Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Journal of Building Engineering 76 (2023) 107017

Contents lists available at ScienceDirect

Journal of Building Engineering


journal homepage: www.elsevier.com/locate/jobe

Aerodynamic optimization for corner modification of


octagonal-shape tall buildings using computational approach
A.H.H. Al-Masoodi b, Y.M. Abbas a, F. Alkhatib d, M. Iqbal Khan a, *, N. Shafiq b,
Mohamed ElGawady c
a
Department of Civil Engineering, College of Engineering, King Saud University, P.O. Box 800, Riyadh, 11421, Saudi Arabia
b
Department of Civil and Environmental Engineering, Universiti Teknologi PETRONAS (UTP), Perak, Malaysia
c
Department of Civil, Architecture and Environmental Engineering, Missouri University of Science and Technology, Rolla, MO, USA
d
Faculty of Civil Engineering and Built Environment, Universiti Tun Hussein Onn Malaysia (UTHM), Johor, Malaysia

A R T I C L E I N F O A B S T R A C T

Keywords: Tall buildings are particularly susceptible to wind loads, which usually govern the design of
Wind load lateral load-resisting systems. Therefore, wind loads must be adequately evaluated in the design
Drag of tall buildings. Aerodynamic modifications are highly effective tools for reducing wind loads.
Aerodynamics This paper investigates the effectiveness of corner modification optimization applied on an
CFD octagonal-plan-shaped model using computational fluid dynamic (CFD) simulation computa­
Tall buildings
tional fluid dynamics associated with finite element analysis to alleviate wind-induced loads.
Optimization
Corner aerodynamic modifications such as chamfered, recessed, rounded, and fins are investi­
gated. The corner modification was limited to a cutting radius of 6 m (12% of the building width)
with a 0.5 m increment. The main considerations for this optimization procedure are top
deflection, inter-story drifts, and the optimal number of additional floors. All corner modifications
improve the building’s performance, except fins corners resulting in adverse effects. In addition,
47 simulation examples from the case study are evaluated, presented, and discussed. With one
additional floor, the optimum shape was able to reduce overall wind loads by 31.67%, resulting in
a reduction in the structural response of 24.89% and 24.18% in maximum top deflection and
inter-story drift, respectively.

1. Introduction
In the modern age, taller and slenderer buildings are constructed due to new construction techniques and materials to develop a
city’s visual identity. However, as the height of a tall building increases, it becomes more sensitive to lateral forces, leading to an
increase in the wind-induced loads and motions that are considered to govern the design of the lateral load [1]. There are important
factors, either locally, such as building shape, or globally, such as surroundings buildings, that impact aerodynamically wind load.
Each building’s shape and surroundings generate a distinct combination of design wind loads. On the other hand, the building’s shape
creates an opportunity to mitigate wind load through outer shape modifications, either major or minor modifications. In that context,
major modification requires significant changes to the building’s shape, significantly impacting the overall architectural form and
structural function (e.g., openings, tapering, twisting, set-backing, etc.). Thus, it is necessary to confirm the significant modification in
early conceptual design [2]. Minor modifications (e.g., corner modification) for the building shape, in contrast, have only a small

* Corresponding author. Department of Civil Engineering, King Saud University, Riyadh, 800-11421, USA.
E-mail address: miqbal@ksu.edu.sa (M.I. Khan).

https://doi.org/10.1016/j.jobe.2023.107017
Received 1 February 2023; Received in revised form 11 May 2023; Accepted 4 June 2023
Available online 15 June 2023
2352-7102/© 2023 Elsevier Ltd. All rights reserved.
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

