qm ∂f 1z L ∂qm M ∂f nz L ∂qm L
∂f ∂z
∂f 1 x ∂x 1 M ∂f nx ∂x 1 ∂f 1 y ∂x 1 M ∂f ny ∂x 1 ∂f 1 z ∂x 1 M ∂f nz ∂x 1
(21)
∂f 1 x ∂x n M ∂f nx L ∂x n ∂f 1 y L ∂x n M ∂f ny L ∂x n ∂f 1z L ∂x n M ∂f nz L ∂x n L
∂f 1 x ∂y 1 M ∂f nx ∂y 1 ∂f 1 y ∂y 1 M ∂f ny ∂y 1 ∂f 1z ∂y 1 M ∂f nz ∂y 1
∂f 1 x ∂y n M ∂f nx L ∂y n ∂f 1 y L ∂y n M ∂f ny L ∂y n ∂f 1z L ∂y n M ∂f nz L ∂y n L
13
∂f 1 x ∂z 1 M ∂f nx ∂z 1 ∂f 1 y ∂z 1 M ∂f ny ∂z 1 ∂f 1z ∂z 1 M ∂f nz ∂z 1
L
L L
L L
L
∂f 1 x ∂z n M ∂f nx ∂z n ∂f 1 y ∂z n M ∂f ny ∂z n ∂f 1 z ∂z n M ∂f nz ∂z n
(22)
268
Based on the symbolic definition of the equilibrium equation of a typical node i, we
269
calculate all the derivatives, simplify the results, and integrate them simultaneously. We
270
may then express the Jacobian of f (q , x , y , z) in a very simple form according to the
271
known matrices defined earlier. This simple and compact form is given in Eq. (23).
272
J = [ A I 3 ⊗ (B diag(q )B t ) ]
273
We next adopt a Newton-Raphson scheme in order to iteratively solve the nonlinear
274
system of equations given in Eq. (15) as follows.
275
f ( q j + 1 , x j + 1 , y j +1 , z j +1 ) = f ( q j , x j , y j , z j ) + J ( q j , x j , y j , z j ) Ψ j
276
where,
(23)
(24)
277
∆q j j ∆x Ψj = j ∆y j ∆z
278
q j +1 = q j + ∆q j
(26-a)
279
x j +1 = x j + ∆x j
(26-b)
280
y j +1 = y j + ∆y j
(26-c)
281
z j +1 = z j + ∆z j
(26-d)
282
Our aim is to find a new state of self-stressed equilibrium by imposing a moderately
283
small variation on the geometry of the tensegrity. We start the iterative process given in
284
Eq. (24) from an equilibrium state so that f (q 0 , x 0 , y 0 , z 0 ) = 0 . Clearly by imposing a slight
285
variation on the geometry (applying Ψ 0 ) we lose the equilibrium at all nodes. This can
(25)
14
286
be mathematically represented through Eq. (27) where p j is the out-of-balance force
287
vector in the jth iteration.
288
f (q j , x j , y j , z j ) = p j
289
Our problem is to find a geometry close to the predefined one in such a way that it
290
enables us to eliminate all out-of-balance forces i.e. p j = 0 . Hence we first calculate an
291
approximate variation in the force density of elements ( ∆q 0 ) using the following
292
equation. Note that at this stage the geometry of the tensegrity is known.
293
∆q 0 = ( A 0 )+ ( −I 3 ⊗ (B diag(q 0 )B t Ψ 0 )
294
In Eq. (28) ( A 0 )+ is the Moore-Penrose pseudo-inverse (Golub and Van Loan, 1996) of
295
the coefficient matrix of the equilibrium equations ( A ) that was calculated for the initial
296
geometry. It should be noted that we have an underdetermined system of equations
297
therefore a pseudo-inverse need to be used to calculate ∆q 0 . Due to the linearization and
298
since the solution is not unique the equilibrium will not be established at this step,
299
therefore:
300
f (q j +1 , x j +1 , y j +1 , z j +1 ) ≠ 0
301
We now have access to a set of variations of the force densities as well as the nodal
302
coordinates. The iterative process commences at this point with the aim of reducing the
303
out-of-balance force vector as far as is required. We first calculate the rectangular
304
Jacobian matrix and the out-of-balance force vector using the current information of
305
force densities and nodal coordinates, after which better approximations to the
306
geometrical and force variables are calculated using the following equation:
(27)
j = 1, K , k
(28)
(29)
15
307
Ψ j = ( J (q j , x j , y j , z j ))+ (−f (q j , x j , y j , z j ))
308
where the plus sign on the Jacobian matrix again refers to a pseudo-inverse operator.
309
The infinity norm of the new out-of-balance force vector is then calculated and
310
compared with a predefined very small value ε as represented in Eq. (31). The above
311
process is iterated until convergence is achieved.
