Ali Kaveh - Meta-Heuristic Algorithms For Optimal Design of Real-Size Structures
Ali Kaveh - Meta-Heuristic Algorithms For Optimal Design of Real-Size Structures
Meta-heuristic
Algorithms for
Optimal Design
of Real-Size
Structures
Meta-heuristic Algorithms for Optimal Design
of Real-Size Structures
Ali Kaveh Majid Ilchi Ghazaan
•
Meta-heuristic Algorithms
for Optimal Design
of Real-Size Structures
123
Ali Kaveh Majid Ilchi Ghazaan
Department of Civil Engineering Department of Civil Engineering
Iran University of Science and Technology Iran University of Science and Technology
Tehran Tehran
Iran Iran
This Springer imprint is published by the registered company Springer International Publishing AG
part of Springer Nature
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
v
vi Preface
Optimization, and Upper Bound Strategy. From these four algorithms, the last one
is highly suitable for large-scale problems. In the cascade optimization strategy,
several optimizers can be used, one followed by another in a specified sequence, to
solve a large-scale problem. In this procedure, the first optimizer starts from a
user-specified design known as “cold-start”. The optimal solution achieved in the
first cascade stage is called a “hot-start” and is used to initiate the second opti-
mization stage. Accordingly, each optimization stage of the cascade procedure
starts from the optimum design achieved in the previous stage. Therefore, each
cascade stage except the first one initiates from a hot-start and produces a new
hot-start for the next stage. In general, the optimization algorithm implemented at
each stage of a cascade process may or may not be the same.
In this book, optimal design of different space structures is performed with
different types of limitations such as strength, buckling, displacement, and natural
frequencies. The considered structures consist of double-layer grids, barrel vaults,
domes, antennas, and steel frames.
This book can be considered as an application of meta-heuristic algorithms to
optimal design of skeletal structures. The present book is addressed to those sci-
entists, engineers, and students who wish to explore the potentials of newly
developed meta-heuristics. The concepts presented in this book are not only
applicable to skeletal structures and finite element models but can equally be used
for designing other systems such as hydraulic and electrical networks. The author
and his graduate students have been involved in various developments and appli-
cations of different meta-heuristic algorithms to structural optimization in the last
two decades. This book contains part of this research suitable for various aspects of
optimization for skeletal structures. This book is likely to be of interest to civil,
mechanical, and electrical engineers who use optimization methods for design, as
well as to those students and researchers in structural optimization who will find it
to be necessary professional reading.
In Chap. 1, a short introduction is provided for the development of optimization
and different meta-heuristic algorithms. Chapter 2 presents an explanation of the
meta-heuristic algorithms utilized in this book. In Chap. 3, optimal design of
well-known structural optimization benchmark problems is discussed. Chapter 4
considers optimum design of large-scale special truss structures. Chapter 5 deals with
optimal design of large-scale double-layer grids. Chapter 6 provides optimal design
of large-scale barrel vaults. Chap. 7 deals with optimal design of dome structures.
Chapter 8 discusses optimal design of steel lattice transmission line towers. Chapter 9
provides optimum seismic design of 3D steel frames. Appendix A provides the
computer codes developed for configuration processing of the structures.
We would like to take this opportunity to acknowledge a deep sense of gratitude
to a number of colleagues and friends who, in different ways, have helped in the
preparation of this book. Professor Ch. Bucher encouraged and supported the first
author to write this book. Our special thanks are due to Ms. Silvia Schilgerius, the
senior editor of the Applied Sciences of Springer, for her constructive comments,
editing, and unfailing kindness in the course of the preparation of this book.
Preface vii
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . ................ 1
1.1 Structural Optimization Using Meta-heuristic Algorithms . . . . . . . 1
1.2 Goals and Organization of the Present Book ................ 3
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ................ 4
2 Optimization Algorithms Utilized in This Book . . . . . . . . . . . . . . . . 7
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Colliding Bodies Optimization Algorithm . . . . . . . . . . . . . . . . . . 7
2.2.1 Theory of Collision Between Two Bodies . . . . . . . . . . . . . 8
2.2.2 Presentation of CBO . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 The Enhanced Colliding Bodies Optimization
Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Vibrating Particles System Algorithm . . . . . . . . . . . . . . . . . . . . . 13
2.4.1 Damped Free Vibration . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4.2 Presentation of VPS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.5 The MDVC-UVPS Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5.1 The Multi-Design Variable Configurations
Cascade Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5.2 The Upper Bound Strategy . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5.3 Presentation of MDVC-UVPS . . . . . . . . . . . . . . . . . . . . . 20
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3 Optimal Design of Usual-Size Skeletal Structures . . . . . . . . . . . . . . . 23
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Numerical Examples with Frequency Constraints . . . . . . . . . . . . . 24
3.2.1 A 72-Bar Space Truss Problem . . . . . . . . . . . . . . . . . . . . 24
3.2.2 A Spatial 120-Bar Dome-Shaped Truss Problem . . . . . . . . 25
3.2.3 A 200-Bar Planar Truss Problem . . . . . . . . . . . . . . . . . . . 28
ix
x Contents
The action of making the best or most effective use of a situation or resource is
called optimization. Optimization problems are studied in different fields and var-
ious steps need to be taken to achieve an optimal solution for a problem. These
steps are as follows: The parameters of the problem, which can be either continuous
or discrete, should be recognized. The objective function(s) and the constraints of
the problem have to be identified. At the end, a suitable optimizer should be chosen
and employed to solve the problem.
Structural optimization is a critical and challenging field that has received
considerable attention by engineers and researchers. These optimization problems
can be classified as follows: (1) obtaining optimal size of structural members (sizing
optimization); (2) finding the optimal form for the structure (shape optimization);
and (3) achieving optimal size and connectivity between structural members
(topology optimization). The main concept of this book is to propose a suitable
optimization technique for size optimization of real-size structures where the search
space has a large size and a great number of design constraints must be controlled.
In the past, the most commonly used optimization techniques were gradient-
based algorithms which utilized gradient information to search the solution space
near an initial starting point [1, 2]. In general, gradient-based methods converge
faster and can obtain solutions with higher accuracy compared to stochastic
approaches. However, the acquisition of gradient information can be either costly or
even impossible to obtain the minima. Moreover, these kinds of algorithms only
guarantee convergence to local minima. Furthermore, a good starting point is quite
vital for a successful execution of these methods. In many optimization problems,
prohibited zones, side limits, and non-smooth or non-convex functions should be
taken into consideration. As a result, these non-convex optimization problems
cannot be solved easily by these methods.
Sizing optimization problems are very popular design problems and can be found
frequently in literature. The contribution of this book is concerned with size opti-
mization of real-size structures with numerous design variables. Various types of
structures with different design constraints are studied. Four efficient optimization
algorithms will be presented and their results are compared to propose a suitable
optimization technique for this class of problems.
The remaining chapters of this book are organized in the following manner:
Chapter 2 presents the rules of the optimization algorithms employed in this
book. These algorithms consist of Colliding Bodies Optimization (CBO) [29],
Enhanced Colliding Bodies Optimization (ECBO) [30], Vibrating Particles System
(VPS) [31], and a hybrid algorithm called MDVC-UVPS [32].
Chapter 3 presents a comparison between the results obtained by proposed
algorithms and some well-known state-of-the-art meta-heuristics for usual-size
skeletal structures. Four truss design examples and two frame design examples with
continuous and discrete sizing variables are studied in this chapter. Different types
of constraints (i.e., natural frequency constraints and strength and displacement
constraints) are considered for benchmarks.
Chapter 4 presents optimal design of three spatial tower trusses. The examples
have 582, 942, and 2386 elements and contain 32, 76, and 220 variables, respec-
tively. All of the structures are designed according to AISC-ASD provisions [33]
and the cross-sectional areas of the bar elements are selected from W-shape profile
list.
Chapter 5 deals with the design optimization of five double-layer grids with
different configurations. The examples contain larger square on square, square on
larger square, square on square, square on diagonal, and diagonal on diagonal
double-layer grids. Strength and slenderness limitations are imposed according to
AISC-LRFD provisions [34].
Chapter 6 deals with size optimization of three double-layer barrel roof struc-
tures. These examples consist of a 384-bar double-layer barrel vault, a 693-bar
double-layer barrel vault, and a 1536-bar double-layer barrel vault. The structures
are subjected to stress, stability, and displacement limitations according to the
provisions of AISC-ASD [31]. The design variables are the cross-sectional areas of
the bar elements which are selected from steel pipe sections.
Chapter 7 examines the abilities of the proposed methods for size optimization
of dome-shaped trusses. Three dome truss design examples with 600, 1180, and
1410 elements are studied in this chapter. Two different constraint cases are con-
sidered for each example (i.e., natural frequency constraints and strength and dis-
placement constraints).
Chapter 8 presents the application of the proposed algorithms for optimization of
steel lattice transmission line towers. Three benchmark structural examples
including 47-bar, 160-bar, and 244-bar power transmission towers are studied in
this chapter. The design variables are the cross-sectional areas of the bar elements
4 1 Introduction
and in all problems, solution candidates are allowed to select discrete values from a
permissible list of cross sections.
Chapter 9 deals with the optimization of 3D steel frames under seismic loads
based on response spectra. Three irregular steel frame problems (i.e., four-story
132-member, four-story 428-member, and twelve-story 276-member steel frames)
are considered to evaluate the performance of the proposed algorithms. The frames
are designed according to the LRFD-AISC design criteria [35]. Load combinations
recommended by ASCE 7-10 [36] are considered and the frames are Intermediate
Moment Frames (IMF).
References
22. Kirkpatrick S, Gelatt CD, Vecchi MP (1983) Optimization by simulated annealing. Science
220(4598):671–680
23. Kaveh A, Talatahari S (2010) A novel heuristic optimization method: charged system search.
Acta Mech 213(3):267–289
24. Kaveh A, Khayatazad M (2012) A new meta-heuristic method: ray optimization. Comput
Struct 112:283–294
25. Kaveh A, Bakhshpoori T (2016) Water evaporation optimization: a novel physically inspired
optimization algorithm. Comput Struct 167:69–85
26. Kaveh A, Dadras A (2017) A novel meta-heuristic optimization algorithm: thermal exchange
optimization. Adv Eng Softw 110:69–84
27. Kaveh A (2017) Advances in metaheuristic algorithms for optimal design of structures, 2nd
edn. Springer International Publishing, Switzerland
28. Kaveh A (2017) Applications of metaheuristic optimization algorithms in civil engineering.
Springer, Switzerland
29. Kaveh A, Mahdavi VR (2014) Colliding bodies optimization: a novel meta-heuristic method.
Comput Struct 139:18–27
30. Kaveh A, Ilchi Ghazaan M (2014) Enhanced colliding bodies optimization for design
problems with continuous and discrete variables. Adv Eng Softw 77:66–75
31. Kaveh A, Ilchi Ghazaan M (2017) Vibrating particles system algorithm for truss optimization
with multiple natural frequency constraints. Acta Mech 228(1):307–322
32. Kaveh A, Ilchi Ghazaan M (2018) A new hybrid meta-heuristic algorithm for optimal design
of large-scale dome structures. Eng Optimiz 50(2):235–252
33. American Institute of Steel Construction (AISC) (1989) Manual of steel construction:
allowable stress design, Chicago, USA
34. American Institute of Steel Construction (AISC) (1994) Manual of steel construction load
resistance factor design, USA
35. AISC 360-10 (2010). Specification for structural steel buildings. American Institute of Steel
Construction, Chicago, Illinois 60601-1802, USA
36. ASCE 7-10 (2010) Minimum design loads for building and other structures
Chapter 2
Optimization Algorithms Utilized
in This Book
2.1 Introduction
The main features and rules of the optimization algorithms utilized in this book are
explained in this chapter. These algorithms consist of Colliding Bodies
Optimization (CBO) [1], Enhanced Colliding Bodies Optimization (ECBO) [2],
Vibrating Particles System (VPS) [3], and a hybrid algorithm called MDVC-UVPS
[4]. All of the algorithms considered here are recently developed and are
multi-agent meta-heuristic methods. These algorithms start with a set of randomly
selected candidate solutions of the optimization problem and according to a series
of simple rules, mainly inspired by the nature, the existing solutions are perturbed
iteratively in order to improve their cost function values. Many other recently
developed meta-heuristic algorithms and their applications can be found in the
recently published books of Kaveh [5, 6].
Collisions between bodies are governed by the laws of momentum and energy.
When a collision occurs in an isolated system between two objects (Fig. 2.1), the
total momentum of the system of objects is conserved. Provided that there are no
net external forces acting upon the objects, the total momentum of the objects
before the collision equals the total momentum of the objects after the collision.
Conservation of total momentum can be expressed by the following equation:
1 1 1 1
m1 v21 þ m2 v22 ¼ m1 v02 02
1 þ m 2 v2 þ Q ð2:2Þ
2 2 2 2
where v1 is the initial velocity of the first object before impact, v2 is the initial
velocity of the second object before impact, v01 is the final velocity of the first object
after impact, v02 is the final velocity of the second object after impact, m1 is the mass
of the first object, m2 is the mass of the second object, and Q is the loss of kinetic
energy due to the impact [7].
where e is the Coefficient Of Restitution (COR) of the two colliding bodies, defined
as the ratio of the relative velocity of separation to the relative velocity of approach:
0
v v0 v0
e¼ 2 1
¼ ð2:5Þ
j v2 v1 j v
According to the coefficient of restitution, there are two special cases of any
collision as follows:
1. A perfectly elastic collision is defined as the one in which there is no loss of kinetic
energy in the collision ðQ ¼ 0 and e ¼ 1Þ. In reality, any macroscopic collision
between objects will convert some kinetic energy to internal energy and other
forms of energy. In this case, after collision, the velocity of separation is high.
2. An inelastic collision is the one in which part of the kinetic energy is changed to
some other forms of energy in the collision. Momentum is conserved in inelastic
collisions (as it is for elastic collisions), but one cannot track the kinetic energy
through the collision since some of it will be converted to other forms of energy.
In this case, the coefficient of restitution is not equal to one ðQ 6¼ 0 and e\1Þ,
and after collision the velocity of separation is low.
For most of the real objects, the value of e is between 0 and 1.
In CBO, each solution candidate is considered as a Colliding Body (CB) and these
massed objects are composed of two main equal groups, i.e., stationary and moving
objects, where the moving objects move to follow stationary objects and a collision
occurs between pairs of objects. This is done for two purposes: (i) to improve the
positions of moving objects; and (ii) to push stationary objects toward better
positions. After the collision, the new positions of colliding bodies are updated
based on their new velocities using the collision laws as discussed in the previous
subsection. The procedure of CBO can be outlined as follows and its pseudocode is
provided in Fig. 2.2.
Step 1 The initial positions of all colliding bodies are determined randomly in
the search space.
10 2 Optimization Algorithms Utilized in This Book
procedure CBO
Initialize algorithm parameters
Initial positions are created randomly
The values of objective function and masses are evaluated
while maximum iterations is not fulfilled
Stationary and moving groups are created
for each CB
The velocity before the collision is evaluated by Eq. (2.7) or Eq. (2.8)
The velocity after the collision is evaluated by Eq. (2.9) or Eq. (2.10)
New location is updated by Eq. (2.12) or Eq. (2.13)
end for
The values of objective function and masses are evaluated
end while
end procedure
where fit(i) represents the objective function value of the ith CB and n is
the number of colliding bodies. In order to select pairs of objects for
collision, CBs are sorted according to their mass in an decreasing order
and they are divided into two equal groups: (i) stationary group and
(ii) moving group (see Fig. 2.3). Moving objects collide with stationary
objects to improve their positions and push stationary objects toward
better positions.
Step 3 The velocity of stationary bodies before collision is zero and moving
objects move toward stationary objects:
vi ¼ 0; i ¼ 1; 2; . . .; n2 ð2:7Þ
vi ¼ xin2 xi ; i ¼ n2 þ 1; n2 þ 2; . . .; n ð2:8Þ
Step 4 The velocities of stationary and moving bodies after the collision (v′i) are
evaluated by
0 ðmi þ n2 þ emi þ n2 Þvi þ n2 n
vi ¼ i ¼ 1; 2; :::; ð2:9Þ
mi þ mi þ n2 2
mi e min vi
0
vi ¼
2
i ¼ n2 þ 1; n2 þ 2; . . .; n ð2:10Þ
mi þ min
2
where iter is the current iteration number and itermax is the total number
of iterations for optimization process.
Step 5 The new position of each stationary CB is
n n
xnew
i ¼ xin2 þ rand v0i ; i ¼ þ 1; þ 2; . . .n ð2:12Þ
2 2
where xnew
i , xi, and v′i are the new position, previous position, and the
velocity after the collision of the ith CB, respectively. rand is a random
vector uniformly distributed in the range of [−1,1] and the sign ‘‘°’’
denotes an element-by-element multiplication. The new position of each
moving CB is calculated by
0
xnew
i ¼ xin2 þ rand vi ; i ¼ n2 þ 1; n2 þ 2; . . .; n ð2:13Þ
In order to get faster and more reliable solutions, Enhanced Colliding Bodies
Optimization (ECBO) is developed [2]. Memory which saves a number of the
best-so-far solutions is used in ECBO to improve the algorithm’s performance.
12 2 Optimization Algorithms Utilized in This Book
procedure ECBO
Initialize algorithm parameters
Initial positions are created randomly
The values of objective function and masses are evaluated
while maximum iterations is not fulfilled
Colliding memory is updated
The population is updated
Stationary and moving groups are created
for each CB
The velocity before the collision is evaluated by Eq. (2.7) or Eq. (2.8)
The velocity after the collision is evaluated by Eq. (2.9) or Eq. (2.10)
New location is updated by Eq. (2.12) or Eq. (2.13)
If rnj < pro
k ← random_int (1,m) /* m is the number of variables */
kth dimension is regenerated randomly in its allowable range
end if
end for
The values of objective function and masses are evaluated
end while
end procedure
Step 6 The velocities of stationary and moving bodies after collision are eval-
uated by Eqs. (2.9) and (2.10), respectively.
Step 7 The new position of each CB is calculated by Eqs. (2.12) and (2.13).
Step 8 To improve the exploration capabilities of the standard CBO and to
prevent premature convergence, a stochastic approach is employed in
ECBO. A parameter like Pro within (0, 1) is introduced to specify
whether a component of each CB must be changed or not. For each
colliding body, Pro is compared with rni (i = 1,2,…,n) which is a ran-
dom number uniformly distributed within (0, 1). If rni < pro, one
dimension of the ith CB is selected randomly and its value is regenerated
by
j j
xij ¼ xmin
j
þ random xmax xmin ð2:14Þ
These forces can be caused by dry friction, or Coulomb friction, between rigid
bodies, by fluid friction when a rigid body moves in a fluid, or by internal friction
between the molecules of a seemingly elastic body. In this section, the free
vibration of single degree of freedom systems with viscous damping is studied. The
viscous damping is caused by fluid friction at low and moderate speeds. Viscous
damping is characterized by the fact that the friction force is directly proportional
and opposite to the velocity of the moving body [8].
The vibrating motion of a body or system of mass m having viscous damping
can be characterized by a block and a spring of constant k as it is shown in Fig. 2.5.
The effect of damping is provided by the dashpot connected to the block, and the
magnitude of the friction force exerted on the plunger by the surrounding fluid is
equal to c_x (c is the coefficient of viscous damping, and its value depends on the
physical properties of the fluid and the construction of the dashpot). If the block is
displaced a distance x from its equilibrium position, the equation of motion can be
expressed as
Before presenting the solutions for this differential equation, the critical damping
coefficient cc is defined as
cc ¼ 2m xn ð2:16Þ
rffiffiffiffi
k
xn ¼ ð2:17Þ
m
c
n¼ ð2:20Þ
2m xn
where q and / are constants generally determined from the initial conditions of the
problem. xD and n are damped natural frequency and damping ratio, respectively.
Equation (2.18) is shown in Fig. 2.6, and the effect of damping ratio on vibratory
motion is illustrated in Fig. 2.7.
In the VPS, the position of each particle is updated by learning from the historically
best position of the entire population, a good particle, and a bad particle. By
controlling the weights of these terms, a proper balance between the diversification
and the intensification inclinations can be achieved. The VPS procedure can be
outlined as follows and its pseudocode is provided in Fig. 2.8.
Step 1 The initial positions of all particles are determined randomly in search
space.
Step 2 The objective function value is calculated for each particle.
Step 3 For each particle, three equilibrium positions with different weights are
defined that the particle tends to approach: (1) the best position achieved
so far across the entire population (HB), (2) a good particle (GP), and
16 2 Optimization Algorithms Utilized in This Book
Fig. 2.7 Free vibration of systems with four levels of damping: a n ¼ 5%, b n ¼ 10%,
c n ¼ 15%, and d n ¼ 20%
procedure VPS
Initialize algorithm parameters
Initial positions are created randomly
The values of objective function are evaluated and HB is stored
while maximum iterations is not fulfilled
for each particle
The GP and BP are chosen
if P<rand
w3=0 and w2=1-w1
end if
for each component
New location is obtained by Eq. (2.22)
end for
Violated components are regenerated by harmony search-based handling
approach
end for
The values of objective function are evaluated and HB is updated
end while
end procedure
(3) a bad particle (BP). In order to select the GP and BP for each
candidate solution, the current population is sorted according to their
objective function values in an increasing order, and then GP and BP are
chosen randomly from the first and second half, respectively.
2.2 Colliding Bodies Optimization Algorithm 17
Figure 2.7 shows the important effect of damping level in the vibration. In order
to model this phenomenon in the optimization algorithm, a descending function that
is proportional to the number of iterations is proposed as follows:
a
iter
D¼ ð2:21Þ
itermax
where iter is the current iteration number and itermax is the total number of iterations
for optimization process. a is a constant and Fig. 2.9 shows the effect of this
parameter on D.
According to the mentioned concepts, the positions are updated by
A ¼ ½w1 :ðHB j xij Þ þ ½w2 :ðGP j xij Þ þ ½w3 :ðBP j xij Þ ð2:23Þ
w1 þ w2 þ w3 ¼ 1 ð2:24Þ
where xji is the jth variable of particle i. w1, w2, and w3 are three parameters to
measure the relative importance of HB, GP, and BP, respectively. Here, rand1,
rand2, and rand3 are random numbers uniformly distributed in the range of [0,1].
The effects of A and D parameters in Eq. (2.22) are similar to that of q and enxn t in
Eq. (2.18), respectively. Also, the value of sinðxD t þ /Þ is considered as unity
(xðtÞ ¼ q enxn t is shown in Fig. 2.6 by dotted line).
A parameter like p within (0, 1) is defined to specify whether the effect of BP
must be considered in updating position or not. For each particle, p is compared
with rand (a random numbers uniformly distributed in the range of [0,1]) and if p <
rand, then w3 = 0 and w2 = 1−w1.
Three essential concepts consisting of self-adaptation, cooperation, and com-
petition are considered in this algorithm. Particles move toward HB so the
self-adaptation is provided. Any particle has the chance to have influence on the
new position of the other one, so the cooperation between the particles is supplied.
Because of the p parameter, the influence of GP (good particle) is more than that of
BP (bad particle), and therefore the competition is provided.
Step 4 The particle moves in the search space to find a better result and may
violate the side constraints. If any component of the system violates a
boundary, it must be regenerated by harmony search-based side con-
straint handling approach [9]. In this technique, there is a possibility like
HMCR (harmony memory considering rate) that specifies whether the
violating component must be changed with the corresponding compo-
nent of the historically best position of a random particle or it should be
determined randomly in the search space. Moreover, if the component of
a historically best position is selected, there is a possibility like PAR
(pitch adjusting rate) that specifies whether this value should be changed
with the neighboring value or not.
Step 5 Steps 2 through 4 are repeated until a termination criterion is fulfilled.
Any terminating condition can be incorporated; however, in this book,
the optimization process is terminated after a fixed number of iterations.
