ArticlePDF Available

Layout optimization of truss structures with modular constraints

Authors:

Abstract and Figures

Truss layout optimization is a well-established technique for designing efficient structures, but the optimized structures are often complex and associated with high manufacturing costs. To address this problem, repetitive modules are often used. However, relying solely on a single type of module may negatively impact the structural efficiency. To mitigate this trade-off, this study presents a novel approach for designing truss structures that incorporates multiple types of modules. Additionally, both the module arrangement and module structures are determined by the optimization approach. To achieve this, a novel mixed-integer linear programming problem is developed, and a heuristic method is proposed to enhance computational efficiency. The proposed approach is validated through several numerical examples, which demonstrate its ability to produce truss designs with low manufacturing costs and high structural efficiency.
Content may be subject to copyright.
Structures 55 (2023) 1460–1469
2352-0124/© 2023 The Author(s). Published by Elsevier Ltd on behalf of Institution of Structural Engineers. This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Layout optimization of truss structures with modular constraints
Yufeng Liu
a
, Zhen Wang
b
,
*
, Hongjia Lu
c
,
*
, Jun Ye
a
,
e
,
*
, Yang Zhao
a
,
d
, Yi Min Xie
c
a
College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310058, China
b
Department of Civil Engineering, Hangzhou City University, Hangzhou 310015, China
c
Centre for Innovative Structures and Materials, School of Engineering, RMIT University, Melbourne 3001, Australia
d
School of Civil Engineering, Shaoxing University, Shaoxing 312000, China
e
Center for Balance Architecture, Zhejiang University, Hang Zhou, 310014, China
ARTICLE INFO
Keywords:
Structure optimization
Modular constraints
Truss layout optimization
Mixed integer linear programming
ABSTRACT
Truss layout optimization is a well-established technique for designing efcient structures, but the optimized
structures are often complex and associated with high manufacturing costs. To address this problem, repetitive
modules are often used. However, relying solely on a single type of module may negatively impact the structural
efciency. To mitigate this trade-off, this study presents a novel approach for designing truss structures that
incorporates multiple types of modules. Additionally, both the module arrangement and module structures are
determined by the optimization approach. To achieve this, a novel mixed-integer linear programming problem is
developed, and a heuristic method is proposed to enhance computational efciency. The proposed approach is
validated through several numerical examples, which demonstrate its ability to produce truss designs with low
manufacturing costs and high structural efciency.
1. Introduction
Truss layout optimization is a technique used to design structures
that comply with certain design constraints, such as design domain,
loading, and support conditions, while also possess optimized properties
such as minimal material usage. However, the optimized structures
obtained through this method can be expensive to manufacture due to
their complex geometries. Additive manufacturing emerges as a prom-
ising cost-saving technique for creating complex designs. However, it
tends to be more suitable for small-scale applications like mechanical
components, rather than large-scale projects such as buildings and
bridges. Therefore, previous studies have considered member classi-
cations [13] and structural complexity constraints [4]. In this paper, an
alternative approach is proposed that involves using repetitive modules
to design truss structures. By using multiple types of modules, efcient
designs with low manufacturing costs can be achieved.
The concept of repetitive modules has attracted much interest in the
eld of continuum topology optimization. The related studies can be
classied into two categories: (1) the design of microstructural mate-
rials; (2) the design of repetitive macro structures. Category (1) aims at
customizing the microstructural architecture of identical unit cells ar-
ranged in a periodic array to create materials with exceptional
properties such as the maximum bulk modulus [5,6], negative Poissons
ratio [7] and negative thermal expansion factor [8]. Reviews for this
category of studies can be found in [6,9,10].
Category (2) focus on designing efcient structures comprised by
identical modules. One of the earliest studies in this area was conducted
by Huang and Xie [11], who combined the periodic constraints with the
Bi-directional Evolutionary Structural Optimization (BESO) technique
to design modular structures with maximum stiffness using a single type
of module. The method was subsequently expanded to include thermal
compliance optimization [12] and frequency optimization [13].
Furthermore, in [14], each periodic structure is allowed to rotate or
mirror in order to increase the optimization search region and identify
structures with better performance. Most of these studies have used pre-
dened non-overlapping sub-domains for the modules, but Rieser [15]
proposed using overlapping sub-domains to reduce the compliance in-
crease caused by the periodic constraints. Nevertheless, all the previous
research concentrated on continuum topology optimization, which is an
approach most suitable for problems with large volume fractions (i.e.,
ratio between structural volume and design domain volume), and may
not perform well when the volume fraction is low. In addition, these
methods often require intensive post-processing to produce practical
results [16]. As an alternative, we propose using the truss layout
* Corresponding authors.
E-mail addresses: wangzhen@zucc.edu.cn (Z. Wang), hongjia.lu@rmit.edu.au (H. Lu), jun_ye@zju.edu.cn (J. Ye).
Contents lists available at ScienceDirect
Structures
journal homepage: www.elsevier.com/locate/structures
https://doi.org/10.1016/j.istruc.2023.06.071
Received 2 February 2023; Received in revised form 31 May 2023; Accepted 14 June 2023
Structures 55 (2023) 1460–1469
1461
optimization approach to design modular structures.
The eld of truss layout optimization has a long history, dating back
to Michells pioneering work in 1904 [17]. Dorn et al. [18] later pro-
posed a numerical approach based on linear programming to identify
optimal truss structures. Subsequently, research in this area has focused
on plastic design with multiple loading conditions [19,20]. Building on
these foundations, Gilbert et al. [21] proposed a member adding
approach that signicantly increases numerical efciency without
compromising solution optimality. Additionally, geometry optimization
methods have been developed to further simplify optimized structures
obtained from layout optimization [22,23]. Recently, [24] also consid-
ered layout optimization with global stability constraints. Despite these
advances, research on the use of repetitive modular design in truss
layout optimization is limited. A recent study [25] investigated a
modular tall building skeleton system assembled from tetrahedral units,
but the pattern of the basic modular units was pre-dened rather than
optimized. Further research in this area is needed to fully explore the
potential of modular design in truss layout optimization.
Beside the gradient-based truss layout optimization methods, meta-
heuristic algorithms, such as genetic algorithm (GA) [2628], shufed
shepherd optimization algorithm (SSOA) [29], adaptive hybrid evolu-
tionary rey algorithm (AHEFA) [30] and hybrid algorithm that
combines different meta-heuristic algorithms [31,32] have also been
applied to truss structure design. For these meta-heuristic algorithms, a
key characteristic is to balance the capabilities of overall exploration
and of local search [33]. Optimizations considering dynamic excitations
[34], natural frequency [35] and other restrictions often need meta-
heuristic algorithms, for the introduction of these related factors often
make the mathematical problem complex and hard to solve with tradi-
tional methods. The expanding leads truss optimization to a more
complete and more inclusive frame [36,37]. However, as suggested by
Sigmund [38], the computational costs for meta-heuristic approaches
can be high. Although strategies like seeding initial population with
feasible solutions in some certain meta-heuristic algorithms have the
potential to improve computational efciency [39], most of the studies
in this eld have focused on size optimization rather than layout
optimization.
In summary, many previous investigations into modular structure
design were restricted to a single type of variable module or a selection
of pre-dened modules. In an effort to ll this gap, our study considers
the truss structural optimization problem that incorporates multiple
module types. The optimization process not only determines the internal
structure of each module but also their spatial arrangement. This is
achieved by formulating a new mixed integer linear programming
(MILP) problem. However, the high computational cost of solving MILP
problems can be an obstacle. To address this issue, we also present a
heuristic two-step approach to reduce the computational costs. The
paper is arranged as follows: Section 2 describes the proposed approach;
Section 3 presents numerical examples; Section 4 provides relevant
discussions, and conclusions are drawn in Section 5.