312
f (q j +1 , x j +1 , y j +1 , z j +1 )
≤ε inf
(30)
ε = 1 × 10 −12
(31)
313
After convergence, we obtain a new set of force densities and nodal coordinates which
314
provide us with a new configuration in a state of self-stressed equilibrium. Note that if
315
the initial variation in geometry is considered moderately small the method normally
316
converges very fast and the final geometry is close to the initial one. However, if a very
317
large variation in geometry is considered the method may converge to a degenerated
318
geometry. As a result, large geometrical variations must be imposed through iterative
319
imposing of small variations and by using this approach, it is also possible to find self-
320
equilibrated configurations even for large geometrical variations.
321
322
6. Proposed computational frame work (restrained tensegrities)
323
The method proposed to this point has been formulated for free standing tensegrities;
324
however it can be generalized in order to reform general pre-stressed pin-jointed
325
frameworks with some restrained nodes. Here, we provide another version of the
326
proposed method for this type of framework. First we consider that all nodes of a pin-
327
jointed pre-stressed framework are divided into two sets, including free and restrained
328
nodes. All variables related to the free and restrained nodes are identified by adding the
16
329
letters f and r as subscripts. Accordingly, we partition the node-member incidence
330
matrix into two parts: the first part, denoted by B f , corresponds to the free nodes; and
331
the second part, denoted by B r , corresponds to the restrained nodes. These are
332
expressed as follows:
333
B B= f Br
334
According to this partitioning, the equilibrium equations given in Eq. (15) need to be
335
modified in order to establish the self-stressed equilibrium at all free nodes. Eq. (33)
336
gives the new set of equilibrium equations as:
337
B f diag(q )B t f f (q , x , y , z ) =
338
The linearized form of the equilibrium equations for free nodes is also as follows:
339
f f (q j +1 , x j +1 , y j +1 , z j +1 ) = f f (q j , x j , y j , z j ) + J f (q j , x j , y j , z j ) Ψ jf
340
For this set of equations we again calculate the rectangular Jacobian matrix where its
341
simplified representation is given in Eq. (35):
342
J f = [ A f I3 ⊗ (B f diag(q)B tf ) ]
343
Note that in Eq. (35) A f is the coefficient matrix of the equilibrium equations related to
344
the free nodes which is mathematically represented as:
345
B f diag(d x ) A f = B f diag(d y ) , d x = B t x , d y = B t y , d z = B t z B f diag(d z )
(32)
B f diag(q )B
t
x 0 y = 0 t B f diag(q )B z 0
(33)
(34)
(35)
17
(36)
346
The solution process for these kinds of frameworks is almost identical to that presented
347
earlier with, however, some slight changes in the variables. We start from a self-stressed
348
state (only related to free nodes) i.e. f f (q 0 , x 0 , y 0 , z 0 ) = 0 . Then the first variation in the
349
force densities is calculated as follows:
350
∆q 0 = ( A 0f ) + ( −I3 ⊗ (B f diag(q 0 )B tf Ψ 0f )
351
in which
(37)
352
∆q j j ∆x j Ψ f = fj ∆y jf ∆z f
353
x jf +1 = x jf + ∆x jf ,
∆x rj = 0 ,
x rj +1 = x 0r
(39-a)
354
y jf +1 = y jf + ∆y jf ,
∆y rj = 0 , y rj +1 = y 0r
(39-b)
355
z jf +1 = z jf + ∆z jf ,
∆z rj = 0 , z rj +1 = z 0r
(39-c)
356
Note that in the above relationships x 0r , y 0r and z 0r are the nodal coordinates of the
357
restrained nodes that are fixed during the iterations (i.e. the geometrical variations
358
related to these nodes are all zero).
359
The new vector of iterated variations (including variations in the force densities and
360
geometrical variations) is then calculated through Eq. (40):
361
Ψ jf = ( J f (q j , x j , y j , z j ))+ ( −f f (q j , x j , y j , z j ))
362
The condition for convergence is also defined as:
(38)
(40)
18
363
f f (q j +1 , x j +1 , y j +1 , z j +1 )
≤ε inf
ε = 1 × 10 −12
(41)
364
365
7. Examples
366
7.1. Example 1
367
In Fig. 1(a) a two-dimensional tensegrity is shown. We study this tensegrity with respect
368
to two cases. In the first case there are no restrained nodes i.e. the tensegrity is free
369
standing, while in the second case the horizontal strut is removed and its two end nodes
370
are restrained (Fig. 1(c)).