The cascade optimization strategy is proposed to use several optimizers, one fol-
lowed by another in a specified sequence, to solve a large-scale problem [10]. In
this procedure, the first optimizer starts from a user-specified design known as
2.2 Colliding Bodies Optimization Algorithm 19
“cold-start”. The optimal solution achieved in the first cascade stage is called a
“hot-start” and is used to initiate the second optimization stage. Accordingly, each
optimization stage of the cascade procedure starts from the optimum design
achieved in the previous stage. Therefore, each cascade stage except the first one
initiates from a hot-start and produces a new hot-start for the next stage. In general,
the optimization algorithm implemented at each stage of a cascade process may or
may not be the same [11, 12].
In the multi-DVC cascade optimization, a series of appropriate Design Variable
Configurations (DVCs) for the optimization problem under consideration is con-
structed to use a different configuration at each cascade optimization stage. The
coarsest DVC which avoids confusing the employed optimizer with huge design
spaces is utilized in the first stage of the cascade procedure. Therefore, the areas of
appropriate design variable values are identified by detecting near-optimum solu-
tions among the relatively limited design options provided. As the number of design
variables processed in the cascade stages becomes larger, more detailed represen-
tation of the full design space is offered and the optimizer is given the opportunity
to improve the quality of the optimal solution reached. In the final cascade stages
utilizing the finest DVC, relatively small adjustments to an already good-quality
design occur in an effort to identify (or at least approach) the globally optimum
design. In summary, the first optimization stage of the cascade procedure serves the
purpose of basic design space exploration, while the last stages aim at fine-tuning
the achieved optimal solution.
An upper bound strategy is proposed as a simple, yet efficient strategy, to reduce the
total number of structural analyses through avoiding unnecessary analyses during
the course of optimization. The key issue in the UBS is to detect those candidate
designs which have no chance to surpass the best design found so far during the
iterations of the optimum design process. After identifying those non-improving
designs, they are directly excluded from the structural analysis stage, resulting in a
significant saving in the computational effort. The current best design can usually be
considered as the upper bound for the forthcoming candidates to eliminate
unnecessary structural analysis and associated fitness computations for those can-
didates that have no chance of surpassing the best solution. Basically, the key
feature in this approach is to define the penalized weight of the current best solution
found during the previous iterations as an upper bound for the net weight of the
newly generated candidate solutions. Thus, any new candidate solution with a net
weight greater than this upper bound will not be analyzed and this will reduce the
computational burden of the optimization algorithm [13].
20 2 Optimization Algorithms Utilized in This Book
According to the above concepts, the following steps are provided to introduce the
MDVC-UVPS algorithm. Pseudocode of this algorithm is provided in Fig. 2.10.
Step 1 A series of design variable configurations are constructed and sorted
according to the size in an increasing order. The coarsest DVC is utilized
in the first stage of the cascade procedure and the initial positions of all
procedure MDVC-UVPS
Initialize algorithm parameters
A series of design variable configurations are defined
The coarsest DVC is selected and initial positions are created randomly based on it
The values of objective function are evaluated and HB is stored
while maximum iterations is not fulfilled
The upper bound is selected
for each particle
The GP and BP are chosen
if P<rand
w3=0 and w2=1-w1
end if
for each component
New location is obtained by Eq. (2.22)
end for
Violated components are regenerated by harmony search-based handling
approach
Net weight is computed
if net weight > upper bound
The position of particle is replaced to its historically best solution
else
The objective function value is calculated
end if
end for
if the termination criterion of current DVC is fulfilled
The next DVC is used
The population is updated
end for
HB is updated
end while
end procedure
References
10. Patnaik SN, Coroneos RM, Hopkins DA (1997) A cascade optimization strategy for solution
of difficult design problems. Int J Numer Methods Eng 40:2257–2266
11. Charmpis DC, Lagaros ND, Papadrakakis M (2005) Multi-database exploration of large
design spaces in the framework of cascade evolutionary structural sizing optimization.
Comput Methods Appl Mech Eng 194:3315–3330
12. Lagaros ND (2014) A general purpose real-world structural design optimization computing
platform. Struct Multidiscip Optim 49(6):1047–1066
13. Kazemzadeh Azad S, Hasançebi O, Kazemzadeh Azad S (2013) Upper bound strategy for
metaheuristic based design optimization of steel frames. Adv Eng Softw 57:19–32
Chapter 3
Optimal Design of Usual-Size Skeletal
Structures
3.1 Introduction
Sizing optimization of truss and frame structures are frequent structural design
problems that are subjected to various constraints such as displacements, stress,
buckling, and natural frequencies. A great number of papers has been published in
literature, where different meta-heuristic search algorithms have been applied to this
class of problems [1, 2]. The aim of this chapter is to examine the ability of the CBO,
ECBO and VPS which have been utilized in the next chapters for comparison with
MDVC-UVPS. The results of well-known state-of-the-art meta-heuristics are also
provided and compared here. The reason for selecting VPS for hybridization is that
the VPS has shown its superiority compared to CBO and ECBO for optimal design of
many structures, Kaveh [1, 2].
There are many design constraints that should be fulfilled in structural opti-
mization. In the literature, the benchmark design examples usually studied with
displacements, stress, and buckling constraints or displacements and frequencies
constraints. Therefore, we decided to follow the same path in this book. Structural
design examples with frequency constraints are studied in the next section and after
that, benchmark examples with strength constraints are optimized.
The optimization problem can formally be stated as
where {X} is the vector containing the design variables; W({X}) presents the weight of
the structure; ng is the number of design groups; nm(i) is the number of members for the
ith group; and qj and Lj denote the material density and the length of the jth member,
respectively. ximin and ximax are the lower and upper bounds of the design variable xi,
respectively. gj({X}) denotes design constraints, and nc is the number of the constraints.
For constraints handling, a penalty approach is utilized. For this purpose, the
objective function (Eq. (3.1)) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. Here, e1 is
set to unity, and e2 is calculated by
iter
e2 ¼ 1:5 þ 1:5 ð3:3Þ
itermax
where iter is the current iteration number, and itermax is the total number of itera-
tions for optimization process [3].
The schematic of a 72-bar space truss is shown in Fig. 3.1 as the first design
example. The elements are divided into 16 groups, considering the symmetry. The
material density is 2767.99 kg/m3 and the elastic modulus is 68.95 GPa for all the
members. Four nonstructural masses of 2268 kg are attached to the nodes 1 through
4. The allowable minimum cross-sectional area of all elements is set to 0.645 cm2.
This example has two frequency constraints. The first frequency is required to be
f1 = 4 Hz and the third frequency is required to satisfy f3 6 Hz.
The optimized designs found by CSS-BBBC (hybridization of Charged System
Search and Big Bang-Big Crunch with trap recognition capability) [4], HALC-PSO
(transplants a Harmony search-based mechanism to Particle Swarm Optimization
with an Aging Leader and Challengers) [5], TWO (Tug of War Optimization) [6],
CBO [7], EBO [7], and VPS [8] are compared in Table 3.1. The CSS-BBBC
obtained the lightest design; however, the best designs of all methods are
approximately identical. Besides, the elastic modulus of 69.8 GPa was used in [4]
that generally results in lighter structures. The average optimized weight and the
standard deviation on average weight of the VPS are less than those of all other
methods. Frequency constraints are satisfied by all methods (see Table 3.2).
The VPS requires 4720 structural analyses to find the optimum solution while
HALC-PSO, CBO, and ECBO require 8000, 4000, and 14,800 structural analyses,
respectively.
3.2 Numerical Examples with Frequency Constraints 25
Figure 3.2 shows the schematic of a 120-bar dome truss. The members are cate-
gorized into seven groups considering the symmetry. The material density is
7971.810 kg/m3 and the modulus of elasticity is 210 GPa for all elements.
Nonstructural masses are attached to all free nodes as follows: 3000 kg at node 1,
500 kg at nodes 2–13, and 100 kg at the remaining nodes. Element cross-sectional
areas can vary between 1 and 129.3 cm2. The frequency constraints are as
f1 9 Hz and f2 11 Hz
Table 3.2 Natural frequencies (Hz) evaluated at the optimum designs of the 72-bar space truss
problem
Frequency Natural frequencies (Hz)
number CSS-BBBC HALC-PSO TWO CBO ECBO VPS
[4] [5] [6] [7] [7] [8]
1 4.000 4.000 4.000 4.000 4.000 4.0000
2 4.000 4.000 4.000 4.000 4.000 4.0002
3 6.004 6.000 6.000 6.000 6.000 6.0000
4 6.2491 6.230 6.259 6.267 6.246 6.2428
5 8.9726 9.041 9.082 9.101 9.071 9.0698
structures, and it is clear that none of the frequency constraints are violated. The
HALC-PSO, CBO, ECBO, and VPS algorithms obtain the optimal solution after
17,000, 3,700, 7700, and 6860 analyses, respectively.
3.2 Numerical Examples with Frequency Constraints 27
Table 3.3 Performance comparison of the spatial 120-bar dome-shaped truss problem
Design variable Areas (cm2)
CSS-BBBC HALC-PSO CBO [7] ECBO [7] VPS [8]
[4] [5]
1 17.478 19.8905 19.7738 19.8290 19.6836
2 49.076 40.4045 40.6757 41.4037 40.9581
3 12.365 11.2057 11.6056 11.0055 11.3325
4 21.979 21.3768 21.4601 21.2971 21.5387
5 11.190 9.8669 9.8104 9.4718 9.8867
6 12.590 12.7200 12.2866 13.0176 12.7116
7 13.585 15.2236 15.1417 15.2840 14.9330
Weight (kg) 9046.34 8889.96 8890.69 8896.50 8888.74
Average N/A 8900.39 8945.64 8920.16 8896.04
optimized
weight (kg)
Standard N/A 6.38 38.33 20.12 6.65
deviation on
average weight
(kg)
Table 3.4 Natural frequencies (Hz) evaluated at the optimum designs of the spatial 120-bar
dome-shaped truss problem
Frequency number Natural frequencies (Hz)
CSS-BBBC [4] HALC-PSO [5] CBO [7] ECBO [7] VPS [8]
1 9.000 9.000 9.000 9.001 9.0000
2 11.007 11.000 11.000 11.001 11.0000
3 11.018 11.000 11.000 11.003 11.0000
4 11.026 11.010 11.010 11.010 11.0096
5 11.048 11.050 11.049 11.052 11.0491
The last structural optimization problem solved in this class is the optimal design of
a 200-bar planar truss schematized in Fig. 3.3. Due to the symmetry, the elements
are divided into 29 groups. The modulus of elasticity and the material density of
members are 210 GPa and 7860 kg/m3, respectively. Nonstructural masses of
100 kg are attached to the upper nodes. A lower bound of 0.1 cm2 is assumed for
the cross-sectional areas. The first three natural frequencies of the structure must
satisfy the following limitations (f1 5 Hz, f2 10 Hz, f3 15 Hz).
Table 3.5 presents the results of the optimal designs utilizing CSS-BBBC [4],
HALC-PSO [5], CBO-PSO (a hybrid of CBO and PSO algorithms) [9], CBO [7],
ECBO [7], and VPS [3]. The weight of the best result obtained by VPS is
3.2 Numerical Examples with Frequency Constraints 29
2156.62 kg that is the best among the compared methods. The average optimized
weight of the HALC-PSO is 2157.14 kg, which is less than those of all other
methods. Table 3.6 reports the natural frequencies of the optimized structures, and
Table 3.5 Performance comparison of the 200-bar planar truss problem
30
Table 3.6 Natural frequencies (Hz) evaluated at the optimum designs of the 200-bar planar truss
problem
Frequency number Natural frequencies (Hz)
CSS-BBBC [4] HALC-PSO [5] CBO-PSO CBO [7] ECBO [7] VPS [3]
1 5.010 5.000 5.003 5.000 5.000 5.0000
2 12.911 12.254 12.281 12.221 12.189 12.2086
3 15.416 15.044 15.125 15.088 15.048 15.0153
4 17.033 16.718 16.613 16.759 16.643 16.6946
5 21.426 21.461 21.331 21.419 21.342 21.4046
Fig. 3.4 Convergence curves for the 200-bar planar truss problem
it is clear that none of the frequency constraints are violated. Comparison of the
convergence rates between CBO, ECBO, and VPS is illustrated in Fig. 3.4.
The VPS requires 16,420 structural analyses to find the optimum solution while
HALC-PSO, CBO-PSO, CBO, and ECBO require 13,000, 9000, 10,500, and
14,700 structural analyses, respectively. It should be noted that the designs found
by VPS at 9,000th, 10,500th, 13,000th, and 14,700th analyses are 2158.35,
2158.06, 2157.74, and 2157.72 kg, respectively.
The schematic and element grouping of a spatial 120-bar dome truss are shown in
Fig. 3.5. This structure is divided into seven groups of elements because of sym-
metry (for the sake of clarity, not all the element groups are numbered in Fig. 3.5).
3.3 Numerical Examples with Strength Constraints 33
The modulus of elasticity is 30,450 ksi (210 GPa) and the material density is
0.288 lb/in.3 (7971.810 kg/m3). The yield stress of steel is taken as 58.0 ksi
(400 MPa). The dome is considered to be subjected to vertical loading at all the
unsupported joints. These loads are taken as −13.49 kips (−60 kN) at node 1,
−6.744 kips (−30 kN) at nodes 2 through 14, and −2.248 kips (−10 kN) at the
remaining nodes. Element cross-sectional areas can vary between 0.775 in.2
34 3 Optimum Design of Usual-Size Skeletal Structures
(5 cm2) and 20.0 in.2 (129.032 cm2). Displacement limitations of ±0.1969 in.
(± 5 mm) are imposed on all nodes in x, y, and z coordinate directions. Constraints
on member stresses are imposed according to the provisions of the AISC [10] as
follows:
The allowable tensile stresses for tension members are calculated as
where E is the modulus of elasticity; ki is the slenderness ratio ðki ¼ kli =ri Þ; Cc
denotes the slenderness ratio dividing the elastic and inelastic buckling regions
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Cc ¼ 2p2 E=Fy ; k is the effective length factor (k is set equal to 1 for all truss
members); Li is the member length; and ri is the minimum radius of gyration.
This truss is optimized by MSPSO (Multistage Particle Swarm Optimization)
[11], TWO (Tug of War Optimization) [12], WEO (Water Evaporation
Optimization) [13], CBO, ECBO, and VPS [3]. Comparison of the optimal designs
is given in Table 3.7. It can be seen that the lightest design (i.e., 33,249.98 lb) and
the best average optimized weight (i.e., 33,253.56 lb) are found by the VPS method.
Figure 3.6 compares the convergence curves of the best results obtained by the
CBO, ECBO, and VPS. The VPS converges to the optimum solution after 8280
analyses. The MSPSO, TWO, WEO, CBO, and ECBO obtain the optimal solution
after 15,000, 16,000, 19,510, 12,080, and 19,800 analyses, respectively.
The schematic of a 3-bay 15-story frame is represented in Fig. 3.7. The applied loads
and the numbering of the member groups are also shown in this figure. The modulus
of elasticity is 29 Msi (200 GPa) and the yield stress is 36 ksi (248.2 MPa).
The effective length factors of the members are calculated as kx 0 for a
sway-permitted frame and the out-of-plane effective length factor is specified as
ky= 1.0. Each column is considered as non-braced along its length, and the
non-braced length for each beam member is specified as one-fifth of the span length.
Limitation on displacement and strength are imposed according to the provisions of
the AISC [14] as follows:
Table 3.7 Performance comparison of the spatial 120-bar dome-shaped truss problem
Element group Optimal cross-sectional areas (in.2)
MSPSO [11] TWO [12] WEO [13] CBO ECBO VPS [3]
1 3.0244 3.0247 3.0243 3.0260 3.0234 3.0244
2 14.7804 14.7261 14.7943 14.8237 14.8569 14.7536
3 5.0567 5.1338 5.0618 5.1576 4.8649 5.0789
4 3.1359 3.1369 3.1358 3.1310 3.1319 3.1371
5 8.4830 8.4545 8.4870 8.3350 8.5716 8.4829
3.3 Numerical Examples with Strength Constraints
Fig. 3.6 Convergence curves for the spatial 120-bar dome-shaped truss problem
DT
R0 ð3:6Þ
H
di
RI 0; i ¼ 1; 2; . . .; ns ð3:7Þ
hi
where di is the inter-story drift, hi is the story height of the ith floor, ns is the total
number of stories, and RI is the inter-story drift index (1/300).
(c) Strength constraints
(
2/c Pn þ /b Mn 1 0; for /PPu n \0:2
Pu Mu
c
ð3:8Þ
/c Pn þ 9/b Mn 1 0; for /PPu n 0:2
Pu 8Mu
c
Pn ¼ Ag : Fy ð3:9Þ
Pn ¼ Ag : Fcr ð3:10Þ
where
8
< Fcr ¼ 0:658k2c Fy ; for kc 1:5
ð3:11Þ
: Fcr ¼ 0:877 Fy ; for kc [ 1:5
k2
c
rffiffiffiffiffi
kl Fy
kc ¼ ð3:12Þ
rp E
where Ag is the cross-sectional area of a member, and k is the effective length factor
that is calculated by Dumonteil [15]:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1:6GA GB þ 4:0ðGA þ GB Þ þ 7:5
k¼ ð3:13Þ
GA þ GB þ 7:5
where GA and GB are stiffness ratios of columns and girders at the two end joints A
and B of the column section, respectively.
Additionally, in this example, the sway of the top story is limited to 9.25 in.
(23.5 cm).
Table 3.8 presents the comparison of the results of different algorithms.
The VPS algorithm yields the least weight for this example, which is 86,985 lb. The
other design weights are 92,723 lb by CSS [16], 93,315 lb by ES-DE (Eagle
Strategy with Differential Evolution) [17], 91,248 lb by DSOS (Discrete Symbiotic
Organisms Search) [18], 93,795 lb by CBO [19] and 86,986 lb by ECBO [19]. The
best design of VPS has been achieved in 19,600 analyses. The CSS, ES-DE, CBO,
and ECBO require 5,000, 10,000, 9520, and 9000 structural analyses to find the
optimum solutions, respectively. It should be noted that the proposed method
achieved about 92,000 lb (the best weight among the other methods except ECBO)
after 10,800 structural analyses. Element stress ratio and inter-story drift evaluated
at the best design optimized by VPS are shown in Fig. 3.8. The maximum stress
ratio is 99.88% and the maximum inter-story drift is 45.41.
3.3 Numerical Examples with Strength Constraints 39
Fig. 3.8 Constraint margins for the best design obtained by VPS for the 3-bay 15-story frame
problem: a element stress ratio and b inter-story drift
40 3 Optimum Design of Usual-Size Skeletal Structures
The last structural optimization problem solved in this chapter is the weight min-
imization of a 3-bay 24-story frame shown in Fig. 3.9. Frame members are col-
lected in 20 groups (16 column groups and 4 beam groups). Each of the four beam
element groups is chosen from all 267 W shapes, while the 16 column element
groups are limited to W14 sections. The material has a modulus of elasticity equal
to E = 29.732 Msi (205 GPa) and a yield stress of fy = 33.4 ksi (230.3 MPa). The
effective length factors of the members are calculated as kx 0 for a
sway-permitted frame and the out-of-plane effective length factor is specified as
ky = 1.0. All columns and beams are considered as non-braced along their lengths.
Similar to the previous example, the frame is designed following the LRFD-AISC
specification and uses an inter-story drift displacement constraint (AISC [14]).
This steel frame structure was optimized by CSS [16], ES-DE [17] and DSOS
[18], CBO [19], ECBO [19], and VPS [3]. Table 3.9 presents a comparison between
these results. The lightest design (i.e., 201,618 lb) is found by ECBO algorithm and
after that, the best design belongs to VPS (i.e., 202,998 lb). Figure 3.10 shows the
convergence curves of the best results found by CBO, ECBO, and VPS. The best
design has been achieved at 16,220 analyses by VPS and it obtained a weight of
209,532 lb after 8800 analyses, which is the best result compared to the
weight achieved by the other method. The CSS, ES-DE, DSOS, CBO, and ECBO
get the optimal solution after 5500, 12,500, 7500, 8280, and 15,360 analyses,
respectively.
In this chapter, the CBO, ECBO, and VPS algorithms are examined in the context
of size optimization of skeletal structure designed for minimum weight. Four
trusses and two frames subjected to different constraints are employed, and their
final results are compared with results of the state-of-the-art algorithms from lit-
erature. The VPS achieved the lightest designs for most of the benchmarks and after
that, ECBO performed better among the compared methods. The small values of the
standard deviation on average weights prove the robustness of these techniques.
Moreover, they show appropriate convergence rates. Although the results found by
CBO are not better than ECBO and VPS but it has no internal parameter to adjust.
Therefore, it can be performed more easily on various problems. Generally, com-
parison of the results proves the efficiency of the proposed algorithms and shows
that they are suitable methods for comparison with the MDVC-UVPS.
3.4 Concluding Remarks 41
Fig. 3.10 Convergence curves for the 3-bay 24-story frame problem
References
1. Kaveh A (2017) Advances in metaheuristic algorithms for optimal design of structures, 2nd
edn. Springer, Switzerland
2. Kaveh A (2017) Applications of metaheuristic optimization algorithms in civil engineering.
Springer, Switzerland
3. Kaveh A, Ilchi Ghazaan M (2017) A new meta-heuristic algorithm: vibrating particles system.
Sci Iranica, Trans A, Civil Eng 24(2):551–566
4. Kaveh A, Zolghadr A (2012) Truss optimization with natural frequency constraints using a
hybridized CSS–BBBC algorithm with trap recognition capability. Comput Struct
102–103:14–27
5. Kaveh A, Ilchi Ghazaan M (2015) Hybridized optimization algorithms for design of trusses
with multiple natural frequency constraints. Adv Eng Softw 79:137–147
6. Kaveh A, Zolghadr A (2017) Truss shape and size optimization with frequency constraints
using Tug of war optimization. Asian J Civil Eng 7(2):311–333
7. Kaveh A, Ilchi Ghazaan M (2014) Enhanced colliding bodies algorithm for truss optimization
with frequency constraints. J Comput Civil Eng 29(6)
8. Kaveh A, Ilchi Ghazaan M (2017) Vibrating particles system algorithm for truss optimization
with multiple natural frequency constraints. Acta Mech 228(1):307–322
9. Kaveh A, Mahdavi VR (2015) A hybrid CBO–PSO algorithm for optimal design of truss
structures with dynamic constraints. Appl Soft Comput 34:260–273
10. American Institute of Steel Construction (AISC) (1989) Manual of steel construction:
allowable stress design
11. Talatahari S, Kheirollahi M, Farahmandpour C, Gandomi AH (2013) A multi-stage particle
swarm for optimum design of truss structures. Neural Comput Appl 23:1297–1309
12. Kaveh A, Zolghadr A (2016) A novel metaheuristic algorithm: tug of war optimization. Int J
Optim Civil Eng 6(4):469–492
13. Kaveh A, Bakhshpoori T (2016) Water evaporation optimization: a novel physically inspired
optimization algorithm. Comput Struct 167:69–85
14. American Institute of Steel Construction (AISC) (2001) Manual of steel construction: load
and resistance factor design
15. Dumonteil P (1992) Simple equations for effective length factors. Eng J AISC 29(3):111–115
44 3 Optimum Design of Usual-Size Skeletal Structures
16. Kaveh A, Talatahari S (2012) Charged system search for optimal design of frame structures.
Appl Soft Comput 12:382–393
17. Talatahari S, Gandomi AH, Yang XS, Deb S (2015) Optimum design of frame structures
using the eagle strategy with differential evolution. Eng Struct 91:16–25
18. Talatahari S (2016) Symbiotic organisms search for optimum design of and grillage system.
Asian J Civil Eng 17(3):299–313
19. Kaveh A, Ilchi Ghazaan M (2015) A comparative study of CBO and ECBO for optimal
design of skeletal structures. Comput Struct 153:137–147
Chapter 4
Optimal Design of Large-Scale Special
Truss Structures
4.1 Introduction
where {X} is a vector containing the design variables; W({X}) presents the weight
of the structure; ng is the number of design groups; nm(i) is the number of members
for the ith group; and qj and Lj denote the material density and the length of the jth
member, respectively. ximin and ximax are the lower and upper bounds of the design
variable xi, respectively. gj({X}) denotes the set of design constraints, and nc is the
number of constraints.
