You are on page 1of 15

Engineering Structures 192 (2019) 166–180

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

A novel shape optimization approach for strained gridshells: Design and T


construction of a simply supported gridshell

Jef Romboutsa,c, , Geert Lombaertb, Lars De Laetc, Mattias Schevenelsa
a
KU Leuven, Faculty of Engineering Science, Department of Architecture, Kasteelpark Arenberg 1 – Box 2431, 3001 Leuven, Belgium
b
KU Leuven, Faculty of Engineering Science, Department of Civil Engineering, Belgium
c
Vrije Universiteit Brussel, Faculty of Engineering, Department of Architectural Engineering, Belgium

A R T I C LE I N FO A B S T R A C T

Keywords: Strained gridshells are reticulated shell structures that are erected from a flat grid of initially straight laths. The
Strained gridshells structural efficiency of a gridshell is determined by its shape, which is traditionally designed using form-finding
Active bending techniques. However, these techniques are primarily used to generate structures in pure tension or compression
Shape optimization for a single load case only. In cases where classic form-finding techniques are not applicable, such as for can-
Form-finding
tilevering gridshells or simply supported gridshells, numerical optimization can be used to find a suitable shape.
Co-rotational beams
In this paper, an optimization procedure is proposed that optimizes the shape of a strained gridshell for a given
grid. The forces applied to erect the gridshell are chosen as the design variables. These erection forces are
optimized to minimize the so-called end-compliance, which is defined as the inner product of the external loads
and the resulting displacements. The method of moving asymptotes is adopted to solve the optimization problem
and implicit dynamic relaxation is used to solve the nonlinear equilibrium equations. Geometric nonlinearity is
taken into account by using co-rotational beam elements to model the gridshell laths. To validate the proposed
approach, a 6 × 6 m2 prototype was built. The results show that this approach allows the structure to be opti-
mized considering multiple load cases, while accounting for practical building constraints, and potential de-
signer constraints.

1. Introduction glass fibre reinforced polymers are gaining popularity because of their
higher Young’s modulus and higher limit strain [1].
Gridshells are shell structures built from discrete elements such as The structural concept of strained gridshells was developed by Frei
wood, composite, or steel bars. Due to the three-dimensional character Otto, who applied it in 1975 for the construction of the Mannheim
of their shape, gridshells are potentially very efficient: when properly Multihalle [2]. This pioneering structure was designed using a detailed
designed, light-weight gridshells can cover very large spans. Based on hanging model. Later, Buro Happold revived the concept of strained
the construction technique used, a distinction can be made between gridshells with the construction of the Japanese Pavilion at the Hann-
unstrained gridshells and strained gridshells. Unstrained gridshells are over Expo in 2000. Shortly after, in 2002, the same engineers used their
composed of bespoke prefabricated nodes and members, usually made experience from the Japanese pavilion for the design and construction
of steel. Curvature is introduced through the nodes or by using already of the Downland gridshell [3]. The most recent large scale timber
curved members. Strained gridshells, on the other hand, are constructed gridshell is the Savill Garden gridshell, also designed by Buro Happold,
from an initially flat grid of continuous laths. The laths are bent on site and constructed in 2006 [4]. The shape of the latter three gridshells was
to generate the gridshell’s curvature. Pinned connections between all determined by fitting an equidistant mesh to a mathematically de-
laths in the two grid directions make sure the grid has no in-plane shear scribed surface. More recently, research on strained gridshells has re-
stiffness, which allows the grid to be deformed into complex curved sulted in the construction of several prototypes. The Ephemerical ca-
shapes. As a consequence, in contrast to unstrained gridshells, strained thedral in Créteil and the Soliday festival gridshell were designed and
gridshells do not require the production of specially customized beams constructed by the Thinkshell research group [5,6] to investigate the
or joints. For strained gridshells, wood is often used because of its ex- applicability of glass fibre reinforced polymers. Again, the shape for
cellent bending characteristics. However, composite materials such as these gridshells was determined geometrically. Gridshells constructed


Corresponding author at: KU Leuven, Faculty of Engineering Science, Department of Architecture, Kasteelpark Arenberg 1 – Box 2431, 3001 Leuven, Belgium.
E-mail address: jef.rombouts@kuleuven.be (J. Rombouts).

https://doi.org/10.1016/j.engstruct.2019.04.101
Received 20 November 2018; Received in revised form 21 March 2019; Accepted 30 April 2019
Available online 09 May 2019
0141-0296/ © 2019 Elsevier Ltd. All rights reserved.
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

by gridshell.it[7–9] and NTNU[10] started from a predefined grid, where the geometric model [25]. Three major approaches exist [26]. The first
the shape of the shell followed the natural equilibrium shape obtained approach is to use the finite element model as the design model, re-
after fixing the supports. sulting in a large optimization problem and mesh irregularity issues.
Like for all shell structures, the structural efficiency of strained The second approach is to separate the two models, using a linear
gridshells is primarily determined by their shape. Gridshells are tradi- combination of basis functions to model the geometry, and using the
tionally designed using form-finding techniques. Originally, hanging parameters of the linear combination as design variables of the opti-
models were used for this purpose. Nowadays, numerical form-finding mization model. This latter approach leads to more smooth designs, but
techniques simulating these physical models are preferred. Lienhard requires a translation between the two models and limits the design
[11] distinguishes between three approaches to design bending-active freedom of the optimizer. More recently, a third approach was devel-
structures, including strained gridshells. In a behavior based approach, oped that uses isogeometric structural analysis to preserve the same
the actual bending of physical elements leads the design. On the other design and analysis model [26].
hand, in a geometry based approach, the shape is determined without In this paper, an optimization procedure is proposed that optimizes
taking the bending behavior of the members into account. The shape is the shape of a strained gridshell for a given grid of straight laths. The
either determined geometrically, or by using form-finding techniques gridshell is modeled using geometrically nonlinear co-rotational beam
that don’t take bending into account. Finally, the integral approach elements, which have 6 degrees of freedom per node. The erection
determines the shape by simulating the bending behavior during form- forces acting on the initially flat grid are chosen as the design variables,
finding. To integrate the bending behavior in the form-finding process, which automatically leads to smooth designs. These erection forces are
material properties have to be taken into account and stiffness matrix optimized to minimize the so-called end-compliance, which is defined
methods such as the Newton-Raphson solver or dynamic equilibrium as the inner product of the external loads and the resulting displace-
methods that take bending into account must be used. A comparison of ments. The design sensitivities are derived analytically to enable the use
these and other form-finding methods is given by Veenendaal and Block of gradient-based optimization. Finally, the proposed optimization ap-
[12]. Dynamic relaxation is often preferred because of its im- proach is applied to design a simply supported gridshell prototype.
plementation simplicity and supposed superior convergence behavior The remainder of this paper is organized as follows. First, Section 2
when dealing with large displacements [7,13]. However, recent re- explains the finite element model. Next, Section 3 describes the opti-
search shows that the Newton-Raphson method can be considered as a mization problem. Subsequently, Section 4 derives the required sensi-
special case of dynamic relaxation [14]. More specifically, if the ori- tivities, which are validated using finite differences in Section 5. Section
ginally explicit dynamic relaxation solver is made implicit by selecting 6 describes the design and construction of a prototype, and finally,
a mass matrix proportional to the stiffness matrix, the two solvers be- Section 7 formulates a conclusion.
come equivalent. Moreover, it has been shown that the resulting im-
plicit dynamic relaxation is faster than classic explicit dynamic re- 2. Finite element model
laxation for bending-active beam structures such as gridshells [15].
Labonnote et al. [16] showed that an accurate simulation of the Like several authors [8,10,27,28], we follow the integrated ap-
structural behavior is crucial for a good prediction of the final shape of proach to design a gridshell; the bending process of a predetermined
a strained gridshell. In an effort to improve the accuracy of dynamic flat grid is simulated. The shape of the gridshell is determined by the
relaxation in dealing with bending-active structures, researchers have external forces applied in the erection process, using for example crane
worked on the development of bending elements for the existing dy- cables or struts. After the gridshell is pushed or pulled in the desired
namic relaxation framework. Barnes [17] and Van Mele et al. [18] shape, the boundary supports are fixed and an additional layer of braces
proposed a beam model with three degrees of freedom per node, Du is added to stabilize the deformed grid. Subsequently, the temporary
Peloux et al. [19,20] and D’Amico et al. [21] presented a model with 4 erection forces are removed and the gridshell deforms slightly under
degrees of freedom per node, and Wakefield [22], D’Amico et al., [7], these changed boundary conditions. Finally, the structure is subjected
Bessini et al. [23] and Rombouts et al. [15] use beam elements with 6 to external loads such as dead loads from cladding, wind loads and
degrees of freedom per node. The latter are generally more accurate service loads (Fig. 9).
[21], and better suited to deal with torsion and anisotropic cross sec- The simulation of the construction process is performed in the same
tions [7]. 4 steps. First, the erection of the flat grid is simulated in a form-finding
Most gridshells carry their loads primarily in compression. Although stage. Second, braces are added and supports are fixed. Third, a re-
this is an efficient way of carrying loads, it requires the supports to laxation stage simulates the removal of the erection forces. Fourth, all
withstand the resulting thrust forces, which is not always possible, relevant load combinations are applied in an analysis stage. The de-
depending on the given boundary conditions. Because traditional form- formations at the end of the form-finding stage are determined by the
finding techniques based on inverse hanging models are only efficient following equilibrium equation:
for generating structures in pure tension or compression, they are
r f = pf − f f = 0 (1)
unusable for the design of cantilevering or simply supported1 gridshells,
where membrane theory for shells is no longer applicable. Moreover, where r f are the residual forces at the end of the form-finding stage, f f
form-finding techniques define the shape for only one load case, and are the corresponding internal forces, and p f are the applied form-
constraints, such as a limit on the bending moments in the laths, cannot finding forces. Similarly, the relaxation stage equilibrium is expressed
be accounted for in the calculation. Therefore, a shape generating as:
method should be developed that overcomes these limitations. This has
r r = pr − f r = 0 (2)
motivated Ding et al. [24] to develop a shape optimization approach for
unstrained gridshells. However, such a method does not yet exist for rr
where are the residual forces at the end of the relaxation stage, are fr
strained gridshells. the corresponding internal forces, and p r are the applied loads, which
A common challenge related to shape optimization is the choice of are equal to zero in the relaxation stage. Finally, Eq. (3) describes
equilibrium in the analysis stage:
r a = pa − f a = 0 (3)
1
the term ‘simply supported’ is used in this paper to describe a situation
where the supports are incapable of carrying the thrust forces exerted by a where raare the residual forces at the end of the analysis stage, are fa
gridshell, for example when a gridshell is constructed on columns or historic the corresponding internal forces, and pa is the external load vector.
walls that cannot withstand horizontal forces. The nonlinear equilibrium equations in the form-finding, relaxation,