371
First case: In this case the tensegrity has 6 nodes and 9 elements and is free standing. An
372
analytical solution for a self-stressed state of this tensegrity is available in the literature
373
(Koohestani and Guest, 2013). In Table 1 the corresponding force density of elements
374
are provided. Clearly, the initial configuration is regular with two axis of symmetry
375
whilst the force densities of the elements are packed into three groups. Now we assume
376
a random set of nodal geometrical variations as given in Table 2 and seek to find a new
377
self-stressed state through the proposed method. The method converges in only 5
378
iterations at which point a new self-stressed state is reached. The new set of force
379
densities (all elements have a different force density value) and nodal coordinates are
380
given in Tables 1-2. The corresponding configuration is depicted in Fig. 1(b) where the
381
new tensegrity is completely irregular. Note that even though the initial nodal variations
382
are assumed to be very large (in comparison with the initial nodal coordinates) the
383
method efficiently converged after only 5 iterations and generated a feasible solution.
384
The convergence history (infinity norm of the out-of-balance nodal forces vs. iteration
385
number) of the numerical procedure is shown in Table 3. In this case, we have 19
386
investigated both pre-stress stability and super-stability of the tensegrity. The results
387
provided in Tables 4-5 (based on the context given in Section 3) verify that the stability
388
is preserved during reshaping process.
389
390
Table 1
391
Set of force densities before and after imposing nodal geometrical variations (example 1,
392
case 1) Element Number 1 2 3 4 5 6 7 8 9
Initial force density ( q 0 ) 2.0000 1.0000 2.0000 2.0000 1.0000 2.0000 -1.0000 -1.0000 -1.0000
Final force density ( q 5 ) 1.5349 1.1844 2.0471 2.4802 1.1893 1.7575 -0.9973 -0.8673 -1.0543
393
394
Table 2
395
Set of nodal coordinates before and after imposing nodal geometrical variations
396
(example 1, case 1) Initial nodal coordinates (self-stressed)
Nodal variations
Final nodal coordinates (self-stressed)
x0
y0
∆x 0
∆y 0
x5
y5
-1.0000 -1.0000 -2.0000 2.0000 1.0000 1.0000
1.5275 -1.5275 0.0000 0.0000 1.5275 -1.5275
0.9138 0.7067 0.5578 0.3134 0.1662 0.6225
0.9879 0.1704 0.2578 0.3968 0.0740 0.6841
-0.2208 -0.2055 -1.3824 2.2706 1.2975 1.5209
2.3323 -1.3794 0.3598 0.3323 1.7633 -0.8372
397
20
398 399
Fig. 1 a) initial self-stressed configuration (free standing); b) configuration after
400
imposing nodal geometrical variations (free standing); c) initial self-stressed
401
configuration (restrained); and d) configurations after imposing nodal geometrical
402
variations (restrained)
403
404
Table 3
405
Infinity norm of the out-of-balance force vector at different iterations (example 1)
f f
inf
inf
(case 1)
1 0.6955
2 0.0836
(case 2)
0.9999
0.0240
Iteration number 3 4 3.6025e-4 3.4055e-9 3.9703e-5
406 21
4.8408e-11
5 7.1054e-15 1.7764e-15
407
408
Table 4
409
Infinitesimal mechanisms and the corresponding work done by the state of self-stress
410
before and after reshaping (example 1, case 1) Before reshaping Node
Free standing (1st infinitesimal mode)
Free standing (2nd infinitesimal mode)
Free standing (3rd infinitesimal mode)
Free standing (4th infinitesimal mode)
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
eig (U tm (I3
⊗ G)U m ) =
[0.0,0.0,0.0,1.3333] ux uy
-0.0222 0.1331 0.0555 0.0555 -0.0222 0.1331 -0.6126 0.1413 -0.2357 -0.2357 -0.6126 0.1413 0.0935 0.5003 0.2969 0.2969 0.0935 0.5003 0.0299 -0.3121 -0.1411 -0.1411 0.0299 -0.3121
-0.0356 -0.0356 -0.0865 -0.5275 -0.5784 -0.5784 0.0173 0.0173 -0.2295 0.2035 -0.0432 -0.0432 -0.2343 -0.2343 -0.3675 0.2164 0.0832 0.0832 -0.5360 -0.5360 -0.4240 -0.0919 0.0201 0.0201
411
412
413
414 22
After reshaping eig (Utm (I3 ⊗ G)U m ) = [0.0,0.0,0.0,1.8061] ux uy
-0.0991 0.3004 0.1132 0.1132 -0.2346 0.3974 -0.3871 0.3329 -0.0045 -0.0016 -0.4413 0.3577 -0.4615 -0.1946 -0.3197 -0.3163 -0.3280 -0.3067 0.0819 0.3579 0.2286 0.2336 0.3198 0.1631
-0.1487 -0.1470 -0.2737 -0.2736 -0.5101 -0.4558 0.2823 0.2853 0.0570 0.4367 0.1377 0.2064 -0.3049 -0.3038 -0.3885 0.0593 0.0514 0.0532 -0.3315 -0.3304 -0.4179 0.2446 0.3033 0.2898
415
Table 5
416
Eigenvalues of the force density matrix before and after reshaping (example 1, case 1)
λ1 λ2 λ3 λ4 λ5 λ6
Eigenvalues before reshaping 0.0000 0.0000 0.0000 2.0000 6.0000 6.0000
Eigenvalues after reshaping 0.0000 0.0000 0.0000 2.5232 5.1391 6.8867
417
418
Second case: In this case the tensegrity has 6 nodes and 8 elements and is not free
419
standing. The initial geometry is the same as first case but we have removed the
420
horizontal strut and restrained both ends. The initial self-stressed state remains
421
constant during this process. Now we apply a new set of nodal geometrical variations
422
and try to find a corresponding self-stressed state using our method. The method again
423
successfully determines a set of force densities corresponding to the initial variations in
424
5 iterations. The results are summarized in Tables 6-7. The final shape and history of the
425
convergence are depicted in Fig. 1(d) and Table 3, respectively. In this case, again the
426
stability is preserved during reshaping; see Table 8 for the results in more detail.