For constraint handling, a penalty approach is utilized. For this purpose, the
objective function (Eq. 4.1) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. Here, e1 is
set to unity, and e2 is calculated by
iter
e2 ¼ 1:5 þ 1:5 ð4:3Þ
itermax
where iter is the current iteration number, and itermax is the total number of iter-
ations for the optimization process.
The constraint conditions for truss structures studied here are briefly explained in
the following. Limitations on stress and stability of truss elements are imposed
according to the provisions of the ASD-AISC [17] as follows.
The allowable tensile stress for tension members are calculated by
where E is the modulus of elasticity; ki is the slenderness ratio ðki ¼ kli =ri Þ; Cc
denotes the slenderness ratio dividing the elastic and inelastic buckling regions
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Cc ¼ 2p2 E=Fy ; k is the effective length factor (k is set to 1 for all truss
members); li is the member length; and ri is the minimum radius of gyration.
According to the provisions of ASD-AISC, the maximum slenderness ratios are
limited to 300 and 200 for tension and compression members, respectively. Nodal
displacements in all coordinate directions must be less than ±3.15 in.
(i.e., ±8 cm).
In this section, three tower truss optimization benchmark problems are optimized
using CBO, ECBO, VPS, and MDVC-UVPS algorithms. These optimization
examples are as follows:
• A spatial 582-bar tower truss
• A spatial 942-bar tower truss
• A spatial 2386-bar tower truss.
The schematic of the 582-bar tower truss is shown in Fig. 4.1. The members of the
truss are grouped into 32 independent sizing variables considering its symmetry
about x- and y-axes. A single load case is considered consisting of lateral loads of
1.12 kips (5.0 kN) applied in both x- and y-directions and vertical loads of
−6.74 kips (−30 kN) applied in z-direction to all free nodes of the tower. A discrete
set of standard steel sections selected from the list of W-shape profiles is considered
for the sizing variables based on the cross-sectional areas and radii of gyration.
Cross-sectional areas of the elements can vary between 6.16 and 215 in.2 (i.e.,
between 39.74 and 1387.09 cm2). Limitations on stress and stability of truss ele-
ments and nodal displacements are defined in Sect. 4.2.
This problem is optimized in three stages by MDVC-UVPS. The number of
design variables in stages 1, 2, and 3 is 8, 15, and 32, respectively. Table 4.1
presents the design variables configurations. Table 4.2 presents the results obtained
by CBO [18], ECBO [18], VPS, and MDVC-UVPS. The best design obtained by
MDVC-UVPS is better than those of the other methods (1,295,038 in.3). The best
volumes found by CBO, ECBO, and VPS are 1,334,994, 1,296,776, and
1,304,569 in.3, respectively. MDVC-UVPS is the most robust optimizer, achieving
the lowest average volume over the independent optimization runs. The CBO,
48 4 Optimum Design of Large-Scale Special Truss Structures
Table 4.1 Design variable configurations utilized for the 582-bar tower problem
Number of design Design variables in the group (design variable
variables in stages configurations)
Stage 1 8 [1 6 9]; [2 4 7 10]; [3 5 8 11]; [12 13 14]; [19 22 25 28
31]; [32]; [15 17 20 23 26 29]; [16 18 21 24 27 30]
Stage 2 15 [1 6 9]; [2 4]; [7 10]; [3 5]; [8 11]; [12]; [13]; [14]; [19 22
25]; [28 31]; [32]; [15 17 20]; [23 26 29]; [16 18 21]; [24
27 30]
ECBO, VPS, and MDVC-UVPS algorithms obtained the optimal solutions after
17,700, 19,700, 17,540, and 15,480 analyses, respectively. Stress ratios and nodal
displacements in all directions evaluated for the best design optimized by
MDVC-UVPS are shown in Fig. 4.2. The maximum stress ratio and the maximum
nodal displacement obtained by MDVC-UVPS are 99.88% and 3.1493 in.
respectively.
4.3 Design Examples 49
Table 4.2 Performance comparison for the spatial 582-bar tower truss problem
Element group Optimal W-shaped sections
CBO [18] ECBO [18] VPS MDVC-UVPS
1 W8 21 W8 21 W8 21 W8 21
2 W14 82 W14 90 W14 90 W14 90
3 W8 28 W8 24 W8 24 W8 24
4 W12 50 W14 61 W21 62 W14 61
5 W8 24 W8 24 W8 24 W8 24
6 W8 21 W8 21 W8 21 W8 21
7 W12 53 W10 49 W10 39 W10 45
8 W12 26 W8 24 W8 24 W8 24
9 W8 21 W8 21 W8 24 W8 21
10 W14 48 W14 43 W10 33 W14 43
11 W8 24 W8 24 W8 24 W8 24
12 W14 61 W12 72 W14 74 W10 68
13 W14 82 W12 72 W10 77 W12 72
14 W12 50 W10 54 W10 49 W10 49
15 W14 74 W12 65 W10 77 W14 82
16 W8 40 W8 31 W8 31 W8 31
17 W12 53 W10 60 W21 62 W21 62
18 W6 25 W8 24 W8 24 W8 24
19 W8 21 W8 21 W10 22 W8 21
20 W8 40 W14 43 W10 49 W14 38
21 W8 24 W8 24 W8 24 W8 24
22 W8 21 W8 21 W8 21 W8 21
23 W12 26 W8 21 W10 22 W6 25
24 W12 26 W8 24 W8 24 W8 24
25 W10 22 W8 21 W8 21 W8 21
26 W10 22 W8 21 W8 21 W8 21
27 W6 25 W8 24 W8 24 W8 24
28 W8 21 W8 21 W8 21 W8 21
29 W8 21 W8 21 W10 22 W8 21
30 W8 24 W8 24 W8 24 W8 24
31 W8 21 W8 21 W10 22 W8 21
32 W6 25 W8 24 W8 24 W8 24
Volume (in.3) 1,334,994 1,296,776 1,304,569 1,295,038
Average optimized volume 1,345,429 1,306,728 1,324,086 1,302,422
(in.3)
Standard deviation on 9116 7536 12,218 4347
average volume (in.3)
50 4 Optimum Design of Large-Scale Special Truss Structures
Fig. 4.2 Constraint margins for the best design obtained by MDVC-UVPS algorithm for the
582-bar tower truss problem: a element stress ratio and b nodal displacements
Figure 4.3 shows the schematic of a 942-bar tower truss. This example has been
investigated by many researchers considering 59 design variables [19]. In this
study, the design variables are increased to 76, and the performance constraints,
material properties, and other conditions are the same as those of the first example.
Figure 4.4 shows the member groups. Three stages with 16, 28, and 76 design
variables are considered to solve this problem using MDVC-UVPS. The design
variable configurations are shown in Table 4.3.
The optimized designs found by different algorithms are compared in Table 4.4.
The volume of the best result obtained by the hybrid algorithm is 3,263,387 in.3
that is the best among the compared methods. The average optimized volume and
the standard deviation on average volume obtained by MDVC-UVPS are less than
those of all other considered methods. This algorithm requires 14,587 structural
analyses to find the optimum solution, while CBO [20], ECBO [20], and VPS [21]
require 29,600, 19,960, and 26,180 structural analyses, respectively. The amount of
saving in structural analyses in each iteration of the MDVC-UVPS is shown in
Fig. 4.5.
The schematic of a 2386-bar tower truss is shown in Fig. 4.6. This example is
studied here for the first time. The performance constraints, material properties, and
other conditions are the same as those of the first example. The elements are divided
4.3 Design Examples 51
into 220 groups and member groups are presented in Fig. 4.7. Four stages are
considered to optimize this example using MDVC-UVPS. The number of design
variable in stages 1, 2, 3, and 4 is 21, 42, 84, and 220, respectively. Table 4.5 lists
the design variable configurations.
Table 4.6 presents the optimum designs obtained by the proposed algorithms.
The lightest design (i.e., 12,165,572 in.3) is achieved by MDVC-UVPS algorithm
52 4 Optimum Design of Large-Scale Special Truss Structures
Table 4.3 Design variable configurations utilized for the 942-bar tower problem
Number of design Design variables in the group (design variable
variables in stages configurations)
Stage 16 [1]; [2–6]; [7–12]; [13–18]; [19–24]; [25–29]; [30–35];
1 [36]; [37–43]; [44–51]; [52–58]; [59]; [60–64]; [65–70];
[71–75]; [76]
Stage 28 [1]; [2 3]; [4–6]; [7–9]; [10–12]; [13–15]; [16–18]; [19–
2 21]; [22–24]; [25 26]; [27–29]; [30–32]; [33–35]; [36];
[37–39]; [40–43]; [44–47]; [48–51]; [52–54]; [55–58];
[59]; [60 61]; [62–64]; [65–67]; [68–70]; [71 72]; [73–
75]; [76]
Table 4.4 Performance comparison for the spatial 942-bar tower truss problem
Element group Optimal W-shaped sections
CBO [20] ECBO [20] VPS [21] MDVC-UVPS
1 W14 145 W12 190 W12 170 W12 170
2 W36 280 W36 230 W36 260 W14 257
3 W24 250 W40 199 W44 262 W24 279
4 W14 257 W24 229 W30 235 W33 241
5 W33 241 W36 150 W36 245 W40 249
6 W44 262 W30 173 W24 229 W14 211
7 W30 211 W24 250 W40 199 W24 192
8 W33 201 W27 258 W14 193 W24 192
9 W24 176 W14 159 W40 174 W14 176
10 W24 162 W30 191 W24 162 W40 174
11 W21 147 W18 158 W14 145 W12 136
12 W12 136 W18 119 W18 119 W24 131
13 W33 221 W24 250 W12 279 W36 230
14 W6 25 W14 30 W8 21 W8 24
15 W10 54 W8 21 W10 22 W8 21
16 W8 21 W8 21 W12 26 W10 22
17 W10 22 W8 21 W10 22 W8 21
18 W8 21 W8 21 W10 22 W10 22
19 W10 22 W8 21 W10 22 W8 21
20 W8 21 W8 21 W10 22 W10 22
21 W8 21 W8 21 W6 25 W8 21
22 W10 22 W8 21 W8 24 W8 21
23 W10 22 W8 21 W10 22 W8 21
24 W14 145 W24 117 W14 145 W21 147
25 W14 34 W12 50 W8 31 W12 30
26 W8 24 W14 30 W8 24 W6 25
27 W8 24 W10 33 W8 24 W6 25
(continued)
4.3 Design Examples 53
Fig. 4.5 Saving in structural analyses using the MDVC-UVPS algorithm for the 942-bar tower
truss problem
after 13,385 analyses. The best design obtained by the CBO [20], ECBO [20], and
VPS [21] is 15,587,709 in.3, 14,086,857 in.3, and 12,989,713 in.3, respectively.
These values are found after 29,970, 29,670, and 29,980 analyses. MDVC-UVPS is
also the most robust optimizer, achieving the lowest average design over the
independent optimization runs. Convergence history diagrams are depicted in
4.3 Design Examples 55
Fig. 4.8 and demonstrate that the intermediate designs found by MDVC-UVPS are
always better than those found by the other considered algorithms.
Truss structures are often designed to carry multiple loading conditions under static
constraints on the nodal displacements, stresses in the members, and critical
buckling loads. This class of problems is highly nonlinear and also widely studied
in the field of practical structural engineering. In this chapter, three tower trusses
with a large number of design variables are studied to test and verify the efficiency
56 4 Optimum Design of Large-Scale Special Truss Structures
Table 4.5 Design variable configurations utilized for the 2386-bar tower problem
Number of design Design variables in the group (design variable
variables in stages configurations)
Stage 21 [1–10]; [11–20]; [21–31]; [32–41]; [42–51]; [52–62];
1 [63–72]; [73–82]; [83–92] [93–103]; [104–113]; [114–
124]; [125–135]; [136–146]; [147–156]; [157–167] [168–
178]; [179–188]; [189–199]; [200–210]; [211–220]
Stage 42 [1:4]; [5:10]; [11:15]; [16:20]; [21:25]; [26:31]; [32:36];
2 [37:41]; [42:46]; [47:51]; [52:56]; [57:62]; [63:67];
[68:72]; [73:77]; [78:82]; [83:87]; [88:92]; [93:97];
[98:103]; [104:108]; [109:113]; [114:118]; [119:124];
[125:129]; [130:135]; [136:140]; [141:146]; [147:151];
[152:156]; [157:161]; [162:167]; [168:172]; [173:178];
[179:183]; [184:188]; [189:193]; [194:199]; [200:204];
[205:210]; [211:215]; [216:220]
Stage 84 [1]; [2–4]; [5–7]; [8–10]; [11 12]; [13–15]; [16 17]; [18–
3 20]; [21 22]; [23–25]; [26–28]; [29–31]; [32 33]; [34–36];
[37 38]; [39–41]; [42 43]; [44–46]; [47 48]; [49–51]; [52
53]; [54–56]; [57–59]; [60–62]; [63 64]; [65–67]; [68 69];
[70–72]; [73 74]; [75–77]; [78 79]; [80–82]; [83 84]; [85–
87]; [88 89]; [90–92]; [93 94]; [95–97]; [98–100]; [101–
103]; [104 105]; [106–108]; [109 110]; [111–113]; [114
115]; [116–118]; [119–121]; [122–124]; [125 126]; [127–
129]; [130–132]; [133–135]; [136 137]; [138–140]; [141–
143]; [144–146]; [147 148]; [149–151]; [152 153]; [154–
156]; [157 158]; [159–161]; [162–164]; [165–167]; [168
169]; [170–172]; [173–175]; [176–178]; [179 180]; [181–
183]; [184 185]; [186–188]; [189 190]; [191–193]; [194–
196]; [197–199]; [200 201]; [202–204]; [205–207]; [208–
210]; [211 212]; [213–215]; [216 217]; [218–220]
Table 4.6 Performance comparison for the spatial 23862-bar tower truss problem
Element group Optimal W-shaped sections
CBO [20] ECBO [20] VPS [21] MDVC-UVPS
1 W14 605 W14 730 W14 665 W14 730
2 W14 730 W14 730 W14 605 W14 730
3 W14 730 W14 730 W14 665 W14 730
4 W14 605 W14 665 W14 665 W14 730
5 W14 730 W14 730 W14 605 W14 730
6 W14 605 W14 730 W14 665 W14 730
7 W40 249 W14 730 W14 665 W14 730
8 W14 730 W40 215 W14 665 W14 730
9 W14 665 W14 665 W14 605 W14 730
10 W14 665 W14 500 W14 665 W14 730
11 W14 665 W12 279 W14 665 W14 455
12 W27 194 W33 318 W14 426 W14 455
13 W27 194 W14 605 W14 665 W14 455
(continued)
4.4 Concluding Remarks 57
Fig. 4.8 Convergence curves for the 2386-bar tower truss problem
of MDVC-UVPS. In the first example, the design found by ECBO, VPS, and
MDVC-UVPS is approximately identical and is about 3% lighter than the result
obtained by the CBO. In the second one, the design obtained by MDVC-UVPS is
4.5, 3.5, and 1% lighter than the best designs achieved by CBO, ECBO, and VPS,
respectively. These values are 28, 15, and 6.5% for the last example. The average
optimized design and the standard deviation on average design of the hybrid
algorithm are less than those of all other compared methods for all of the examples.
Moreover, this algorithm comes close to the optimum design rapidly.
References 63
References
1. Lee KS, Geem ZW (2004) A new structural optimization method based on the harmony
search algorithm. Comput Struct 82:781–798
2. Camp CV (2007) Design of space trusses using Big Bang-Big Crunch optimization. J Struct
Eng 133:999–1008
3. Rahami H, Kaveh A, Gholipour Y (2008) Sizing, geometry and topology optimization of
trusses via force method and genetic algorithm. Eng Struct 30:2360–2369
4. Lamberti L (2008) An efficient simulated annealing algorithm for design optimization of truss
structures. Comput Struct 86:1936–1953
5. Kaveh A, Talatahari S (2009) A particle swarm ant colony optimization for truss structures
with discrete variables. J Constr Steel Res 65:1558–1568
6. Li LJ, Huang ZB, Liu F (2009) A heuristic particle swarm optimization method for truss
structures with discrete variables. Comput Struct 87:435–443
7. Wu CY, Tseng KY (2010) Truss structure optimization using adaptive multi-population
differential evolution. Struct Multidiscip Optim 42(4):575–590
8. Dede T, Bekiroglu S, Ayvaz Y (2011) Weight minimization of trusses with genetic algorithm.
Appl Soft Comput 11:2565–2575
9. Gandomi AH, Talatahari S, Yang XS, Deb S (2013) Design optimization of truss structures
using cuckoo search algorithm. Struct Des Tall Spec 22(17):1330–1349
10. Kaveh A, Mahdavi VR (2014) Colliding bodies optimization method for optimum discrete
design of truss structures. Comput Struct 139:43–53
11. Kaveh A, Ilchi Ghazaan M (2014) Enhanced colliding bodies optimization for design
problems with continuous and discrete variables. Adv Eng Softw 77:66–75
12. Ahrari A, Atai AA, Deb K (2014) Simultaneous topology, shape and size optimization of
truss structures by fully stressed design based on evolution strategy. Eng Optimiz 47(8):37–41
13. Hasançebi O, Kazemzadeh Azad S (2015) Adaptive dimensional search: a new metaheuristic
algorithm for discrete truss sizing optimization. Comput Struct 154:1–16
14. Bekdaş G, Nigdeli SM, Yang XS (2015) Sizing optimization of truss structures using flower
pollination algorithm. Appl Soft Comput 37:322–331
15. Ho-Huu V, Nguyen-Thoi T, Vo-Duy T, Nguyen-Trang T (2016) An adaptive elitist
differential evolution for optimization of truss structures with discrete design variables.
Comput Struct 165:59–75
16. Kaveh A, Bakhshpoori T (2016) A new metaheuristic for continuous structural optimization:
water evaporation optimization. Struct Multidiscip Optim 54(1):23–43
17. American Institute of Steel Construction (AISC) (1989) Manual of steel construction:
allowable stress design, Chigago, USA
18. Kaveh A, Ilchi Ghazaan M (2014) Enhanced colliding bodies optimization for design
problems with continuous and discrete variables. Adv Eng Softw 77:66–75
19. Hasançebi O (2008) Adaptive evolution strategies in structural optimization: enhancing their
computational performance with applications to large-scale structures. Comput Struct
86:119–132
20. Kaveh A, Ilchi Ghazaan M (2016) Optimum design of large-scale truss towers using cascade
optimization. Acta Mech 227(9):2645–2656
21. Kaveh A, Ilchi Ghazaan M (2017) MATLAB codes for vibrating particles system algorithm.
Int J Optim Civil Eng 7(3):355–366
Chapter 5
Optimal Design of Double-Layer Grids
5.1 Introduction
Double-layer grids belong to the category of space structures and consist of two
planar networks of members forming the top and bottom layers parallel to each
other and interconnected by vertical and inclined web members. Double-layer grids
are characterized by ball joints with no moment or torsional resistance; therefore, all
members can only resist tension or compression. Even in the case of connection by
comparatively rigid joints, the influence of bending or torsional moment is
insignificant. A wide variety of double-layer grids can be formed either by altering
the direction of the top and bottom layers with respect to each other or by changing
the relative positions of the nodal points of the top and bottom layers. Additional
variations can also be introduced by utilizing layers of different sizes [1].
Double-layer grids are ideally suited for covering exhibition pavilions, assembly
halls, swimming pools, hangars, churches, bridge decks, and many types of
industrial buildings in which large unobstructed areas are required.
In the past decades, a number of meta-heuristic algorithms have been developed
and used for structural optimization problems, e.g., see Kaveh [2]. Double-layer
grids have a great number of structural elements, and therefore optimization tech-
niques can be rewardingly employed to achieve economic and efficient designs of
them. Here, five different types of double-layer grids are studied and optimized
utilizing the colliding bodies optimization (CBO) [3], enhanced colliding bodies
optimization (ECBO) [4], vibrating particles system (VPS) [5], and a hybrid
algorithm called MDVC-UVPS [6]. The cross-section areas of the grid elements are
considered as discrete design variables and all of them are selected from a list of
tube sections available in AISC-LRFD [7]. Strength, stability, and displacement
constraints are considered for each example.
Here, the aim of the optimization problem is to find a set of design variables that
result in a double-layer grid with the minimum weight while satisfying certain
constraints. This can be expressed as
Find fX g ¼ x1 ; x2 ; ::; xng
X
ng X
nmðiÞ
to minimize W ð f X gÞ ¼ xi qj Lj
i¼1 j¼1 ð5:1Þ
(
gj ðfX gÞ 0; j ¼ 1; 2; . . .; nc
subjected to :
ximin xi ximax
where {X} is the vector containing the design variables; ng is the number of design
groups; W({X}) presents the weight of the structure; nm(i) is the number of
members for the ith group; qj and Lj denote the material density and the length of
the jth member, respectively. ximin and ximax are the lower and upper bounds of the
design variable xi, respectively. gj({X}) denotes design constraints; and nc is the
number of constraints.
For handling the constraints, a penalty approach is utilized. For this purpose, the
objective function (Eq. 5.1) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. Here, e1 is
set to unity and e2 is calculated by
iter
e2 ¼ 1:5 þ 1:5 ð5:3Þ
itermax
where iter is the current iteration number and itermax is the total number of iterations
for the optimization process.
The constraint conditions for grid structures are briefly explained in the
following.
Displacement constraint:
di dmax
i ; i ¼ 1; 2; . . .; nn ð5:4Þ
pu pr ; pr ¼ /c Fcr Ag ; /c ¼ 0:85
8 Fy qffiffiffiffi
>
< ð0:658Fe ÞFy ; KL 4:71 E
r Fy p2 E ð5:6Þ
Fcr ¼ qffiffiffiffi ; Fe ¼
> ðKL 2
: 0:877Fe ; r Þ
KL E
r [ 4:71 Fy
KL
200; for compression members
r ð5:7Þ
KL
300; for tension members
r
Each example has been solved 20 times independently, and 1000 iterations are
considered as the terminal condition. A population of 20 particles is considered for
each algorithm and the other algorithm parameters are set the same as the values
proposed in [4–6]. The optimization algorithms are coded in MATLAB and the
structures are analyzed using the direct stiffness method by our own codes.
The larger square on square double-layer grid contains 520 members and
165 nodes shown in Fig. 5.1. The bottom layer is simply supported at the nodes
illustrated in Fig. 5.2a. Each top layer joint is subjected to a concentrated vertical
load of 46 kN. Cross-sectional areas of the members are categorized into 20 groups
as shown in Fig. 5.2. In order to optimize this structure by MDVC-UVPS, two
stages are considered. The design variable configuration utilized for the first stage is
listed as follows: [1], [2 4], [3 5], [6 8], [7 9], [10 11], [12 13], [14 15], [16], [17
18], and [19 20].
The results found by CBO, ECBO, VPS, and MDVC-UVPS algorithms are
summarized in Table 5.2. VPS achieves the lightest design (i.e., 60,018 kg).
MDVC-UVPS has better performance in terms of the average optimized weight and
standard deviation on average weights which are 63,130, and 1932 kg, respectively.
The best designs obtained by CBO, ECBO, and MDVC-UVPS are 64,513, 61,119,
and 61,456 kg, respectively. The maximum stress ratios for the best designs of the
CBO, ECBO, VPS, and MDVC-UVPS are 94.39, 96.76, 96.90, and 98.33%,
respectively. Convergence histories are depicted in Fig. 5.3. The required number
of structural analyses to achieve the best design by CBO, ECBO, VPS, and
MDVC-UVPS are 3400, 19,820, 19,560, and 2804 analyses, respectively.