2. Formulation for truss layout optimization with modular
constraints
2.1. Review of layout and geometry optimization
Numerical truss layout optimization methods provide a powerful
means to generate highly efcient truss-like structures with given con-
ditions (e.g., domain, loading and supporting conditions). The standard
process involves a series of steps, as shown in Fig. 1(ac). Firstly, the
design domain, load and support conditions are specied (Fig. 1(a));
secondly, nodes are generated inside the design domain and potential
members are created by interconnecting these nodes, generating a
‘ground structure (Fig. 1(b)); thirdly, the optimal layout is identied
(Fig. 1(c)) by solving the following problem:
objective :min
a,qV=lTa(1a)
subject to :Bq =f(1b)
σ
caq
σ
ta(1c)
where V represents the volume of structure. l and a are the length vector
and cross-sectional area vector, respectively; B is known as the equi-
librium matrix, q is the internal force vector and f is the external force
vector;
σ
c
and
σ
t
are the allowable compressive and tensile stresses of the
material. The objective function (1a) represents the optimization goal of
minimum structural volume, constraint (1b) represents the mechanical
equilibrium constraint, and (1c) represents the stress constraint. a and q
are taken as the design variables.
Problem (1) uses the rigid-plastic material assumption. According to
Rozvany [20], in single load condition, the optimal results obtained by
plastic design are consistent with elastic optimal designs. Therefore, the
layout optimization approach is frequently used during the conceptual
design stage.
In addition, Problem (1) is a linear programming (LP) problem, thus
can be tackled effectively using modern interior point solvers. However,
the optimized structures are usually complex. To address this issue, the
geometry optimization [40] can be used. In this approach, the node
coordinates are treated as additional design variables on the basis of
layout optimization. The optimization formulation is written as follow:
objective :min
a,q,x,yV=l(x,y)Ta(2a)
subject to :B(x,y)q=f(2b)
σ
caq
σ
ta(2c)
where x =[x
1
, x
2
, , x
d
]
T
and y =[y
1
, y
2
, , y
d
]
T
are vectors of nodal x-
and y-coordinates with d denoting the number of nodes.
Problem (2) uses the optimized structures from Problem (1) as
starting points. In addition, geometrical modications such as merging
close nodes and creating crossovers are performed in-between the iter-
ations. Due to the non-convexity of Problem (2), the returned structural
Fig. 1. Workow of the traditional layout optimization approach: (a) design domain, loading and support conditions; (b) generate the ground structure; (c) the
optimized structure and (d) rationalize the structure by moving the nodes.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1462
volume may uctuate up or down slightly compared to the input
structure. For more details of the geometry optimization, the research
conducted by He and Gilbert [40] can be referred.
2.2. Modular constraints
In the current study, the truss structures comprised by multiple types
of modules will be designed. For clarity, the process is demonstrated
through a simple cantilever example shown in Fig. 2. In this example, we
rst divide the design domain into a number of identical sub-regions
(Fig. 2(b)), which are used as the potential slots for the structural
modules. In the next step, identical node grids and ground structures are
constructed in each slot (Fig. 2(c)). In the nal step, by solving the
optimization problem, the module arrangement (i.e., the type of module
in each slot) and the inner structure of each module are determined
(Fig. 2(d)).
In this process, the number of module types is pre-dened as p. To
apply the modular constraints, the module type variable tk,ζ is dened to
represent the existence of ζ-th type of module on the k-th slot. tk,ζ is a
binary variable (i.e., tk,ζ {0,1}) and tk,ζ=1 donates that the ζ-th type
of module is assigned to the k-th slot. To ensure that each slot adopts
only one type of module, the following constraint is used:
tk,1+tk,2++tk,ζ+tk,p=1,k {1,2, ...,
η
}(3)
where
η
represents the number of module slots.
In addition, the following constraint is used to guarantee that the
slots adopting the same module type have the same structure:
Am,1+M× (tk,2+tk,3++tk,p) aiAm,1M× (tk,2+tk,3++tk,p)
Am,2+M× (tk,1+tk,3++tk,p) aiAm,2M× (tk,1+tk,3++tk,p)
Am,p+M× (tk,1+tk,2++tk,p1) aiAm,pM× (tk,1+tk,2++tk,p1)
(4)
where a
i
is cross-sectional area of the i-th member in the structure, Am,ζ
donates the area of the m-th member for the ζ-th module type. Note that
index m is a local member index associated with the ground structure of
each module and i is a global member index associated with the ground
structure for the design problem. The corresponding index m for the i-th
member can be obtained in the modular ground structure construction
process (Fig. 2(c)). M is a pre-dened large number, which is commonly
used to activate and deactivate constraints in MILP problems [41].
To sum up, by combing Problem (1) and constraints (3) and (4), the
truss layout optimization problem with modular constraints is:
objective :min
a,q,t,Am
V=lTa(5a)
subject to :Bq =f(5b)
σ
caq
σ
ta(5c)
tk,1+tk,2++tk,p=1,k {1,2, ..., q}(5d)
Am+MTi>=aie>=AmMTi,i {1,2, ..., n}(5e)
where A
m
=[A
m,1
, A
m,2
, , A
m,p
]
T
; n is the number of members in the
structure; e= [1,1,1,]is an all-one vector; Ti is dened by:
Ti=
0tki,2tki,p
tki,10 tki,p
tki,1tki,2 0
,
where ki is the slot number of the i-th member.
In Problem (5), a, q, A
m
and T
i
are considered as design variables.
Specically, a, q, Am are continuous variables while t= [tk,1,tk,3,,tk,p]
are binary variables. In addition, the objection function (5a) and the
constraints (5b) to (5e) are all linear. This characteristic classies
Problem (5) a MILP problem. Therefore, we employ the reputable
commercial solver Gurobi [42] to solve Problem (5).
2.3. Geometry optimization of modular truss structure
The geometry optimization approach (i.e., Problem (2)) described in
Section 2.1 can be used on the modular structure as well. The optimi-
zation formulation is:
objective :min
a,q,x,y,t,Am
V=l(x,y)Ta(6a)
subject to :B(x,y)q=f(6b)
σ
caq
σ
ta(6c)
tk,1+tk,2++tk,p=1,k {1,2, ..., q}(6d)
Am+MTi>=aie>=AmMTi,i {1,2, ..., n}(6e)
Similar to Problem (2), Problem (6) is also a non-linear non-convex
problem. Thus, the structure obtained from Problem (5) is used as the
start point of Problem (6). Considering that, tk,i are all xed constants in
Problem (6), making all the variables in Problem (6) continuous. It is
worth noting that Problem (6) is only carried out on the optimized
structures which have much less members than the full ground structure
used in Problem (5). Therefore, the computational cost is negligible
compared to Problem (5).
2.4. Heuristic two-step approach for improving numerical efciency
MILP problems are usually companied by high computational costs.
Although in Problem (5), ground structures are constructed within the
module slots, resulting in much fewer potential members compared to
the traditional layout optimization problem, the computational costs
still increase rapidly as the design domain size expands. Therefore, a
heuristic approach is proposed to increase the computational efciency.
This approach is based on an assumption as following:
Assumption 1.The optimized module arrangements obtained using low-
or high-complexity modules are similar, with high-complexity modules being
those with a denser node grid in the module ground structure.
Based on this assumption, the two-step approach illustrated in Fig. 3
is proposed for the efciency improvement. In the rst step, the inter-
mediate modules (i.e., with low-complexity) are used (Fig. 3(b)) in
solving Problem (5) to obtain the module arrangements (Fig. 3(c)),
Fig. 2. Workow for truss layout optimization with repetitive modular constraints: (a) design domain, loading and support conditions; (b) divide the design domain
into potential module slots; (c) generate the ground structure for each module slot and (d) solve the problem and obtain the optimized structure.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1463
where Problem (5) is a MILP. In the second step (Fig. 3(d-e)), the ob-
tained module arrangement is used with the target modules (i.e., with
high-complexity) to obtain a more-efcient structure. Since the binary
variables are now constant, Problem (5) in the second solving step be-
comes a LP problem which can be solved with much less computational
costs. Considering that the MILP problem costs the majority of the
solving time, the computational costs of the second step is therefore
negligible. Therefore, compared to the direct solving approach in Sec-
tion 2.2, this heuristic approach can signicantly reduce the computa-
tional costs by reducing the numerical size of the MILP problem in the
rst step.