427
428
429
430
431
23
432
Table 6
433
Set of force densities before and after imposing nodal geometrical variations (example 1,
434
case 2) Element Number 1 2 3 4 5 6 7 8
Initial force density ( q 0 ) 2.0000 1.0000 2.0000 2.0000 1.0000 2.0000 -1.0000 -1.0000
Final force density ( q 5 ) 0.7939 0.8369 2.2859 3.6742 1.5160 1.2771 -0.4931 -1.4056
435
436
437
Table 7
438
Set of nodal coordinates before and after imposing nodal geometrical variations
439
(example 1, case 2) Initial nodal coordinates (self-stressed)
Nodal variations
Final nodal coordinates (self-stressed)
x0
y0
∆x 0
∆y 0
x5
y5
-1.0000 -1.0000 -2.0000 2.0000 1.0000 1.0000
1.5275 -1.5275 0.0000 0.0000 1.5275 -1.5275
0.5306 0.8324 0.0000 0.0000 0.2992 0.4526
0.4226 0.3596 0.0000 0.0000 0.4243 0.4294
-0.4558 -0.0520 -2.0000 2.0000 1.2471 1.4577
1.9224 -1.1951 0.0000 0.0000 1.9091 -1.1877
440
441
442
443
24
444
Table 8
445
Infinitesimal mechanisms and the corresponding work done by the state of self-stress
446
before and after reshaping (example 1, case 2) Before reshaping Node
Second case (restrained)
1 2 5 6
eig
(U tm (I 3
⊗ G)U m ) = 1.2000
After reshaping eig (U tm (I 3 ⊗ G)U m ) = 1.3197
ux
uy
ux
uy
0.4183 -0.4183 0.4183 -0.4183
-0.2739 -0.2739 0.2739 0.2739
-0.4607 0.2864 -0.4650 0.2893
0.3700 0.4668 -0.1834 -0.1321
447
448
7.2. Example 2
449
In this example reshaping of a truncated tetrahedron tensegrity is investigated. This
450
tensegrity is one of the most well-known and has been widely studied. The tensegrity
451
has 12 nodes and 24 elements (including 18 cables and 6 struts) and is free standing. A
452
set of force densities corresponding to a self-stressed state is extracted from an
453
analytical solution available in Tibert and Pellegrino (2002) and summarized in Table 9.
454
We also calculate the nodal coordinates of this tensegrity using the approach introduced
455
in Koohestani and Guest (2013). The configuration obtained and the nodal coordinates
456
(multiplied by 10) are depicted in Fig. 2(a) and provided in Table 10. The reshaping of
457
this tensegrity is carried out based on the random set of nodal geometrical variations
458
given in Table 10. The proposed method converges after 9 iterations (the history is
459
summarized in Table 11) and forms a new irregular truncated tetrahedron with a
460
feasible set of force densities, as shown in Fig. 2(b). Note that all elements now have
461
different force densities. The corresponding nodal coordinates are provided in Table 10.
462
We also check the viability of our results by calculating the first four eigenvalues of the 25
463
force density matrix all of which are of the order 1e-15. This also verifies the super-
464
stability of the tensegrity after reshaping (see Table 12).