167
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

elements is given next in Section 2.1.

2.1. Co-rotational beam elements

Co-rotational beam elements are specifically developed to handle


large displacements [30]. They allow arbitrarily large translations and
rotations, as long as the strains remain small. This is achieved by de-
coupling rigid body motion and local beam deformations. In this sec-
tion, the basis of the co-rotational beam element as described by Cris-
field [29] is reviewed. A more extensive explanation can be found in
[29,31]. The considered beam element is based on the Euler–Bernoulli
beam theory. Furthermore, the material is assumed to be homogeneous,
isotropic, and linearly elastic (Fig. 1).
To distinguish between rigid body motion and local beam de-
formations, a local co-rotated reference configuration is fitted to the
actual deformed beam. Local deformations are traced by comparing the
actual deformed beam to the co-rotated reference beam. The co-rotated
reference beam is represented by a set of local element basis vectors
Te = [e x e y ez], where e x lies in the direction of the vector connecting
the beam ends. This set of element basis vectors is constructed from the
known current beam end orientations, which are in turn represented by
the nodal basis vectors Ta = [ax a y a z] and Tb = [bx b y bz ] (Fig. 2). Note
that ax and bx point in the direction of the beam’s centerline, while a y
and a z , and b y and bz point in the direction of the cross section’s
principal axes. For the actual procedure to construct Te , the reader is
referred to the original paper [29]. For an undeformed beam, both sets
of nodal basis vectors are equal to the element basis vectors.
In each step of the nonlinear solver, the nodal basis vectors are
updated using Rodrigues’ rotation formula [31,32]. For beam end a we
get:

T ta = R (Δλ a) T ta− Δt (4)


where T ta
collects the current nodal basis vectors, T ta− Δt
collects the
nodal basis vectors of the previous iteration, and R(λ ) is an orthogonal
rotation matrix describing the rotation around a given pseudo-vector
λ = {α β γ }T over a given angle λ = ‖λ‖, where the double vertical lines
denote the Euclidean norm:
sin(λ ) 1 − cos(λ )
R (λ ) = I + S (λ ) + S (λ ) S (λ )
λ λ2 (5)

where I is a three by three identity matrix, and S is a skew-symmetric


matrix defined as:

⎡ 0 −γ β ⎤
S(λ ) = ⎢ γ 0 − α⎥
⎢− β α 0 ⎥
⎣ ⎦ (6)

]T
In Eq. (4), the rotation vector Δλ a = [Δαa Δβa Δγa is the rotation vector
describing the change in orientation of beam end a between this

Fig. 1. The design and construction of strained gridshells goes through four
stages.

and analysis stage (Eqs. ()()()(1)–(3)) are solved using implicit dynamic
relaxation. For more details on implicit dynamic relaxation, the reader
is referred to [15]. Co-rotational beam elements [29] are used to model
Fig. 2. Definition of the nodal basis vectors (ax , a y, a z) and (bx , b y , bz ) as well as
the grid laths. A short description of the used co-rotational beam
the element basis vectors (e x , e y , ez) of a deformed beam element.

168
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

iteration and the previous one. In our case, these angle changes are
calculated by the implicit dynamic relaxation solver. The updating
procedure for the nodal basis vectors of beam end b is analogous.
The deformation of the beam is described by seven independent
local beam deformations2, recorded at the beam ends. An axial de-
formation δ gives rise to a normal force N, and six rotational de-
formations θ cause six corresponding bending moments M. These ro-
tational deformations describe the beam’s twist as well as bending
around the cross section’s two principal axes at the beam ends.
The local rotational deformations θ are expressed in terms of the
nodal and element basis vectors:

2sin(θx , a) = −(a z)T e y + (ez)T a y 2sin(θx , b) = −(bz )T e y + (ez)T b y 2sin(θy, a)


= −(a z)T e x + (ez)T ax 2sin(θy, b) = −(bz )T e x + (ez)T bx 2sin(θz, a) Fig. 3. Orientation of the local beam deformations (δ θx ,a θy,a θz,a θx ,b θy,b θz,b)
= −(a y)T e x + (e y)T ax 2sin(θz, b) = −(b y)T e x + (e y)T bx (7) and corresponding internal forces (N Mx ,a My,a Mz,a Mx ,b My,b Mz,b) .