impact on the architecture and structural designs (e.g., rounding, chamfering, slotting, recessing, etc.). The architects can thus modify
the form of the building’s corners at a later conceptual design stage.
The effects of three wind force components, along wind (drag force), across the wind (lift force), and torsional moment, applied on a
tall building are mainly controlled by the geometric form, as shown in Fig. 1. The geometric shape also determines the formation of the
vortex-shedding forces. Therefore, at the early design stage of a project, it was suggested that performing aerodynamic modifications
could minimize wind effects on vortex-shedding turbulence with the height of supertall buildings (e.g., Burj Kalifa [3], which has a
height of 828 m). According to Equation (1), the building’s width and shape would significantly impact the wind’s force. In this
formula, St is the Strouhal number, f is vortex shedding frequency, D is the structure’s cross-sectional area, and V is the flow velocity. St
is a nondimensional number that ranges between 0.1 and 0.3, or approximately 0.14 for square shapes and 0.2 for circular shapes [4,
5].
St = fD/V (1)
Tall buildings’ outer shapes are typically aerodynamically “bluff” and have sharp corners. Many numerical and experimental wind
engineering studies have been conducted to investigate wind loads for tall buildings of various shapes [6–13]. Moreover, modifications
in a minor or major way for the tall building can help to the reduction in the aerodynamic effect. Minor modification of tall buildings is
preferred over major ones to preserve the architectural form and structural system. Table 1 summarizes the corner modification
methods and the main finding of previous studies based on aerodynamic modifications of tall buildings. Corner modification, imposing
a slight modification on building corners, is a common type of minor modification. The chamfered, rounded and recessed corners are
among the most popular corner modifications for studying the wind aerodynamic effect in tall buildings [14]. Kowk [15] established
that the chamfered corner slots and horizontal corner modifications significantly impacted along-wind and across-wind responses.
Chamfered and rounded corner modifications are considered to be the smallest changes in the building architecture and are suitable for
modifying the cross-section of the building that is reportedly designed from the perspective of wind resistance [16]. Another study
reported that this modification could reduce the drag forces [17]. Moreover, recessed corner modification reported that a 7.5%
recession ratio was most effective for reducing the aerodynamic forces coefficients [18]. Furthermore, Kwok and Bailey [19] estab­
lished that the fin corner modifications increase along-wind response while reducing across-wind response.
The early study used aeroelastic and HFFB to study tall buildings’ aerodynamic mitigations [15,19]. These methods are reliable, but
they are only helpful in comparing a limited number of feasible building shapes, and it is expensive for repeated investigation [15]. A
recent method that is used for aerodynamic mitigations is computational fluid dynamics (CFD) [20,21]. This method is integrated into
a parametric environment and connected to finite element analysis software for assessment response of the tall building structure. This
integrating CFD with an optimization algorithm is more helpful in exploring the broader alternative structures of a tall building to find
the optimal shape [8]. This encouraged researchers to work on applying aerodynamic optimization of tall buildings [14,21,22]. Sanyal
and Dalui [23] reported that the most effective in mitigating wind load and overturning moment is the rounded one using the CFD
method. Also, Dai et al. established by utilizing the CFD method that the rounded and chamfered roofs showed higher velocities and
lower turbulence intensities than the recessed roofs [21]. Aerodynamic modifications lead to decreased concrete volume, which can
improve sustainability due to lowering embedded carbon [24–26].
The current study is focused on the CFD analysis of an octagonal-shaped building subjected to various aerodynamic modifications
at the corners. The study is currently being done for modifications such as chamfered, recessed (single & double), rounded, and fins.
The study is carried out for one-direction wind (at angles of 0). The primary goal of this paper was to comprehend and compare the
efficiency behavior of each type of corner modification as applied to an octagonal tall building using a performance-based aero­
dynamic optimization approach, as well as to opt for the optimum design among those investigated, which is the novelty of this paper.
Furthermore, unlike seismic engineering, this research develops the application of performance-based wind design of tall buildings,
which has not been sufficiently investigated and adopted in wind engineering. The aerodynamic efficiency of the different modifi­
cations was studied, presented, and discussed based on structural performance, such as top deflection and inter-story drift, as well as an
additional floor for the optimum design to maintain the same total area.
Aerodynamic adjustments can efficiently lower wind load and structural costs but reduce the saleable floor space. Thus, addressed
aerodynamic adjustments and proposed evaluating the wind loadings and wind-induced responses of original shape buildings (sharped
corners) with another corner aerodynamic modification, such as recessed and chamfered corners, while the increase in the number of
stories to maintain the same total useable floor area of the entire building with an original design.

Fig. 1. Force directions and periodic vortex shedding applied on a tall building.

2
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Table 1
Corner modification, method, and main findings of previous studies focused on minor modifications of tall buildings.