465
466
Table 9
467
Set of force densities before and after imposing nodal geometrical variations (example
468
2) Element Number 1 2 3 4 5 6 7 8 9 10 11 12
Initial force density ( q 0 ) 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000
Final force density ( q 9 ) 0.5307 1.4747 1.0284 0.6277 1.7149 0.6160 0.8636 1.1055 0.6438 1.1197 0.6853 1.7006
Element Number 13 14 15 16 17 18 19 20 21 22 23 24
469
470
471
472
473
474
475
476
26
Initial force density ( q 0 ) 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 -0.5746 -0.5746 -0.5746 -0.5746 -0.5746 -0.5746
Final force density ( q 9 ) 1.0084 0.8532 0.7765 0.7037 1.5288 0.8610 -0.4720 -0.4614 -0.4604 -0.7455 -0.4642 -0.6265
477
Table 10
478
Set of nodal coordinates before and after imposing nodal geometrical variations
479
(example 2) Initial nodal coordinates (self-stressed) y0 1.0386 -1.0281 -0.2698 -1.5127 -2.6805 -2.4113 -1.0585 1.0128 0.2283 2.6958 1.5542 2.4313
x0 2.2633 2.4396 1.0871 0.5608 0.9890 -0.9941 -2.6605 -2.7455 -1.9136 -0.6831 0.2657 1.3914
z0 1.6190 1.3466 2.7509 -2.4939 -0.8118 -1.4210 0.7895 0.5083 2.2601 -1.0431 -2.5171 -0.9874
Nodal variations ∆x 0 0.1920 0.1389 0.6963 0.0938 0.5254 0.5303 0.8611 0.4849 0.3935 0.6714 0.7413 0.5201
∆y 0 0.3477 0.1500 0.5861 0.2621 0.0445 0.7549 0.2428 0.4424 0.6878 0.3592 0.7363 0.3947
∆z 0 0.6834 0.7040 0.4423 0.0196 0.3309 0.4243 0.2703 0.1971 0.8217 0.4299 0.8878 0.3912
Final nodal coordinates (self-stressed) x9 2.4249 2.7984 1.7289 0.7617 1.4453 -0.5446 -1.8948 -2.0784 -1.4931 -0.0684 0.9128 1.8562
y9 1.4323 -0.8165 0.3056 -1.3020 -2.6153 -1.6724 -0.7495 1.5274 0.7654 3.1395 2.1836 2.8104
z9 2.2074 2.0798 3.0658 -2.4350 -0.3283 -1.1597 0.9753 0.7517 3.2343 -0.6298 -1.6022 -0.5569
480
481
482
Table 11
483
Infinity norm of the out-of-balance force vector in different iterations (example 2)
f
inf
1 0.7111
2 0.0440
Iteration number 3 4 5 6.2670e-5 1.400e-10 1.1990e-10
484
27
… …
9 2.41e-14
485 486
Fig. 2 a) initial truncated tetrahedron tensegrity; and b) new form after imposing nodal
487
geometrical variations
488
489
Table 12
490
Eigenvalues of the force density matrix for the truncated tetrahedron tensegrity before
491
and after reshaping
λ1 λ2 λ3 λ4 λ5 λ6 λ7 λ8 λ9 λ10 λ11 λ12
Eigenvalues before reshaping 0.0000 0.0000 0.0000 0.0000 3.0000 3.0000 3.0782 3.0782 3.0782 4.6234 4.6234 4.6234
492
493
28
Eigenvalues after reshaping 0.0000 0.0000 0.0000 0.0000 2.0685 2.3620 2.4700 3.4987 3.6714 4.5320 4.7996 5.8231
494
7.3. Example 3
495
In Fig. 3(a) a 12-plex cylindrical tensegrity is shown. This tensegrity has 24 nodes and
496
36 elements including 24 cables and 12 struts. All 12 nodes that are located in x-y plane
497
(z=0) are considered to be restrained. For this type of tensegrity (in general for an n-
498
plex) an analytical solution corresponding to a state of self-stressed equilibrium is
499
available (Tibert and Pellegrino, 2002). The self-stressed equilibrium is reached when
500
all nodes at the top of the tensegrity are rotated about the z-axis by an angle
501
θ = (π / 2) − (π / n) = 5π / 12 . The force density of cables and struts are accordingly
502
obtained as 1, 2 sin(π / 12) and − 2 sin(π / 12) as given in Table 13. We impose a set of
503
nodal geometrical variations to the free nodes (nodes at z=2) as defined in Table 14. The
504
proposed method successfully forms an irregular tensegrity as shown in Fig. 3(b) in only
505
5 iterations.
506 507 508 509
Fig. 3 a) an initial restrained 12-plex cylindrical tensegrity; and b) new form after imposing nodal geometrical variations to top free nodes
29
510
511
Note that this outstanding performance is achieved even though the assumed variations
512
are moderately large. The final nodal coordinates and infinity norm of the out-balance
513
force vector are provided in Tables 14-15. This tensegrity has one infinitesimal
514
mechanism which has been stiffened by the single state of self-stress (see the results
515
given in Table 16 before and after reshaping).