It is worth noting that Assumption 1 may lead to sub-optimal solu-
tions. However, it is found that the gap between the results obtained by
the heuristic approach and the global optimal results (i.e., obtained with
the approach in Section 2.2) is small, which will be demonstrated with
the examples in Section 3.
3. Numerical examples
In this section, several numerical examples are presented to
demonstrate the effectiveness of the proposed approaches. In these
classic examples which employ commonly used boundary conditions,
the self-weight of the structure is not considered. The allowable tensile
stress and compressive stress are both taken as
σ
t
=
σ
c
=1 MPa. The
calculation is conducted in a workstation with an Intel i7-12700 k
(3.61Ghz) CPU and 32 GB running memory. The parameter complexity
of modules, which is expressed as n
x
×n
y
, represents that each module
has n
x
nodes in horizontal direction and n
y
nodes in vertical direction.
For the termination criteria of Gurobi [43], 1% gap of objective function
is used for all the examples. As Gurobi utilizes a gradient-based solving
Fig. 3. Workow for the heuristic two-step approach: (a) divide the design domain into modules, (b) generate ground structure with intermediate module
complexity, (c) solve the intermediate optimization problem to identify the module arrangement, (d) generate ground structure with target module complexity, (e)
solve the target problem and obtain the nal optimized structure.
Fig. 4. Simply supported example: (a) case
description; (b) solution obtained with 2 ×
2 complexity modules, V =0.544 V
0
; (c)
solution obtained using the module
arrangement of (b) but switching to mod-
ules with 4 ×4 complexity, V =0.537 V
0
;
(d) geometry optimized result of (c), V =
0.536 V
0
; (e) solution obtained with direct
solving approach using 4 ×4 complexity
modules, V =0.539 V
0
; V is the structural
volume, the results are normalized against
the structural volume in Fig. 5(a); members
with the same color share the same module
type.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1464
algorithm, the solving process is deterministic, and the results are global
optimal. This characteristic effectively eliminates any potential inu-
ence from the solving algorithm.
3.1. A simply supported domain subjected to a point load
Firstly, we use a simple example to demonstrate the proposed
approach. The case description is shown in Fig. 4(a). In this example we
consider the module type number p=2 and adopt the heuristic two-step
approach in Section 2.4. In the rst step, the design domain is discretized
into 32 module slots as shown in Fig. 4(a) and the module arrangement
is identied by solving Problem (5) using 2 ×2 intermediate complexity
modules (i.e., Fig. 4(b)). In the second step, the module arrangement is
xed and 4 ×4 target module complexity is considered, and the opti-
mized structure is shown in Fig. 4(c). The geometry optimization
approach described in Section 2.3 is then used to further rationalize the
structure (Fig. 4(d)). For comparison, the result obtained by directly
using 4 ×4 modules with Problem (5) is shown in Fig. 4(e). Compare
Fig. 4(d) to Fig. 4(e), the volume difference is 1.47%, which is consistent
with Assumption 1. In terms of computational costs, Fig. 4(d) takes 72 s
of CPU time while Fig. 4(e) takes 275580 s. By using the two-step
approach, the solving time is reduced by 99.97% but the obtained so-
lutions are similar, highlighting the effectiveness of the proposed
approach.
To analyze the inuence of module type number p, we further
consider structures with p=1and 4. The optimized structures are
shown in Fig. 5(ab). Note that Fig. 5(a) is a result from the traditional
layout optimization algorithm that does not contains integer variables.
Since only one type of module is used in Fig. 5(a), much of the structural
material is non-fully stressed, causing an 86.68% increase in volume
compared to Fig. 4(d). When p is increased to 4, an extremely simple
truss structure is obtained (Fig. 5(b)), in which two modules consist of
only one inclined member, one module consists of only one horizontal
member, and one module is empty. Note that Fig. 5 (b) is obtained using
the heuristic two-step approach, with a CPU cost of 1325 s. From Figs. 4
and 5, it is clear that increasing the number of module types signicantly
reduces the structural volume, highlighting the benets of using multi-
ple types of modules.
3.2. Inuence of intermediate module complexity
The second example is a cantilevered structure, as illustrated in
Fig. 6. The design domain has dimensions L =6 m and H =3 m and is
subject to a vertical unit load F =1 N applied at the top right corner and
two pin supports at the two corners on the left boundary.
In this example, we consider the module type number p=4. The
intermediate and target module complexities are set to 2 ×2 and 4 ×4.
By using the heuristic two-step approach, the optimized structure is
shown in Fig. 7(a). In addition, the optimized structure obtained directly
using 4 ×4 modules are shown in Fig. 7(b). Compare Fig. 7(a) to (b), the
volume of Fig. 7(a) is 6.1% higher than that of Fig. 7(b), while the
associated CPU time is only 1/800 of the latter. However, Fig. 7(a) and
(b) display distinct geometries. While Fig. 7(b) contains a few empty
regions, all the slots are lled with structures in Fig. 7(a). A possible
reason for this is that the complexities of the intermediate modules are
too low to provide a close estimate of the optimal module arrangement,
causing the differences in both inner structures of modules and module
type arrangements in Fig. 7(a) and (b).
To verify the assumption, the problem is solved again with the in-
termediate module complexity switched to 3 ×3. With the modication,
the obtained result become identical to Fig. 7(b), and the CPU time is 3%
of Fig. 7(b). This suggests that intermediate modules with too low
complexity may lead to sub-optimal solutions. However, in this case, the
effect is minor since the volume difference between Fig. 7(a) and (b) is
only 6.1%.
Interestingly, two modules in Fig. 7(b) form a super module, a mini-
cantilever that duplicates three times, leading to only two types of
modules for manufacturing (i.e., a horizontal bar and the mini-
cantilever). To investigate the potential material saving from using a
more complex structure for the mini-cantilever, we consider 6 ×6 and 8
×8 modules and the optimized results are shown in Fig. 8. It can be seen
that using more complex mini-cantilever structures resulted in minor
volume decreases (i.e., less than 1%). This suggests that once the module
arrangement is determined, the inuence of the module complexity on
the optimized structural volume is minor. It is worth noting that Fig. 8(c)
has slightly higher volume than Fig. 8(a). This is because the nodes of
the 6 ×6 grid is not a subset of 8 ×8 grid, which further makes the
search region of the optimization problem in Fig. 8(a) not a subset of
Fig. 8(c).
3.3. The beam example
In this section, we consider a beam structure with aspect ratio equal
to 1:5 (height: length). The case description is shown in Fig. 9(a), in
which a unit load F =1 N is applied at the center of the bottom edge. The
number of the module types p=4 and the target module complexity is 4
×4. The problem is solved using the heuristic two-step approach (i.e.,
using 3 ×3 intermediate module) and the direct approach. These two
approaches lead to the same optimized structure as shown in Fig. 10.
However, the heuristic two-step approach (2362 s) takes only 1.96% the
Fig. 5. Simply supported example with different number of module types: (a) with one type of modules, V =V
0
; (b) with four types of modules, V =0.363 V
0
, where
V represents the structural volume. The results are normalized against the structural volume in Fig. 5(a).
Fig. 6. Schematic diagram of the cantilever case.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1465
CPU time of the direct approach (120498 s), demonstrating the nu-
merical efciency of the proposed approach. Despite this, here the
structural efciency can be further improved. Firstly, in Fig. 10, modules
with mirror topologies are considered as two types; secondly, the
modules located at the top corners (i.e., marked with red dashed lines)
are zero-stressed, but the limited number of module types forces the two
slots to be lled with structures (i.e., an additional module type is
required for using empty modules). Therefore, to address these issues, a
problem that take advantage of the symmetry is considered, as shown in
Fig. 9(b).
In Fig. 9(b), half of the problem is considered by adding roller sup-
ports at the right boundary. The heuristic two-step approach is used with
2 ×2 intermediate modules and 4 ×4 target modules. The module type
number p=3,4and 5 are considered and the optimized results are
shown in Fig. 11(the structure is supplemented according to symmetry).