Refs. Corner Modification Method Finding

[19] Vertical fins, vented fins, BLWT Fins increase alon1g-wind response while reducing across-wind response.
corner slots
[15] Chamfered, corner slots, BLWT There was a significant impact on along-wind and across-wind responses.
horizontal slots
[16] Chamfered and rounded BLWT It was also observed that a super high-rise building of 600 m can be designed from the perspective of wind
resistance by modifying a suitable cross-section of the building.
[27] Chamfered and recessed BLWT mitigate across-wind
force for normal wind
incidence angle
[17] Chamfered and rounded BLWT The separated shear layers approach the side surface with corner cutting and corner rounding, thus
promoting reattachment and reduction of drag forces.
[28] recessed and chamfered BLWT The structural responses and the effects on economic rewards must be taken into consideration for deciding
the corner modification can be chosen.
[29] Chamfered and recessed BLWT Investigated across-wind aerodynamic damping and developed a pair of formulas for cross-wind force PSD,
moment coefficient, and shear forces
[18] recessed CFD According to the model results, two types of corner recession significantly impact the aerodynamic
coefficients of base moment and torque, and a 7.5% recession ratio was most effective for reducing the
aerodynamic forces coefficients.
[8] corner configuration CFD the author has found that the proposed model shows a reduction of about 30% in both along-wind and
across-wind response through a two-surface chamfering which is not more than 20% of the building width.
[30] Chamfered BLWT It is illustrated that chamfered corner effectively reduces most wind force coefficients Cs to less than 0.9.
[14] recessed, chamfered, and CFD Corner chamfered is the most efficient for reducing along-wind loads, with a base moment reduction.
rounded
[23] recessed, chamfered, and CFD The rounded one is the most effective in mitigating wind load and overturning moment. The main issue with
rounded rounded corners is to increase in airflow at the corner zones
[31] recessed and chamfered BLWT Recessed and chamfered corners both help to reduce vortex shedding compared with the squire section.
[22] Chamfered CFD the chamfered corner may reduce drag force and moment by up to 25% and enhance the structure
performance by up to 12%.
[21] recessed, chamfered, and CFD The rounded and chamfered roofs showed higher velocities and lower turbulence intensities than the
rounded roofs recessed roofs,
[20] Chamfered CFD Chamfered corners enhance to reduce the force and moment coefficients, and 30◦ AOA is optimum.
[2] Rounded CFD Aerodynamic improvement of up to 14% of structure performance

*BLWT- Boundary-Layer Wind Tunnel, CFD-Computational Fluid Dynamics.

2. CFD validation with experimental results of CAARC building


Validation and verification of relevant models are required to ensure the accuracy and validity of numerical results. The
Commonwealth Advisory Aeronautical Research Council (CAARC) standard tall building model is used for calibrating numerical
results [32,33], as shown in Fig. 2. In this study, we used TJ University’s wind tunnel test results with 0 wind direction as a reference to
accomplish the sensitivity analysis and assess the mean wind pressure coefficient (Cp) of the CFD method of the study [32].
An entire building model can be recognized in numerical simulation, but it is necessary to reduce tasks by reducing grid number and
computational domain size. The conditions effects of blockage effects must be less than 5% during the computational setting to achieve
maximum similarities between the numerical model and the wind tunnel test model [32], as shown in Equation (2).
δ = A0 /A (2)

where δ represents the blockage ratio, A0 indicates the area of the building wall perpendicular to the wind direction, and A is the area
of the computational domain perpendicular to the wind direction, as illustrated in Fig. 1.

Fig. 2. Configuration of the computational domain and measuring points of the CAARC model [32].

3
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

In this study, The Outflow boundary condition was employed to simulate steady and incompressible wind flow. The computational
domain’s wall and the surface of the tall CAARC standard building have been defined using wall boundary conditions [32] and the
CAARC model was performed with Butterfly in Grasshopper tool. SIMPLE pressure–velocity coupling algorithms were adopted for the
solver setting, and the second-order upwind scheme was adopted to discretize momentum, turbulent kinetic energy, and turbulence
dissipation rate. The processor used is an Intel (R) Core (TM) i7-9750H CPU running at 2.60 GHz with 32.0 GB RAM and parallel
decomposition. The CAARC standard building was simulated in a rectangular building model with similar length, width, and height
dimensions of 30.48, 45.72, and 182.88 m, respectively. Due to sufficiency for model calibration [32,34], 20 pressure coefficient Cp
wind pressure taps were chosen at 2/3 height of the building from the ground, as shown in Fig. 1.
As an outcome, the pressure of the wind coefficient Cp for these taps has been determined with respect to a benchmark static point
located at the flow inlet within the central location of the building x-axis and at the relevant height of the building. The tunnel’s
computational domain was replicated identically to the one from the reference paper with 12.7 m/s of velocity. The computational
domain’s dimensions were 900 m × 600 m x 400 m (L × B × Z), as shown in Fig. 2. The building stands along the tunnel’s x-axis and
within 300 m of the inlet point of flow, allowing enough space for the air flow to develop fully within an acceptable calculated blockage
ratio of 3.54%, which is less than the suggested limit of 5% [32,35,36]. L1 and L2 were set to 300 m and 569.52 m to allow wind to
develop appropriately.
The wall function was utilized in each tested model to satisfy the physics in the near wall zone by encompassing the inner region
between the wall and the turbulence fully developed region. The mesh size grid was also set consistently for the models with a
maximum grid size of 0.00054 H [32], equivalent to 0.98755 m and 678,600 cell numbers. To limit iteration numbers within an
acceptable computational time, the convergence threshold of residuals was set to 0.0001, as suggested by Frank et al. [37].