where θx , θy , and θz are the deformation angles of the beam ends


[29]. The full internal force vector f is assembled from the element
around the local x-, y-, and z-axis respectively (Fig. 3).
forces f e , with e = 1… m , where m is the number of elements.
The axial deformation δ is simply the difference between the ele-
The element tangent stiffness matrix is obtained from differentiation
ment length at the current time lt and the initial element length l 0 :
of the internal forces f e :
δ = lt − l 0 (8)
∂f e ∂f e ∂f¯ e ∂f e ∂T e
with the length defined as the Euclidian distance between beam ends a Ke = = e e + = T eK eT eT + Keg = K ee + Keg
∂x e ¯
∂f ∂x ∂T e ∂x e (16)
and b:

l= (x a − xb)2 + (ya − yb )2 + (z a − z b)2 (9) The resulting stiffness matrix consists of an elastic part K ee and a geo-
metric part Keg . The derivation of the geometric stiffness matrix Keg can
where x a is the x-coordinate of node a, xb is the x-coordinate of node b, be found in the original paper [29]. The element stiffness matrix Ke for
and so on. each element e = 1… m is used to assemble the full tangential stiffness
The local internal forces are expressed in terms of the local de- matrix K .
formations. For an element e:
f¯ e = K
¯ eu¯ e (10)
3. Optimization problem
e
where f̄ is the vector collecting all local element forces (Fig. 3):

f̄ e = [N Mx ,a My,a Mz, a Mx , b My,b Mz,b]T The aim of the proposed shape optimization approach is to mini-
(11)
mize the so-called end-compliance C of a strained gridshell under
and ū e is the vector collecting all local deformations (Fig. 3): multiple load cases, given buildability and design-related constraints.
The end-compliance C is defined as the inner product of the external
ū e = [δ θx ,a θy,a θz,a θx ,b θy,b θz,b]T (12)
loads and the resulting displacements measured in the analysis stage.
and K e is the local element stiffness matrix: The shape of the gridshell is determined by the erection forces. These
erection forces are modeled in the form-finding stage as so-called form-
⎡ EA 0 0 0 0 0 0 ⎤ finding loads, and are taken as the design variables in the optimization
⎢ 0 GJ 0 0 − GJ 0 0 ⎥
process. When only one load case is considered, the unconstrained
⎢ 0 0 4EIy 0 0 2EIy 0 ⎥
1⎢ optimization problem is formulated as follows:
Ke = 0 ⎢ 0 0 0 4EIz 0 0 2EIz ⎥

l ⎢ 0 − GJ 0 0 GJ 0 0 ⎥ min C = paT (x a − xr)
⎢ 0 0 2EIy 0 0 4EIy 0 ⎥ f
⎢ ⎥
pv (17)
⎣ 0 0 0 2EIz 0 0 4EIz ⎦ (13)
e where x a is the position vector after the analysis stage, xr is the position
The global element forces f e are related to the local forces f̄ as follows:
vector after the relaxation stage, pa are the externally applied loads,
f e = T ef̄ e (14) and p fv is the selection of form-finding forces taken as the design vari-
ables. This selection is made using a selection matrix L fv , which selects
fe ]T
where = [fx ,a f y,a fz,a m x ,a m y,a mz,a fx ,b f y,b fz,b m x ,b m y,b mz,b is the
vector containing forces and moments in the global x-, y-, and z-axis for the variable degrees of freedom from all degrees of freedom of the finite
beam end “a” and beam end “b”, and T e is a transformation matrix element model in the form-finding stage:
defined as
p fv = L fv p f (18)
∂u¯ e T
Te = ⎛ e ⎞
⎝ ∂x ⎠ (15) Three constraint equations are considered in this paper. First, box
constraints on the design variables limit the erection forces for the
where x e = [x a ya z a αa βa γa xb yb zb αb βb γb]T is a vector collecting
construction of the gridshell. Second, geometrical constraints on the
translational and rotational coordinates of beam end “a” and beam end
node positions during the erection stage give the designer control over
“b” in the global x-, y-, and z-direction. For the actual calculation of the
the final shape. Third, a constraint on the internal bending moments
transformation matrix T e , the reader is referred to the original paper
makes sure the members do not exceed their maximum bending capa-
city. For simplicity, this constraint is applied only to the form-finding
2
Referred to as ‘strains’ in the original paper. However, because the con- stage, as it is expected that most of the bending in the laths is due to the
cerning quantities can not be regarded as strains in the strict mechanical sense, erection of the gridshell. The constrained optimization problem is for-
the term ‘deformations’ is used here instead. mulated as:

169
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

Fig. 4. Design stages of a braced bending-active arch.

Table 1
Comparison of the derivatives of the end-compliance C of the braced arch shown in Fig. 4. Analytical derivatives are compared to the forward finite difference
derivatives for a decreasing increment ∊.
K ∂C / ∂p fv [m]

Analytical Analytical −0.3839 −0.8803 −1.6586 −2.3481 −1.6586 −0.8803 −0.3839


Numerical −0.3849 −0.8827 −1.6630 −2.3537 −1.6630 −0.8827 −0.3849

∊ [N] ∂C / ∂p fv [m]

Finite differences 10−1 −0.3833 −0.8760 −1.6436 −2.3182 −1.6436 −0.8760 −0.3833
10−2 −0.3847 −0.8820 −1.6611 −2.3502 −1.6611 −0.8820 −0.3847
10−3 −0.3849 −0.8827 −1.6629 −2.3534 −1.6629 −0.8827 −0.3849
10−4 −0.3849 −0.8827 −1.6631 −2.3537 −1.6631 −0.8827 −0.3849
10−5 −0.3849 −0.8827 −1.6631 −2.3538 −1.6631 −0.8827 −0.3849
10−6 −0.3849 −0.8827 −1.6631 −2.3538 −1.6631 −0.8827 −0.3849
10−7 −0.3849 −0.8827 −1.6631 −2.3538 −1.6631 −0.8827 −0.3849
10−8 −0.3848 −0.8827 −1.6631 −2.3537 −1.6630 −0.8827 −0.3848
10−9 −0.3845 −0.8817 −1.6624 −2.3538 −1.6624 −0.8826 −0.3844
10−10 −0.3819 −0.8815 −1.6576 −2.3493 −1.6631 −0.8771 −0.3831

Table 2
r
Comparison of the derivatives of the y-coordinate x y,5 of the middle node of the braced arch shown in Fig. 4. Analytical derivatives are compared to the forward finite
difference derivatives for a decreasing increment .∊

K r
∂x5,y / ∂p fv [m/N]

Analytical Analytical 0.0052 0.0112 0.0175 0.0207 0.0175 0.0112 0.0052


Numerical 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052

∊ [N] r
∂x5,y / ∂p fv [m/N]

Finite differences 10−1 0.0052 0.0112 0.0175 0.0207 0.0175 0.0112 0.0052
10−2 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−3 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−4 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−5 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−6 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−7 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−8 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−9 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052
10−10 0.0052 0.0112 0.0176 0.0208 0.0176 0.0112 0.0052

2 2
min C = paT (x a − xr)s. t. p fv,min ⩽ p fv ⩽ p fv,max x rc,min ⩽ x rc ⩽ f
mf ⩽ m max (20)
f
pv

x rc,max f
m min f
⩽ mf ⩽ m max (19) where the superscript 2 denotes the square of each element of the
vector. In this paper, only bending moments at the beam ends are
where p fv,min and p fv,max are respectively the lower and upper bounds for considered, as these are immediately available from the finite element
the design variables, x rmin and x rmax are respectively the vectors col- model.
lecting the lower and upper bounds for x rc , which is the vector collecting When multiple load cases are taken into account, a nonlinear finite
the positional values for the constrained degrees of freedom, and mf is element analysis has to be performed for each load case. Moreover, the
the vector with internal bending moments, with m min f f
and m max as the optimization problem is reformulated as a “min–max” problem. The
corresponding minimum and maximum values. To reduce the number objective function becomes:
of constraint equations, the bending moment constraint can be re-
f f
formulated as follows, assuming m min = −m max :

170
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

Table 3
f
Comparison of the derivatives of the bending moment M4,b at the right beam end of the fourth beam element (from left to right) of the braced arch shown in Fig. 4.
Analytical derivatives are compared to the forward finite difference derivatives for a decreasing increment ∊.
K f / ∂p f [m]
∂M4,b v

Analytical Analytical 0.0941 0.2288 0.4441 0.2640 0.1366 0.0627 0.0239


Numerical 0.0943 0.2293 0.4449 0.2649 0.1372 0.0630 0.0240

∊ [N] f / ∂p f [m]
∂M4,b v

Finite differences 10−1 0.0942 0.2291 0.4451 0.2646 0.1369 0.0629 0.0240
10−2 0.0943 0.2293 0.4449 0.2649 0.1372 0.0630 0.0240
10−3 0.0943 0.2293 0.4449 0.2649 0.1372 0.0631 0.0240
10−4 0.0943 0.2293 0.4449 0.2649 0.1372 0.0631 0.0240
10−5 0.0943 0.2293 0.4449 0.2649 0.1372 0.0631 0.0240
10−6 0.0943 0.2293 0.4449 0.2649 0.1372 0.0631 0.0240
10−7 0.0943 0.2293 0.4449 0.2649 0.1372 0.0631 0.0240
10−8 0.0943 0.2293 0.4449 0.2649 0.1372 0.0631 0.0240
10−9 0.0943 0.2294 0.4450 0.2650 0.1373 0.0631 0.0240
10−10 0.0942 0.2293 0.4449 0.2656 0.1379 0.0640 0.0248

directly because the supported nodes should be free to move in the


formfinding stage. If the designer wants to control the final position of
the supports, he has to define position constraints for the supported
nodes. Third, also the area covered by the gridshell depends on the final
shape found by the optimizer. Again, position constraints can give the
designer some control. However, it might be necessary to change the
grid size, layout and/or support locations iteratively to fulfill specific
requirements.