Since in Fig. 11 the modules with mirror topologies are regarded as one
type, an additional empty module can be used, leading to the more
simplied structures and lower material usages (i.e., Fig. 11(b) has
Fig. 7. Cantilever case: (a) result obtained using heuristic two-step method with 2 ×2 intermediate modules, V =V
0
; (b) result obtained via directly using 4 ×4
modules, V =0.942 V
0
, where V represents the structural volume, the results are normalized against the structural volume in Fig. 7(a).
Fig. 8. Results of the cantilever case with more complicated modules (a) optimized structure with 6 ×6 target modules, V =0.939 V
0
; (b) geometry optimized result
of (a), V =0.937 V
0
; (c) optimized structure with 8 ×8 target modules, V =0.940 V
0
; (d) geometry optimized result of (c), V =0.936 V
0
, where V represents the
structural volume, the results are normalized against the structural volume in Fig. 7(a).
Fig. 9. Simply supported beam: (a) the full case; (b) the half case with symmetrical boundary conditions.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1466
7.03% lower volume compared to Fig. 10). It is noteworthy that Fig. 11
(b) still features four modules with horizontal zero-stressed members
due to the restriction on the number of module types, but this issue is
resolved when p is increased to 5, as shown in Fig. 11(c). In addition, as p
increases from 3 to 5, the volume of the optimized structure and the
layout complexity reduces, which is consistent with the conclusion ob-
tained from Fig. 4 and Fig. 5.
3.4. Inuence of the number of modules
To explore the impact of varying module slot discretization, we
conducted another cantilever example in this section. The problem
description is shown in Fig. 12(a). By taking advantage of the symmetry,
half of the problem is investigated. This example engages an interme-
diate module complexity of 2 ×2 and a target module complexity of 4 ×
4. We utilized four module types (p=4) while varying the module
numbers across cases.
Firstly, the result involving 4 ×2 modules, as shown in Fig. 12(b),
closely mirrors the result obtained from classical layout optimization
without modular constraints. This is understandable as all four modules
in the lower half exhibit different patterns, suggesting that the modular
constraints are not active in this case. Upon increasing the module
number to 8 ×4 (Fig. 12(c)) and 12 ×6 (Fig. 12(d)), the structural
volume rises by 33.7% and 47.0% respectively. This volume increase
might seem counterintuitive since, in most optimization problems, a
more rened node grid often yields improved results. Nonetheless, its
important to note that the number of module types does not increase
with the number of modules. As a result, the modular constraints
become more restrictive with the increase in module numbers. In this
scenario, the adverse effect of modular constraint restrictions surpasses
the benets offered by the rened node grid, leading to an increase in
structural volume.
3.5. Modules with 1:2 aspect ratio
In this section, we consider a building bracing design problem, which
targets at identifying the optimized structure subjected to horizontal
wind loads as shown in Fig. 13(a). The module type number p=4 and
the aspect ratio of each module slot is 2:1. In the heuristic two-step
approach, 2 ×2 intermediate modules and 6 ×6 target modules are
used and the optimized result is shown in Fig. 13(b).
For comparison, the result adopting only one type of modules is
Fig. 10. Optimized solution of the simply supported beam example, V =V
0
, where V represents the structural volume.
Fig. 11. Simply supported beam example with symmetrical boundary condition: (a) solution with 3 types of modules, V =1.120 V
0
; (b) solution with 4 types of
modules and, V =0.930 V
0
; (c) solution with 5 types of modules, V =0.701 V
0
, where V represents the structural volume, the results are normalized against the
structural volume in Fig. 10.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1467
shown in Fig. 13(c). It can be found that Fig. 13(b) have a volume of only
27.54% of Fig. 13(c). The volume reduction obtained via using more
modular types is larger than the previous examples in Fig. 5 and Fig. 11.
This is because in this example, the design problem is dominated by the
bending effect caused by the horizontal loads, with the maximum
bending moment occurring at the bottom edge and minimum bending
moment at the top. Consequently, using a single type of module for both
top and bottom results in a signicant material waste. However,
allowing for multiple types of modules, as seen in Fig. 13(b), allows for
the strategic placement of heavier modules at the bottom and lighter
modules at the top, resulting in a signicant reduction in material usage.
4. Discussions
From the results shown in Section 3, it can be seen that signicant
volume reductions can be achieved by using multiple types of modules
(compared to one type of module), suggesting the superiority of the
proposed approach. Meanwhile, the complexity of modules has a rela-
tively minor inuence on the optimized structural volume (as shown in
Figs. 7 and 8). In addition, it can be observed that the proposed heuristic
two-step approach can achieve similar results as the direct approach but
with much less computational costs. However, it should be pointed out
that even with the heuristic approach, the computational costs increase
Fig. 12. The cantilever case: (a) the case illustration (b) case with 4 ×2 modules, V =V
0;
(c) case with 8 ×4 modules, V =1.337 V
0
; (d) case with 12 ×6 modules, V
=1.470 V
0
; where V represents the structural volume, the results are normalized against the structural volume in Fig. 12(b).
Fig. 13. The horizontal wind load case (a) design domain, loading and supporting conditions; (b) solution with 4 types of modules, V =0.275 V
0
. (c) Solution with
only one type of modules, V =V
0
, where V represents the structural volume, results are normalized against structural volume in Fig. 13(c).
Y. Liu et al.
Structures 55 (2023) 1460–1469
1468
rapidly when more module slots are used, or 3D problem is considered.
Therefore, the future studies can focus on further improving the
computational efciency. This problem may be solved or eased by
abandoning the global optimality and turning to non-linear mixed
integer programming solver as suggested in [44].
In this study, we focused primarily on integrating only modular
constraints into the layout optimization problem for obtaining results
close to the global optimum. Given that our formulation is rooted in
truss optimization, it holds the potential to be combined with other truss
optimization approaches that take into consideration other factors like
stability, natural frequency, and dynamic loads. Notably, in some ex-
amples, the internal structures of modules can still be complex, sug-
gesting that further structural simplication techniques may be
benecial. An effective method has been proposed by Fairclough [45],
which holds promise for integration with our current algorithm.
5. Conclusions
This study explores the design of efcient modular truss structures by
optimizing both module arrangements and module structures concur-
rently. This approach presents a distinct advantage over previous
methods that restrict designs to a single type of module or a selection of
pre-dened modules. To facilitate this, we propose a novel mixed-
integer linear programming (MILP) formulation based on the tradi-
tional layout optimization approach. To mitigate the computational
costs associated with MILP problems, we employ a two-step heuristic
approach. In the rst step, we use a sparse node grid to determine
module arrangements. The second step then further renes the struc-
tural efciency of the solution from the rst step by utilizing a denser
node grid. The efcacy of this novel approach is underscored by several
illustrative examples, which yield the following conclusions:
(1) By increasing the number of module types, signicant material
savings can be achieved. In one example, a 72.46% reduction in
material usage is achieved by switching from one type of module
to four.
(2) The proposed heuristic two-step approach can lead to similar
optimized results as directly solving the MILP problem. However,
the computational cost of the former is only ~ 1% of the latter,
highlighting the efciency of the heuristic approach.
(3) In the cantilever example, it is observed that when the module
type number remains constant and the number of modules is
relatively low (e.g., 72), an increase in the number of modules
tends to have a counterproductive effect. This occurs as the re-
strictions imposed by modular constraints surpass the benets
gained from a rened node grid, resulting in an overall increase in
structural volume.
(4) Geometry optimization can be applied to further rationalize the
structure from the MILP problem to reduce structure complexity,
especially for these cases which have complicated inner struc-
tures in the modules.
Declaration of Competing Interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inuence
the work reported in this paper.
Acknowledgement
The authors would like to acknowledge the project supported by the
National Natural Science Foundation of China (No. 52078452,
52208215), the Basic Public Research Project of Zhejiang Province (No.
LGG22E080005, LQ22E080008) and the Australian Research Council
(No. FL190100014).
References
[1] Achtziger W, Stolpe M. Truss topology optimization with discrete design
variablesguaranteed global optimality and benchmark examples. Struct
Multidiscip Optim 2006;34(1):120.
[2] Mela K. Resolving issues with member buckling in truss topology optimization
using a mixed variable approach. Struct Multidiscip Optim 2014;50(6):103749.