2.1. Mesh grid size


Mesh grid sizes were determined based shortest building width (B), which is equal to 30.45 m in the CAARC building, and the
number of refinements (n) [Equation (3)]. For more refinement, it was required to increase n according to Frank et al. [37] for CFD
simulation guidelines. The mesh size has been evaluated for turbulence mode purposes using Realizable (k-ε), the most accurate
turbulence mode mentioned in Meng et al. [37]. Three refinements have been done for mesh types, as shown in Fig. 3. Thus, a fine grid
was used to investigate the turbulence mode for more accuracy, as illustrated in Fig. 4.
( )
B
Mesh grid size = √ ̅̅
̅ (3)
10 n− 1 2
As shown in Fig. 7 in the below subsection, the most accurate turbulence mode is RNG (k − ε), so the mesh grid sizes were re-
evaluated based on that mode for this simulation study. To further verify mesh grid size with RNG, 5 numbers of refinement has
been performed, namely, coarse, medium, fine, X-fine, and XX-fine, as shown in Table 2. As indicated in Fig. 5, the absolute deviations
in wind pressure coefficient points of all surfaces have been calculated and verified based on XX-fine as representing the most
refinement with the longest computational time (3.9Hrs). The fine mesh was adopted based on the result, as it has similar accuracy
with decreasing time-consuming e up to 60% compared to the XX-fine mesh type.

2.2. Turbulence models


Three turbulence models were evaluated on wind pressure coefficients about TJ experimental results: RNG (k − ε), realizable (k −
ε), and standard (k − ε) as illustrated in Fig. 6. Other variables, such as wind profile and speed, remained constant during the validation
of turbulence models. The mesh size of the fine grid was also set consistently for the three models, as shown in the subsection.
Fig. 7 illustrates the results of the three evaluated turbulence models. It is indicated that the total absolute deviation of RNG (k − ε)

Fig. 3. Absolute deviation of Cp for different mesh types based on TJ experimental results.

4
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 4. Fine mesh grid for turbulence mode investigation.

Table 2
Mesh type and refinements.

Mesh Type Coarse Medium Fine X-fine XX-fine

Refinement No.(n) 1 2 3 4 5
Grid size(m) 3.045 2.15 1.52 1.08 0.76

Fig. 5. Absolute deviation of Cp for different mesh types based on XX-fine mesh type.

was 9.3%, showing reasonable agreement with the TJ model, which is within ±15%, as recommended by Meng et al. [37]. In addition,
for points 1–5 in RNG and realizable, absolute deviations on Surface-I are below 5%, but other surfaces are still between 5% and 15%.
However, the absolute deviation of standard models achieved less accuracy, with a 29% difference in TJ experimental result. Thus,

Fig. 6. Cp values for different turbulence models with TJ experimental results.

5
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

RNG (k − ε) is the most accurate, which can be used for this study simulation. However, all three models displayed similar behavior to
the experimental result, reflecting reliable numerical output validity, as shown in Fig. 6.