4. Sensitivity analysis

This section describes the sensitivity analysis required for gradient-


based optimization.
Fig. 5. Initial grid and boundary conditions of a simply supported gridshell.
4.1. Direct differentiation of the end-compliance

The objective function is determined from the analysis stage,


whereas the design variables act in the form-finding stage. Therefore,
the objective function only depends indirectly on the design variables.
The following expression shows how the end-compliance C is a function
of the form-finding forces:

C (̂ p fv ) = C (x a (xf (p f (p fv )), xr (xf (p f (p fv )))) (22)

where xf , xr , xa
are the position vectors at the end of the form-finding,
relaxation, and analysis stage respectively, p f are all form-finding
forces, and p fv are the variable form-finding forces. Note that the po-
sitions x a after the analysis stage do not depend on the positions xr at
the end of the relaxation stage, because equilibrium is independent of
previous loading conditions, as long as the finite element model does
not change. Differentiation of the end-compliance with respect to the
variable form-finding forces gives:
Fig. 6. Top view of the initial grid, showing the form-finding forces of the
dC ∂C d x a ∂C d xr
optimized gridshell. = + r
f ∂x a f ∂x d p fv
dpv dpv (23)

⎛⎡ ⎤⎞ where, from differentiation of Eq. (17), ∂C / ∂x a and ∂C / ∂xr are given by:
min max(C) = max ⎜ ⎢C1 C2… Ci… Cn⎥ ⎟ with Ci = pia T (x ia − xr)
f
pv
⎝⎣ ⎦⎠ (21) ∂C
= paT
∂x a (24)
where pia is the external load vector of load case i, and x ia is position
∂C
vector at the end of the corresponding analysis stage. = −paT
∂xr (25)
Some limitations and practical considerations should be mentioned.
First, the computation time required for the nonlinear finite element and d x a / d p fv and d xr / d p fv are elaborated using the chain rule:
analysis in each iteration and every load case imposes practical limits
d xa ∂x a d xf
on the mesh size and the number of load cases which can be considered. =
Second, the designer cannot choose the final location of the supports
d p fv ∂xf d p fv (26)

171
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

Fig. 8. Convergence history of the end-compliance for load case 1 and 2.

Fig. 9. Maximum occurring bending moment around the laths’ weak axis,
compared to the maximum allowed bending moment for every iteration.

d xf ∂xf d p f
=
d p fv ∂p f d p fv (28)
where ∂xf / ∂p f is found by differentiation of the equilibrium equation
governing the form-finding stage (Section 4.1.2), and d p f / d p fv is found
by differentiation of Eq. (18):
dpf T
= L fv
d p fv (29)

4.1.1. Differentiation of the equilibrium governing the analysis stage


In Eq. (26), d x a / d xf is found by differentiation of equilibrium Eq.
(3):
dr a d pa df a
= − =0
dx f d xf d xf (30)
where d pa / d xf
is zero, as we assume that the position vector after form-
finding does not influence the magnitude of the loads in the analysis
stage, nor the degrees of freedom on which the loads act. On the other
hand, the dependency of the internal forces f a at the end of the analysis
stage on the position vector xf at the end of the form-finding stage is
Fig. 7. All design stages of the optimized gridshell. given by:

f ̂ (xf ) = f a (xf , x a (xf ))


a
(31)
d xr ∂xr d xf
=
d p fv ∂xf d p fv (27) Differentiation of the internal forces fa results in:
df a ∂f a ∂f a ∂x a
where ∂x a / ∂xf and ∂xr / ∂xf are found by differentiation of the equili- = +
d xf ∂xf ∂x a ∂xf (32)
brium equations governing the corresponding design stage. This is done
next for the analysis stage in Section 4.1.1. The differentiation of the Next, combining Eqs. (30) and (32) gives:
equilibrium equation of the relaxation stage proceeds along the same ∂f a ∂x a ∂f a
=− f
lines. Applying the chain rule again to d xf / d p fv gives: ∂x a ∂xf ∂x (33)
where ∂f a/ ∂x a equals the tangent stiffness matrix at the end of the

172
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

Fig. 10. Comparison of displacements (full scale). Color indicates vertical displacement, ranging up to 0.9 m (dark red). (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.)

analysis stage. Eq. (33) is rewritten as: ∂f aF


= KaF
∂x aF (40)
∂x a ∂f a
Ka =− f
∂xf ∂x (34) where KaF is the tangent stiffness matrix for the deformed geometry at
where ∂f a/ ∂xf
is still unknown. The geometry at the end of the form- the end of the analysis stage, assembled for all degrees of freedom of the
finding stage influences the internal forces f a at the end of the analysis unconstrained finite element model. Further application of the chain
stage through the boundary supports that are added to fix the supported rule gives for d x aF/ d xf :
nodes in their form-finding stage positions. The geometry after form- d x aF ∂x aF d xfa
finding also influences the internal forces f a after the analysis stage =
d xf ∂xfa d xf (41)
through the initial lengths l a0 of the braces, which directly depend on the
nodal positions at the end of the form-finding stage. Therefore, the where:
internal force vector f a at the end of the analysis stage is an indirect ∂x aF T
function of the position vector xf after the form-finding stage: = LFfa
∂xfa (42)

a
(xf ) = f a (f aF (x aF (xfa (xf )), l a0 (x Ff (xf )))) (35) where LFfa
is a selection matrix selecting the removed degrees of
freedom from all degrees of freedom. Moreover,
where f aF
is the internal force vector at the end of the analysis stage for
all degrees of freedom of the unconstrained finite element model, x aF is d xfa
= L ffa
the position vector describing the deformed geometry at the end of the d xf (43)
analysis stage for all degrees of freedom, xfa is the position vector of the
where L ffa selects the removed degrees of freedom from the degrees of
degrees of freedom removed between the form-finding stage and the
freedom in the form-finding stage. The next derivative in Eq. (39),
relaxation stage by fixing the supports, and x Ff is the position vector at
∂f aF/ ∂l a0 , is found according to Eqs. ()()()(44)–(46). Because changes in
the end of the form-finding stage for all degrees of freedom. The deri-
length do not affect the interconnection of the elements, the same as-
vative of the internal forces (d f a/ d xf ) is given by:
sembly procedure can be used for the partial derivatives of the internal
df a ∂f a d f a forces as for the internal forces themselves. Differentiation with respect
= a Ff
dx f ∂f F d x (36) to the initial length l 0j of an element j:
v
where ∂f a/ ∂f aF is given by: d f aF df e
dl 0j
= ∑ Ce
dl 0j
∂f a e=1 (44)
= LFa
∂f aF (37)
where v is the number of elements, Ce
is a rectangular assembly matrix
where LFa is a selection matrix defined such that: for element e, consisting of zeros and ones and f e are the internal forces
of element e. Differentiation of Eq. (14) gives:
f a = LFa f aF (38)
df e dTe ¯ e d f¯ e
Returning to Eq. (36), d f aF/ d xf is further elaborated using the chain j
= j
f + Te j
dl 0 dl 0 dl 0 (45)
rule:
d f aF ∂f a d x aF ∂f a d l a where d T e/ dl 0j is zero. Differentiation of Eq. (10) gives:
f
= Fa f
+ aF 0f
dx ∂x F d x ∂l 0 d x (39) df ¯e ¯e e
dK e
= ¯ e d u¯
u¯ + K
where, by definition: dl 0j dl 0j dl 0j (46)

173
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

Fig. 12. Connections of initial grid.

Fig. 13. Final gridshell.