[3] Lu H, Xie YM. Reducing the number of different members in truss layout
optimization. Struct Multidiscip Optim 2023;66(3):52.
[4] Cai Q, He L, Xie YM, Feng R, Ma J. Simple and effective strategies to generate
diverse designs for truss structures. Structures 2021;32:26878.
[5] Challis VJ, Roberts AP, Wilkins AH. Design of three dimensional isotropic
microstructures for maximized stiffness and conductivity. Int J Solids Struct 2008;
45(1415):413046.
[6] Cadman JE, Zhou S, Chen Y, Li Q. On design of multi-functional microstructural
materials. J Mater Sci 2012;48(1):5166.
[7] Andreassen E, Lazarov BS, Sigmund O. Design of manufacturable 3D extremal
elastic microstructure. Mech Mater 2014;69(1):110.
[8] Sigmund O, Torquato S. Design of materials with extreme thermal expansion using
a three-phase topology optimization method. J Mech Phys Solids 1997;45(6):
103767.
[9] Wu J, Sigmund O, Groen JP. Topology optimization of multi-scale structures: a
review. Struct Multidiscip Optim 2021;63(3):145580.
[10] Osanov M, Guest JK. Topology optimization for architected materials design. Ann
Rev Mater Res 2016;46(1):21133.
[11] Huang X, Xie YM. Optimal design of periodic structures using evolutionary
topology optimization. Struct Multidiscip Optim 2008;36(6):597606.
[12] Chen Y, Zhou S, Li Q. Multiobjective topology optimization for nite periodic
structures. Comput Struct 2010;88(1112):80611.
[13] Zuo ZH, Xie YM, Huang X. Optimal topological design of periodic structures for
natural frequencies. J Struct Eng 2011;137(10):122940.
[14] Thomas S, Li Q, Steven G. Finite periodic topology optimization with oriented unit-
cells. Struct Multidiscip Optim 2021;64(4):176579.
[15] Rieser J, Zimmermann M. Topology optimization of periodically arranged
components using shared design domains. Struct Multidiscip Optim 2022;65(1):18.
[16] Deaton JD, Grandhi RV. A survey of structural and multidisciplinary continuum
topology optimization: post 2000. Struct Multidiscip Optim 2013;49(1):138.
[17] Michell AGM. LVIII. The limits of economy of material in frame-structures. Philos
Mag 1904;8(4348):58997.
[18] Herbert J, Dorn WS, Gomory RE, Greenberg. Automatic design of optimal
structures. J Mec 1964;3(1):2552.
[19] Hemp WS, Chan H. Optimum structures. 1973.
[20] Rozvany GIN, Sokol T, Pomezanski V. Fundamentals of exact multi-load topology
optimization - stress-based least-volume trusses (generalized Michell structures) -
Part I: Plastic design. Struct Multidiscip Optim 2014;50(6):105178.
[21] Gilbert M, Tyas A. Layout optimization of large-scale pin-jointed frames. Eng
Comput 2003;20(78):104464.
[22] Ye J, Kyvelou P, Gilardi F, Lu H, Gilbert M, Gardner L. An end-to-end framework
for the additive manufacture of optimized tubular structures. IEEE Access 2021;9:
16547689.
[23] He L, Gilbert M, Johnson T, Pritchard T. Conceptual design of AM components
using layout and geometry optimization. Comput Math Appl 2019;78(7):230824.
[24] Weldeyesus AG, Gondzio J, He L, Gilbert M, Shepherd P, Tyas A. A Truss geometry
and topology optimization with global stability constraints. Struct Multidiscip
Optim 2020;62(4):172137.
[25] Angelucci G, Mollaioli F, Tardocchi R. A new modular structural system for tall
buildings based on tetrahedral conguration. Buildings 2020;10(12):240.
[26] Azid IA, Kwan ASK, Seetharamu KN. An evolutionary approach for layout
optimization of a three-dimensional truss. Struct Multidiscip Optim 2002;24(4):
3337.
[27] Chen S-Y, Shui X-F, Huang H. Improved genetic algorithm with two-level
approximation using shape sensitivities for truss layout optimization. Struct
Multidiscip Optim 2017;55(4):136582.
[28] Nguyena X-H, Lee J. Sizing, shape and topology optimization of trusses with energy
approach. Struct Eng Mech 2015;56(1):10721.
[29] Kaveh A. Size/layout optimization of truss structures using shufed shepherd
optimization method. Period Polytech-Civ Eng 2020;64(2):40821.
[30] Lieu QX, Do DTT, Lee J. An adaptive hybrid evolutionary rey algorithm for
shape and size optimization of truss structures with frequency constraints. Comput
Struct 2018;192:99112.
[31] Azad SK. Enhanced hybrid metaheuristic algorithms for optimal sizing of steel truss
structures with numerous discrete variables. Struct Multidiscip Optim 2017;55(6):
215980.
[32] Nguyen-Van Sy, Nguyen KT, Luong VH, Lee S, Lieu QX. A novel hybrid differential
evolution and symbiotic organisms search algorithm for size and shape
optimization of truss structures under multiple frequency constraints. Expert Syst
Appl 2021;184:115534.
[33] Kaveh A, Zolghadr A. Comparison of nine meta-heuristic algorithms for optimal
design of truss structures with frequency constraints. Adv Eng Softw 2014;76:930.
[34] Kazemzadeh Azad S, Bybordiani M, Kazemzadeh Azad S, Jawad FKJ. Simultaneous
size and geometry optimization of steel trusses under dynamic excitations. Struct
Multidiscip Optim 2018;58(6):254563.
[35] Nguyen-Van Sy, Nguyen KT, Dang KD, Nguyen NTT, Lee S, Lieu QX. An
evolutionary symbiotic organisms search for multiconstraint truss optimization
under free vibration and transient behavior. Adv Eng Softw 2021;160:103045.
Y. Liu et al.
Structures 55 (2023) 1460–1469
1469
[36] Lieu QX. A novel topology framework for simultaneous topology, size and shape
optimization of trusses under static, free vibration and transient behavior. Eng
Comput 2022;38(6):125.
[37] Dang KD, Nguyen-Van Sy, Thai S, Lee S, Luong VH, Lieu QX. A single step
optimization method for topology, size and shape of trusses using hybrid
differential evolution and symbiotic organisms search. Comput Struct 2022;270:
106846.
[38] Sigmund O. On the usefulness of non-gradient approaches in topology
optimization. Struct Multidiscip Optim 2011;43(5):58996.
[39] Azad SK. Seeding the initial population with feasible solutions in metaheuristic
optimization of steel trusses. Eng Optimiz 2018;50(1):89105.
[40] He L, Gilbert M. Rationalization of trusses generated via layout optimization. Struct
Multidiscip Optim 2015;52(4):67794.
[41] Soleimani-damaneh M. Modied big-M method to recognize the infeasibility of
linear programming models. Knowledge-Based Syst 2008;21(5):37782.
[42] Gonz´
alez M, L´
opez-Espín JJ, Aparicio J. A parallel algorithm for matheuristics: a
comparison of optimization solvers. Electronics 2020;9(9):1541.
[43] Optimization IG. Gurobi Optimizer Reference Manual. 2014.
[44] Belotti P, Bonami P, Fischetti M, Lodi A, Monaci M, Nogales-G´
omez A, et al. On
handling indicator constraints in mixed integer programming. Comput Optim Appl
2016;65(3):54566.
[45] Fairclough HE, He L, Pritchard TJ, Gilbert M. LayOpt: an educational web-app for
truss layout optimization. Struct Multidiscip Optim 2021;64(4):280523.
Y. Liu et al.
... However, most studies in both continuum and truss optimisation tend to rely on predefined modules or utilise only one type of module. Recently, a novel approach that concurrently optimises both the module structure and arrangement was introduced [46]. This method, which incorporates multiple types of modules, significantly mitigates the increase in structural volume typically caused by modular constraints (i.e., compared to those optimised solutions without modular constraints considered). ...