3. Aerodynamic optimization procedure


The Aerodynamic Optimization Procedure can be employed to investigate various corner configurations such as rounding,
chamfering, single recessing, double recessing, fins, and so on, as shown in Fig. 8. It should be observed that, besides improving
aerodynamic performance, the building form is commonly constrained by other architectural and structural criteria. Thus, a suitable
selection of design parameters ensures that the optimum configuration meets architectural and structural design aims. In this study, the
Aerodynamic optimization Technique attempts to mitigate the drag force, lift, and/or torsional moment, as indicated in Fig. 1. Thus,
the top deflection of the building can be reduced, which must be less than H/500 as specified in the Eurocode design code standard. It
can be done by searching for the best corner modification by considering the area reduction to maintain the same useable floor area for
the entire building [22]. In the current study, the shape of corners is controlled by one geometric design variable, corner cutting radius
(R), which was programmed by using the optimization algorithm for changing the shape of the corners for all types of corner
modification in the fastest way (cutting radius ranges from 0 to 6 m with 0.5 increments).
Every round of this optimization cycle involves a particular procedure, such as computational fluid dynamics (CFD), fluid-structure
interaction (FSI), and finite element analysis (FEA), as shown in Fig. 8. The stages are illustrated and can be summarized as follows:

- A three-dimensional parametric model of the building’s geometries with an octagonal shape.


- Corner cutting radius (R) is identified as a building corner variable of aerodynamic modification types, which ranges from 0 to 6 m
with 0.5 m variations
- The computational component (CFD-FSI-FEA) investigates all aerodynamic modification types.
- Optimal aerodynamic shape is obtained from top deflection and the inert-drift story of the building with maintaining the same floor
area as the original building shape.

4. Design scenario
The following design studied in this investigation was an initial application of the proposed design process on Tower 23-Marina
(392 m), Dubai’s third tallest building, as illustrated in Fig. 9. Tower 23-Marina was carefully chosen because of its octagonal ge­
ometry, which can be effective for corner modification investigation in this study by using one typical floor with a height of 360 m.
Furthermore, it presents a challenging exercise to incorporate minor aerodynamic modifications that maintain the intended purpose of
the architectural form and structural design to further mitigate wind effects with a less expensive and practical computational method.
According to the previous studies (See Table 1) and the specific building configuration of this case study, the current research
focuses on one parameter (i.e., corner fillet Radius, R) to be implemented for all types of corner modification for aerodynamic opti­
mization, except for Fins Corner, which used the length of the Fin as variable input (Fig. 10). These modifications allow for aero­
dynamic modification while maintaining the tall building’s overall architectural style shape and structural function. All corners were
evaluated with a variable Corner Fillet Radius (R) ranging from 0 to 6 m with 0.5 m increments, which can be modified from 0 to 12%
of building weight, as illustrated in Fig. 11. However, the Fin corner length (L) was only investigated with 0.5 m and 1 m due to its
reverse effect on wind load. The building layout plan is measured for all the variables of corner modification type for an economic
purpose [22], as shown in Fig. 12.

4.1. CFD process


A wind tunnel was established for evaluating the models of corner modification type. According to [38], the tunnel domain size was
established depending on building height (H), estimating 2.3 m for windward, sides, and top and 10 H for leeward, as shown in Fig. 13.
A blockage ratio of 1.03% was determined using a Python script, ensuring that it was less than the recommended blockage ratio of 5%

Fig. 7. Comparison in Cp values between numerical results of turbulence mode based on TJ experimental result.

6
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 8. Aerodynamic optimization procedure.

Fig. 9. (a) 23-Marina Tower, Dubai, and (b) Floor plan of the tower.

[9]. Another script was written to calculate mesh cell sizes in relation to the dimensions of the building. As investigated in the CFD
validation section, fine mesh size was used to balance the accuracy of the findings and the computational time, as illustrated in Fig. 13.

4.2. Coupling of the fluid-structure interaction


The fluid-structure interaction (FSI) coupling algorithm was created in Python to generate CFD output pressures into FE input
loads. Fig. 14 illustrates the FSI coupling algorithm that developed the parametric model of the building, then, the FSI translation
procedure commenced by converting the pressure applied by the building mesh surfaces into vector forces as a single normal force on
each mesh area. The collected vectors are then combined vertically and horizontally for each story height to generate one point load

7
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 10. Corner aerodynamic modification of the tall octagonal building.

Fig. 11. Modification percentage of the building size.

Fig. 12. Area reduction due to Corner aerodynamic modification.

(resultant force), which can be separated in both the X and Y directions. A torsional moment is applied at the middle of each story
diaphragm.

4.3. FE analysis
The force reactions were transferred to the finite element program (ETABS) for analysis for evaluating the structural performance to

8
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 13. Virtual wind tunnel with fine mesh for the case study.