Fig. 14. Pinned connections between the wooden columns and the ground.

d l a0 ∂l a d x Ff
f
= 0f
dx ∂x F d xf (49)

where ∂l a0/ ∂x Ff
is given below by Eq. (50). The coordinates at the end of
the form-finding stage are used to calculate the length of the braces at
the start of the analysis stage. Differentiation of Eq. (9) gives:

∂l 0j
Fig. 11. All construction stages of the optimized gridshell.
∂x Ff
1
where the derivatives of K̄ e and ū e are found by differentiation of Eqs. = ((x af − xbf )(LFx a − LFx b) + (yaf − ybf )(LFya − LFy b) + (z af − z bf )
l 0j
(13), and (8) and (12):
(LFz a − LFz b)) (50)
¯e
K
¯e
∂K ⎧ j , if e = j where LFxa selects the x-coordinate of beam node a:
j
= l0
∂l 0 ⎨
⎩ 0, otherwise (47) x af = LFx a x Ff (51)

Equivalently, LFya and LFza select the y- and z-coordinate of node a, and
∂u¯ e [−1 0 0 0 0 0 0]T , if e = j LFx b , LFy b and LFzb select the x-, y-, and z-coordinate of node b. Finally:
j
=⎧
∂l 0 ⎨
⎩ 0, otherwise (48)
d x Ff T
= LFf
Application of the chain rule to the term d l a0/ d xf in Eq. (39) gives: d xf (52)

174
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

Fig. 15. Ball joint connections between the wooden columns and the grid laths.

where LFf selects the form-finding degrees of freedom from all degrees
of freedom.
Inserting Eqs. (50) and (52) into (49) gives d l a0/ d xf . From Eqs. (47)
and (48) we can find ∂f aF/ ∂l 0 using Eqs. ()()()(44)–(46). Inserting Eqs.
(42) and (43) into Eq. (41) gives d x aF/ d xf . Subsequently, inserting these
derivatives, together with Eq. (40), into Eq. (39) gives d f aF/ d xf , which
can be used in Eq. (36), together with Eq. (37), to find d f a/ d xf . This
derivative is required in Eq. (34) to find ∂x a / ∂xf .

4.1.2. Differentiation of the equilibrium governing the form-finding stage


In Eq. (28), ∂xf / ∂p f is found by differentiation of the form-finding
equilibrium Eq. (1):
dr f dpf df f
= − =0
dpf dpf dpf (53)
where:
dpf
=I
dpf (54)
where I is an identity matrix. The internal forces ff are related to the
external loads p f as follows:

f ̂ (p f ) = f f (xf (p f ))
f
(55)
Differentiation of the internal forces f f results in:
df f ∂f f ∂xf Fig. 16. Predicted deformation (full scale) of the optimized gridshell under 4
f
= point loads. Vertical displacements, indicated in color, range up to 0.1 m.
dp ∂xf ∂p f (56)
Next, combining Eqs. (53), (54) and (56), and using the fact that ∂f f / ∂xf
combined with Eq. (34) in Eq. (26) to find d x a / d p fv . Next, d xr / d p fv is
equals the tangent stiffness matrix Kf at the end of the form-finding
calculated in the same way from Eq. (27). Finally, d x a / d p fv and d xr / d p fv
stage gives:
are inserted into Eq. (23), together with Eq. (24) and (25) to find
∂xf dC / dp fv . This procedure corresponds to the direct differentiation
Kf =I
∂p f (57) method. Using this method, the following systems of equations have to
be solved:
Inserting Eq. (29) and Eq. (57) into Eq. (28) gives d xf / d p fv , which is

175
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

vector λCf is calculated by solving the following adjoint equation:


T
dC
(Kf )T λCf = ⎛ f ⎞
⎝ dx ⎠ (64)
where the adjoint method is used again to find dC / dxf :
dC ∂f a ∂f r
= (λCa )T f + (λCr )T f
dx f ∂x ∂x (65)
where ∂f a/ ∂xf and ∂f r/ ∂xf are calculated as before (Section 4.1.1), and
λCa and λCr are found by solving the following adjoint equations:
∂C T
(Ka)T λCa = −⎛ a ⎞
Fig. 17. Measuring points for the point load analysis. ⎝ ∂x ⎠ (66)
T
∂C
d xf dpf (Kr)T λCr = −⎛ r ⎞
Kf = ⎝ ∂x ⎠ (67)
d p fv d p fv (58)
where ∂C / ∂x a
and ∂C / ∂xr
are calculated using Eqs. (24) and (25). The
d xa df a systems of Eqs. (64), (66), and (67) that need to be solved in the adjoint
Ka f = method have only one right-hand side. However, the differentiation of
dpv d p fv (59)
each constraint equation requires additional systems of equations to be
d xr df r solved.
Kr f
=
dpv d p fv (60)
4.3. Direct differentiation of the nodal positions
where the number of right-hand sides for each system of equations is
equal to the number of design variables. In the case where the number To calculate the sensitivities of the position constraints, we have to
of design variables is higher than the number of equations in the op- differentiate the selected positions x rc at the end of the relaxation stage
timization problem, the adjoint differentiation method can be used to with respect to the design variables p fv :
reduce the computational effort for the calculation of the design sen- d x rc ∂x rc d xr
sitivities. The application of this method for the differentiation of the =
d p fv ∂xr d p fv (68)
end-compliance is discussed next in Section 4.2.
where:
4.2. Adjoint differentiation of the end-compliance ∂x rc
= Lrc
∂xr (69)
First, consider the following expression for the derivative dC / dp fv :
where Lcf is a selection matrix defined such that:
dC ∂C d xf
f
= x rc = Lrc xr (70)
dpv ∂xf d p fv (61)
The last term in Eq. (68) is given by Eq. (27).
Inserting Eq. (57) gives:

dC ∂C f −1 d p f 4.4. Adjoint differentiation of the nodal positions


f
= K
dpv ∂xf d p fv (62)
If the adjoint method is used, the derivative of the constrained po-
In the adjoint method, Eq. (63) is reformulated by introducing an ad- sitions x rc becomes:
joint vector λCf :
d x rc dpf
= (λ xf )T
dC dpf d p fv d p fv (71)
f
= (λCf )T f
dpv dpv (63)
where d p f / d p fv is calculated as before, using Eq. (29), and the adjoint
where d p f / d p fv is calculated as before, using Eq. (29), and the adjoint vector λ xf is calculated by solving the following adjoint equation:

Table 4
Comparison of vertical displacements predicted and measured at 13 nodes (Fig. 17) for 4 point loads (Fig. 16).
Point Load case 3 Load case 4 Load case 5 Load case 6

Predicted Measured Predicted Measured Predicted Measured Predicted Measured

1 −0.013 −0.037 0.000 −0.014 0.002 −0.005 0.001 0.000


2 −0.012 −0.030 0.018 0.008 0.016 0.021 0.001 0.000
3 −0.004 −0.029 −0.070 −0.091 −0.010 - 0.000 0.000
4 −0.012 −0.020 0.017 0.012 0.018 0.017 0.002 0.000
5 −0.004 −0.032 −0.012 −0.015 −0.025 −0.045 0.001 0.000
6 −0.032 −0.045 0.043 0.046 0.040 0.057 0.001 0.000
7 0.016 −0.014 −0.001 −0.030 −0.005 0.002 −0.001 0.000
8 −0.006 −0.048 −0.089 −0.167 −0.029 −0.037 −0.001 0.006
9 0.029 −0.019 0.012 −0.035 −0.007 0.000 −0.002 0.000
10 −0.032 −0.026 0.039 0.056 0.045 0.053 0.006 0.000
11 0.016 −0.004 −0.002 0.006 −0.003 −0.005 −0.069 −0.058
12 −0.006 −0.055 −0.034 −0.039 −0.096 −0.148 0.001 0.005
13 0.029 −0.041 0.001 0.004 0.008 −0.023 −0.002 0.000

176
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

r T (76) and (77).


dx
(Kf )T λ xf = ⎛ fc ⎞ ⎜ ⎟

⎝ dx ⎠ (72)
4.6. Adjoint differentiation of the bending moments
where d x rc/ d xf is calculated as follows:
d x rc df r The above steps again correspond to the direct method. Using the
= (λ xr )T f
d xf dx (73) adjoint method, the derivative of bending moment mif becomes:

where d f r/ d xf is calculated analogous to d f a/ d xf in Section 4.1.1, and dmif dpf