... A novel algorithm is introduced to identify optimised module arrangements using data clustering techniques, while an iterative optimisation scheme is employed to identify the modular structure. Compared to the previous approach in [46], our new method can significantly reduce computational cost with only a minor variation in structural volume. The paper is structured as follows: Section 2 reviews the related optimisation methods; Section 3 explains the principles and details of the new proposed algorithm; Section 4 demonstrates the effectiveness of the proposed approach through different numerical examples; Section 5 includes related discussions and conclusions are summarised in Section 6. ...
... In the authors' previous work [46], the optimisation of modular structures utilising the mixed integer linear programming approach is presented as: ...
Article
Full-text available
Modular structures offer significant potential for reducing manufacturing and construction costs. However, designing such structures with multiple module types poses a challenge. It requires consideration of both the arrangement of the modules and the internal structure of each module. In previous studies, mixed integer programming was employed to address this complexity. Although this approach can yield near-global optimum solutions, the substantial computational cost has limited its application in large-scale 3D problems. To address this challenge, an efficient iterative method is proposed in this study. This method starts with the result of standard layout optimisation and progressively steers it toward a modular structure. During this process, modular constraints are gradually enforced, and a data clustering method is used to determine the module arrangement. Due to the omission of integer variables, the computational cost is significantly reduced. Several numerical examples are presented to demonstrate the effectiveness of the proposed approach. Results indicate that the new method significantly reduces computational costs compared to previous methods while maintaining comparable structural efficiency, thus enabling applications to large-scale 3D problems.
... The ground structure approach is typically used for layout optimization. The truss-like ground structure is a highly interconnected truss consisting of several potential bars that connect nodes specified in a prescribed design domain [17]. The generated ground structure provides data that can be used for further analysis and optimization of truss structures. ...
Conference Paper
Full-text available
This paper suggests the design strategies of precast concrete panels fabricated by 3D-printing technology. 3D-printed precast panels offer high shape-flexibility, reduce workforce demand, and reduce environmental impacts. However, they must not be too heavy to transport and install. The consideration of weight emphasizes the need to perform structural optimization. This research presents truss-like structural modeling based on the properties of 3D-printed concrete along with the optimal design methodology using the Bi-directional Evolutionary Structural Optimization (BESO) algorithm to achieve the minimum weight. Studies on numerical examples are also conducted to understand the relationship between the proposed strategies and the optimal results achieved.
... Secondly, discrete topology optimization is mainly used for truss structure optimization and determines the number and position of bars for the truss and its respective cross-sectional areas [14]. The process, also known as layout optimization, involves four stages: (a) specifying the design domain, loads, and supports; (b) discretizing the domain through a grid of nodes and interconnecting them with potential members, thus defining a ground structure; (c) optimizing to identify the subset of members forming the optimal structural layout; and (d) using geometry optimization to improve the solution [15,16]. ...
Article
Full-text available
As the operation of buildings becomes more efficient, the carbon emissions generated by other phases of the building’s life cycle should also be mitigated to address the climate crisis. In this sense, structural systems play an essential role in the total embedded carbon of construction. This paper presents an approach to the conceptual design development of truss structures based on material quantity and embedded carbon. For this, a multi-objective optimization process enables the integration of different criteria, such as structural performance, shape complexity, utilization ratio, and design rationalization. The procedure is implemented in Rhino/Grasshopper using a parametric model that the designer can adjust according to the project requirements. The procedure was applied to two study cases consisting of long-span roof structures. The results show that the mass and embedded carbon can be decreased by over 50% after implementing the present approach. They also indicate that material quantity and embedded emissions tend to increase when augmenting cross-section rationalization; however, displacements have the opposite response. Furthermore, it was found that some topologies perform better regarding the two first objectives (material quantity and embedded emissions). The proposed workflow allowed for the assessment of different rationalization levels of the design to maintain a reduction in these variables while enabling a more suitable truss for construction, helping improve the energy efficiency of buildings driven by a design rationalization perspective.
... However, a seamless workflow is needed for Architecture, Engineering, and Construction (AEC) specialists to optimize their design, since the available methods heavily depend on commercial software and technical knowledge. Most of the literature relies either on standalone analysis tools, commercial software, or algorithms coded specifically as part of the research [11][12][13][14][15]. For example, in [13], the authors optimized prestressed truss structures by considering shape and size variables by following AISC-ASD specifications and employing the parameter-free Jaya algorithm via custom computer software coded for the study. ...
Article
Full-text available
Featured Application Artificial intelligence (AI)-based structural design optimization based on reliability in building information modeling (BIM) projects. Abstract Providing safe, environmentally conscious, and cost-effective designs is the primary duty of civil engineers. To this end, many different algorithms and methods have been developed in parallel with the progress of digital technologies over the past decades. Techniques such as AI-based Metaheuristic Algorithms (MAs), Reliability Analysis, and Building Information Modelling (BIM) are some of those methods that serve this purpose. The present study focuses on establishing a design optimization methodology by implementing the techniques in the open literature on one software environment to create a robust engineering and architectural workflow. The methodology involves multiple stages such as (i) creating parametric trusses employing Visual Programming (VP) software Dynamo (Version: 3.0.4); (ii) performing a First-order Reliability Method (FORM) analysis which includes a Finite Element Method (FEM) analysis as a Limit State Function (LSF); (iii) employing MAs to achieve optimum design variables under uncertain design constraints; (iv) testing the methodology with various real-word examples and scenarios; (v) creating an optimized model on Robot Structural Analysis 2024 (RSA) software in real time in the purpose of further adjustments. The results demonstrated that creating a design optimization workflow by utilizing a BIM environment can enhance the design process by easing the storing, sharing, and utilizing of design data by different branches capable of performing different complicated tasks successfully.
... Efficient design of truss structures considering truss layout optimum design is a known technique and at the same time is very challenging and complex that requires high building costs. Using a novel method and utilizing multiple types of modules, mixed-integer linear programming issue was created, and with choosing a heuristic method and applying it to the created problem, indicated that computational efficiency of the proposed approach was enhanced 3 . Jaya optimization algorithm is a well-known metaheuristic method, that its capability in optimum design of prestressed steel trusses was investigated to study the optimum design of shape and size decision variables in a simultaneous manner. ...
Article
Full-text available
Optimum design of truss structures can be a challenging and difficult field of study specially if an optimum design problem is comprised of continuous and discrete decision variables such as in truss size and layout optimization problems. Usually metaheuristic approaches are selected to solve these kind of problems and there are many different metaheuristic methods that have been used in dealing with truss optimization problems, which provide optimum solutions for those problems. However, among these various proposed metaheuristics, it is very seldom to face and find a method that is based on technologies such as computer networks. This paper aims to utilize a metaheuristic algorithm which is based on some principals of connected banking systems, which is called the Connected Banking System Optimizer(CBSO) algorithm. The performance of the CBSO is tested against three benchmark truss size/layout optimum design problems namely 15, 18 and 25 bar truss structures. Besides the truss size/layout problems, CEC-BC-2017 test functions with ten dimensions and also three other constrained optimum design problems are also investigated to examine the performance of the CBSO in dealing with these standard unconstrained and constrained test functions. The results of the CBSO optimum designs are presented and compared with some of the available studies in the literature. Statistical analyses are done to have a fair evaluation of the CBSO algorithm’s efficiency and its ability in dealing with the benchmark truss size/layout optimization problems. The CBSO outperforms its rivals and presents lighter weight truss structures. There is no need for configuration of extra parameters in the CBSO algorithm when solving the examined benchmark truss structures, which makes it a robust parameter-free optimization algorithm.
... Additive manufacturing (AM) or 3D printing, a robotic layer-by-layer production process, has seen rapid development and widespread use across multiple sectors in recent decades [1][2][3][4][5][6][7]. This technology has recently been introduced to the construction industry, offering a range of potential benefits and uses, such as greater automation, increased geometric flexibility, reduced material consumption, hybrid construction, strengthening and repair [8][9][10][11][12][13][14]. A particular advantage of AM that is drawing increasing attention is its potential to facilitate production in remote locations, where conventional construction capacity is restricted [15][16][17]. ...