Figure 5.4 shows the 3D view of a square on larger square grid. This structure has
672 members and 205 nodes and the bottom layer is simply supported at the nodes
shown in Fig. 5.5a. Each top layer joint is subjected to a concentrated vertical load
Fig. 5.2 Top view of the 520-bar double-layer grid problem and member groups: a all members
with simple supports, b bottom layer members, c top layer members, and d web members
of 30 kN. The cross-sectional areas of the members are categorized into 22 groups
as depicted in Fig. 5.5. In order to optimize this structure by MDVC-UVPS, two
stages are considered. The design variable configuration utilized for the first stage is
listed as follows: [1 2], [3 4], [5 6], [7 8], [9], [10 12], [11 13], [14 16], [15 17], [18
19], [20 21], and [22].
Table 5.3 lists the optimal designs found by different methods. MDVC-UVPS
obtained the lightest design compared to other methods that is 53,552 kg.
Moreover, the average optimized weight and the standard deviation on average
weight of MDVC-UVPS (58,589 and 3626 kg) are less than those of all other
methods. The best designs found by the CBO, ECBO, and VPS are 55,621, 54,569,
and 53,704 kg, respectively. The maximum stress ratios for the best designs of the
5.3 Design Problems 71
Table 5.2 Performance comparison for the 520-bar double-layer grid problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 EST 4 ST 4 ST 5 ST 6
2 EST 4 ST 5 ST 5 EST 4
3 ST 4 DEST 3 ST 3½ EST 4
4 ST 4 ST 3½ ST 3½ ST 3
5 DEST 5 EST 5 EST 4 ST 6
6 ST 3½ ST 3½ EST 3 EST 4
7 EST 3½ EST 5 DEST 2½ EST 4
8 EST 4 DEST 4 EST 5 DEST 4
9 EST 6 ST 8 EST 6 EST 6
10 EST 6 EST 6 ST 8 EST 8
11 EST 10 EST 5 EST 5 ST 5
12 EST 8 DEST 5 DEST 5 EST 6
13 EST 8 ST 12 EST 12 ST 10
14 ST 10 DEST 6 DEST 6 EST 12
15 DEST 8 DEST 8 DEST 6 DEST 8
16 ST 6 ST 6 ST 6 ST 5
17 ST 4 ST 4 ST 4 ST 4
18 ST 3½ ST 3½ ST 3½ ST 4
19 EST 3 ST 3 ST 3½ ST 5
20 ST 3½ ST 2½ ST 2½ ST 5
Weight (kg) 64,513 61,119 60,018 61,456
Average optimized weight (kg) 66,906 63,463 63,360 63,130
Standard deviation on average weight (kg) 7167 4052 2446 1932
Fig. 5.3 Convergence curves for the 520-bar double-layer grid problem
72 5 Optimal Design of Double-Layer Grids
Fig. 5.5 Top view of the 672-bar double-layer grid problem and member groups: a all members
with simple supports, b bottom layer members, c top layer members, and d web members
5.3 Design Problems 73
Table 5.3 Performance comparison for the 672-bar double-layer grid problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 ST 4 ST 4 ST 5 ST 4
2 ST 6 ST 5 ST 4 ST 5
3 ST 2½ ST 3½ ST 4 ST 3½
4 ST 2½ EST 1½ EST 1½ EST 1½
5 ST 2½ ST 4 ST 6 ST 3
6 EST 6 EST 4 EST 6 DEST 4
7 EST 4 EST 6 EST 6 DEST 4
8 EST 6 DEST 4 DEST 5 EST 6
9 EST 5 ST 6 EST 5 ST 6
10 ST 4 ST 3½ ST 3½ ST 3½
11 ST 6 EST 6 ST 5 DEST 4
12 EST 5 EST 4 EST 4 ST 5
13 EST 5 DEST 4 EST 6 EST 6
14 DEST 4 DEST 4 ST 6 EST 5
15 EST 8 EST 6 DEST 4 DEST 4
16 DEST 4 DEST 4 EST 6 EST 5
17 ST 5 DEST 5 ST 8 DEST 5
18 ST 4 ST 4 ST 4 ST 4
19 ST 4 ST 4 ST 3½ ST 3½
20 ST 3½ ST 3½ ST 3½ ST 3½
21 ST 3½ ST 3 ST 3½ EST 2½
22 ST 2½ ST 2½ ST 2½ ST 2½
Weight (kg) 55,621 54,569 53,704 53,552
Average optimized weight (kg) 62,287 59,164 60,850 58,589
Standard deviation on average weight (kg) 9853 5597 5985 3626
CBO, ECBO, VPS, and MDVC-UVPS are 96.59, 95.73, 99.96, and 98.91%,
respectively. Convergence histories are demonstrated in Fig. 5.6. It should be noted
that MDVC-UVPS requires 3772 structural analyses to find the optimum solution
while CBO, ECBO, and VPS require 4640, 14,480, and 19,640 structural analyses,
respectively.
74 5 Optimal Design of Double-Layer Grids
Fig. 5.6 Convergence curves for the 672-bar double-layer grid problem
The design of square on square double-layer grid shown in Fig. 5.7 is considered as
the third example. This structure has 800 members and 221 nodes and the bottom
layer is simply supported at the nodes illustrated in Fig. 5.8a. Each top layer joint is
subjected to a concentrated vertical load of 30 kN. Cross-sectional areas of the
members are categorized into 24 groups as shown in Fig. 5.8. In order to optimize
this structure by MDVC-UVPS, two stages are considered. The design variable
configuration utilized for the first stage is listed as follows: [1], [2 4], [3 5], [6 8], [7
9], [10], [11], [12 14], [13 15], [16 18], [17 19], [20 21 22], and [23 24].
Table 5.4 summarizes the results obtained by CBO, ECBO, VPS, and
MDVC-UVPS methods. MDVC-UVPS has better performance in terms of the best
weight, average optimized weight, and standard deviation on average weight which
are 53,590, 57,679, and 3524 kg, respectively. The best designs obtained by CBO,
ECBO, and VPS are 55,714, 53,673, and 53.714 kg, respectively. The maximum
stress ratios for the best designs of the CBO, ECBO, VPS, and MDVC-UVPS are
Fig. 5.8 Top view of the 800-bar double-layer grid problem and member groups: a all members
with simple supports, b bottom layer members, c top layer members, and d web members
99.94, 93.86, 91.87, and 94.71%, respectively. Convergence histories are depicted
in Fig. 5.9. The required number of structural analyses to achieve the best design by
CBO, ECBO, VPS, and MDVC-UVPS are 11,520, 16,860, 15,600, and 5122
analyses, respectively.
Figure 5.10 shows the 3D view of a square on the diagonal grid. This grid has
1016 members and 320 nodes and simple support conditions are employed for the
bottom layer at the nodes demonstrated in Fig. 5.11a. Each top layer joint is sub-
jected to a concentrated vertical load of 30 kN. The elements are divided into
76 5 Optimal Design of Double-Layer Grids
Table 5.4 Performance comparison for the 800-bar double-layer grid problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 EST 3½ ST 4 ST 4 ST 4
2 ST 6 ST 5 ST 5 ST 5
3 ST 2 EST 2 EST 1½ ST 1½
4 ST 3½ ST 3 EST 3 ST 3
5 ST 2½ EST 2 ST 3½ ST 2½
6 ST 3 ST 2 EST 1½ ST 2
7 EST 3 EST 3½ ST 5 ST 3
8 ST 2½ ST 3 ST 4 DEST 2
9 EST 3 EST 3½ EST 3 ST 5
10 ST 5 ST 3 ST 2 DEST 3
11 ST 8 EST 5 ST 6 DEST 4
12 ST 3½ ST 3½ ST 3½ ST 3½
13 ST 4 ST 6 ST 6 ST 6
14 ST 5 ST 6 ST 6 ST 5
15 ST 6 ST 6 ST 6 ST 5
16 ST 6 ST 6 ST 6 ST 6
17 DEST 4 EST 5 EST 6 DEST 4
18 EST 5 EST 6 EST 5 DEST 4
19 EST 5 DEST 4 DEST 4 DEST 5
20 EST 3½ ST 4 ST 4 ST 4
21 ST 3½ ST 3½ ST 3½ ST 3½
22 ST 3 ST 3½ ST 3 ST 3
23 ST 2½ ST 2½ ST 2½ ST 2½
24 ST 2½ ST 2½ ST 2½ ST 2½
Weight (kg) 55,714 53,673 53,714 53,590
Average optimized weight (kg) 61,464 58,953 57,912 57,679
Standard deviation on average weight (kg) 10,127 4643 4102 3524
25 groups and the member groups are presented in Fig. 5.11. Two stages with 11
and 25 variables are considered by MDVC-UVPS algorithm. The first DVC is as
follows: [1 2], [3 4 5], [6 7 8], [9 10 11], [12], [13 15], [14 16], [17 19], [18 20], [21
22 23], and [24 25].
Table 5.5 presents the optimum designs obtained by proposed algorithms. The
lightest design (i.e., 65,826 kg) is achieved by MDVC-UVPS algorithm after
4234 analyses. The best designs obtained by CBO, ECBO, and VPS are 74,849,
67,839, and 67,229 kg, respectively. These values are found after 9760, 15,760,
and 15,220 analyses. MDVC-UVPS obtains the lost values of average optimized
weight and standard deviation on average weight which are equal to 70,488, and
5018 kg, respectively. The maximum values of the stress ratio for CBO, ECBO,
5.3 Design Problems 77
Fig. 5.9 Convergence curves for the 800-bar double-layer grid problem
VPS, and MDVC-UVPS are 93.01, 93.99, 96.07, and 97.10%, respectively.
Convergence history diagrams are depicted in Fig. 5.12.
Figure 5.13 shows the 3D view of a diagonal on diagonal grid. This structure has
1520 members and 401 nodes and the bottom layer is simply supported at the
nodes shown in Fig. 5.14a. Each top layer joint is subjected to a concentrated
vertical load of 16 kN. The cross-sectional areas of the members are categorized
into 31 groups as depicted in Fig. 5.14. In order to optimize this structure by
MDVC-UVPS, two stages are considered. The design variable configuration
78 5 Optimal Design of Double-Layer Grids
Fig. 5.11 Top view of the 1016-bar double-layer grid problem and member groups: a all
members with simple supports, b bottom layer members, c top layer members, and d web
members
utilized for the first stage is listed as follows: [1 2], [3 4 5], [6 7 8], [9 10 11], [12],
[13 14], [15 16], [17 18], [19 20], [21], [22], [23 24 25], [26 27 28], and [29 30 31].
Table 5.6 presents a comparison between the results of the optimal designs
found by different methods. MDVC-UVPS obtained the lightest design compared
to other methods that is 79,571 kg. Moreover, the average optimized weight and the
standard deviation on the average weight of MDVC-UVPS (85,398 and 4407 kg)
are less than those of all other methods. The best designs found by the CBO,
ECBO, and VPS are 93,174, 82,254, and 82,357 kg, respectively. The maximum
5.3 Design Problems 79
Table 5.5 Performance comparison for the 1016-bar double-layer grid problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 EST 5 EST 5 ST 6 DEST 4
2 DEST 3 EST 5 ST 5 DEST 3
3 ST 3½ ST 3 ST 3½ ST 3½
4 ST 2½ ST 3½ ST 2½ ST 2½
5 ST 2½ ST 2½ ST 4 ST 3
6 ST 2 ST 2 EST 1 EST 1½
7 ST 2 DEST 2 EST 2 EST 1½
8 ST 2½ DEST 2 DEST 2 EST 2½
9 DEST EST 2 EST 3 ST 3½
2½
10 DEST ST 6 DEST DEST 2
2½ 2½
11 ST 1½ ST 2 EST 12 DEST 2½
12 DEST 5 EST 8 DEST 5 EST 8
13 EST 3½ EST 3½ ST 4 EST 4
14 EST 3½ ST 5 ST 5 ST 4
15 EST 4 ST 4 ST 5 ST 5
16 ST 6 EST 5 DEST 4 ST 4
17 ST 5 ST 5 EST 4 ST 6
18 EST 4 EST 5 EST 4 ST 6
19 EST 5 EST 5 EST 4 EST 6
20 ST 8 ST 8 DEST 4 EST 6
21 ST 6 ST 5 ST 6 ST 5
22 ST 3 ST 3 ST 3½ ST 3½
23 EST 6 EST 2½ EST 2½ EST 2½
24 ST 3½ ST 5 ST 2½ ST 2½
25 EST 1½ ST 4 EST 1½ ST 2½
Weight (kg) 74,849 67,839 67,229 65,826
Average optimized weight (kg) 79,422 73,042 72,366 70,488
Standard deviation on average weight (kg) 8154 9158 5545 5018
stress ratio for the best designs of the CBO, ECBO, VPS, and MDVC-UVPS are
99.59, 99.40, 99.94, and 99.92%, respectively. Convergence histories are demon-
strated in Fig. 5.15. It should be noted that MDVC-UVPS requires 3142 structural
analyses to find the optimum solution while CBO, ECBO, and VPS require 4360,
18,000, and 12,120 structural analyses, respectively.
80 5 Optimal Design of Double-Layer Grids
Fig. 5.12 Convergence curves for the 1016-bar double-layer grid problem
Fig. 5.14 Top view of the 1520-bar double-layer grid problem and member groups: a all members
with simple supports, b bottom layer members, c top layer members, and d web members
Table 5.6 Performance comparison for the 1520-bar double-layer grid problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 EST 5 DEST 5 EST 5 ST 6
2 ST 5 EST 5 ST 5 DEST 3
3 ST 2½ ST 2½ EST 2½ ST 2½
4 EST 1½ EST 1½ EST 1½ ST 2½
5 ST 1½ ST 2 ST 1½ ST 2½
6 EST 1¼ ST 2 ST 1¼ EST 2½
7 ST 3 EST 2½ ST 8 EST 2½
8 ST 4 EST 3 EST 5 EST 2½
(continued)
82 5 Optimal Design of Double-Layer Grids
Fig. 5.15 Convergence curves for the 1520-bar double-layer grid problem
References
1. Lan TT (2005) Space frame structures. Handbook of structural engineering. CRC Press, Boca
Raton, FL
2. Kaveh A (2017) Advances in metaheuristic algorithms for optimal design of structures, 2nd
edn. Springer, Switzerland
3. Kaveh A, Mahdavi VR (2014) Colliding bodies optimization: a novel meta-heuristic method.
Comput Struct 139:18–27
4. Kaveh A, Ilchi Ghazaan M (2014) Enhanced colliding bodies optimization for design problems
with continuous and discrete variables. Adv Eng Softw 77:66–75
5. Kaveh A, Ilchi Ghazaan M (2016) Vibrating particles system algorithm for truss optimization
with multiple natural frequency constraints. Acta Mech 228(1):307–332
6. Kaveh A, Ilchi Ghazaan M (2018) A new hybrid meta-heuristic algorithm for optimal design of
large-scale dome structures. Eng Optimiz 50(2):235–252
7. American Institute of Steel Construction (AISC) (1994) Manual of steel construction load
resistance factor design. USA
Chapter 6
Optimal Design of Double-Layer Barrel
Vault Space Structures
6.1 Introduction
The brick architecture of the Orient and the masonry construction of the Romans
provide numerous examples of the structural use of barrel vaults. During recent
years, architects and engineers have rediscovered the advantages of barrel vaults as
viable and often highly suitable forms for covering not only low-cost industrial
buildings, warehouses, large-span hangars, and indoor sports stadiums, but also
large cultural and leisure centers. The impact of industrialization on production of
prefabricated barrel vaults has proved to be the most significant factor leading to
lower cost for these structures [1].
Barrel vaults are given different names depending on the way their surface is
formed. The earlier types of barrel vaults were constructed as single-layer struc-
tures. Nowadays, with the increase of the spans, double-layer systems are often
preferred. While the members of single-layer barrel vaults are mainly under the
action of flexural moments, those of double-layer barrel vaults are almost exclu-
sively under the action of axial forces and the elimination of bending moments
leads to a full utilization of strength of all the elements. Double-layer barrel vaults
are generally statically indeterminate. In such systems, due to the rigidity, the risk
of instability can almost be eliminated. The use of this type of barrel vaults
enhances the stiffness of the vault structure and provides structural systems of great
potential, capable of having spans in excess of 100 m [2, 3].
In the recent decades, different meta-heuristic algorithms have been developed
and applied to structural optimization problems. The growing popularity of these
techniques arises from (i) the lack of dependency on gradient information; (ii) in-
herent capability to deal with both discrete and continuous design variables; and
(iii) incorporating global search features to produce reasonable solutions for com-
plicated problems [4]. Double-layer barrel vaults have a great number of structural
elements, and utilizing optimization techniques has a considerable influence on
their economy and efficient structural configuration. For optimal design of
double-layer barrel vaults, Kaveh and Eftekhar presented an improved hybrid Big
Bang-Big Crunch (IBBBC) algorithm [5]. In another study optimal design of some
single-layer barrel vaults and a double arch barrel vault were investigated by Kaveh
et al. [6]. In several studies, the optimal design of a practical model of a braced
barrel vault has been studied by various researchers. Hasançebi and Çarbaş used ant
colony search method [7], Hasançebi et al. employed a reformulation of the ant
colony optimization [8] and Hasançebi and Kazemzadeh Azad utilized a refor-
mulation of Big Bang-Big Crunch algorithm and adaptive dimensional search
method [4, 9, 10].
In this chapter, three double-layer barrel roof structures are optimized to
investigate the performance of the CBO [11], ECBO [12], VPS [13] and
MDVC-UVPS [14] meta-heuristic algorithms. The structures are subjected to
stress, stability, and displacement limitations according to the provisions of
AISC-ASD [15]. The design variables are the cross-sectional areas of the bar
elements which are selected from a list of steel pipe sections.
where {X} is the vector containing the design variables; W({X}) presents the weight
of the structure; nm(i) is the number of members for the ith group; qj and Lj denote
the material density and the length of the jth member, respectively. ximin and ximax
are the lower and upper bounds of the design variable xi, respectively. gj({X})
denotes design constraints; and nc is the number of constraints.
For constraint handling, a penalty approach is utilized. For this purpose, the
objective function (Eq. 6.1) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. Here, e1 is
set to unity and e2 is calculated by:
iter
e2 ¼ 1:5 þ 1:5 ð6:3Þ
itermax
where iter is the current iteration number and itermax is the total number of iterations
for optimization process.
The constraint conditions for barrel vaults studied here are briefly explained in
the following. Limitation on stress and stability of barrel vault elements are
imposed according to the provisions of the AISC-ASD [15] as follows:
The allowable tensile stresses for tension members are calculated by
where E is the modulus of elasticity; ki is the slenderness ratio ðki ¼ kli =ri Þ; Cc
denotes the slenderness ratio dividing the elastic and inelastic buckling regions
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðCc ¼ 2p2 E=Fy Þ; k is the effective length factor (k is set 1 for all truss members);
Li is the member length; and ri is the minimum radius of gyration.
AISC-ASD recommends the maximum slenderness ratio of the elements to be
restricted to 300 and 200 for tension and compression members, respectively.
Three double-layer barrel vault problems with 384, 693, and 1536 bar elements are
considered to evaluate the performance of CBO, ECBO, VPS, and MDVC-UVPS
algorithms. The design variables are the cross-sectional areas of the bar elements
which are selected from the list of steel pipe sections of AISC-LRFD [16]. These
pipe sections are shown in Table 6.1. ST, EST, and DEST abbreviations stand for
standard weight, extra strong, and double-extra strong, respectively. Each example
has been solved 30 times independently and a maximum of 1000 iterations is
considered as the termination condition. A population of 20 particles is considered
88 6 Optimal Design of Double-Layer Barrel Vault Space Structures
for each algorithm and the other algorithm parameters are set similar to the values
proposed in [12–14]. The optimization algorithms are coded in MATLAB and the
structures are analyzed using the direct stiffness method by our own codes.
The first design problem deals with the size optimization of a spatial 384-bar barrel
vault consisting of two rectangular nets as shown in Fig. 6.1. In order to make the
structure stable, the angles of the bottom nets are placed at the center of one of the
above nets. The two nets are connected using diagonal elements [3]. The span of
the barrel vault is 24.82 m, its rise is 5.12 m and its length is 26.67 m. The depth of
the structure, i.e., the distance between the top and bottom layers, is equal to
1.35 m. This structure consists of 111 pinned joints and 384 bar elements, which
are grouped into 31 independent sizing variables as identified in Fig. 6.1. The
structural material properties are assumed as follows: The modulus of elasticity is
considered to be 30,450 ksi (210,000 MPa), the yield stress of steel is taken as 58
ksi (400 MPa), and the density of steel is equal to 0.288 lb per cubic inch
(7833.413 kg/m3). All connections are assumed as ball jointed and the supports are
considered at the two external edges of the top layer of the barrel vault. Vertical
concentrated loads of −20 kips (−88.964 kN) are applied to all free joints (non-
support joints) of the top layer. Strength and slenderness limitations are according
to AISC-ASD provision [15], which are discussed earlier. Displacement constraints
of ±0.1969 in (5 mm) are imposed on all nodes in x, y and z directions.
Table 6.2 presents the comparison of the results of different algorithms.
The VPS algorithm yields the least weight for this example, which is 62,455.30 lb
(28,329.24 kg). The other design weights are 69,448.52 lb (31,501.32 kg) by CBO,
62,486.02 lb (28,343.18 kg) by ECBO, and 62,735.42 lb (28,456.31 kg) by
MDVC-UVPS. The best design of the VPS has been achieved in 12,780 analyses.
CBO, ECBO, and MDVC-UVPS require 4320, 15,980, and 3460 structural anal-
yses to find their optimum solutions, respectively. Figure 6.2 shows the conver-
gence curves of the best results found by CBO, ECBO, VPS, and MDVC-UVPS.
The considered 693-bar braced barrel vault consists of 259 joints and 693 members
with 23 independent design variables. The free span of the barrel vault is 19.03 m,
its rise is 5.75 m and its length is 22.9 m. The geometry and the member grouping
scheme of the structure is shown in Fig. 6.3. The structural material properties are
assumed as follows: The modulus of elasticity is taken as 29,000 ksi
(203,893.6 MPa), the yield stress of steel is assumed to be 36 ksi (253.1 MPa), and
the density of steel is 0.288 lb/in3 (7833.413 kg/m3). It is assumed that the barrel
90 6 Optimal Design of Double-Layer Barrel Vault Space Structures
Fig. 6.1 a 3D view, b plan view with group numbers c flattened cross-sectional view of the
384-bar double-layer barrel vault
6.3 Design Examples 91
Table 6.2 Performance comparison for the 384-bar braced barrel vault problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 ST 1/2 ST 1/2 ST 3/4 ST 1/2
2 EST 2 ST 2 1/2 EST 2 1/2 EST 2
3 EST 2 EST 2 EST 2 1/2 EST 2
4 ST 3 ST 1 1/2 EST 1 1/2 ST 1 1/2
5 DEST 2 1/2 EST 4 DEST 3 DEST 3
6 ST 2 1/2 ST 1 1/2 ST 1 1/2 ST 1 1/2
7 ST 12 ST 12 ST 12 ST 12
8 DEST 4 ST 10 EST 8 DEST 5
9 DEST 5 ST 12 EST 10 EST 10
10 ST 12 DEST 8 EST 10 EST 10
11 DEST 5 DEST 5 DEST 5 DEST 5
12 DEST 6 EST 8 DEST 5 ST 12
13 DEST 3 ST 6 ST 6 ST 6
14 EST 3 1/2 EST 3 1/2 DEST 3 ST 4
15 ST 2 1/2 ST 2 1/2 ST 2 1/2 EST 2 1/2
16 EST 6 ST 5 ST 5 ST 4
17 EST 6 EST 4 DEST 3 ST 6
18 EST 2 EST 1 1/2 EST 1 1/2 EST 1 1/2
19 EST 2 ST 1 1/4 ST 1 1/4 ST 1 1/4
20 EST 2 1/2 EST 1 1/2 EST 1 1/2 EST 1 1/2
21 EST 4 EST 1 1/2 EST 1 1/2 EST 1 1/2
22 ST 3 1/2 ST 1 1/4 EST 1 1/2 ST 1 1/4
23 EST 1 1/2 EST 1 1/2 EST 1 1/2 EST 1 1/2
24 ST 3 1/2 EST 2 1/2 EST 2 1/2 ST 3 1/2
25 ST 2 1/2 ST 2 1/2 EST 2 1/2 EST 2
26 DEST 4 ST 2 1/2 EST 1 1/2 EST 2
27 EST 3 DEST 2 ST 3 ST 3 1/2
28 EST 2 EST 1 1/2 EST 1 1/2 EST 2
29 ST 2 1/2 ST 2 1/2 EST 2 EST 2
30 ST 3 EST 1 1/2 EST 2 EST 2
31 ST 2 1/2 EST 1 1/2 EST 1 1/2 EST 2
Weight (lb) 69,448.52 62,486.02 62,455.30 62,735.42
(31,501.32 kg) (28,343.18 kg) (28,329.24 kg) (28,456.31 kg)
Average optimized 123,397 65,785 67,900 65,738
weight (lb) (55,971.93 kg) (29,839.57 kg) (30,798.92 kg) (29,818.26 kg)
Standard deviation on 103,837 3386 2913 2882
average weight (lb) (47,099.67 kg) (1535.86 kg) (1321.31 kg) (1307.25 kg)
92 6 Optimal Design of Double-Layer Barrel Vault Space Structures
Fig. 6.2 Convergence curves for the 384-bar barrel vault problem
vault is loaded by uniformly distributed vertical loads applied to the top of the roof,
and supports are considered at the two external edges of the top and bottom layers.