= (λ mf )T
the adjoint vector λ xr is calculated by solving the following adjoint d p fv d p fv (84)
equation:
where d p f / d p fv is calculated as before, using Eq. (29), and the adjoint
dxr T vector λ mf is calculated by solving the following adjoint equation:
(Kr)T λ xr = −⎛ rc ⎞ ⎜ ⎟

⎝ dx ⎠ (74)
f T
dm
where d x rc/ d xr is given by Eq. (69). (Kf )T λ mf = ⎜⎛ fi ⎟⎞
⎝ dx ⎠ (85)
4.5. Direct differentiation of the bending moments where dmif / dxf is calculated directly using Eqs. (77), (79), (15), (82)
and (52).
Calculating the sensitivities of the bending moment constraints re-
quires the differentiation of the bending moments, which is done next 5. Finite difference check of the design sensitivities
in element-wise notation. The ith element mif of vector mf is an internal
element force, and is an indirect function of the design variables: In this section, the analytical expressions for the design sensitivities
mî f (p fv ) = mif (f¯ e (u¯ e (x e (x Ff (xf (p fv )))))) given above are checked numerically. A simple two-dimensional ex-
(75)
ample is chosen to validate the derivatives. A bending-active arch is
Consequently, differentiation of mif gives: erected from a 4 m long straight flexible beam, with roller supports at
its ends (Fig. 4). One additional roller support in the middle eliminates
dmif ∂mif d f¯ e
= rigid body motion. The beam has a diameter of 10 mm, Young’s mod-
d p fv ∂f¯ e d p fv (76) ulus: E = 33 GPa, and is discretized into 8 elements. An initial erection
where: force of 5 N is applied vertically on each node. These forces are re-
garded as the design variables. After the beam is erected, it is braced
∂mif e
and the beam ends are pinned. The bracing elements have a diameter of
= L f¯ f
∂f¯ e mi (77) 1 mm, and Young’s modulus: E = 210 GPa. Removing the erection
e
where L f̄ f selects mif
e
from the corresponding internal force vector f̄ , forces allows the beam to relax. Next, the relaxed arch is loaded with a
mi
point load of 500 N in the middle.
depending on which beam end and around which axis the considered
The derivatives of the end-compliance C with respect to the form-
bending moment acts. Application of the chain rule to the term d f¯ e/ d p fv
finding loads p f are calculated using both the direct and adjoint
in Eq. (76) gives:
method. As they should, both methods give exactly the same results.
d f¯ e ∂f¯ e d u¯ e Therefore, these results are simply referred to as analytically calculated
=
dpv f ∂u¯ e d p fv (78) derivatives in what follows. These are compared to the same derivatives
calculated using a forward finite difference expression. For a general
where: derivative of f (x ) to x, assuming a small increment ∊:
∂f¯ e ¯e df f (x + ∊) − f (x )
=K ≈
∂u¯ e (79) dx ∊ (86)
where K̄ e is the element stiffness matrix defined in Eq. (13). Applying where we let the increment ∊ range from 10−1 to 10−10 (with the same
the chain rule to the term d u¯ e / d p fv in Eq. (78) gives: unit as x). It is observed from Table 1 that the analytically calculated
d u¯ e ∂u¯ e d x e derivatives vary slightly from the converged values of the numerically
=
dpv f ∂x e d p fv (80) calculated derivatives. Closer investigation has shown that this in-
accuracy is primarily caused by the fact that the analytical expression
where ∂u¯ e / ∂x e is given by Eq. (15), and: for the tangent stiffness matrix given by Crisfield [29] is not perfectly
consistent. Therefore, a comparison using numerically calculated stiff-
d xe ∂x e d x Ff
f
= ness matrices for the form-finding, relaxation and analysis stage is also
dpv ∂x Ff d p fv (81) included in Table 1, where d f/ d x was calculated using equation (86),
where: with ∊ = 10−10 m. Numerically calculating the stiffness matrices clearly
results in an improvement of the accuracy of the analytical derivatives.
∂x e
= LFe However, we observed that the inaccuracy caused by analytically cal-
∂x Ff (82) culating the tangent stiffness matrix has limited influence on the con-
where LFe is a selection matrix. The last term in Eq. (81) is given by: vergence behavior of the optimization problem. Therefore, we advise to
use the analytical expression for the stiffness matrix, as the numerical
d x Ff ∂x Ff d xf calculation is computationally more demanding.
=
d p fv ∂xf d p fv (83) Table 2 compares the analytically and numerically calculated deri-
vatives for the vertical position x5,r y of the middle node in the relaxation
where ∂x Ff/ ∂xf
is defined in Eq. (52), and d xf / d p fv is derived in Section
stage, validating the sensitivity expressions given in Sections 4.3 and
4.1.2. With d x / d p fv from Eqs. (29) and (57), d x Ff/ d p fv is found with Eq.
f
4.4 for the geometry constraint in (19). Finally, Table 3 compares the
(83), which can be used in Eq. (81) to find d x e/ d p fv . This is plugged into analytically and numerically calculated derivatives for a bending mo-
Eq. (80), together with Eq. (15) to get d u¯ e / d p fv , which is used in Eq. ment M4,f b at the right end of the fourth element (left side of the middle
(78) with ∂f¯ e/ ∂u¯ e from Eq. (79), finally leading to dmif / dp fv with Eqs. node) in the form-finding stage, required for the bending moment