Article
Full-text available
Wire arc additive manufacturing (WAAM) offers the potential for automated construction in remote locations such as in polar regions, where the climate is cold and harsh and labour is scarce. To facilitate such applications, an experimental study into the mechanical properties of WAAM steel plates at polar temperatures has been conducted. Following microstructural and geometric characterisation, 48 as-built and machined WAAM tensile coupons made of normal and high strength steel wires were tested in the temperature range of − 100 ℃ to 20 ℃. The low temperature environment was achieved by means of a liquid nitrogen cooling chamber, while the deformations of the coupons were monitored using digital image correlation (DIC). The stress-strain curves and key mechanical properties of the WAAM material are presented and discussed. It is shown that the Young's modulus, strength and ductility of the WAAM steels increased with decreasing temperature to varying degrees, with the clearest ascending trend observed for the ultimate strength. Fractography was carried out on the tested coupons, revealing principally ductile fracture at low temperatures, though coupled with somewhat brittle features. Finally, predictive equations that can accurately capture the low temperature mechanical properties of the WAAM steels are proposed.
... However, the structures obtained by this method usually come with higher manufacturing costs due to the flexibility in layout, size, and shape of the structure, generating often complex geometries. To this end, additional constraints may be added to limit the complexity of the final structure, by utilizing a select set of modules for the layout of the structure [8]. ...
Preprint
Full-text available
The optimal structural design is imperative to minimize material consumption and reduce the environmental impact of construction. Given the complexity in the formulation of structural design problems, the process of optimization is commonly performed using artificial intelligence (AI) global optimization, such as Genetic Algorithm (GA). However, the integration of AI-based optimization together with visual programming (VP) in building information modeling (BIM) projects warrants further investigation. This study proposes a workflow by combining structure analysis, VP, BIM, and GA to optimize trusses. The methodology encompasses several steps including: (i) generation of parametric trusses in Dynamo VP; (ii) performing the finite element modeling (FEM) using Robot Structural Analysis (RSA); (iii) retrieving and evaluating the FEM results interchangeably between Dynamo and RSA; (iv) finding the best solution using GA; and (v) importing the optimized model into Revit, enabling the user to perform simulations and engineering analysis, such as life cycle assessment (LCA) and quantity surveying. The feasibility of the proposed workflow was tested on benchmark problems and compared with open literature. The outcomes of this study offer insight into the opportunities and limitations of combining VP, GA, FEA, and BIM for structural optimization applications, particularly to enhance structural efficiency and sustainability in construction.
... However, the structures obtained by this method usually come with higher manufacturing costs due to flexibility in the layout, size, and shape of the structure, which often generates complex geometries. To this end, additional constraints may be added to limit the complexity of the final structure by utilizing a select set of modules for the layout of the structure [8]. ...
Article
Full-text available
The optimal structural design is imperative in order to minimize material consumption and reduce the environmental impacts of construction. Given the complexity in the formulation of structural design problems, the process of optimization is commonly performed using artificial intelligence (AI) global optimization, such as the genetic algorithm (GA). However, the integration of AI-based optimization, together with visual programming (VP), in building information modeling (BIM) projects warrants further investigation. This study proposes a workflow by combining structure analysis, VP, BIM, and GA to optimize trusses. The methodology encompasses several steps, including the following: (i) generation of parametric trusses in Dynamo VP; (ii) performing finite element modeling (FEM) using Robot Structural Analysis (RSA); (iii) retrieving and evaluating the FEM results interchangeably between Dynamo and RSA; (iv) finding the best solution using GA; and (v) importing the optimized model into Revit, enabling the user to perform simulations and engineering analysis, such as life cycle assessment (LCA) and quantity surveying. This methodology provides a new interoperable framework with minimal interference with existing supply-chain processes, and it will be flexible to technology literacy and allow architectural, engineering and construction (AEC) professionals to employ VP, global optimization, and FEM in BIM-based projects by leveraging open-sourced software and tools, together with commonly used design software. The feasibility of the proposed workflow was tested on benchmark problems and compared with the open literature. The outcomes of this study offer insight into the opportunities and limitations of combining VP, GA, FEA, and BIM for structural optimization applications, particularly to enhance structural efficiency and sustainability in construction. Despite the success of this study in developing a workable, user-friendly, and interoperable framework for the utilization of VP, GA, FEM, and BIM for structural optimization, the results obtained could be improved by (i) increasing the callback function speed between Dynamo and RSA through specialized application programming interface (API); and (ii) fine-tuning the GA parameters or utilizing other advanced global optimization and supervised learning techniques for the optimization.
Article
The article examines the question of the effectiveness of rational steel combined sprengel trusses compared to: typical. Modern approaches to the rational design of combined sprengel steel structures, in particular trusses, are considered. Particular attention is paid to reducing the metal content and labor intensity of manufacturing by optimizing geometric parameters, choosing rational cross-sections of elements, using high-strength materials, and regulating the stress-strain state (SSS) of structures. Methods for regulating SSS are presented, which include creating eccentricities in nodes and supports, as well as changing the cross-sections of individual elements. Maximum moment in the stiffening girder above the second support by 18%, which will further increase its efficiency. The prospects for the implementation of adaptive structures capable of changing their geometry depending on the loads using active elements have been investigated. The advantages of such solutions, including reducing the mass of structures and the technological complexity of their implementation, have been determined. The issue of reusing steel elements of building structures, which contributes to reducing CO 2 emissions, construction costs and resource conservation, has been considered. The results obtained contribute to increasing the efficiency and environmental friendliness of steel structures, offering promising solutions for sustainable construction in the future.
Article
Full-text available
Despite the long history of the truss layout optimization approach, its practical applications have been limited, partly due to high manufacturing costs associated with complex optimized structures consisting of members with different cross-sectional areas and member lengths. To address this issue, this study considers optimizing truss structures comprising limited types of members. On this topic, two distinct problems are considered, wherein the first problem, members of the same type share the same cross-sectional area (i.e., section-type problem); and in the second problem, members of the same type share the same cross-sectional area and length (i.e., member-type problem). A novel post-processing approach is proposed to tackle the target problems. In this approach, the optimized structures from the traditional layout and geometry optimization approaches are used as the starting points, members of which are then separated into groups by the k-means clustering approach. Subsequently, the clustered structures are geometrically optimized to reduce the area and length deviations (i.e., the differences between member area/length values and the corresponding cluster means). Several 2D and 3D examples are presented to demonstrate the capability of the proposed approaches. For the section-type problem, the area deviations can be reduced to near 0 for any given cluster number. The member-type problem is relatively more complex, but by providing more clusters, the maximum length deviation can be reduced below the target thresholds. Through the proposed clustering approach, the number of different members in the optimized trusses can be substantially decreased, thereby significantly reducing manufacturing costs.
Article
Full-text available
This article proposes a novel topology framework for simultaneously optimizing topology, size and shape of truss structures with multiple constraints under static, free vibration and transient responses for the first time. To achieve such a purpose, the topology pseudo-area variable of members is newly proposed discretely assigning to either 10-310310^{-3} or 1 to respectively represent the absence or presence of a member. This suggestion aims at not only evading the numerical instability due to the singularity of global stiffness matrix when solving equilibrium equations in finite element analyses but also saving the computational effort owing to the intact preserve of FE model structure. The objective function of this study is to minimize the structural weight. The cross-sectional area of truss members is taken discrete/continuous design variables into account, whilst nodal coordinates are treated as continuous ones. In addition, kinematic stability, displacement, stress, Euler buckling loading, natural frequency and transient behavior are dealt with as constraints. The derivative-free adaptive hybrid evolutionary firefly algorithm is utilized as an optimizer to resolve such optimization problems including mixed continuous-discrete variables. A large number of benchmark examples are tested to verify the validity of the presented paradigm. Obtained outcomes indicate that the present methodology is effective and robust in searching better high-quality optimal solutions against many existing algorithms in the literature.