The applied loads are considered as follows: a uniform dead load (DL) pressure of
35 kg/m2; a positive wind load (WL) pressure of 160 kg/m2; and a negative wind
load (WL) pressure of 240 kg/m2. For design purposes, these loads are combined
follows:
• Load case 1: 1.5(DL + WL) = 1.5(35 + 160) = 292.5 kg/m2 (2.87 kN/m2)
• Load case 2: 1.5(DL – WL) = 1.5(35 – 240) = –307.5 kg/m2 (3.00 kN/m2)
Stress and slenderness constraints are considered according to AISC-ASD [15]
which are discussed earlier. The nodes are subjected to the displacement limits
of ± 0.1 in (0.254 cm) in every direction.
Table 6.3 presents a comparison between the obtained results. The lightest
design (i.e., 9091.1 lb (4123.65 kg)) is found by MDVC-UVPS algorithm and after
that the best design belongs to VPS (i.e., 9201.4 lb (4173.68 kg)). Figure 6.4 shows
the convergence curves of the best results found by CBO, ECBO, VPS, and
MDVC-UVPS. MDVC-UVPS converges to the optimum solution after 4120
analyses. CBO, ECBO, and VPS obtain their optimal solutions after 4400, 16,720,
and 9800 analyses, respectively. Element stress ratios at the best design optimized
by MDVC-UVPS are shown in Fig. 6.5. The maximum stress ratio is 81.12%.
The last design example is the size optimization of a 1536-bar double-layer barrel
vault with 413 joints, a span of 40 m, a length of 50 m, and 35 independent variable
6.3 Design Examples 93
Fig. 6.3 a 3D view, b plan view with group numbers of the top layer and c flattened
cross-sectional view with the group number of bracing and the bottom layer elements of the
693-bar double-layer barrel vault
94 6 Optimal Design of Double-Layer Barrel Vault Space Structures
Table 6.3 Performance comparison for the 693-bar barrel vault problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 ST 4 ST 4 EST 3 ST 4
2 ST 1 ST 1 ST 1 ST 1
3 ST 1 1/4 ST 3/4 ST 3/4 ST 3/4
4 ST 1 1/4 ST 1 ST 1 ST 1
5 ST 3/4 ST 3/4 ST 3/4 ST 3/4
6 EST 3 ST 3 ST 3 1/2 ST 3 1/2
7 ST 1 ST 1 ST 1 ST 1
8 ST 3/4 ST 1 ST 3/4 ST 1
9 ST 1 1/2 ST 1 ST 1 ST 1
10 ST 3/4 ST 3/4 ST 3/4 ST 3/4
11 ST 3 EST 2 ST 3 EST 2 1/2
12 ST 1 ST 1 1/4 EST 1 1/4 ST 1
13 ST 1 1/4 EST 2 EST 1 ST 1 1/2
14 ST 1 1/4 ST 1 ST 1 ST 1
15 ST 3/4 ST 3/4 ST 3/4 ST 3/4
16 ST 2 ST 1 EST 1 1/2 EST 1 1/4
17 ST 1 1/2 ST 1 ST 1 ST 1
18 EST 1 1/2 ST 3 EST 1 1/2 EST 2
19 ST 1 1/2 ST 1 ST 1 ST 1
20 ST 3/4 ST 3/4 EST 3/4 ST 3/4
21 ST 2 1/2 ST 3/4 ST 1 ST 1
22 ST 1 ST 3/4 ST 1 ST 1
23 ST 3/4 ST 3/4 ST 3/4 ST 3/4
Weight (lb) 10,221.8 9240.5 9201.4 9091.1
(4,636.53 kg) (4191.42 kg) (4173.68 kg) (4123.65 kg)
Average optimized 15,563 9577 9823 9475
weight (lb) (7,059.26 kg) (4344.05 kg) (4455.64 kg) (4297.79 kg)
Standard deviation on 3976 505 598 765
average weight (lb) (1803.48 kg) (229.06 kg) (271.25 kg) (347.00 kg)
groups [17]. The geometric details and member groups are presented in Fig. 6.6.
The modulus of elasticity is taken as 30,450 ksi (210,000 MPa), the yield stress of
steel is equal to 58 ksi (400 MPa), and the density of materials is considered to be
0.288 lb/in3 (7833.413 kg/m3). The supports are fixed at the two external edges of
the top layer of the structure and all joints of the top layer are subjected to con-
centrated vertical loads of 5 kips. The design constraints (including stress and
slenderness limitation) are considered according to AISC-ASD [15] which are
discussed earlier. The nodal displacements are limited to ±0.1969 in (5 mm) in
every direction.
6.3 Design Examples 95
Fig. 6.4 Convergence curves for the 693-bar barrel vault problem
Fig. 6.5 Stress ratios for the best design obtained by MDVC-UVPS for the 693-bar barrel vault
problem
Table 6.4 presents the results of the optimal designs of different optimization
algorithms. The weight of the best result obtained by MDVC-UVPS is 122,852 lb
(55,724.73 kg) that is the best among the compared methods. The average opti-
mized weight of this method is 146,229 lb (66,328.36 kg) which is less than those
of all other methods. Comparison of the convergence curves of CBO, ECBO, VPS,
and MDVC-UVPS is illustrated in Fig. 6.7. MDVC-UVPS requires 4762 structural
analyses to find the optimum solution while CBO, ECBO, and VPS require 4540,
15,060, and 18,080 structural analyses, respectively.
96 6 Optimal Design of Double-Layer Barrel Vault Space Structures
Fig. 6.6 a 3D view, b top plan view of quarter of the barrel vault with the related group numbers
and c flattened cross-sectional view of the 1536-bar double-layer barrel vault
6.3 Design Examples 97
Table 6.4 Performance comparison for the 1536-bar barrel vault problem
Element group Sections
CBO ECBO VPS MDVC-UVPS
1 DEST 3 EST 6 DEST 4 DEST 4
2 ST 4 DEST 3 ST 3 1/2 ST 4
3 ST 4 EST 3 1/2 ST 4 EST 3 1/2
4 DEST 2 1/2 EST 4 DEST 3 EST 5
5 EST 6 DEST 4 EST 6 EST 5
6 DEST 3 DEST 5 EST 5 EST 5
7 EST 1 1/4 ST 1/2 EST 1/2 ST 1/2
8 EST 2 ST 1 1/2 EST 2 ST 1 1/2
9 EST 2 1/2 ST 1 1/2 ST 3 1/2 ST 1 1/2
10 ST 2 ST 1 ST 1 1/2 ST 1 1/4
11 EST 2 EST 1 1/2 ST 1 1/4 ST 2 1/2
12 EST 4 ST 2 1/2 ST 3 EST 2 1/2
13 ST 1 1/4 EST 3 1/2 ST 3 1/2 EST 3
14 ST 5 EST 4 ST 5 DEST 3
15 ST 8 EST 2 1/2 DEST 3 ST 6
16 ST 8 DEST 2 1/2 DEST 3 ST 5
17 EST 4 ST 3 1/2 ST 3 DEST 2 1/2
18 DEST 2 1/2 EST 1 1/2 EST 2 1/2 EST 2
19 ST 3 1/2 EST 1 1/2 EST 1 1/2 EST 2
20 EST 2 ST 1 1/2 ST 2 1/2 ST 1 1/2
21 DEST 2 1/2 ST 1 1/2 EST 1 1/2 ST 1 1/2
22 EST 2 ST 1 1/2 EST 2 ST 1 1/2
23 ST 2 1/2 ST 1 1/2 ST 1 1/2 EST 2
24 ST 1 1/2 ST 1 1/2 DEST 2 1/2 ST 1 1/2
25 EST 2 1/2 ST 1 1/2 ST 1 1/2 EST 2
26 ST 3 1/2 DEST 2 ST 3 ST 2 1/2
27 EST 2 EST 1 1/2 ST 1 1/2 EST 1 1/2
28 EST 2 ST 1 1/2 ST 1 1/2 ST 1 1/2
29 EST 1 1/2 EST 1 1/2 EST 2 ST 1 1/2
30 ST 2 ST 1 1/4 ST 2 EST 1 1/2
31 EST 1 1/2 ST 2 1/2 EST 1 1/2 EST 1 1/2
32 ST 3 1/2 ST 1 1/4 EST 1 1/2 ST 2
33 EST 2 EST 2 ST 1 1/2 EST 1 1/2
34 ST 2 ST 1 1/2 EST 1 1/2 ST 2
35 EST 1 1/2 EST 1 1/2 ST 1 1/2 ST 2
Weight (lb) 152,229 128,111 129,117 122,852
(69,049.91 kg) (58,110.17 kg) (58,566.49 kg) (55,724.73 kg)
Average optimized 215,621 149,002 147,089 146,229
weight (lb) (97,804.04 kg) (67,586.17 kg) (66,718.45 kg) (66,328.36 kg)
Standard deviation on 36,322 16,775 14,644 14,552
average weight (lb) (16,475.38 kg) (7,609.01 kg) (6,642.41 kg) (6600.68 kg)
98 6 Optimal Design of Double-Layer Barrel Vault Space Structures
Fig. 6.7 Convergence curves for the 1536-bar barrel vault problem
Barrel vaults are effective semicylindrical structural systems that are commonly
used to provide long-span and economical roofs for multipurpose facilities
including warehouses, rail stations, pools, sport centers, airplane hungers, and
community centers.
In this chapter, three barrel vault optimization problems with discrete variables
are considered under stress, stability, and displacement limitations. In the 384-bar
barrel vault problem, the design found by the ECBO, VPS, and MDVC-UVPS are
approximately identical and are about 10% lighter than the result obtained by CBO.
In the 693-bar barrel vault problem, the design obtained by MDVC-UVPS is 12.4,
1.6, and 1.2% lighter than the best design achieved by the CBO, ECBO, and VPS,
respectively. These values are 24, 4.3, and 5 for the 1536-bar barrel vault problem.
A suitable average optimized weight and also an acceptable standard deviation from
the mean value of the independent runs show the robustness of the proposed
algorithms. Also, comparison of the convergence curves indicates that the
MDVC-UVPS comes close to the optimum design rapidly.
References
1. Makowski ZS (2006) Analysis, design and construction of braced barrel vaults. Elsevier
Applied Science Publishers, London, New York
2. Kaveh A, Mirzaei B, Jafarvand A (2014) Optimal design of double layer barrel vaults using
improved magnetic charged system search. Asian J Civil Eng 15(1):135–154
3. Kaveh A, Moradveisi M (2016) Optimal design of double-layer barrel vaults using CBO and
ECBO algorithms. Iran J Sci Technol, Trans Civil Eng 40(3):167–178
References 99
7.1 Introduction
Domes are one of the oldest and well-established structural forms and have been
used in architecture since the earliest times. These structures are of special interest
to engineers as they enclose large spaces with small surfaces and have proven to be
very economical in terms of consumption of constructional materials [1]. The main
aim of this chapter is frequency constraint optimization of dome truss structures;
however, all the domes are also optimized considering strength, stability, and
displacement constraints. Structural optimization considering natural frequency
constraints is believed to represent nonlinear and non-convex search spaces with
several local optima [2]. In this class of problems, large generalized eigenproblems
should be solved in order to find the natural frequencies of the structure. The size of
the structure affects the dimensions of the matrices involved and thus the required
computational time and effort. On the other hand, as the number of optimization
variables increases, more and more structural analyses are needed to be performed
in order to reach a near-optimal solution [3].
Structural optimization considering natural frequency constraints has been
studied since the 1980s [2] and was approached using mathematical programming
and meta-heuristic algorithms. Lin et al. [4] studied the minimum weight design of
structures under simultaneous static and dynamic constraints proposing a bi-factor
algorithm based on the Kuhn–Tucker criteria. Konzelman [5] considered the
problem using some dual methods and approximation concepts for structural
optimization. Grandhi and Venkayya [6] utilized an optimality criterion based on
uniform Lagrangian density for resizing and scaling procedure to locate the con-
straint boundary. Wang et al. [7] proposed an optimality criteria algorithm for
combined sizing–layout optimization of three-dimensional truss structures. In this
method, the sensitivity analysis helps to determine the search direction and the
optimal solution is achieved gradually from an infeasible starting point with a
minimum weight increment, and the structural weight is indirectly minimized.
In this chapter, the main aim of the optimization problem is to minimize the weight
of the structure under multiple frequency constraints while the design variables are
only the cross-sectional areas of bars (sizing optimization). Each variable should be
chosen within a permissible range.
The mathematical formulation can be expressed as follows:
where {X} is a vector containing the design variables; W({X}) presents the weight of
the structure; ng is the number of design groups; nm(i) is the number of members for
7.2 Frequency Constraint Optimization Problem 103
the ith group; qj and Lj denote the material density and the length of the jth member,
respectively. xj is the jth natural frequency of the structure and x*j is its upper bound;
xk is the kth natural frequency of the structure and x*k is its lower bound; ximin and
ximax are the lower and upper bounds of the design variable xi, respectively.
For constraint handling, a penalty approach is utilized. For this purpose, the
objective function (Eq. 7.1) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. In this
chapter, e1 is set to unity and e2 is calculated by
iter
e2 ¼ 1:5 þ 1:5 ð7:3Þ
itermax
where iter is the current iteration number and itermax is the total number of iterations
for optimization process.
The efficiencies of CBO, ECBO, VPS, and MDVC-UVPS are studied through three
dome truss structures. These examples contain
• A 600-bar dome truss
• A 1180-bar dome truss
• A 1410-bar dome truss
Two constraint cases are considered for each example. In Case 1, natural fre-
quency constraints are incorporated. In Case 2, limitation on stresses and stability of
truss elements are considered according to the provisions of the ASD-AISC [17] as
follows.
The allowable tensile stresses for tension members are calculated by
where E is the modulus of elasticity; ki is the slenderness ratio ðki ¼ kli =ri Þ; Cc
denotes the slenderness ratio dividing the elastic and inelastic buckling regions
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(Cc ¼ 2p2 E=Fy ); k is the effective length factor (k is set to 1 for all truss
members); li is the member length; and ri is the minimum radius of gyration.
In this design code provisions, the maximum slenderness ratio is limited to 300
for tension members, and this limit is recommended to be 200 for compression
members. Nodal displacements in all the coordinate directions must be less
than ±8 cm.
The 600-bar single-layer dome structure is schematized in Fig. 7.1. The entire
structure is composed of 216 nodes and 600 elements. Figure 7.2 shows a sub-
structure in more detail for nodal numbering and coordinates. Each of the elements
of this substructure is considered as a design variable. Thus, this is a size opti-
mization problem with 25 variables. The elastic modulus is 200 GPa and the
material density is 7850 kg/m3 for all elements. The yield stress of steel is taken as
400 MPa. A nonstructural mass of 100 kg is attached to all free nodes. The dome is
considered to be subjected to vertical loading at all the unsupported joints for the
second constraint case. According to symmetry, one only needs to define loading
conditions of the labeled nodes shown in Fig. 7.2. These loads are −100, −500,
−1500, −2500, −3500, −4500, −5500, −5000, and 0 kN, respectively. The
allowable minimum and maximum cross-sectional areas of all elements are set to
1 10−4 and 100 10−4 m2, respectively. There are two constraints on the
Table 7.1 presents the optimum designs obtained by CBO, ECBO [18], VPS [19],
and MDVC-UVPS [20]. It can be seen that the best design was found by
MDVC-UVPS. Also, the average optimized weight and standard deviation on
average weight of this method are the lowest which are 6119.95 and 16.23 kg.
Table 7.2 reports the natural frequencies of the optimized structures and it is clear
that none of the frequency constraints are violated. Convergence histories for these
techniques are depicted in Fig. 7.3. The required structural analyses to achieve the
best designs by the CBO, ECBO, VPS, and MDVC-UVPS, respectively, are
17,940, 19,020, 25,040, and 17,513 analyses, respectively. It should be noted that
MDVC-UVPS obtained the best design of VPS after 10,295 analyses.
106 7 Optimum Design of Dome-Shaped Trusses
Table 7.1 Performance comparison for the 600-bar dome truss problem—constraint case 1
Element number (nodes) Areas (cm2)
CBO ECBO VPS [19] MDVC-UVPS
[18] [20]
1 (1–2) 1.2404 1.4305 1.3155 1.2575
2 (1–3) 1.3797 1.3941 1.2299 1.3466
3 (1–10) 5.2597 5.5293 5.5506 4.9738
4 (1–11) 1.2658 1.0469 1.3867 1.4025
5(2–3) 17.2255 16.9642 17.4275 17.3802
6 (2–11) 38.2991 35.1892 40.1430 37.9742
7(3–4) 12.2234 12.2171 12.8848 13.0306
8 (3–11) 15.4712 16.7152 15.5413 15.9209
9 (3–12) 11.1577 12.5999 12.2428 11.9419
10 (4–5) 9.4636 9.5118 9.3776 9.1643
11 (4–12) 8.8250 8.9977 8.6684 8.4332
12 (4–13) 9.1021 9.4397 9.1659 9.2375
13 (5–6) 6.8417 6.8864 7.1664 7.2213
14 (5–13) 5.2882 4.2057 5.2170 5.2142
15 (5–14) 6.7702 7.2651 6.5346 6.7961
16 (6–7) 5.1402 6.1693 5.4741 5.2078
17 (6–14) 5.1827 3.9768 3.6545 3.4586
18 (6–15) 7.4781 8.3127 7.6034 7.6407
19 (7–8) 4.5646 4.1451 4.2251 4.3690
20 (7–15) 1.8617 2.4042 1.9717 2.1237
21 (7–16) 4.8797 4.3038 4.5107 4.5774
22 (8–9) 3.5065 3.2539 3.5251 3.4564
23 (8–16) 2.4546 1.8273 1.9255 1.7920
24 (8–17) 4.9128 4.8805 4.7628 4.8264
25 (9–17) 1.2324 1.5276 1.6854 1.7601
Weight (kg) 6182.01 6171.51 6120.01 6115.10
Average optimized weight (kg) 6226.37 6191.50 6158.11 6119.95
Standard deviation on average 60.12 39.08 28.49 16.23
weight (kg)
Table 7.2 Natural frequencies (Hz) evaluated at the optimum designs of the 600-bar dome truss
problem
Frequency number Natural frequencies (Hz)
CBO ECBO [18] VPS [19] MDVC-UVPS [20]
1 5.000 5.002 5.000 5.000
2 5.000 5.003 5.000 5.000
3 7.000 7.001 7.000 7.000
4 7.000 7.001 7.000 7.000
5 7.001 7.002 7.000 7.000
7.3 Design Examples 107
Fig. 7.3 Convergence curves for the 600-bar dome truss problem—constraint case 1
The results found by the proposed algorithms are summarized in Table 7.3.
MDVC-UVPS achieves the lightest design (i.e., 7338.37 kg). It also has better
performance in terms of the average optimized weight and the standard deviation on
average weight, which are 7694.13 and 122.67 kg, respectively. The best designs
obtained by CBO, ECBO, and VPS are 7464.76, 7463.78, and 7379.63 kg,
respectively. The maximum stress ratios for the best designs of CBO, ECBO, VPS,
and MDVC-UVPS are 99.99, 98.89, 99.99, and 99.99%, respectively. The required
numbers of structural analyses to achieve the best design by CBO, ECBO, VPS,
and MDVC-UVPS are 18,520, 16,400, 18,720, and 15,406 analyses, respectively.
For the second example, size optimization of a 1180-bar dome truss structure is
considered. The configuration of the structure is depicted in Fig. 7.4. The structure
consists of 400 nodes and 1180 elements. A substructure is illustrated in Fig. 7.5 in
more detail for nodal numbering. Each of the elements of this substructure is
considered as a design variable. Thus, this is a size optimization problem with 59
variables. Table 7.4 summarizes the coordinates of the nodes in Cartesian coordi-
nate system. The elastic modulus is 200 GPa and the material density is 7850 kg/
m3 for all elements. The yield stress of steel is taken as 400 MPa. A nonstructural
mass of 100 kg is attached to all free nodes. The dome is considered to be subjected
to vertical loading at all the unsupported joints for the second constraint case.
According to symmetry, only we need to define loading conditions of labeled nodes
108 7 Optimum Design of Dome-Shaped Trusses
Table 7.3 Performance comparison of the 600-bar dome truss problem—constraint case 2
Element number (nodes) Areas (cm2)
CBO ECBO VPS MDVC-UVPS
1 (1–2) 1.0007 1.1545 1.1552 1.0023
2 (1–3) 2.6144 4.6875 4.5998 4.5959
3 (1–10) 1.0267 1.4021 1.1522 5.368
4 (1–11) 1.0036 1.0025 1.0004 1.0208
5(2–3) 4.5976 4.8782 4.611 4.5967
6 (2–11) 1.1161 2.7684 1.3882 1.3896
7(3–4) 4.752 5.1739 4.752 4.8037
8 (3–11) 4.7657 4.7672 5.0116 4.7688
9 (3–12) 4.6437 1.4847 1.5148 1.4038
10 (4–5) 5.0566 5.7881 5.4851 5.0869
11 (4–12) 5.8028 5.7286 5.5706 5.576
12 (4–13) 8.041 3.8644 3.4961 3.4982
13 (5–6) 6.5194 6.2631 6.0269 6.0143
14 (5–13) 7.588 6.8744 6.8286 7.0211
15 (5–14) 4.688 4.7951 4.686 4.6903
16 (6–7) 7.0441 7.0731 7.0437 7.0399
17 (6–14) 8.463 8.5365 9.257 8.5434
18 (6–15) 5.7627 5.7799 5.8364 5.7635
19 (7–8) 8.5035 8.814 8.5035 8.5032
20 (7–15) 10.9017 10.9549 10.839 10.8387
21 (7–16) 7.7492 7.8738 7.7491 7.7481
22 (8–9) 6.1231 6.2894 6.0047 6.0053
23 (8–16) 13.7034 13.6983 13.6386 13.6388
24 (8–17) 5.4526 5.4853 5.4506 5.4509
25 (9–17) 12.4331 12.5248 12.4329 12.4327
Weight (kg) 7464.76 7463.78 7379.63 7338.37
Average optimized weight (kg) 8093.16 7739.41 7701.80 7694.13
Standard deviation on average weight 934.45 154.61 487.09 122.67
(kg)
shown in Fig. 7.5. These loads are −1000, −2500, −3500, −3500, −3500, −2000,
−1000, −100, −500, 0, −2000, −3000, −3500, −3500, −2500, −1500, −100, −200,
−500, and −200 kN, respectively. The allowable minimum and maximum
cross-sectional areas of all elements are set to 1 10−4 and 100 10−4 m2,
respectively. There are two constraints on the natural frequencies, which are
x1 7 and x3 9 Hz. This problem is optimized in three stages by
MDVC-UVPS. The numbers of design variables considered in the first and second
stages are 8 and 18, respectively. These DVCs contain
7.3 Design Examples 109
Table 7.4 Coordinates of the nodes of the 1180-bar dome truss problem
Node number Coordinates (x, y, z) Node number Coordinates (x, y, z)
1 (3.1181, 0.0, 14.6723) 11 (4.5788, 0.7252, 14.2657)
2 (6.1013, 0.0, 13.7031) 12 (7.4077, 1.1733, 12.9904)
3 (8.8166, 0.0, 12.1354) 13 (9.9130, 1.5701, 11.1476)
4 (11.1476, 0.0, 10.0365) 14 (11.9860, 1.8984, 8.8165)
5 (12.9904, 0.0, 7.5000) 15 (13.5344, 2.1436, 6.1013)
6 (14.2657, 0.0, 4.6358) 16 (14.4917, 2.2953, 3.1180)
7 (14.9179, 0.0, 1.5676) 17 (14.8153, 2.3465, 0.0)
8 (14.9179, 0.0, −1.5677) 18 (14.4917, 2.2953, −3.1181)
9 (14.2656, 0.0, −4.6359) 19 (13.5343, 2.1436, −6.1014)
10 (12.9903, 0.0, −7.5001) 20 (3.1181, 0.0, 13.7031)
Table 7.5 presents a comparison between the results of the optimal designs.