516 517 518
Table 13
519
Set of force densities before and after imposing nodal geometrical variations (example
520
3) Element Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
Initial force density ( q 0 ) 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 1.0000 0.5176 0.5176 0.5176 0.5176 0.5176 0.5176
Final force density ( q 5 ) 1.0128 0.9922 0.9714 0.9291 0.8672 0.5920 0.6583 0.5887 0.5743 0.5774 0.5973 0.7432 1.4110 0.9939 0.9311 0.8907 0.9162 0.4359
Element Number 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
521
522
30
Initial force density ( q 0 ) 0.5176 0.5176 0.5176 0.5176 0.5176 0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176 -0.5176
Final force density ( q 5 ) 0.3890 0.7026 0.6701 0.6660 0.6696 1.2051 -0.9788 -0.9235 -0.8859 -0.8792 -0.5599 -0.4712 -0.6780 -0.6662 -0.6636 -0.6732 -1.1046 -1.2430
523
Table 14
524
Set of nodal coordinates before and after imposing nodal geometrical variations
525
(example 3) Initial nodal coordinates (self-stressed) y0 0.2588 -0.2588 -0.7071 -0.9659 -0.9659 -0.7071 -0.2588 0.2588 0.7071 0.9659 0.9659 0.7071 1.0000 0.8660 0.5000 0.0000 -0.5000 -0.8660 -1.0000 -0.8660 -0.5000 0.0000 0.5000 0.8660
x0 0.9659 0.9659 0.7071 0.2588 -0.2588 -0.7071 -0.9659 -0.9659 -0.7071 -0.2588 0.2588 0.7071 0.0000 0.5000 0.8660 1.0000 0.8660 0.5000 0.0000 -0.5000 -0.8660 -1.0000 -0.8660 -0.5000
z0 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 2.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
Nodal variations ∆x 0 0.9659 0.9659 0.7071 0.2588 -0.2588 -0.7071 -0.9659 -0.9659 -0.7071 -0.2588 0.2588 0.7071 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
∆y 0 0.2588 -0.2588 -0.7071 -0.9659 -0.9659 -0.7071 -0.2588 0.2588 0.7071 0.9659 0.9659 0.7071 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
∆z 0 0.0000 0.4000 0.8000 1.2000 1.6000 2.0000 2.0000 1.6000 1.2000 0.8000 0.4000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
Final nodal coordinates (self-stressed) x5 1.8470 1.8604 1.4015 0.5891 -0.3901 -1.3571 -2.0386 -2.0389 -1.5278 -0.5836 0.5111 1.4101 0.0000 0.5000 0.8660 1.0000 0.8660 0.5000 0.0000 -0.5000 -0.8660 -1.0000 -0.8660 -0.5000
y5 0.4999 -0.4283 -1.2634 -1.7823 -1.8571 -1.4884 -0.4717 0.4724 1.4298 1.9995 1.9999 1.4278 1.0000 0.8660 0.5000 0.0000 -0.5000 -0.8660 -1.0000 -0.8660 -0.5000 0.0000 0.5000 0.8660
z5 2.0152 2.3279 2.6827 3.0661 3.4826 4.0777 4.0954 3.6000 3.1962 2.8037 2.4252 2.0445 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
526
527
528
Table 15
529
Infinity norm of the out-of-balance force vector in different iterations (example 3)
f
inf
1 0.6941
Iteration number 2 3 4 0.0811 8.8969e-4 4.0463e-7
530
31
5 2.0372e-14
531
Table 16
532
Infinitesimal mechanisms and the corresponding work done by the state of self-stress
533
before and after reshaping Before reshaping eig (Utm (I 3
Node ux
1 2 3 4 5 6 7 8 9 10 11 12
0.0673 -0.0673 -0.1838 -0.2511 -0.2511 -0.1838 -0.0673 0.0673 0.1838 0.2511 0.2511 0.1838
After reshaping
⊗ G)U m ) = 0.2173 uz uy
-0.2511 -0.2511 -0.1838 -0.0673 0.0673 0.1838 0.2511 0.2511 0.1838 0.0673 -0.0673 -0.1838
-0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255 -0.1255
eig (U tm (I 3 ⊗ G)U m ) = 0.1936 ux uz uy
-0.0676 0.0657 0.1769 0.2433 0.2507 0.1982 0.0752 -0.0754 -0.2071 -0.2811 -0.2758 -0.1931
0.2522 0.2452 0.1769 0.0652 -0.0672 -0.1982 -0.2808 -0.2815 -0.2071 -0.0753 0.0739 0.1931
0.1245 0.0979 0.0810 0.0705 0.0643 0.0600 0.0737 0.0724 0.0822 0.0955 0.1109 0.1273
534
535
536
8. Concluding remarks
537
The generation of irregular tensegrities through conventional form-finding methods is a
538
challenging research topic. We have proposed a novel numerical approach in order to
539
address this issue through reshaping of existing regular/irregular tensegrities. The
540
reshaping method uses an underdetermined linearized form of the nonlinear
541
equilibrium equations and a geometrical variation approach to form an irregular self-
542
equilibrated tensegrity. The performance of the proposed method has been studied
543
through the examples of free standing and restrained tensegrities. The method promises
544
to generate very accurate results within an iterative computational framework with a
545
remarkable efficiency and very good convergence rate. Since the geometrical variation
32
546
can be applied either in a random or completely guided manner, both random irregular
547
configurations and tensegrities with some specific geometrical properties can be
548
achieved, making the reshaping method very attractive for designing real world
549
tensegrities with special characteristics for particular applications. Our method does not
550
check the stability of a tensegrity during the reshaping process, however, since we work
551
based on small geometric variations, the stability of a tensegrity is normally preserved
552
during reshaping.