Article
Full-text available
Building structures from identical components organized in a periodic pattern is a common design strategy to reduce design effort, structural complexity and cost. However, any periodic pattern will impose certain design restrictions often leading to lower structural efficiency and heavier weight. Much research is available for periodic structures with connected components. This paper addresses minimal compliance design for periodic arrangements of unconnected components. The design problem discussed here is relevant for many applications where a tightly nested, space-saving arrangement of identical components is required. We formulate an optimal design problem for a component being part of a periodic arrangement. The orientation and position of the component relatively to its neighbours are prescribed. The component design is computed by topology optimization on a design domain possibly shared by several neighbouring components. Additional constraints prevent components from overlapping. Constraint aggregation is employed to reduce the computational cost of many local constraints. The effectiveness of the method is demonstrated by a series of 2D and 3D examples with an ever-smaller distance between the components. Moreover, problem-specific ranges with only little to no increase in compliance are reported.
Article
Full-text available
Although additive manufacturing (AM) has been maturing for some years, it has only recently started to capture the interest of the cost-sensitive construction industry. The research presented herein is seeking to integrate AM into the construction sector through the establishment of an automated end-to-end framework for the generation of high-performance AM structures, combining sophisticated optimization techniques with cutting edge AM methods. Trusses of tubular cross-section subjected to different load cases have been selected as the demonstrators of the proposed framework. Optimization studies, featuring numerical layout and geometry optimization techniques, are employed to obtain the topology of the examined structures, accounting for practical and manufacturing constraints. Cross-section optimization is subsequently undertaken, followed by a series of geometric operations for the design of free-form joints connecting the optimized members. Solid models of the optimized designs are then exported for wire arc additive manufacturing (WAAM). Following determination of the optimal printing sequence, the trusses are printed and inspected. The efficiency of the optimized designs has been assessed by means of finite element modelling and compared against equivalent conventional designs. More than 200% increases in efficiency (reflected in the capacity-to-mass ratios) were achieved for all optimized trusses (when compared to their equivalent reference designs), demonstrating the effectiveness of the proposed optimization framework.
Article
Full-text available
Periodic topology optimization has been suggested as an effective means to design efficient structures which address a range of practical constraints, such as manufacturability, transportability, replaceability and ease of assembly. This study proposes a new approach for design of finite periodic structures by allowing variable orientation states of individual unit-cells. In some design instances of periodic structures, the unit-cell may exhibit certain geometric features allowing multiple possible assembly orientations (e.g. facing up or down). For the first time, this work incorporates such assembly flexibility within the periodic topology optimization, which enables to greatly expand the conventional periodic design space and take more advantage of structural periodicity. Given its broad applications, a methodology for the design of more efficient periodic structures while maintaining the same degree of periodic constraint may be of significant benefit to engineering practice. In this study, several numerical examples are presented to demonstrate the effectiveness of this new approach for both static and vibratory criteria. Brute force analysis is also utilized to compare all possible assembly configurations for several periodic structures with a small number of unit-cells. A heuristic approach is suggested for selecting more beneficially oriented configurations in periodic structures with a large number of unit-cells for which an exhaustive search may be computationally infeasible. It is found that in all the presented cases, the oriented periodic structures outperform the conventional non-oriented (or namely translational) periodic counterparts. Finally, an educational MATLAB code is provided for replication of the design results in this paper.
Article
Full-text available
A new interactive truss layout optimization web-app has been developed for educational use. This has been designed to be used on a range of devices, from mobile phones to desktop PCs. Truss designs are first generated via numerical layout optimization and then rationalized via geometry optimization. It is then shown that these designs can be simplified using a computationally inexpensive process that allows the user to control the trade-off between complexity and structural volume. The process involves the use of smooth Heaviside representations of member existence variables, with nodal slack forces employed that allow unstable intermediate truss structures. Full details of the web-app are provided in this contribution, from underlying formulation to cloud computing implementation. A range of numerical examples are used to demonstrate the efficacy of the web-app, and to show how it can potentially be used in educational and practical engineering settings.
Article
Full-text available
Structural topology optimization plays an important role in obtaining conceptual designs in the preliminary design stage. However, traditional structural optimization methods can only generate one optimized design for the material distribution under certain constraints. The optimized structure could have some disadvantages, such as an unattractive appearance, difficulty in manufacturing, or high construction cost. Therefore, it is more practical to produce multiple designs that not only have high structural performance but also have substantially different forms from which the designer can choose. Two strategies were explored in this study for generating diverse truss structures, namely, the penalizing length method (PLM) and the modifying ground structure method (MGSM). Using the proposed PLM, it is possible to delete unneeded bars in the optimized structure, such as very slender bars, and the cross-sectional areas of the remaining bars will be automatically redistributed to ensure structural nodal stability. In addition, by generating overlapping potential bars in the ground structure, the structural instability problem caused by pin joints can be overcome. Two-dimensional and three-dimensional numerical examples were provided to indicate the effectiveness of the proposed methods. The numerical results showed that the proposed methodologies can generate diverse structures while maintaining structural performance.
Article
In this article, a single step optimization approach for topology, size and shape of trusses subjected to multiple static and free vibration constraints is developed. For that aim, a topology pseudo-area variable based on a penalty parameter is discretely assigned by either 10⁻³ or 1 which represents the absence or attendance of a truss member. This helps to not only dodge the singularity of global stiffness matrix as resolving the equilibrium equation system, but also preserve the intact finite element model structure without unnecessarily and repeatedly done time-consuming performances in finite element analyses. The members’ cross-sectional area serves as discrete/continuous size variables, whilst nodes’ spatial coordinates are considered as continuous shape ones. The structural weight is minimized with multiple restrictions on kinematic stability, stress, displacement, natural frequency and Euler buckling loading. A hybrid differential evolution and symbiotic organisms search is employed as an optimizer and refined to tackle both continuous and discrete variables. Eight well-known examples for simultaneous topology, size and shape optimization of 2D and 3D trusses imposed by multiple static and free vibration constraints are tested to verify the reliability and the robustness of the suggested paradigm. The current method can result in competitive high-quality solutions against other state-of-the-art algorithms in most of the examined examples. In addition, the current approach can be also extended to apply for simultaneous topology, size and shape optimization of large-scale trusses in practice.
Article
In this paper, an evolutionary symbiotic organisms search algorithm as a hybridization of differential evolution and symbiotic organisms search is developed for shape and size optimization of truss structures with free vibration and transient behavior under multiple constraints. For that aim, a mutation operator combined by two global differential evolution operators and a novel symbiotic organisms search operator is proposed to reinforce the exploration ability of the proposed algorithm. This newly suggested symbiotic organisms search operator is relied upon the symbiotic relationship in such a way that an arbitrary organism can receive benefits from both mutualisms and commensalisms. A threshold is automatically integrated into the mutation step to switch from the exploration capability to the exploitation one. In addition, an elitist scheme is applied to the selection phase to purify the most potential candidates for the next symbiotic ecosystem. Accordingly, the present algorithm can result in a high-quality optimal solution with a better convergence evolution. 26 mathematical functions and 7 well-known benchmark problems regarding size and shape truss optimization under multiple constraints are tested to verify the effectiveness of the proposed methodology. Obtained outcomes have indicated that the developed algorithm outperforms both original algorithms and many existing approaches in the literature. To further illustrate the ability of the proposed paradigm, two among the above five examples under transient excitations with strength, displacement, and buckling constraints are then optimized.
Article
Although a large number of metaheuristic algorithms and their variants have been proposed for many engineering optimization problems, no paradigms hybridized by differential evolution (DE) and symbiotic organisms search (SOS) to concurrently improve the optimal solution quality and the convergence speed have been published thus far, especially for size and shape optimization of truss structures with multiple frequency constraints. Therefore, this article aims to propose a novel optimization algorithm as a cross-breed of the DE and the SOS, named HDS, for such problems. This algorithm can simultaneously and effectively enhance both global and local searching abilities by utilizing newly developed operators hybridized from the DE and SOS. An automatically adapted parameter is suggested for a better trade-off between those two capabilities. Furthermore, an elitist scheme is used in the selection phase to extract the best solutions for the next generation. As a consequence, the proposed methodology results in high-quality optimal solutions with a lower computational effort in comparison with two original methods, even many other optimization paradigms available in the literature. 26 benchmark mathematical functions are examined first. 5 numerical examples of shape and size optimization of truss structures are then investigated to validate the feasibility of the current paradigm.