177
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

constraint. assumed constant for the calculation of the wind load, and the vertical
component of the wind load is neglected. Only one wind direction is
6. Prototype considered during the optimization, because the response to other wind
directions is assumed equivalent due to the symmetry of the structure.
To validate the proposed optimization approach, a cantilevering, Moreover, a nonlinear finite element analysis has to be performed for
simply supported gridshell is designed and built. The method of moving each considered load case in every iteration, which strongly affects the
asymptotes [33] is chosen as the optimization algorithm, and implicit overall required computation time.
dynamic relaxation [15] is used to solve the nonlinear equilibrium Because the number of constraints is higher than the number of
equations. design variables, the direct differentiation method was used to calculate
the derivatives. A convergence criterion has been used to stop the op-
6.1. Optimization timization when the change in end-compliance corresponding to the
most compliant load case between two iterations was lower than
An initially flat grid of 6 m by 6 m is composed of 13 wooden laths 0.01 Nm. Using the method of moving asymptotes, convergence was
in each direction, placed 0.5 m apart (Fig. 5). The laths are made of reached after 54 iterations. Fig. 6 shows the distribution of the opti-
Yellow Pine, and have a cross section of 36 mm by 12 mm. The Young’s mized form-finding forces. The resulting gridshell is shown in all design
modulus E = 13 GPa and shear modulus G = 1.05 GPa were de- stages in Fig. 7. It can be observed directly that the difference in shape
termined experimentally. between the relaxation stage and the analysis stage for both load cases
The connections are modelled using short beam elements with zero is hardly visible, which indicates that the deformations under these
torsional stiffness, which allows the laths to rotate freely in the plane of loads remain small, meaning that the optimization has succeeded in
the grid. Since the laths in the x- and y-directions are placed on top of minimizing the end-compliance for the given conditions. A more in-
each other, the length of the connection elements is taken equal to the depth discussion of the gridshell’s performance is given in the next
lath thickness (12 mm) to account for the resulting eccentricity. subsection.
Moreover, the connection elements are given a Young’s modulus Fig. 8 shows the convergence history of the end-compliance due to
E = 210 GPa, and cross-section radius r = 10 mm. both considered load cases. The initial design has a compliance of
Four nodes of the grid are supported, leaving a strip of 1 m wide to 353.6 Nm and 16.6 Nm for load case 1 and 2 respectively. At con-
cantilever on each side (Fig. 5). The boundary conditions are chosen so vergence, the end-compliance corresponding to load case 1 reduced to
that the gridshell is simply supported, i.e., no thrust forces are taken by 50.5 Nm, whereas the compliance for load case 2 increased slightly to
the supports. One node is pinned, 2 nodes are modelled as roller sup- 18.2 Nm, showing that the optimizer only considers the load case that
ports, allowing movement in x- and y- direction (horizontally), and the leads to the highest compliance.
node opposite to the pinned node is only allowed to move in the x- Fig. 9 shows the maximum bending moment around all laths’ weak
direction. axis at the end of the form-finding stage for all 54 iterations. The
Two of the three types of constraints discussed in Section 3 are maximum bending moment occurring in the final design equals
considered in the current optimization problem. First, box constraints 45.0 Nm, which satisfies the constraint of 45 Nm. The bending moments
enforce the magnitude of the erection forces to be lower than 1 kN. after the relaxation stage are maximum 45.2 Nm, and the bending
Second, a constraint on the bending moment around the lath’s weak moments after the analysis phase are maximum 50.4 Nm and 48.1 Nm
axis makes sure these bending moments stay below 45 Nm. This limit for load case 1 and 2 respectively, which shows that most bending is
was chosen based on experiments of several sample laths. A safety introduced during the erection of the gridshell. Because of the added
margin is included to take the variability of the material properties into safety margin, this slight violation of the constraint should not be
account. problematic for the construction of the gridshell.
During the form-finding stage, a vertical form-finding load is ap-
plied every three nodes of the initially flat grid, resulting in a total of 25 6.2. Assessment of the gridshell’s structural performance
loads. These act as the design variables in the optimization and are
updated each iteration. For the first iteration, all form-finding loads are Already in the previous subsection, Fig. 7 showed that the de-
set to 100 N. For aesthetic reasons, the design is forced to be symmetric. formations caused by both considered load cases remain small for the
Therefore, the form-finding forces are artificially symmetrized by optimized gridshell. In this subsection, the structural efficiency of the
taking only 6 of the 25 form-finding loads as design variables, and optimized gridshell is demonstrated by subjecting it to a load case
copying them over each octant. All supported nodes are free to move equivalent to load case 1, but with an increased service load of 160 N/
horizontally in this stage. m2 to amplify the structural response. Since the gridshell was optimized
In the relaxation stage, 2 additional layers of the same lath type are for a similar load case, it is expected to perform well for this load case
added on top of the gridshell along the diagonals of the grid to brace the too. The optimized gridshell is compared to the same grid laid out flat,
structure and fix the shape. The previous connecting elements are used and a gridshell obtained by choosing a constant value for the form-
to model the connections between these additional laths and the ex- finding loads p f . The magnitude of these loads was changed iteratively
isting grid. The boundary conditions shown in Fig. 5 are implemented until the structural height of this gridshell was similar to the structural
in this stage. No loads are applied and the relaxation of the structure is height of the optimized gridshell. This last case resembles a gridshell
simulated. designed using a standard, uninformed form-finding approach. Al-
In the analysis stage, the relaxed gridshell is subjected to all relevant though it is obvious that a flat grid will deform more than a curved grid,
load cases. For this prototype, 2 load cases are taken into account. Load it is considered here to get an idea of the flexibility of the laths and their
case 1 consists of the self weight and an equally distributed service load response under the applied loads.
of 80 N/m2 , applied on the nodes, where the contributory area is cal- Fig. 10 shows the displacements for the three considered cases. The
culated for the flat grid, assuming a continuous surface. The second colors indicate the vertical displacements after applying the loads,
load case consists of the self weight of the gridshell and a wind load which range up to around 0.9 m (dark red) for the flat grid, 0.4 m
corresponding to a wind at 17 m/s in the x-direction. The wind load is (green) for the gridshell designed with a uniform form-finding force,
calculated based on Eurocode 1 part 1–4 [34]. For an exact calculation and only 0.1 m (light blue) for the optimized gridshell. The slight
of the wind load the calculation has to be performed in each iteration, asymmetry introduced by the layering of the laths and the different
and the sensitivities to the design variables should be determined, types of supports causes the flat grid to deform primarily in one di-
which is far from trivial. Therefore, the geometry of the gridshell is rection. Nevertheless, Fig. 10 clearly shows the structural efficiency of

178
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

the optimized gridshell. Not only does the gridshell designed with a the predicted and measured displacements are observed, the numerical
uniform form-finding load deform notably more than the optimized model does capture the overall behavior of the gridshell. The observed
gridshell, the bending moment around the weak axis of the laths locally inaccuracy is caused by an accumulation of building errors, measuring
reaches up to 130 Nm in the form-finding stage, which is much higher errors, and modeling approximations. Especially the assumption of
than the allowed 45 Nm. Consequently, a time-consuming trial and material linearity for the wood, which behaves quite non-linearly in
error approach would be needed to adapt the form-finding forces to reality, overestimates the stiffness of the gridshell, which could explain
obtain a feasible design. However, this would lead to a reduction of the the underestimation of the vertical displacements for load case 3, 4, and
structural height of this gridshell, and therefore also its structural ef- 5, especially at the loaded node.
ficiency. Contrarily, the optimized gridshell achieves a high structural
height, while still satisfying the bending constraint. 7. Conclusion

6.3. Construction This paper presents a shape optimization approach for strained
gridshells, where the objective is to minimize the end-compliance,
After the optimized design was obtained, the gridshell was built by using the erection forces as the design variables. In every iteration step,
students in the context of a master thesis project. The construction four design stages are modeled using co-rotational beam elements, si-
process followed the design stages considered in the optimization al- mulating the construction process of the gridshell. The design sensi-
gorithm (Fig. 11). First, the initial grid was assembled. Next, the flat tivities are derived to enable the use of a gradient-based optimization
grid was deformed to approximate the shape from the form-finding algorithm.
stage of the optimized design. Although the forces were used as the A prototype of a simply supported, cantilevering strained gridshell
design variables in the optimization process, a displacement based ap- was designed using the proposed shape optimization approach, with the
proach was followed during construction. Wooden struts were used to method of moving asymptotes [33] as the optimization algorithm, and
control the vertical position of the nodes loaded by the form-finding implicit dynamic relaxation [15] to solve the nonlinear equilibrium
loads, and additional lash straps were used to easily deform the grid equations. The prototype shows that the proposed optimization ap-
into position. In the following stage, diagonal laths were added, for proach is capable of increasing the overall stiffness of the gridshell with
which the holes were drilled on the spot to make sure the laths fitted the respect to multiple load cases, as compared to a gridshell of similar
erected grid. Next, all temporary supports were removed and the structural height designed by means of a classical form-finding proce-
gridshell was placed on its final supports, which allowed it to relax its dure. Moreover, it is shown that constraints on the internal bending
shape. Finally, the gridshell was subjected to a load test, which will be moments, and the erection forces can be applied successfully, ensuring
discussed in more detail in the next subsection. the optimized design is buildable in reality. Additional load tests on the
The laths were connected using steel bolts. Longer bolts were ap- built structure show that although the simulation of the final gridshell’s
plied at the nodes where diagonals would cross (Fig. 12), as they would displacements is somewhat inaccurate due to a combination of mod-
eventually connect four layers of laths. Wooden columns were used to eling, construction, and measuring errors, it still captures the overall
materialize the modeled boundary conditions (Fig. 13). A tripod fixed structural behavior.
one node, a bipod constrained the opposite node in x- and z- direction, In conclusion, the proposed optimization approach proves to be an
and a simple column was used for the other supported nodes. A hinge- effective tool for the design of structurally efficient strained gridshells,
like connection with the ground was achieved by putting the columns which is especially useful in cases where classic form-finding techni-
just a few centimeters in the ground, and using ground anchors to resist ques fail, such as simply supported, or cantilevering gridshells.
tension forces (Fig. 14). The connections of the columns with the actual
grid were made using an extended bolt that is free to rotate, essentially Acknowledgements
working as a ball joint (Fig. 15).
We would like to thank C. Mommeyer, W. Bols, P.-J. Van den
6.4. Comparison calculations and measurements Abbeele, and T. Raymaekers, for their contribution to the design and
construction of the prototype. This work was supported by the Research
In this subsection, the structural performance of the built gridshell is Foundation Flanders (FWO; Grant No. G0C2315N).
compared to the theoretical model by applying four point loads and
measuring and modeling the structural response. Fig. 16 shows the References
position and magnitude of the applied loads, as well as the resulting
deformations as predicted by the numerical model. Colors indicate the [1] Douthe C, Caron JF, Baverel O. Gridshell structures in glass fibre reinforced poly-
vertical displacements, which range up to around 0.1 m (dark red). mers. Constr Build Mater 2010;24(9):1580–9. https://doi.org/10.1016/j.
conbuildmat.2010.02.037.
Because we wanted the load test to be non-destructive, the magni- [2] Liddell I. Frei otto and the development of gridshells. Case Stud Struct Eng
tude of the loads was adapted to the resulting displacements. 2015;4:39–49. https://doi.org/10.1016/j.csse.2015.08.001.
Consequently, only the point load in the middle has a magnitude of [3] Harris R, Romer J, Kelly O, Johnson S. Design and construction of the downland
gridshell. Build Res Inform 2003;31(6):427–54. https://doi.org/10.1080/
1 kN, as the gridshell acts very stiff for this symmetric load case. The 0961321032000088007.
more distant the loading point was from the center of the gridshell, the [4] Harris R, Haskins S, Roynon J. The savill garden gridshell design and construction.
more the magnitude had to be reduced to limit the corresponding de- Struct Eng 86(17).
[5] Peloux LD, Baverel O, Caron J-F, Tayeb F. From shape to shell: a design tool to
formations. materialize freeform shapes using gridshell structures. Proceedings of the design
When the load was applied to the actual structure, significant per- modelling symposium Berlin. 2013.
manent deformations were observed after loading; a phenomenon [6] Peloux LD, Tayeb F, Caron J-F, Baverel O. The ephemeral cathedral of creteil: a
350m2 lightweight gridshell structure made of 2 kilometers of gfrp tubes.
which has also been reported by Labonnote et al. [16]. Moreover, be-
Proceedings of CIGOS. 2015.
cause the built shape of the gridshell deviates from the modeled shape, [7] D’Amico B, Kermani A, Zhang H. Form finding and structural analysis of actively
even in its unloaded state, only relative vertical displacements were bent timber grid shells. Eng Struct 2014;81:195–207.
measured. 13 measuring points were distributed over the gridshell [8] D’Amico B, Kermani A, Zhang H, Pugnale A, Colabella S, Pone S. Timber gridshells:
numerical simulation, design and construction of a full scale structure. Structures
nodes (Fig. 17). 2015;3:227–35. https://doi.org/10.1016/j.istruc.2015.05.002.
Table 4 compares the relative vertical displacements measured from [9] Pone S, Colabella S, Parenti B, Lancia D, Fiore A, D’Amico B, et al. Construction and
the built gridshell to those predicted by the numerical model for all 4 form-finding of a post-formed timber grid-shell. Proceedings of the international
conference on structures and architecture. 2013.
considered point loads. Although relatively large differences between