553
554
References Ali, N.B.H., Smith, I.F.C., 2010. Dynamic behavior and vibration control of a tensegrity structure. Int. J. Solids Struct. 47, 1285–1296. Calladine, C.R., Pellegrino, S., 1991. First-order infinitesimal mechanisms. Int. J. Solids Struct. 27, 505-515. Chen, Y., Feng, J., Wu, Y., 2012 a. Prestress stability of pin-jointed assemblies using ant colony systems. Mech. Res. Commun. 41, 30–36. Chen, Y., Feng, J., Wu, Y., 2012 b. Novel Form-Finding of Tensegrity Structures Using Ant Colony Systems. ASME J. Mech. Robot. 4 (3), art. no. 031001.1-7. Connelly, R., 1982. Rigidity and energy. Invent. Math. 66, 11–33.
Connelly, R., Whiteley, W., 1996. Second-order rigidity and prestress stability for tensegrity frameworks. SIAM J. Discrete Math. 9, 453–491. Estrada, G.G., Bungartz, H.J., Mohrdieck, C., 2006. Numerical form-finding of tensegrity structures. Int. J. Solids Struct. 43, 6855–6868.
33
Fraternali, F., Senatore, L., Daraio, C., 2012. Solitary waves on tensegrity lattices. J. Mech. Phys. Solids 60, 1137–1144.
Golub, G.H., Van Loan, C.F., 1996. Matrix Computations, third ed. Johns Hopkins University Press, Baltimore.
Kanno, Y., 2012. Topology optimization of tensegrity structures under self-weight loads. J. Oper. Res. Soc. Jpn. 55, 125–145.
Kanno, Y., 2013a. Exploring new tensegrity structures via mixed integer programming. Struct. Multidisc. Optim. 48, 95-114.
Kanno, Y., 2013b. Topology optimization of tensegrity structures under compliance constraint: a mixed integer linear programming approach. Optim. Eng. 14, 61-96.
Koohestani, K., 2012. Form-finding of tensegrity structures via genetic algorithm. Int. J. Solids Struct. 49, 739-747.
Koohestani, K., 2013. A computational framework for the form-finding and design of tensegrity structures. Mech. Res. Commun. 54, 41–49.
Koohestani, K., Guest, S.D., 2013. A new approach to the analytical and numerical formfinding of tensegrity structures. Int. J. Solids Struct. 50, 2995-3007.
Li, Y., Feng, X.Q., Cao, Y.P., Gao, H., 2010. A Monte Carlo form-finding method for large scale regular and irregular tensegrity structures. Int. J. Solids Struct. 47, 1888–1898.
Luo,Y., Xu, X., Lele,T., Kumar, S., Ingber, D.E., 2008. A multi-modular tensegrity model of an actin stress fiber. J. Biomech. 41, 2379–2387.
Masic, M., Skelton, R.E., Gill, P.E., 2005. Algebraic tensegrity form-finding. Int. J. Solids Struct. 42, 4833–4858.
Micheletti, A., Williams, W.O., 2007a. A Marching Procedure for Form-finding for
Tensegrity Structures. J. Mech. Mater. Struct. 2, 857-882.
34
Micheletti, A., Williams, W.O., 2007b. Shape-change of Tensegrity Systems by Controlling
Edge-Lengths, IASS, December 3-6, Venice, Italy.
Moored, K.W., Kemp, T.H., Houle, N.E., Bart-Smith, H., 2011. Analytical predictions, optimization, and design of a tensegrity-based artificial pectoral fin. Int. J. Solids Struct. 48, 3142–3159.
Motro, R., 1984. Forms and forces in tensegrity systems. In: Nooshin, H. (Ed.), Proceedings of Third International Conference on Space Structures. Elsevier, Amsterdam, pp. 180–185.