MDVC-UVPS obtained the lightest design compared to other methods that is
37,451.77 kg. Moreover, the average optimized weight and the standard deviation
on average weight of MDVC-UVPS are less than those of all other methods which
are 37,545.53 and 64.85 kg. Table 7.6 shows the optimized structural frequencies
(Hz) for various methods. None of the frequency constraints were violated. The
MDVC-UVPS algorithm requires 19,391 structural analyses to find the optimum
solution while CBO, ECBO, and VPS require 19,960, 19,860, and 24,780 structural
analyses, respectively. MDVC-UVPS obtained the best design of ECBO after 5873
analyses. The amount of saving in structural analyses at each iteration of the
MDVC-UVPS is shown in Fig. 7.6.
Table 7.7 lists the optimal designs found by different methods. MDVC-UVPS
obtained the lightest design compared to other methods that is 17,909.10 kg.
Moreover, the average optimized weight and the standard deviation on average
weight of MDVC-UVPS (18,417.05 and 427.44 kg) are less than those of the other
methods. The best designs found by the CBO, ECBO, and VPS are 19,869.30,
7.3 Design Examples 111
Table 7.5 Performance comparison for the 1180-bar dome truss problem—constraint case 1
Element number (nodes) Areas (cm2)
CBO ECBO [18] VPS [20] MDVC-UVPS
[20]
1 (1–2) 13.0171 7.6678 6.8743 7.3691
2 (1–11) 10.4346 11.1437 10.0230 9.3399
3 (1–20) 3.0726 1.8520 4.4140 2.7203
4 (1–21) 12.6969 14.5563 13.5515 13.2822
5 (1–40) 3.5654 4.9499 1.8303 3.6758
6 (2–3) 6.5190 6.8095 7.0824 6.1391
7 (2–11) 7.4233 6.6803 6.3960 7.0964
8 (2–12) 6.3471 6.7889 6.5646 6.0208
9 (2–20) 2.3013 1.0630 2.3705 2.1225
10 (2–22) 12.1936 9.1602 13.2621 12.3488
11 (3–4) 7.2877 6.9891 7.0922 6.8578
12 (3–12) 7.0961 6.9881 6.8079 5.7773
13 (3–13) 6.5669 6.9555 6.3815 6.9931
14 (3–23) 7.8257 7.5443 7.3122 7.3355
15 (4–5) 8.6812 9.5431 8.7221 10.5464
16 (4–13) 5.7888 6.9123 6.3680 6.9589
17 (4–14) 21.1342 8.9891 7.3159 8.0977
18 (4–24) 10.0502 6.8926 11.5749 7.7738
19 (5–6) 12.9279 12.6128 14.7985 12.4614
20 (5–14) 9.3212 8.1983 5.5174 7.8154
21 (5–15) 10.1260 11.8358 15.7381 10.2039
22 (5–25) 10.1358 9.7321 8.3419 8.9262
23 (6–7) 15.8585 19.1650 17.5000 16.5275
24 (6–15) 9.9672 10.4682 10.3084 9.0166
25 (6–16) 14.8493 14.1178 15.1958 13.8204
26 (6–26) 11.4909 11.14567 10.9395 11.4021
27 (7–8) 26.2359 23.4125 24.9421 24.2631
28 (7–16) 13.8812 15.5167 13.9614 14.5494
29 (7–17) 18.8857 16.6613 18.4153 17.7753
30 (7–27) 14.0257 15.9631 14.4945 15.4594
31 (8–9) 33.8826 37.0532 36.3529 34.1372
32 (8–17) 25.7142 22.2937 19.6608 19.1254
33 (8–18) 24.8644 22.7409 23.7259 24.1954
34 (8–28) 19.8498 23.5624 22.0297 21.5899
35 (9–10) 53.2630 47.7652 47.3286 49.4717
36 (9–18) 22.7771 22.5066 22.9442 26.2915
37 (9–19) 35.4230 34.6418 30.8229 33.7558
38 (9–29) 57.5480 31.6492 33.1098 29.7608
(continued)
112 7 Optimum Design of Dome-Shaped Trusses
Table 7.6 Natural frequencies (Hz) evaluated at the optimum designs of the 1180-bar dome truss
problem
Frequency number Natural frequencies (Hz)
CBO ECBO [18] VPS [20] MDVC-UVPS [20]
1 7.000 7.000 7.000 7.000
2 7.001 7.001 7.001 7.001
3 9.000 9.000 9.000 9.000
4 9.000 9.000 9.000 9.000
5 9.005 9.064 9.177 9.005
7.3 Design Examples 113
Fig. 7.6 Saving in structural analyses using the MDVC-UVPS algorithm in the 1180-bar dome
truss problem—constraint case 1
18,860.43, and 18,903.65 kg, respectively. The maximum stress ratios for the best
designs of CBO, ECBO, VPS, and MDVC-UVPS are 81.11, 86.52, 93.50, and
99.99%, respectively. Convergence history diagrams are depicted in Fig. 7.7.
MDVC-UVPS requires 14,456 structural analyses to find the optimum solution
while CBO, ECBO, and VPS require 18,220, 19,740, and 19,880 structural anal-
yses, respectively.
Table 7.7 Performance comparison of the 1180-bar dome truss problem—constraint case 2
Element number (nodes) Areas (cm2)
CBO ECBO VPS MDVC-UVPS
1 (1–2) 4.8705 5.9493 5.7015 9.4521
2 (1–11) 7.3783 5.1128 7.2102 3.8492
3 (1–20) 1.2699 1.7992 1.0055 8.1978
4 (1–21) 6.1376 5.4755 3.4166 4.2423
5 (1–40) 3.542 4.6291 3.8978 8.1978
6 (2–3) 6.1148 7.084 5.1114 9.4521
7 (2–11) 5.4713 6.473 5.064 3.8492
8 (2–12) 9.4118 7.1685 5.1467 4.8473
9 (2–20) 9.6458 11.6397 9.7255 8.1978
10 (2–22) 4.7464 6.5289 4.6638 4.2423
11 (3–4) 9.2156 9.4237 9.7013 9.4521
12 (3–12) 6.9436 5.8957 10.1884 4.8473
13 (3–13) 5.1492 5.4008 8.9843 5.5341
14 (3–23) 4.0869 5.4558 4.5727 7.112
(continued)
114 7 Optimum Design of Dome-Shaped Trusses
Fig. 7.7 Convergence curves for the 1180-bar dome truss problem—constraint case 2
Figure 7.8 shows the 1410-bar double-layer dome truss structure. The entire
structure is composed of 390 nodes and 1410 elements. Figure 7.9 shows a sub-
structure in more detail for nodal numbering. Each of the elements of this sub-
structure is considered as a design variable. Thus, this is a size optimization
problem with 47 variables. Table 7.8 presents the coordinates of the nodes in
Cartesian coordinate system. The elastic modulus is 200 GPa and the material
density is 7850 kg/m3 for all elements. The yield stress of steel is taken as
400 MPa. A nonstructural mass of 100 kg is attached to all free nodes. The dome is
116 7 Optimum Design of Dome-Shaped Trusses
considered to be subjected to vertical loading at all the unsupported joints for the
second constraint case. According to symmetry, only we need to define loading
conditions of labeled nodes shown in Fig. 7.9. These loads are −200, −600, −1000,
−1500, −2000, −2500, 0, −400, −1000, −1200, −1500, −2000, and −1000 kN,
respectively. The allowable minimum and maximum cross-sectional areas of all
elements is set to 1 10−4 and 100 10−4 m2, respectively. There are two
7.3 Design Examples 117
Table 7.8 Coordinates of the nodes of the 1410-bar double-layer dome truss problem
Node number Coordinates (x, y, z) Node number Coordinates (x, y, z)
1 (1.0, 0.0, 4.0) 8 (1.989, 0.209, 3.0)
2 (3.0, 0.0, 3.75) 9 (3.978, 0.418, 2.75)
3 (5.0, 0.0, 3.25) 10 (5.967, 0.627, 2.25)
4 (7.0, 0.0, 2.75) 11 (7.956, 0.836, 1.75)
5 (9.0, 0.0, 2.0) 12 (9.945, 1.0453, 1.0)
6 (11.0, 0.0, 1.25) 13 (11.934, 1.2543, −0.5)
7 (13.0, 0.0, 0.0)
Table 7.9 summarizes the results obtained by CBO, ECBO [18], VPS [20], and
MDVC-UVPS [20] methods. MDVC-UVPS has a better performance in terms of
the best weight, average optimized weight, and standard deviation on average
Table 7.9 Performance comparison for the 1410-bar dome truss problem—constraint case 1
Element number (nodes) Areas (cm2)
CBO ECBO [18] VPS [20] MDVC-UVPS
[20]
1 (1–2) 1.0073 7.7765 5.6333 5.8499
2 (1–8) 2.5808 6.2173 4.7628 4.5115
3 (1–14) 24.3407 23.9162 37.7385 19.4823
4 (2–3) 6.6750 11.2399 7.4927 8.8480
5 (2–8) 3.8881 2.5775 3.1824 5.0084
6 (2–9) 5.0607 1.8559 1.0193 1.3568
7 (2–15) 78.9781 16.9202 8.9475 17.4331
8 (3–4) 9.2944 13.7947 10.4272 9.1098
9 (3–9) 2.6585 5.4502 4.1398 2.8712
10 (3–10) 3.5399 2.9751 3.1408 3.5473
11 (3–16) 10.2473 13.7811 15.4194 12.3768
12 (4–5) 9.6820 9.3870 8.9931 10.1099
(continued)
118 7 Optimum Design of Dome-Shaped Trusses
Table 7.10 Natural frequencies (Hz) evaluated at the optimum designs of the 1410-bar dome
truss problem
Frequency number Natural frequencies (Hz)
CBO ECBO [18] VPS [20] MDVC-UVPS [20]
1 7.000 7.008 7.001 7.000
2 7.001 7.008 7.002 7.001
3 9.000 9.001 9.000 9.000
4 9.000 9.012 9.000 9.000
5 9.000 9.012 9.000 9.000
Fig. 7.10 Convergence curves for the 1410-bar dome truss problem—constraint case 1
weight that, respectively, are 10,345.12, 10,393.83, and 39.15 kg, respectively.
Table 7.10 reports the natural frequencies of the optimized structures and it is clear
that none of the frequency constraints are violated. Convergence histories for VPS
and MDVC-UVPS are depicted in Fig. 7.10. The required numbers of structural
analyses to achieve the best designs by the CBO, ECBO, VPS, and MDVC-UVPS,
respectively, are 19,000, 19,420, 25,700, and 17,750 analyses, respectively.
MDVC-UVPS obtained the best design of VPS after 10,697 analyses.
Table 7.11 presents the optimum designs obtained by the proposed algorithms. The
lightest design (i.e., 7661.64 kg) is achieved by MDVC-UVPS algorithm after
16,308 analyses. The best designs obtained by CBO, ECBO, and VPS are 8413.46,
7860.01, and 7848.68 kg, respectively. These values are found after 18,940,
120 7 Optimum Design of Dome-Shaped Trusses
Table 7.11 Performance comparison of the 1410-bar dome truss problem—constraint Case 2
Element number (nodes) Areas (cm2)
CBO ECBO VPS MDVC-UVPS
1 (1–2) 5.1214 5.217 4.6048 4.8489
2 (1–8) 2.2479 2.213 1.5208 1.5104
3 (1–14) 1 4.0413 1.4229 4.3939
4 (2–3) 5.6721 5.3523 4.785 4.8489
5 (2–8) 2.5777 2.8635 2.3714 2.3413
6 (2–9) 1.6817 1.8832 2.2803 1.6246
7 (2–15) 1.4126 1.0007 6.0836 4.3939
8 (3–4) 6.8558 6.4681 5.037 4.8489
9 (3–9) 2.1922 1.2068 2.1952 2.1707
10 (3–10) 2.0673 1.738 1.6864 1.6765
11 (3–16) 8.9218 12.5144 2.9786 4.3939
12 (4–5) 6.4513 6.3101 5.8296 7.6688
13 (4–10) 2.5147 1.7218 2.4275 2.4287
14 (4–11) 2.3745 2.4362 4.4668 1.8282
15 (4–17) 4.273 3.5615 3.0016 5.5832
16 (5–6) 6.5994 6.1832 6.1684 7.6688
17 (5–11) 3.3831 2.7977 2.5737 2.5749
18 (5–12) 2.7308 4.1412 4.5709 3.6629
19 (5–18) 8.5163 4.1542 4.2362 5.5832
20 (6–7) 7.834 7.9148 8.7333 7.6688
21 (6–12) 3.6101 5.894 3.3266 3.7234
22 (6–13) 5.0307 3.3083 5.439 3.1638
23 (6–19) 6.127 6.6223 5.8551 5.5832
24 (7–13) 3.8352 3.6804 3.7713 3.64
25 (8–9) 5.3726 4.8207 4.6028 6.1741
26 (8–14) 2.0258 1.5864 1.5129 1.5104
27 (8–15) 5.5215 2.5913 2.3505 2.3413
28 (8–21) 3.6576 1.0843 4.334 4.0242
29 (9–10) 5.638 5.9325 8.0424 6.1741
30 (9–15) 1.7705 3.0351 1.5699 1.6246
31 (9–16) 2.3381 1.2356 2.5573 2.1707
32 (9–22) 3.316 1.708 7.4354 4.0242
33 (10–11) 6.4184 4.8743 4.8246 6.3156
34 (10–16) 5.0152 3.429 1.6796 1.6765
35 (10–17) 2.9268 1.9623 3.3532 2.4287
36 (10–23) 5.7701 2.7079 2.4308 4.8511
37 (11–12) 8.4621 5.0557 5.1426 6.3156
38 (11–17) 1.925 4.1289 1.9981 1.8282
39 (11–18) 3.0442 3.4292 2.5741 2.5749
(continued)
7.3 Design Examples 121
19,840, and 19,860 analyses. MDVC-UVPS obtained values for average optimized
weight and standard deviation on average weight which are equal to 8106.52 and
244.08 kg, respectively. The maximum values of the stress ratio for CBO, ECBO,
VPS, and MDVC-UVPS are 96.96, 99.41, 99.53, and 99.99%, respectively.
References
1. Lan TT (2005) Space frame structures. Handbook of structural Engineering. CRC Press,
Boca Raton, FL, pp 24–31
2. Bellagamba L, Yang TY (1981) Minimum mass truss structures with constraints on
fundamental natural frequency. AIAA J 19:1452–1458
3. Kaveh A, Zolghadr A (2016) Optimal design and analysis of large-scale domes with
frequency constraints. Smart Struct Syst 18(4):733–754
4. Lin JH, Chen WY, Yu YS (1982) Structural optimization on geometrical configuration and
element sizing with static and dynamic constraints. Comput Struct 15:507–515
5. Konzelman CJ (1986) Dual methods and approximation concepts for structural optimization.
M.Sc. Thesis, Department of Mechanical Engineering, University of Toronto, Canada
6. Grandhi RV, Venkayya VB (1988) Structural optimization with frequency constraints.
AIAA J 26:858–866
7. Wang D, Zha W, Jiang J (2004) Truss optimization on shape and sizing with frequency
constraints. AIAA J 42:622–630
8. Sedaghati R (2005) Benchmark case studies in structural design optimization using the force
method. Int J Solids Struct 42:5848–5871
9. Lingyun W, Mei Z, Guangming W, Guang M (2005) Truss optimization on shape and sizing
with frequency constraints based on genetic algorithm. Comput Mech 35:361–368
10. Gomes HM (2011) Truss optimization with dynamic constraints using a particle swarm
algorithm. Expert Syst Appl 38:957–968
11. Kaveh A, Zolghadr A (2012) Truss optimization with natural frequency constraints using a
hybridized CSS–BBBC algorithm with trap recognition capability. Comput Struct
102–103:14–27
12. Miguel LFF, Fadel Miguel LF (2012) Shape and size optimization of truss structures
considering dynamic constraints through modern meta-heuristic algorithms. Expert Syst Appl
39:9458–9467
13. Zuo W, Bai J, Li B (2014) A hybrid OC–GA approach for fast and global truss optimization
with frequency constraints. Appl Soft Comput 14:528–535
14. Kaveh A, Javadi SM (2014) Shape and size optimization of trusses with multiple frequency
constraints using harmony search and ray optimizer for enhancing the particle swarm
optimization algorithm. Acta Mech 225:1595–1605
15. Kaveh A, Ilchi Ghazaan M (2015) Hybridized optimization algorithms for design of trusses
with multiple natural frequency constraints. Adv Eng Softw 79:137–147
16. Hosseinzadeh Y, Taghizadieh N, Jalili S (2016) Hybridizing electromagnetism-like mech-
anism algorithm with migration strategy for layout and size optimization of truss structures
with frequency constraints. Neural Comput Appl 27(4):953–971
17. American Institute of Steel Construction (AISC) (1989) Manual of steel construction:
allowable stress design, Chicago, USA
18. Kaveh A, Ilchi Ghazaan M (2017) Optimal design of dome truss structures with dynamic
frequency constraints. Struct Multidiscip Optim 53(3):605–621
19. Kaveh A, Ilchi Ghazaan M (2016) Vibrating particles system algorithm for truss optimization with
multiple natural frequency constraints. Acta Mech 228(1):307–322. (Published online, 1–16)
20. Kaveh A, Ilchi Ghazaan M (2018) A new hybrid meta-heuristic algorithm for optimal design
of large-scale dome structures. Eng Optimiz 50(2):235–252
Chapter 8
Optimal Design of Steel Lattice
Transmission Line Towers
8.1 Introduction
Lattice towers are used for power lines of all voltages and are the most common
type for high-voltage transmission lines. The design optimization of these structures
has always been a difficult task due to a large number of design variables. Some
studies have already been performed in the context of optimization of transmission
line tower structures. For instance, Rao [1] utilized a derivative-free nonlinear
optimization technique for minimum weight design of high-voltage transmission
line towers under a set of control parameters, including geometrical parameters as
well as tensions in conductors and ground-wires. Kaveh et al. [2] employed a
migration genetic algorithm for optimization of transmission towers and they
trained neural networks as analyzers to take part of the computational load. París
et al. [3] studied the shape optimization of a transmission line tower, subjected to
multiple load cases and code constraints. Guo and Li [4] presented an adaptive
genetic algorithm for different optimization models of steel transmission towers
considering size, layout, and topology design variables. Chunming et al. [5] utilized
a genetic algorithm for optimization of a transmission tower, where cross-sectional
areas and material types of the members were selected as design variables. Tort
et al. [6] integrated the simulated annealing optimization algorithm into the com-
mercial PLS-TOWER software to optimize steel lattice towers for minimum weight
using both size and layout design variables.
In this chapter, the efficiency of Colliding Bodies Optimization (CBO) [7],
Enhanced Colliding Bodies Optimization (ECBO) [8], Vibrating Particles System
(VPS) [9], and a hybrid algorithm called MDVC-UVPS [10] are investigated in
optimum design of three latticed steel towers. The procedure considers discrete
values of cross-sectional areas.
For the sizing optimization of transmission line towers, the cross-section areas of
truss bars are often considered as discrete design variables; therefore, all of them are
selected from a list of discrete cross sections based on production standards. The
optimization problem aims to minimize the weight of the structure while satisfying
strength and serviceability requirements.
Size optimization of a transmission line tower with its members being collected
in ng design groups can be formulated as follows:
Find fXg ¼ ½x1 ; x2 ; . . .; xng
Png P
nmðiÞ
to minimize WðfXgÞ ¼ xi qj Lj
i¼1 j¼1 ð8:1Þ
gj ðfXgÞ 0; j ¼ 1; 2; . . .; nc
subjected to:
ximin xi ximax
where {X} is the vector containing the design variables; W({X}) presents the weight
of the structure; nm(i) is the number of members for the ith group; qj and Lj denote
the material density and the length of the jth member, respectively. ximin and ximax
are the lower and upper bounds of the design variable xi, respectively. gj({X})
denotes the jth design constraint; and nc is the number of constraints.
For constraint handling, a penalty approach is utilized. For this purpose, the
objective function (Eq. 8.1) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. Here, e1 is
set to unity and e2 is calculated by
iter
e2 ¼ 1:5 þ 1:5 ð8:3Þ
itermax
where iter is the current iteration number and itermax is the total number of iterations
of the optimization process.
The following three benchmark structural examples are considered in this section:
• A 47-bar power transmission tower
• A 160-bar power transmission tower
• A 244-bar power transmission tower
8.3 Design Problems 125
The design variables are the cross-sectional areas of the bar element and in all
problems, solution candidates are allowed to select discrete values from a per-
missible list of cross sections (real numbers are rounded to the nearest integer in
each iteration). Each example has been solved thirty times independently and a
maximum of 1000 iterations is considered as the termination condition.
A population of 20 particles is considered for each algorithm and the other algo-
rithm parameters are set similar to the values proposed in [8–10]. The optimization
algorithms are coded in MATLAB and the structures are analyzed using the direct
stiffness method by our own codes.
The first design example demonstrated in Fig. 8.1 has 47 members and 22 nodes.
The cross-sectional areas of the members were categorized into 27 groups, as
follows: (1) A1 = A3, (2) A2 = A4, (3) A5 = A6, (4) A7, (5) A8 = A9, (6) A10,
(7) A11 = A12, (8) A13 = A14, (9) A15 = A16, (10) A17 = A18, (11) A19 = A20,
(12) A21 = A22, (13) A23 = A24, (14) A25 = A26, (15) A27, (16) A28,
(17) A29 = A30, (18) A31 = A32, (19) A33, (20) A34 = A35, (21) A36 = A37,
(22) A38, (23) A39 = A40, (24) A41 = A42, (25) A43, (26) A44 = A45, and
(27) A46 = A47. The cross-sectional areas were chosen from the 64 discrete values
listed in Table 8.1. The material of the members has a Young’s modulus of
206.842 kN/mm2 (30,000 ksi) and a density of 8303.97 kg/m3 (0.3 lb/in3). The
nodes of structure are subjected to a combination of three loading cases: (1) 6.0 kips
acting in the positive x-direction and 14.0 kips acting in the negative y-direction at
nodes 17 and 22, (2) 6.0 kips acting in the positive x-direction and 14.0 kips acting
in the negative y-direction at node 17, and (3) 6.0 kips acting in the positive x-
direction and 14.0 kips acting in the negative y-direction at node 22. The first
loading case corresponds to the load applied by the two power lines to the tower at
an angle, while the second and third loading cases occur when one of the two lines
snaps.