179
J. Rombouts, et al. Engineering Structures 192 (2019) 166–180

[10] Dyvik SH, Mork JH, Nilsen M, Luczkowski M. Modular kinematic timber gridshell; a the design of actively bent structures. Eng Struct 2016;117:518–27. https://doi.org/
simple scheme for constructing advanced shapes. Proceedings of the IASS annual 10.1016/j.engstruct.2016.03.034.
symposium. 2016. [22] Wakefield D. Dynamic relaxation analysis of pretensioned networks supported by
[11] Lienhard J, Alpermann H, Gengnagel C, Knippers J. Active bending, a review on compression arches Ph.D. thesis London: City University; 1980.
structures where bending is used as a self-formation process. Int J Space Struct [23] Bessini J, Lázaro C, Monleón S. A form-finding method based on the geometrically
2013;28(34):187–96. exact rod model for bending-active structures. Eng Struct 2017;152:549–58.
[12] Veenendaal D, Block P. An overview and comparison of structural form finding https://doi.org/10.1016/j.engstruct.2017.09.045.
methods for general networks. Int J Solids Struct 2012;49:3741–53. https://doi. [24] Ding C, Seifi H, Dong S, Xie YM. A new node-shifting method for shape optimization
org/10.1016/j.ijsolstr.2012.08.008. of reticulated spatial structures. Eng Struct 2017;152:727–35. https://doi.org/10.
[13] Liew A, Van Mele T, Block P. Vectorised graphics processing unit accelerated dy- 1016/j.engstruct.2017.09.051.
namic relaxation for bar and beam elements. Structures 2016;8:111–20. https:// [25] Haftka RT, Grandhi RV. Structural shape optimization—a survey. Comput Meth
doi.org/10.1016/j.istruc.2016.09.002. Appl Mech Eng 1986;57(1):91–106. https://doi.org/10.1016/0045-7825(86)
[14] Rombouts J, Lombaert G, De Laet L, Schevenels M. On the equivalence of dynamic 90072-1.
relaxation and the Newton-Raphson method. Int J Numer Meth Eng [26] Wall WA, Frenzel MA, Cyron C. Isogeometric structural shape optimization. Comput
2018;113(9):1531–9. https://doi.org/10.1002/nme.5707. Meth Appl Mech Eng 2008;197:2976–88. https://doi.org/10.1016/j.cma.2008.01.
[15] Rombouts J, Lombaert G, De Laet L, Schevenels M. A fast and accurate dynamic 025.
relaxation approach for form finding and analysis of bending-active structures. Int J [27] Douthe C, Baverel O, Caron JF. Form-finding of a gridshell in composite materials. J
Space Struct. Int Assoc Shell Spatial Struct 2006;47(1):53–62.
[16] Labonnote N, Mork JH, Dyvik SH, Nilsen M, Rønnquist A, Manum B. Experimental [28] Mesnil R, Ochsendorf J, Douthe C. Stability of pseudo-funicular elastic grid shells.
and numerical study of the structural performance of a timber gridhsell. Int J Space Struct 2015;30(1):27–36.
Proceedings of the world conference on timber engineering. 2016. [29] Crisfield MA. A consistent co-rotational formulation for non-linear, three-dimen-
[17] Barnes MR. Form finding and analysis of tension structures by dynamic relaxation. sional, beam-elements. Comput Meth Appl Mech Eng 1990;81:131–50. https://doi.
Int J Space Struct 1999;14(2):89–104. https://doi.org/10.1260/ org/10.1016/0045-7825(90)90106-V.
0266351991494722. [30] Belytschko T, Schwer L, Klein MJ. Large displacement, transient analysis of space
[18] Van Mele T, De Laet L, Veenendaal D, Mollaert M, Block P. Shaping tension frames. Int J Numer Meth Eng 1977;11(1):65–84. https://doi.org/10.1002/nme.
structures with actively bent linear elements. Int J Space Struct 2013;28(3 1620110108.
4):127–35. https://doi.org/10.1260/0266-3511.28.3-4.127. [31] Crisfield MA. Non-linear finite element analysis of solids and structures vol. 2.
[19] Du Peloux L, Tayeb F, Lefevre B, Baverel O, Caron J-F. Formulation of a 4-DoF Wiley; 1997.
torsion/bending element for the formfinding of elastic gridshells. In: Proceedings of [32] Felippa CA, Haugen B. A unified formulation of small-strain corotational finite
the international association for shell and spatial structures (IASS) symposium elements: I. Theory. Comput Meth Appl Mech Eng 2005;194:2285–335. https://doi.
2015, IASS symposium 2015: future visions, Amsterdam; August 2015. org/10.1016/j.cma.2004.07.035.
[20] Lefevre B, Tayeb F, Peloux LD, Caron J-F. A 4-degree-of-freedom kirchhoff beam [33] Svanberg K. The method of moving asymptotes—a new method for structural op-
model for the modeling of bending–torsion couplings in active-bending structures. timization. Int J Numer Meth Eng 1987;24(2):359–73. https://doi.org/10.1002/
Int J Space Struct 2017;32(2):69–83. https://doi.org/10.1177/ nme.1620240207.
0266351117714346. [34] European Commitee for Standardization. Actions on structures: General actions -
[21] D’Amico B, Zhang H, Kermani A. A finite-difference formulation of elastic rod for Wind actions, Eurocode 1 Part 1-4; 1991.

180

You might also like