Fig. 14. (a) Parametrical model, (b) Vectors flow toward the building, (c) resultant force of wind load each story, (d) Structural Model, (e) Deformed model due to
wind load.

wind effect as illustrated in Fig. 14 (d) & (e). Structural elements, including the lateral resisting system, were all assumed to simplify
the original building design, and they maintained the floor plan constant throughout the structure’s height. A rigid diaphragm was
considered to achieve maximum translation of forces and displacements to the lateral structural system without deforming the
contribution from slab stiffness. For each floor, forces were applied to the center of the diaphragm, representing the center of mass on
that floor, so that, in addition to the applied torsional moment, eccentricity between the center of rigidity and the center of mass is
considered by FE analysis for torsional displacement.

5. Optimization results and discussions


5.1. CFD generated wind forces
The line graphs illustrate the wind forces applied on the tall building, along-wind load, across-wind load, and torsional moment for
all the types of corner aerodynamic modification from 0 to 6 m in R, where corner cutting radius 0 represents the basic model with no
corner modification. Fig. 15 illustrates the effect of the corner-cutting radius of the original model, with the corner-cutting radius
increasing from 0 m to 2.5 m, the along-wind load for all the types showed a sharp decrease of about 17%–24%. After that, they would
conversely increase slightly, then reduce, and finally conversely decrease. As seen from Fig. 15, the drag force reduction reaches its
maximum at 5 m of R with the chamfered corner and for both recessed corners among the corner modification types, about 33% and
29.5%, respectively. Then, it would conversely increase if the drag force continues to grow over 5 m in R, which is the same findings as

9
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Zhang et al. [18] using the high-frequency force balance (HFFB) technique. In addition, the lowest drag force reduction showed on the
rounded corner at about 16% at 3.5 m in R. It continued with a gradual increase to 14% at 5 m. Beyond 5 m, both recessed corners
experienced a rapid increase to less than 19% of drag force reduction, while rounded corners reduced to 17.5%. The regularity of the
vortex shedding and leeward negative pressure that apply to the octagonal shape may explain the ineffective to reduce drag force,
further increasing the corner cutting radius over 5 m.
On the contrary, both across-wind load and torsional moment have an insignificant effect, which went downward and upward
between − 1500 kN and +1500 kN and − 2500 kN m and +2500 kN m sequentially (Figs. 16 and 17) for all the types of corner
aerodynamic modification. It was because the oncoming wind was applied in one direction toward the tall building with a constant
layout.
Fig. 18 represents the generated forces of the fin corner by CFD in along-wind, across-wind directions, and torsional moment,
respectively, where the along-wind force showed a slight increase from o to 0.5 m of fin length (L) and continued sharply increase
about 24% in comparison to the original shape at 1 m in L. Similar to the above corner modification, across-wind forces and torsional
moment have no significant change compared with along-wind force.

5.2. Tall building performance constraints


Fig. 19 shows the maximum top deflection of the tall building. It is also noticed that the behavior of maximum top deflection has the
same impact with along-wind force plots. However, the aerodynamic improvement can be observed for the optimal maximum top
deflection with the chamfered corner at 5 m of R, about 27% compared with the original shape. However, the rounded corner was the
lowest among the corner modifications, with 14% of aerodynamic improvement (see Fig. 20).
Similarly, for aerodynamic improvement, Figure. Twenty plots maximum inter-story drifts due to wind forces. It indicates the
highest improvement in the aerodynamic responses with the chamfered corner with an R of 5 m, up to 26% compared to the original
shape. In comparison, the rounded corner with an R of 3.5 m, was the lowest improvement among these corner modifications, about up
to 13.5%.

5.3. Optimum design


Based on the results above, it is noticed that the optimum shape among all the types and their R variable is chamfered corner shape
with an R of 5 m. Based on the flowchart in Fig. 8, we need to maintain the same useable area of the optimum shape by adding an extra
floor. Moreover, the useable area per floor is calculated roughly in the case study as about 1200 m2. As illustrated in Fig. 12, the
reduced area of the optimum shape is about 1080 m2 (12 m2 per floor), then the optimum shape model is added to one floor. A final
check is carried out through the reduction of maximum top deflection, and inter-story drift is 350 mm and 1.2 × 103, respectively,
which is the optimum design among other types with 90 floors, as illustrated in Figs. 22 and 23. Thus, the chamfered corner with R of 5
m with an extra one floor (91 floor) is considered the optimum design for both structural performance and useable area of the floor, so
the aerodynamic improvement due to reduction in wind-induced motion and forces will result in a considerable saving in the required
building materials, damping systems and consequently the building cost. Fig. 21 depicts the generated aerodynamic forces by CFD in
wind load for both the original building shape with 90 floors and the optimum shape with 91 floors. It is noticed that the total optimal
loads are reduced by 31.6% in comparison to the original shape.
In addition, to assess the effectiveness of the suggested optimization methodology, the displacement profile of the original building
was compared to the optimal shape. Fig. 22 illustrates the displacement profile for the optimum and original designs due to wind load.
Compared to the original, the optimum design achieved a maximum top deflection reduction of up to 24.9%.
Likewise, the inter-story drift of the optimum and original designs are compared, as shown in Fig. 23 for wind directions. The
maximum inter-story drifts can be observed on floors between level 77 to level 81, as illustrated in Fig. 21. Furthermore, the optimum
design achieved a maximum inter-story drift reduction of up to 24.9% compared to the original.