Both stress and buckling constraints should be satisfied for all of the members of
the tower. Allowable tensile and compressive stresses are taken as 137.895 MPa
(20 ksi) and 103.421 MPa (15.0 ksi), respectively. Moreover, the Euler buckling
compressive stress for a member with a cross-sectional area of Ai is calculated as
follows:
KEAi
rcr
i ¼ L2 i ¼ 1; 2; 3; . . .; 47 ð8:4Þ
i
where K is a constant parameter that should be selected based on the type of the
cross-sectional geometry; E is the Young’s modulus of the material; and Li is the
length of member i. The buckling constant K is assumed as 3.96 [11].
126 8 Optimal Design of Steel Lattice Transmission Line Towers
Table 8.2 shows the optimum designs found by different algorithms. The lightest
design (i.e., 2374.09 lb) is achieved by MDVC-UVPS algorithm after 4867 anal-
yses. The weights of the best designs obtained by CBO, ECBO, and VPS are
2386.44, 2375.35, and 2374.81 lb, respectively. These values are found after 9760,
Table 8.1 List of available cross-sectional areas for the 47-bar power transmission tower problem
No. Area (mm2) Area (in.2) No. Area (mm2) Area (in.2) No. Area (mm2) Area (in.2) No. Area (mm2) Area (in.2)
8.3 Design Problems
Table 8.2 Performance comparison for the 47-bar power transmission tower problem
Design variable Areas (in.2)
CBO ECBO VPS MDVC-UVPS
1 3.84 3.84 3.84 3.84
2 3.38 3.38 3.38 3.38
3 0.785 0.766 0.766 0.766
4 0.196 0.111 0.111 0.111
5 0.994 0.785 0.785 0.785
6 1.8 1.99 1.99 1.99
7 2.13 2.13 2.13 2.13
8 1.228 1.228 1.228 1.228
9 1.563 1.563 1.563 1.563
10 2.13 2.13 2.13 2.13
11 0.111 0.111 0.111 0.111
12 0.111 0.141 0.111 0.111
13 1.8 1.8 1.8 1.8
14 1.8 1.8 1.8 1.8
15 1.563 1.457 1.457 1.457
16 0.442 0.442 0.442 0.563
17 3.63 3.63 3.63 3.63
18 1.457 1.457 1.457 1.457
19 0.307 0.307 0.307 0.25
20 3.09 3.09 3.09 3.09
21 1.266 1.266 1.266 1.228
22 0.307 0.307 0.307 0.391
23 3.84 3.84 3.84 3.84
24 1.563 1.563 1.563 1.563
25 0.111 0.111 0.111 0.111
26 4.59 4.59 4.59 4.59
27 1.457 1.457 1.457 1.457
Weight (lb) 2386.44 2375.35 2374.81 2374.09
Average optimized weight (lb) 2462.76 2415.51 2415.07 2413.46
Standard deviation on average 44.79 41.01 35.65 38.21
weight (lb)
1 in.2 = 6.4516 cm2
16,240, and 15,540 analyses. The average optimized weight and standard deviation
on average weight achieved by MDVC-UVPS are 2413.46 and 38.21 lb, respec-
tively. Figure 8.2 shows the convergence curves of the best results found by CBO,
ECBO, VPS, and MDVC-UVPS.
8.3 Design Problems 129
Fig. 8.2 Convergence curves for the 47-bar power transmission tower problem
The second design example is the 160-bar tower structure shown in Fig. 8.3. The
members of the structure are categorized into 38 design groups. The geometrical
information including member’s connectivity, design groups, and coordinates of
nodes can be found in [12]. The material of the members has a Young’s modulus of
2.047 106 kgf/cm2 and a density of 7850 kg/m3. This tower is subjected to the
combination of eight loading conditions as listed in Table 8.3. Moreover, the design
constraints consist of buckling stresses of the structural members. For a member
under compressive force, the buckling stress can be calculated as follows:
For kl/r 120
ðkl=rÞ2
rall ¼ 1300 ð8:5Þ
24
r = {0.47, 0.57, 0.67, 0.77, 0.87, 0.97, 0.97, 1.06, 1.16, 1.26, 1.15, 1.26, 1.36, 1.46,
1.35, 1.36, 1.45, 1.55, 1.75, 1.95, 1.74, 1.94, 2.16, 2.36, 2.57, 2.35, 2.56, 2.14, 2.33,
2.97, 2.54, 2.93, 2.94, 2.94, 2.92, 3.54, 3.96, 3.52, 3.51, 3.93, 3.92, 3.92} (cm).
The optimal designs found by different methods are presented in Table 8.4. The
weight of the best result obtained by MDVC-UVPS is 1336.71 kg that is the best
among the compared methods. The average optimized weight of this method is
1364.56 kg, which is less than those of all other methods. Comparison of the
convergence rates of CBO, ECBO, VPS, and MDVC-UVPS is illustrated in
Fig. 8.4. MDVC-UVPS requires 4518 structural analyses to find its optimum
8.3 Design Problems 131
Table 8.3 Loading conditions for the 160-bar power transmission tower problem
Condition Node Fx Fy Fz Condition Node Fx Fy Fz
1 52 −868 0 −491 5 52 −917 0 −491
37 −996 0 −546 37 −951 0 −546
25 −1091 0 −546 25 −1015 0 −546
28 −1091 0 −546 28 −636 0 −428
2 52 −493 1245 −363 6 52 −917 1259 −491
37 −996 0 −546 37 −572 0 −428
25 −1091 0 −546 25 −1015 1303 −546
28 −1091 0 −546 28 −1015 0 −546
3 52 −917 0 −491 7 52 −917 0 −491
37 −951 0 −546 37 −951 0 −546
25 −1015 0 −546 25 −1015 0 −546
28 −1015 0 −546 28 −636 1303 −428
4 52 −917 0 −546 8 52 −498 1460 −363
37 −572 1259 −428 37 −951 0 −546
25 −1015 0 −546 25 −1015 0 −546
28 −1015 0 −546 28 −1015 0 −546
Table 8.4 Performance comparison for the 160-bar power transmission tower problem
Design variable Areas (cm2)
CBO ECBO VPS MDVC-UVPS
1 19.03 19.03 19.03 19.03
2 5.27 5.27 5.27 5.27
3 19.03 19.03 19.03 19.03
4 5.27 5.27 5.27 5.27
5 19.03 19.03 19.03 19.03
6 5.75 5.75 5.75 5.75
7 15.39 15.39 15.39 15.39
8 5.75 5.75 5.75 5.75
9 13.79 13.79 13.79 13.79
10 5.75 5.75 5.75 5.75
11 5.75 5.75 5.75 5.75
12 12.21 12.21 13.79 12.21
13 6.25 6.25 6.25 6.25
14 5.75 5.75 5.75 5.75
15 4.79 3.88 2.66 3.88
16 6.25 7.44 7.44 7.44
17 4.79 1.84 1.84 1.84
18 8.66 8.66 8.66 8.66
19 2.66 2.66 2.66 2.66
20 3.07 3.07 3.07 3.07
(continued)
132 8 Optimal Design of Steel Lattice Transmission Line Towers
Fig. 8.4 Convergence curves for the 160-bar bar power transmission tower problem
8.3 Design Problems 133
solution while CBO, ECBO, and VPS require 5820, 11,420, and 11,080 structural
analyses, respectively.
The 244-bar transmission tower shown in Fig. 8.5 is studied as the final design
problem to demonstrate the efficiency of the compared algorithms. Members of the
structure are categorized into 26 groups as given by Saka [13]. The cross-sectional
areas were chosen from the 45 discrete values listed in Table 8.5. The modulus of
elasticity and yield stress of steel are assumed to be 210 kN/mm2 and 233.3 N/mm2,
Table 8.5 List of available cross sections for the 244-bar power transmission tower problem
No. Section A in.2 (mm2) r in. (mm) Num. Section A in.2 (mm2) r in. (mm)
1 L661 11.0 (7096.76) 1.17 (29.72) 24 L 3 1/2 3 1/2 1/2 3.25 (2096.77) 0.683 (17.35)
2 L 6 6 7/8 9.73 (6277.41) 1.17 (29.72) 25 L 3 1/2 3 1/2 7/16 2.87 (1851.61) 0.684 (17.37)
3 L 6 6 3/4 8.44 (5445.15) 1.17 (29.72) 26 L 3 1/2 3 1/2 3/8 2.48 (1600.00) 0.687 (17.45)
4 L 6 6 5/8 7.11 (4587.09) 1.18 (29.97) 27 L 3 1/2 3 1/2 5/16 2.09 (1348.38) 0.690 (17.53)
5 L 6 6 9/16 6.43 (4148.38) 1.18 (29.97) 28 L 3 1/2 3 1/2 1/4 1.69 (1090.32) 0.694 (17.63)
6 L 6 6 1/2 5.75 (3709.67) 1.18 (29.97) 29 L 3 3 1/2 2.75 (1774.19) 0.584 (14.83)
7 L 6 6 7/16 5.06 (3264.51) 1.19 (30.23) 30 L 3 3 7/16 2.43 (1567.74) 0.585 (14.86)
8 L 6 6 3/8 4.36 (2812.90) 1.19 (30.23) 31 L 3 3 3/8 2.11 (1361.29) 0.587 (14.91)
9 L 6 6 5/16 3.65 (2354.83) 1.20 (30.48) 32 L 3 3 5/16 1.78 (1148.38) 0.589 (14.96)
10 L 5 5 7/8 7.98 (5148.38) 0.973 (24.71) 33 L 3 3 1/4 1.44 (929.03) 0.592 (15.04)
11 L 5 5 3/4 6.94 (4477.41) 0.975 (24.77) 34 L 3 3 3/16 1.09 (703.22) 0.596 (15.14)
12 L 5 5 5/8 5.86 (3780.64) 0.978 (24.84) 35 L 2 1/2 2 1/2 1/2 2.25 (1451.61) 0.487 (12.37)
13 L 5 5 1/2 4.75 (3064.51) 0.983 (24.97) 36 L 2 1/2 2 1/2 3/8 1.73 (1116.13) 0.487 (12.37)
14 L 5 5 7/16 4.18 (2696.77) 0.986 (25.04) 37 L 2 1/2 2 1/2 5/16 1.46 (941.93) 0.489 (12.42)
15 L 5 5 3/8 3.61 (2329.03) 0.990 (25.15) 38 L 2 1/2 2 1/2 1/4 1.19 (767.74) 0.491 (12.47)
16 L 5 5 5/16 3.03 (1954.83) 0.944 (25.25) 39 L 2 1/2 2 1/2 3/16 0.902 (581.93) 0.495 (12.57)
17 L 4 4 3/4 5.44 (3509.67) 0.778 (19.76) 40 L 2 2 3/8 1.36 (877.42) 0.389 (9.88)
18 L 4 4 5/8 4.61 (2974.19) 0.779 (19.79) 41 L 2 2 5/16 1.15 (741.93) 0.390 (9.91)
19 L 4 4 1/2 3.75 (2419.35) 0.782 (19.86) 42 L 2 2 1/4 0.938 (605.16) 0.391 (9.93)
20 L 4 4 7/16 3.31 (2135.48) 0.785 (19.94) 43 L 2 2 3/16 0.715 (461.29) 0.394 (10.00)
21 L 4 4 3/8 2.86 (1845.16) 0.788 (20.02) 44 L 2 2 1/8 0.484 (312.26) 0.398 (10.11)
22 L 4 4 5/16 2.40 (1548.38) 0.791 (20.09) 45 L 1 1/4 1 1/4 3/16 0.434 (280.00) 0.244 (6.198)
23 L 4 4 1/4 1.94 (1251.61) 0.795 (20.19)
8 Optimal Design of Steel Lattice Transmission Line Towers
8.3 Design Problems 135
Table 8.6 The load cases and displacement bounds for the 244-bar bar power transmission tower
problem
Loading conditions Joint number Loads (kN) Displacement
limitations
(mm)
X Z X Z
1 1 10 −30 45 15
2 10 −30 45 15
17 35 −90 30 15
24 175 −45 30 15
25 175 −45 30 15
2 1 – −360 45 15
2 – −360 45 15
17 – −180 30 15
24 – −90 30 15
25 – −90 30 15
respectively [14]. The load cases considered and the bounds imposed on the dis-
placements are shown in Table 8.6. Values of the allowable tensile and compres-
sive stresses are calculated based on ASD-AISC code [15]. Moreover, as per the
recommendation of ASD-AISC, the maximum slenderness ratios are set to 200 and
300 for compression and tension members, respectively.
Table 8.7 shows the comparison of the results of different algorithms.
MDVC-UVPS algorithm yields the least volume design for this example, which is
0.755011 m3. The other designs are 0.813555 m3 by CBO, 0.768159 m3 by ECBO,
and 0.764982 m3 by VPS. The best design of the hybrid method has been achieved
after 6665 analyses. CBO, ECBO, and VPS require 30,000, 4020, 9260, and 19,480
structural analyses to find their optimum solutions, respectively. Figure 8.6 shows
the convergence curves of the best results found by CBO, ECBO, VPS, and
MDVC-UVPS.
In this chapter, size optimization of transmission line towers under multiple loading
cases is studied. Three steel lattice towers with 26, 27, and 38 discrete variables are
considered. Numerical results demonstrate that the MDVC-UVPS achieved the
lightest designs in all examples. The average and standard deviation values
achieved from independent runs also confirm the efficiency of the hybrid method in
optimal design of real-size structures. Moreover, comparison of the convergence
curves and the required structural analyses indicate that MDVC-UVPS comes close
to the optimum design rapidly.
Table 8.7 Performance comparison for the 244-bar bar power transmission tower problem
136
Design Sections
variable CBO ECBO VPS MDVC-UVPS
1 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16
2 L 4 4 1/4 L 4 4 1/4 L 4 4 1/4 L 4 4 1/4
3 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4
4 L 4 4 5/16 L 3 3 7/16 L 3 3 3/8 L 5 5 7/16
5 L 3 3 7/16 L 2 2 3/8 L 2 1/2 2 1/2 3/8 L 2 2 3/8
6 L 5 5 7/8 L 5 5 7/8 L 5 5 5/16 L 4 4 3/4
7 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16
8 L 4 4 3/4 L 5 5 3/8 L 5 5 7/16 L 4 4 3/4
9 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4
10 L 2 2 3/8 L 2 2 3/8 L 2 2 3/8 L 2 2 3/8
11 L 3 1/2 3 1/2 5/16 L 5 5 3/4 L 5 5 3/8 L 5 5 1/2
12 L 5 5 3/4 L 4 4 5/8 L 4 4 5/8 L 4 4 3/8
13 L 2 1/2 2 1/2 3/8 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4
14 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8
15 L 6 6 3/4 L 6 6 5/8 L 6 6 7/8 L 6 6 3/4
16 L 3 1/2 3 1/2 5/16 L 3 1/2 3 1/2 5/16 L 3 1/2 3 1/2 5/16 L 3 1/2 3 1/2 5/16
17 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8
18 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8
19 L 3 3 3/8 L 3 3 3/8 L 2 1/2 2 1/2 1/2 L 3 3 3/8
20 L 6 6 5/8 L 6 6 9/16 L 6 6 1/2 L 6 6 9/16
21 L 3 3 3/16 L 3 3 3/16 L 3 3 3/16 L 3 3 3/16
22 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4
23 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4 L 2 2 1/4
24 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8 L 2 2 1/8
25 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16
26 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16 L 1 1/4 1 1/4 3/16
Volume (m3) 0.813555 0.768159 0.764982 0.755011
Average optimized volume (m3) 0.899143 0.830075 0.833456 0.818619
Standard deviation on average volume (m3) 0.1462 0.1184 0.1207 0.1053
8 Optimal Design of Steel Lattice Transmission Line Towers
References 137
Fig. 8.6 Convergence curves for the 244-bar bar power transmission tower problem
References
1. Rao GV (1995) Optimum designs for transmission line towers. Comput Struct 57(1):81–92
2. Kaveh A, Gholipour Y, Rahami H (2008) Optimal design of transmission towers using
genetic algorithm and neural networks. Int J Space Struct 23(1):1–9
3. París J, Martínez S, Navarrina F, Colominas I, Casteleiro M (2010) Structural optimization of
high tension towers. In: 2nd International conference on engineering optimization, Portugal
4. Guo HY, Li ZL (2011) Structural topology optimization of high-voltage transmission tower
with discrete variables. Struct Multidiscip Optim 43(6):851–861
5. Chunming W, Tingting S, Bin M, Jing G (2012) Research on the optimal layout of high
strength steel in the transmission tower. Phys Procedia 33:619–625
6. Tort C, Şahin S, Hasançebi O (2017) Optimum design of steel lattice transmission line towers
using simulated annealing and PLS-TOWER. Comput Struct 15(179):75–94
7. Kaveh A, Mahdavi VR (2014) Colliding bodies optimization: a novel meta-heuristic method.
Comput Struct 139:18–27
8. Kaveh A, Ilchi Ghazaan M (2014) Enhanced colliding bodies optimization for design
problems with continuous and discrete variables. Adv Eng Softw 77:66–75
9. Kaveh A, Ilchi Ghazaan M (2016) Vibrating particles system algorithm for truss optimization
with multiple natural frequency constraints. Acta Mech 228(1):307–332
10. Kaveh A, Ilchi Ghazaan M (2018) A new hybrid meta-heuristic algorithm for optimal design
of large-scale dome structures. Eng Optimiz 50(2):235–252
11. Lee KS, Geem ZW, Lee SH, Bae KW (2005) The harmony search heuristic algorithm for
discrete structural optimization. Eng Optimiz 37(7):663–684
12. Groenwold AA, Stander N, Snyman JA (1999) A regional genetic algorithm for the discrete
optimal design of truss structures. Int J Numer Methods Eng 44:749–766
13. Saka MP (1990) Optimum design of pin-jointed steel structures with practical applications.
J Struct Eng 116(10):2599–2620
14. Toğan V, Daloğlu AT (2006) Optimization of 3D trusses with adaptive approach in genetic
algorithms. Eng Struct 28(7):1019–1027
15. American Institute of Steel Construction (AISC) (1998) Manual of steel construction:
allowable stress design, Chicago, USA
Chapter 9
Optimal Seismic Design of 3D Steel
Frames
9.1 Introduction
steel frames. The optimized designs obtained by the proposed algorithm are better
than those found by the standard differential evolution algorithm and also very
competitive with literature. The overall convergence behavior is significantly
enhanced by the hybrid optimization strategy. Artar and Daloglu [6] used a method
based on genetic algorithm for minimum weight design of composite steel frames
with semirigid connections and column bases. Their results show that consideration
of the contribution of concrete on the behavior of the floor beams enables a lighter
and more economical design for steel frames with semirigid connections and col-
umn bases. Kaveh et al. [7] employed cuckoo search algorithm to optimize steel
frame structures under seismic loading based on response spectral and equivalent
static analyses. The effect of lateral seismic loading distribution on the achieved
optimum designs was also investigated. Results show similar weights for optimum
designs using spectral and equivalent static analyses; however, different material
distribution and seismic behaviors are observed. Carbas [8] enhanced the perfor-
mance of the firefly algorithm by suggesting two new expressions for the attrac-
tiveness and randomness parameters of the algorithm and optimized two steel space
frame design examples by this method. Kaveh and Bolandgerami [9] indicated the
efficiency of the cascade enhanced colliding body optimization for optimum design
of large-scale space steel frames. Kaveh et al. [10] employed nine meta-heuristic
algorithms to study the effect of the change in the ductility type on the structural
weight. Results show that the Ordinary Moment Frame (OMF) can produce lighter
designs in most cases, in spite of larger base shear. Kaveh et al. [11] studied three
different types of lateral resisting steel moment frames consisting of ordinary
moment frame, intermediate moment frame, and special moment frame. Optimum
seismic design was performed for 3D steel moment frames with different types of
lateral resisting systems.
The contribution of this chapter is concerned with optimization of steel frames
under seismic loads based on response spectral. Frame members are selected from
available set of steel sections for producing practically acceptable designs according
to Load and Resistance Factor Design-American Institute of Steel Construction
(LRFD-AISC) specification. Three irregular steel frame problems are considered
to evaluate the performance of the CBO, ECBO, VPS and MDVC-UVPS algo-
rithms [12].
The design should be carried out in such a way that the frame satisfies the following
constraint:
1. Strength constraints: Each frame member should have sufficient strength to
resist the internal forces developed due to factored external loading.
2. Serviceability constraints: Beam deflections and lateral displacement of the
frame should be less than the limits specified in the code.
3. Geometric constraints: Three types of geometric constraints are considered to
satisfy practical requirements. The first type ensures that the depths of the
W-shaped sections selected for the columns of two consecutive stories should be
either equal to each other or the one in the above story should be smaller than
the one in the below story. The second and third types of constraint make sure
that if a beam is connected to flange of a column, the flange width of the beam
should be less than or equal to the flange width of the column in the connection
and if a beam is connected to the web of a column, the flange width of the beam
should be less than or equal to (D-2tb), where D and tb are the depth and the
flange thickness of the W-shaped section selected for column, respectively.
Figure 9.1 and the following formula clarify the geometric constraints.
ðDÞi
1 i ¼ 2; . . .; ns ð9:1Þ
ðDÞi1
ðBf Þbi
1 i ¼ 1; . . .; nf ð9:2Þ
ðBf Þci
ðBf Þbi
1 i ¼ 1; . . .; nw ð9:3Þ
ðDÞi 2ðtb Þi
where ns, nf, and nw are the number of stories, the total number of joints where
beams are connected to the flange of a column, and the total number of joints where
beams are connected to the web of a column, respectively. (Bf)bi and (Bf)ci are the
flange width of W-section selected for the beam and column at joint i, respectively.
Sizing optimization of a steel frame with its members being collected in ng design
groups can be formulated as follows:
Find fXg ¼ ½x1 ; x2 ; ::; xng
Png P
nmðiÞ
to minimize WðfXgÞ ¼ xi qj Lj
ð9:4Þ
i¼1 j¼1
gj ðfXgÞ 0; j ¼ 1; 2; . . .; nc
subjected to:
ximin xi ximax
where {X} is the vector containing the design variables; W({X}) presents the weight
of the structure; nm(i) is the number of members for the ith group; qj and Lj denote
the material density and the length of the jth member, respectively. ximin and ximax
are the lower and upper bounds of the design variable xi, respectively. gj({X})
denotes design constraints; and nc is the number of the constraints.
For constraints handling, a penalty approach is utilized. For this purpose, the
objective function (Eq. 9.4) is redefined as follows:
where P({X}) is the penalized cost function or the objective function to be mini-
mized and t denotes the sum of the violations of the design constraints. In this
chapter, e1 is set to unity and e2 is calculated by
iter
e2 ¼ 1:5 þ 1:5 ð9:6Þ
itermax
where iter is the current iteration number and itermax is the total number of iterations
for optimization process.
9.3 Design Examples 143
Optimum seismic designs of three steel space frames are investigated in this sec-
tion. These design examples contain:
• The four-story 132-member steel frame
• The four-story 428-member steel frame
• The twelve-story 276-member steel frame
The algorithms select suitable sections from the complete W-section list given in
AISC which consists of 273 sections starting from W6 8.5 to W14 730. These
sections with their properties are used to prepare a design pool. The sequence
numbers assigned to this pool that are sorted with respect to the area of sections are
considered as design variables. In other words, the design variables represent a
selection from a set of integer numbers between 1 and 273. A population of 20
agents is used for all the algorithms. In the first and the last two examples, the
optimization process is terminated after 150 and 200 iterations, respectively.