Motro,R. 2003. Tensegrity: structural systems for the future. London, UK: ButterworthHeinemann.
Pagitz, M., Mirats Tur, J.M., 2009. Finite element based form-finding algorithm for tensegrity structures. Int. J. Solids Struct. 46, 3235–3240.
Paul, C., Lipson, H., Cuevas, F.J.V., 2005. Evolutionary form-finding of tensegrity structures. In: Proceedings of the 2005 Conference on Genetic and Evolutionary Computation. ACM, Washington.
Pellegrino, S., 1986. Mechanics of kinematically indeterminate structures, Ph.D dissertation, University of Cambridge, UK.
Pellegrino, S., Calladine, C.R., 1986. Matrix analysis of statically and kinematically indeterminate frameworks. Int. J. Solids Struct. 22, 409-428.
Rhode-Barbarigos, L., Ali, N.B.H., Motro, R., Smith, I.F.C., 2010. Designing tensegrity modules for pedestrian bridges. Eng. Struct. 32, 1158–1167.
Rieffel, J., Cuevas, F.V., Lipson, H., 2009. Automated discovery and optimization of large irregular tensegrity structures. Comput. Struct. 87, 368–379.
Schek, H.J., 1974. The force density method for form finding and computation of general networks. Comput. Methods Appl. Mech. Engrg. 3, 115–134.
35
Stamenovic, D., Ingber, D.E., 2009. Tensegrity-guided self assembly: from molecules to living cells. Soft Matter 5, 1137–1145.
Stewart, G.W., 1973. Introduction to matrix computations. Academic Press, New York.
Sultan, C., 2009. Tensegrity: 60 years of art, science, and engineering. In Advances in applied mechanics, vol. 43 (eds H. Aref & E. van der Giessen), pp. 69–145. San Diego, CA: Elsevier Academic Press Inc.
Sultan, C., 2014. Tensegrity deployment using infinitesimal mechanisms. Int. J. Solids
Struct. 51, 3653-3668.
Sultan, C., Corless, M., Skelton, R.E., 1999. Reduced prestressability conditions for tensegrity structures. In: Proceedings of 40th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 12–15 April 1999, St Louis, MO, AIAA.
Sultan, C., Corless, M., Skelton, R.E., 2002. Symmetrical reconfiguration of tensegrity
structures. Int. J. Solids Struct. 39, 2215-2234.
Tibert, A.G., Pellegrino, S., 2002. Deployable tensegrity reflectors for small satellites. J.Spacecr. Rockets 39, 701–709.
Tran, H.C., Lee, J., 2010. Advanced form-finding of tensegrity structures. Comput. Struct. 88, 237–246.
Tran, H.C., Lee, J., 2013. Form-finding of tensegrity structures using double singular value decomposition. Eng. Comput. 29, 71-86.
Xu, X., Luo, Y., 2010. Form-finding of nonregular tensegrities using genetic algorithm. Mech. Res. Commun. 37, 85–91.
Zhang L., Maurin, B., Motro, R., 2006. Form-finding of nonregular tensegrity systems. ASCE J. Struct. Eng. 132, 1435–1440.
36
Zhang, J.Y., Ohsaki, M., 2006. Adaptive force density method for form-finding problem of tensegrity structures. Int. J. Solids Struct. 43, 5658–5673. Zhang, J.Y., Ohsaki, M., 2007. Stability conditions for tensegrity structures. Int. J. Solids Struct. 44, 3875–3886. Zhang, J.Y., Ohsaki, M., 2012. Self-equilibrium and stability of regular truncated tetrahedral tensegrity structures. J. Mech. Phys. Solids. 60, 1757–1770. Zhang, L.Y., Li, Y., Cao, Y.P., Feng, X.Q., 2014. Stiffness matrix based form-finding method of tensegrity structures. Eng. Struct. 58, 36-48. 555
556
557
558
559
560
561
562
563
564
565
566
567
37
568
569
Figures Captions
570
571
Fig. 1 a) initial self-stressed configuration (free standing); b) configuration after
572
imposing nodal geometrical variations (free standing); c) initial self-stressed
573
configuration (restrained); and d) configurations after imposing nodal geometrical
574
variations (restrained)
575
576
577
Fig. 2 a) initial truncated tetrahedron tensegrity; and b) new form after imposing nodal
578
geometrical variations
579
580
581
Fig. 3 a) an initial restrained 12-plex cylindrical tensegrity; and b) new form after
582
imposing nodal geometrical variations to top free nodes
583
38
584
Highlights
585
We propose a novel iterative approach for the reshaping of tensegrities
586
Reshaping is performed using a geometrical variation approach
587
An underdetermined linearized form of nonlinear equilibrium equations is used
588
The method generates very accurate results with a remarkable convergence rate
589
39