The frames are subjected to gravity and earthquake loads and are designed
according to the LRFD-AISC design criteria [13]. Load combinations recom-
mended by ASCE 7–10 are considered in Ref. [14] and the frames are Intermediate
Moment Frames (IMF). The gravity loads are as follows: the design dead load is
2.88 kN/m2, the design live load is 2.39 kN/m2, and the ground snow load is
0.755 kN/m2 [15]. Iran has a long history of seismicity and has experienced
destructive earthquakes since ancient times. The history and the existence of faults
indicate the seismicity of metropolitan Tehran and this city are considered as high
seismicity area. Therefore, the response spectrum function of Tehran is utilized in
this study [16]. Figure 9.2 shows the design response spectrum as acceleration
versus period. The details of this figure are reported in Table 9.1. Occupancy
importance and redundancy factor are considered as 1. Modal and directional
combinations of response spectrum load cases are performed by CQC and SRSS
methods, respectively. In all the seismic load combinations, 100% of the design
seismic force for one direction is added to the 30% of the design seismic force for
the perpendicular direction. The vertical seismic load effect is also considered (SDS
is set to 1.05). For the first two examples, story drift is limited to 8.75 cm and for
the last problem, it is limited to 7 cm.
In the following examples, the height of each story is taken as 3.5 m and each
frame member is modeled as a line element. A992 steel is used with modulus of
elasticity equal to E = 200 GPa, yield stress of Fy = 250 MPa, and weight per unit
volume of q = 7850 kgf/m3. Direct analysis and general second-order methods are
used for the models. Tau-b fixed is employed for stiffness reduction method.
144 9 Optimum Seismic Design of 3D Steel Frames
The plan and 3D views of the first example are shown in the Figs. 9.3 and 9.4,
respectively. This is an irregular steel frame with 70 joints and 132 members which
are collected in 16 independent design variables. The member grouping is given in
Table 9.2. In order to optimize this structure by MDVC-UVPS, three stages are
considered. The design variable configuration utilized for the first and second stages
are listed as follows: first stage: [1 2 3 4], [5 6 7 8], [9 10 11 12], and [13 14 15 16];
second stage: [1 2], [3 4], [5 6], [7 8], [9 10], [11 12], [13 14], and [15 16].
The optimal designs found by the different algorithms are compared in Table 9.3
that shows also the corresponding structural weights and required number of
structural analyses. The MDVC-UVPS obtained 304.89 kN after 1775 analyses
which is better than 362.83 kN found by CBO after 2900 analyses, 325.98 kN
achieved by ECBO after 2740 analyses and 349.83 kN obtained by VPS after 2720
analyses. The maximum values of the stress ratio for CBO, ECBO, VPS, and
MDVC-UVPS are 85, 95, 95, and 98%, respectively. Convergence history dia-
grams are depicted in Fig. 9.5. It can be seen that the intermediate designs found by
MDVC-UVPS are always better than those found by other algorithms.
Fig. 9.3 Plan view of the four-story 132-member steel frame problem
146 9 Optimum Seismic Design of 3D Steel Frames
Fig. 9.4 3D view of the four-story 132-member steel frame problem with a U-shape form
The second design example is selected as a space four-story steel frame with 215
joints and 428 members. Figures 9.6 and 9.7 show the plan and 3D views of the
structure. This is a size optimization problem with 20 variables and the member
grouping of the frame members is listed in Table 9.4. Three stages with 5, 10, and
20 variables are considered by MDVC-UVPS algorithm. These DVCs are as
9.3 Design Examples 147
Table 9.3 Performance comparison of the four-story 132-member steel frame problem
Element group Optimal W-shaped sections
CBO ECBO VPS MDVC-UVPS
1 W21 68 W18 60 W18 97 W18 55
2 W14 82 W16 45 W18 86 W18 46
3 W14 74 W12 50 W18 76 W16 67
4 W14 99 W10 88 W16 77 W14 48
5 W16 50 W21 55 W18 50 W24 55
6 W14 68 W18 65 W16 57 W16 57
7 W12 58 W12 45 W12 50 W16 50
8 W10 88 W12 53 W12 58 W14 48
9 W14 22 W16 26 W14 22 W16 26
10 W12 26 W16 26 W12 26 W16 26
11 W14 22 W12 26 W14 22 W8 21
12 W10 22 W8 24 W10 22 W10 22
13 W12 30 W14 26 W14 26 W16 26
14 W14 26 W8 28 W12 30 W16 31
15 W12 22 W10 22 W12 19 W10 22
16 W6 25 W12 22 W12 22 W14 22
Weight (kN) 362.83 325.98 349.83 304.89
Number of structural analyses 2900 2740 2720 1775
Fig. 9.5 Convergence curves for the four-story 132-member steel frame problem
148 9 Optimum Seismic Design of 3D Steel Frames
Fig. 9.6 Plan view of the four-story 428-member steel frame problem
Fig. 9.7 3D view of the four-story 428-member steel frame problem with one plane of symmetry
9.3 Design Examples 149
Table 9.4 Member grouping of the four-story 428-member steel frame problem
Story Corner column Side column Inner column Side beam Inner beam
1 1 5 9 13 17
2 2 6 10 14 18
3 3 7 11 15 19
4 4 8 12 16 20
Table 9.5 Performance comparison of the four-story 428-member steel frame problem
Element group Optimal W-shaped sections
CBO ECBO VPS MDVC-UVPS
1 W30 90 W14 61 W27 84 W16 50
2 W21 83 W12 58 W14 68 W14 68
3 W14 68 W12 53 W14 68 W14 53
4 W14 90 W10 54 W12 72 W14 48
5 W14 38 W24 62 W18 65 W21 68
6 W14 61 W21 68 W18 65 W21 68
7 W12 87 W18 60 W18 60 W21 50
8 W12 53 W16 67 W16 36 W16 45
9 W18 106 W40 167 W24 62 W27 84
10 W16 100 W40 149 W24 62 W18 65
11 W16 89 W36 160 W14 74 W18 86
12 W14 82 W33 169 W14 61 W14 211
13 W14 26 W16 26 W16 26 W14 26
14 W10 19 W12 30 W12 26 W14 26
15 W24 55 W16 26 W14 30 W16 26
16 W5 16 W12 30 W12 30 W14 26
17 W16 31 W12 26 W16 26 W12 26
18 W12 35 W12 30 W10 26 W12 26
19 W8 24 W10 26 W12 22 W12 22
20 W16 31 W10 22 W10 22 W8 21
Weight (kN) 1247.72 1170.9 1115.68 1064.17
Number of structural 3700 3660 3520 1931
analyses
follows: first stage: [1 2 3 4], [5 6 7 8], [9 10 11 12], [13 14 15 16], and [17 18
19 20]; second stage: [1 2], [3 4], [5 6], [7 8], [9 10], [11 12], [13 14], [15 16], [17
18], and [19 20].
Table 9.5 presents the optimum designs obtained by proposed algorithms. The
lightest design (i.e., 1064.17 kg) is achieved by MDVC-UVPS algorithm after 1931
analyses. The best designs obtained by the CBO, ECBO, and VPS are 1247.72,
150 9 Optimum Seismic Design of 3D Steel Frames
Fig. 9.8 Constraint margins for the best design obtained by MDVC-UVPS algorithm in the
four-story 428-member steel frame problem: a element stress ratio and b story drift
1170.9 and 1115.68 kN, respectively. These values are found after 3700, 3660 and
3520 analyses. Figure 9.8 demonstrates the existing stress ratios and inter-story
drifts for the best designs of MDVC-UVPS. The maximum stress ratio is 93% while
maximum drift is 8.71 cm. The maximum stress ratios for the best designs of the
CBO, ECBO, and VPS are 87, 80, and 90%, respectively.
The plan and 3D views of the twelve-story frame in Figs. 9.9 and 9.10 show an
irregular steel space frame composed of 130 joints and 276 members. All members
are categorized into 24 design groups. The member grouping is given in Table 9.6.
Three stages with 8, 16, and 24 variables are considered by MDVC-UVPS algo-
rithm. These DVCs are as follows: first stage [1 2 3 4], [5 6], [7 8 9 10], [11 12], [13
14 15 16], [17 18], [19 20 21 22], and [23 24]; second stage: [1 2], [3 4], [5], [6], [7
8], [9 10], [11], [12], [13 14], [15 16], [17], [18], [19 20], [21 22], [23], and [24].
In Table 9.7, the minimum weight designs of the twelve-story 276-member steel
frame problem obtained by the proposed algorithms are compared. The
MDVC-UVPS obtained the lightest design compared to other methods that is
1644.31 kN. This algorithm requires 1576 structural analyses to find the optimum
solution while CBO, ECBO, and VPS require 3580, 3140, and 3440 structural
analyses, respectively. The maximum values of the stress ratio for the CBO, ECBO,
VPS, and MDVC-UVPS are 84, 97, 79, and 94%, respectively. Convergence his-
tory diagrams are depicted in Fig. 9.11 and demonstrated that the intermediate
designs found by MDVC-UVPS are always better than those found by other
9.3 Design Examples 151
Fig. 9.9 Plan view of the twelve-story 276-member steel frame problem
Optimum design of real-size steel space frames under design code provisions is a
complicated optimization problem due to the presence of large numbers of highly
nonlinear constraints and discrete design variables. Main purpose of this chapter is
to propose a suitable meta-heuristic optimization algorithm for optimum design of
this class of problems. Three irregular steel frame optimization problems are con-
sidered and the stress constraints of LRFD-AISC, maximum lateral displacement
constraints, and geometric constraints are imposed on all frames. Numerical results
demonstrate that the MDVC-UVPS achieves the lightest designs in all the con-
sidered examples. This algorithm also shows a good convergence rate and comes
close to the optimum design rapidly. In the first example, the best weight found by
this method is 19, 7, and 14% lighter than the weight of the best design obtained
using the CBO, ECBO, and VPS, respectively. In the second example, the optimum
152 9 Optimum Seismic Design of 3D Steel Frames
design produced by the hybrid method is 17, 10, and 5% lighter than the ones
attained by other techniques, respectively. These values are, respectively, 15, 11,
and 4% for the last example. Besides, average computational time required for
MDVC-UVPS is about 35% less than the other algorithms.
9.4 Concluding Remarks 153
Table 9.6 Member grouping of the twelve-story 276-member steel frame problem
Story Corner column Side column Side beam Inner beam
1–2 1 7 13 19
3–4 2 8 14 20
5–6 3 9 15 21
7–8 4 10 16 22
9–10 5 11 17 23
11–12 6 12 18 24
Table 9.7 Performance comparison of the twelve-story 276-member steel frame problem
Element group Optimal W-shaped sections
CBO ECBO VPS MDVC-UVPS
1 W30 173 W33 130 W33 141 W33 118
2 W27 194 W24 162 W33 130 W21 101
3 W24 146 W24 146 W33 118 W14 132
4 W24 131 W18 192 W30 132 W14 99
5 W16 67 W14 120 W14 68 W14 90
6 W14 145 W10 100 W12 79 W12 170
7 W36 231 W33 152 W33 201 W40 235
8 W33 241 W33 152 W30 211 W40 199
9 W27 258 W27 194 W30 191 W36 232
10 W24 306 W27 194 W27 194 W33 221
11 W14 176 W24 335 W24 335 W33 169
12 W12 230 W24 250 W24 250 W27 217
13 W16 40 W21 57 W21 57 W21 62
14 W16 36 W18 65 W16 67 W24 62
15 W16 45 W21 55 W21 55 W21 55
16 W18 40 W21 55 W18 65 W18 60
17 W16 45 W14 48 W12 50 W16 31
18 W8 21 W12 50 W12 53 W14 48
19 W27 217 W16 40 W16 40 W18 40
20 W24 250 W16 40 W16 45 W16 45
21 W36 182 W12 53 W12 53 W12 53
22 W33 169 W12 50 W12 50 W12 50
23 W8 31 W40 264 W27 129 W33 141
24 W12 26 W8 67 W8 67 W12 35
Weight (kN) 1896.06 1824.64 1716.41 1644.31
Number of structural 3580 3140 3440 1576
analyses
154 9 Optimum Seismic Design of 3D Steel Frames
Fig. 9.11 Convergence curves for the twelve-story 276-member steel frame problem
Fig. 9.12 Saving in structural analyses using the MDVC-UVPS algorithm in the twelve-story
276-member steel frame problem
References
4. Hasançebi O, Carbas S (2014) Bat inspired algorithm for discrete size optimization of steel
frames. Adv Eng Softw 67:173–185
5. Talatahari S, Gandomi AH, Yang XS, Deb S (2015) Optimum design of frame structures
using the eagle strategy with differential evolution. Eng Struct 9:16–25
6. Artar M, Daloglu AT (2015) Optimum design of composite steel frames with semi-rigid
connections and column bases via genetic algorithm. Steel Compos Struct 19(4):1035–1053
7. Kaveh A, Bakhshpoori T, Azimi M (2015) Seismic optimal design of 3D steel frames using
cuckoo search algorithm. Struct Des Tall Special Build 24(3):210–227
8. Carbas S (2016) Design optimization of steel frames using an enhanced firefly algorithm. Eng
Optim 48(12):2007–2025
9. Kaveh A, Bolandgerami A (2017) Optimal design of large-scale space steel frames using
cascade enhanced colliding body optimization. Struct Multidiscip Optim 55(1):237–256
10. Kaveh A, Ghafari MH, Gholipour Y (2017) Optimum seismic design of steel frames
considering the connection types. J Construct Steel Res 130:79–87
11. Kaveh A, Ghafari MH, Gholipour Y (2017) Optimal seismic design of 3D steel moment
frames: different ductility types. Struct Multidiscip Optim 56(6):1353–1367
12. Kaveh A, Ilchi Ghazaan M (2018) Optimum seismic design of 3D irregular steel frames using
modern meta-heuristic algorithms. J Comput Civil Eng, 32(3), 04018015
13. AISC 360-10 (2010) Specification for structural steel buildings. American Institute of Steel
Construction, Chicago, Illinois 60601-1802, USA
14. ASCE 7-10 (2010) Minimum design loads for building and other structures, Chicago, USA
15. Aydoğdu İ, Akın A, Saka MP (2016) Design optimization of real world steel space frames
using artificial bee colony algorithm with levy flight distribution. Adv Eng Softw 92:1–14
16. ICSRDB (2015) Iranian code of practice for seismic resistant design of buildings (Standard
2800), 4rd edn, PN S 253, Building and Housing Research Center of Iran
Appendix A
Configuration Processing
A.1 Introduction
For large systems, configuration processing is one of the most tedious and
time-consuming parts of the analysis. Different methods have been proposed for
configuration processing and data generation, among which the formex algebra of
Nooshin [1, 2] is perhaps the most general tool for this purpose. Behravesh et al. [3]
employed set theory and showed that some concepts of set algebra can be used to
build up a general method for describing the interconnection patterns of structural
systems. There are many other references on the field of data generation; however,
most of them are prepared for specific classes of problem. For example, many
algorithms have been developed and successfully implemented on mesh or grid
generation, a complete review of which may be found in a paper by Thacker [4] and
a book by Thomson et al. [5].
Graph theoretical methods for the formation of structural and finite element
models are developed by Kaveh [6, 7]. In all these methods, a submodel is
expressed in algebraic form and then the functions are used for the formation of the
entire model. The main functions employed consist of translation, rotation,
reflection, and projection, or combination of these functions. On the other hand,
many structural models can be viewed as the graph products of two or three
subgraphs, known as their generators. Many properties of structural models can be
obtained by considering the properties of their generators. This simplifies many
complicated calculations, particularly in relation with eigen solution of regular
structures, as shown by Kaveh and Rahami [8, 9] and Kaveh and Mirzaie [10]. Four
undirected and directed graph products are presented in [11] for the formation of
structural models. The undirected products are extensively used in graph theory and
combinatorial optimization, however, the directed products are more suitable for the
formation of practical structural models.
Different structures are studied in this book and all of them, except the 3D
frames, are analyzed using the direct stiffness method by our own codes. In this
appendix, MATLAB codes for generation of the coordinates and connectivities of
three different types of structures utilized in this book are developed. These
A.2 Examples
A.2.1 Example 1
Fig. A.2 a bottom layer, b top layer, and c web members of the 1016-bar double-layer grid
160 Appendix A: Configuration Processing
% Bottom-layer elements
% elements: connections at elements
elements = [];
elements = [elements; [1:NUM_X-1; 2:NUM_X]'];
for ny=1:NUM_Y
firstNodeBottomRow = (2 * NUM_X + 1) * (ny - 1) + 1;
firstNodeTopRow = (2 * NUM_X + 1) * (ny - 1) + NUM_X + 1;
for nx=1:NUM_X
elements = [elements; [firstNodeBottomRow + nx - 1, firstNodeTopRow +
nx - 1]];
Appendix A: Configuration Processing 161
% Top-layer elements
firstNodeTopLayer = (2 * NUM_X + 1) * NUM_Y + NUM_X + 1;
for ny=1:NUM_Y
firstNode = firstNodeTopLayer + (ny - 1) * NUM_X;
elements = [elements; [firstNode:firstNode+NUM_X-2;
firstNode+1:firstNode+NUM_X-1]'];
if ny ~= NUM_Y
elements = [elements; [firstNode:firstNode+NUM_X-1;
firstNode+NUM_X:firstNode+2*NUM_X-1]'];
end
end
% Web elements
for ny=1:NUM_Y
firstNodeBot = (2 * NUM_X + 1) * (ny - 1) + 1;
firstNodeTop = firstNodeTopLayer + (ny - 1) * NUM_X;
elements = [elements; [firstNodeBot:firstNodeBot+NUM_X-1;
firstNodeTop:firstNodeTop+NUM_X-1]'];
for nx=1:NUM_X
elements = [elements; [firstNodeBot + NUM_X + nx - 1, firstNodeTop +
162 Appendix A: Configuration Processing
nx - 1]];
elements = [elements; [firstNodeBot + NUM_X + nx, firstNodeTop + nx –
1]];
end
elements = [elements; [firstNodeBot+2*NUM_X+1:firstNodeBot+3*NUM_X;
firstNodeTop:firstNodeTop+NUM_X-1]'];
end
A.2.2 Example 2
The double-layer barrel vault presented in Fig. A.3 consists of 693 members and
259 joints. The geometry and the member’s labels of the structure are shown in
Fig. A.4.
In geometry, a circular segment is a region of a circle which is “cut off” from the
rest of the circle by a chord. Let R be the radius of the circle, h the central angle,
c the chord length, s the arc length, h the height of the segment, and d the height of
the triangular portion (Fig. A.5); therefore, we have
h c2
R ¼ hþd ¼ þ ðA:1Þ
2 8h
c
h ¼ 2 arctan ðA:2Þ
2d
For the top layer of the barrel vault, hT = 5.75 m and cT = 22.9 m. Therefore, RT
and hT are 14.275 m and 1.862 rad, respectively. For the bottom layer, hB = 4.25 m
and cB = 19.03 m. Thus, RB and hB are 12.776 m and 1.680 rad, respectively.
The connections at elements and node coordinates of the double-layer barrel
vault can be generated by the following MATLAB code.
Fig. A.4 a flatten cross-sectional view and b plan view of the 693-bar double-layer barrel vault
Fig. A.5 A circular segment is enclosed between a chord (the dashed line) and the arc whose
endpoints equal the chord
164 Appendix A: Configuration Processing
% Top-layer elements
% elements: connections at elements
elements = [];
for nz=1:NUM_Z+1
firstNode = (nz - 1) * (NUM_X + 1) + 1;
elements = [elements; [firstNode:firstNode+NUM_X-1;
firstNode+1:firstNode+NUM_X]'];
if nz < NUM_Z+1
for nx=1:NUM_X
if (nx <= NUM_X/2 && nz <= NUM_Z/2) ||
(nx > NUM_X/2 && nz > NUM_Z/2)
elements = [elements; [firstNode + nx, firstNode + nx +
NUM_X]];
else
elements = [elements; [firstNode + nx - 1, firstNode + nx +
NUM_X + 1]];
end
if nx ~= NUM_X
elements = [elements; [firstNode + nx, firstNode + nx +
NUM_X + 1]];
end
end
end
end
% Bottom-layer elements
for nz=1:NUM_Z+1
firstNode = (NUM_X + 1) * (NUM_Z + 1) + (nz - 1) * NUM_X + 1;
elements = [elements; [firstNode:firstNode+NUM_X-2;
firstNode+1:firstNode+NUM_X-1]'];
end
% Web elements
for nz=1:NUM_Z+1
firstNodeBot = (NUM_X + 1) * (NUM_Z + 1) + (nz - 1) * NUM_X + 1;
firstNodeTop = (nz - 1) * (NUM_X + 1) + 1;
for nx=1:NUM_X-1
elements = [elements; [firstNodeBot + nx - 1, firstNodeTop + nx]];
elements = [elements; [firstNodeBot + nx, firstNodeTop + nx]];
end
end
166 Appendix A: Configuration Processing
A.2.3 Example 3
The dome truss shown in Fig. A.6 consists of 600 elements and 216 joints. Fig. A.7
presents a substructure in more detail for nodal numbering and coordinates. The
following MATLAB code is developed to generate the dome from its substructure.
% Node coordinates
% nodes: node coordinates
nodes = [];
for ns=1:NUM_S
nodes = [nodes; [RADIUS * cos((ns-1) * THETA); RADIUS * sin((ns-1) *
THETA); Z_COORDS]'];
end
% Elements
% elements: connections at elements
elements = [];
for ns=1:NUM_S
firstNodeCur = (ns - 1) * NUM_NODE + 1;
if ns ~= NUM_S
firstNodeNext = firstNodeCur + NUM_NODE;
else
firstNodeNext = 1;
end
elements = [elements; [firstNodeCur:firstNodeCur+NUM_NODE-2;
firstNodeCur+1:firstNodeCur+NUM_NODE-1]'];
elements = [elements; [firstNodeCur, firstNodeCur + 2]];
elements = [elements; [firstNodeCur:firstNodeCur+NUM_NODE-2;
firstNodeNext:firstNodeNext+NUM_NODE-2]'];
elements = [elements; [firstNodeCur+2:firstNodeCur+NUM_NODE-1;
firstNodeNext+1:firstNodeNext+NUM_NODE-2]'];
elements = [elements; [firstNodeCur, firstNodeNext + 1]];
end
168 Appendix A: Configuration Processing
References
1. Nooshin H (1975) Algebraic representation and processing of structural configurations.
Comput Struct 5(2):119–130
2. Nooshin H (1984) Formex configuration processing in structural engineering. Elsevier
Applied Science Publishers Ltd, Essex, U.K.
3. Behravesh A, Kaveh A, Nani M, Sabet S (1988) A set theoretical approach for configuration
processing. Comput Struct 30(6):1293–1302
4. Thacker WC (1980) A brief review of techniques for generating irregular computational grids.
Int J Numer Methods Eng 15(9): 1335–1341
5. Thompson JF, Warsi ZU, Mastin CW (1985) Numerical grid generation: foundations and
applications. Elsevier, Amsterdam
6. Kaveh A (1993) A graph-theoretical approach to configuration processing. Comput Struct 48
(2):357–363
7. Kaveh A (2004) Structural mechanics: graph and matrix methods. 3rd edn. Research Studies
Press, Somerset, UK
8. Kaveh A, Rahami H (2006) Block diagonalization of adjacency and Laplacian matrices for
graph product; applications in structural mechanics. Int J Numer Methods Eng 68(1):33–63
9. Kaveh A, Rahami H (2005) A unified method for eigendecomposition of graph products. Int J
Numer Methods Biomed Eng 21(7):377–388
10. Kaveh A, Mirzaie R (2008) Minimal cycle basis of graph products for the force method of
frame analysis. Int J Numer Methods Biomed Eng 24(8):653–669
11. Kaveh A, Koohestani K (2008) Graph products for configuration processing of space
structures. Comput Struct 86(11–12):1219–1231