Fig. 15. Total drag force of corner modification types.

10
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 16. Total lift force of corner modification types.

Fig. 17. Total of the torsional moment of corner modification types.

6. Conclusions
The current research proposes a rigorous aerodynamic optimization procedure that combines computational tools such as a
parametric model, computational fluid dynamics, fluid-dynamic interaction, and structural finite element analysis. It shows its effi­
ciency during optimizing the building’s top deflection and inter-story drift due to optimizing wind loads. The case study employed in
this research was a 90-story supertall building (360 m) with an octagonal geometry shape (suitable for corner modification study). The
proposed method and techniques were able to alleviate the impact of wind loads on the tall building by modifying the radius and length
of the corner of the building, with really no impact on the original architectural form and structural design. The effects of the corner
aerodynamic modifications of the building were compared those of to the original shape, and the following conclusions were drawn:
- A significant aerodynamic improvement of the building was presented, although the applied wind load was in one direction only in
this study.
- The along-wind load for all corner aerodynamic modification types was significantly reduced compared to the original shape. On
the contrary, both the across-wind load and the torsional moment have insignificant effects due to the uniformity of the building
section and the constant width in the octagonal shape when the R is changed.
- Fin corners are ineffective in this study and have an adverse impact on tall building performance. These types should be used with
caution to avoid any negative consequences.
- The reduction area of the floor layout due to aerodynamic modification of the optimum design needs to be considered by adding the
optimum number of additional floors to maintain the same total area of the original design.
- With one additional floor due to corner modification in optimum design, the aerodynamic improvement is about 31.67%. Thus, the
maximum top deflection and inter-story drift reduction are 24.89% and 24.18%, respectively.
- The corner aerodynamic optimization, with the combination of computational techniques, assists engineers and architects in
making decisions by incorporating computational technological tools in an automated manner during the preliminary design
stages.

11
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 18. Wind forces of fins corner type.

Fig. 19. Maximum top structural displacement of corner modification types.

Credit author statement


A.H.H. Al-Masoodi: Conceptualization, Methodology, Software, Original draft preparation.
Y.M. Abbas: Visualization, Investigation, Data curation, Software, Validation.
F. Alkhatib: Visualization, Investigation, Data curation.
M. Iqbal Khan: Supervision, Validation, Writing- Reviewing and Editing.
N. Shafiq: Conceptualization, Methodology, Original draft preparation.
Mohamed ElGawady: Visualization, Writing- Reviewing and Editing.

Declaration of competing interest


The authors declare no conflict of interest.

12
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 20. Maximum inter-story drifts of corner modification types.

Fig. 21. Imposed wind loads on the building.

Fig. 22. Top structural displacement.

13
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

Fig. 23. Inter-story drifts.

Data availability

Data will be made available on request.

Acknowledgment
The authors extend their appreciation to Researcher Supporting Project number (RSPD2023R692), King Saud University, Riyadh,
Kingdom of Saudi Arabia.

References
[1] A.M. Almusharaf, M. Elnimeiri, A Performance-Based Design Approach for Early Tall Building Form Development, 2010.
[2] F. Alkhatib, et al., Computational aerodynamic optimization of wind-sensitive irregular tall buildings, Buildings 12 (7) (2022) 939.
[3] W. Baker, The world’s tallest building-Burj Dubai, UAE, in: Proceedings of the International Conference on Tall Buildings, 2004. Seoul, South Korea.
[4] A.H. Al-masoodi, et al., The aerodynamic performance of tall buildings by utilizing aerodynamic modifications-A review study, in: IOP Conference Series: Earth
and Environmental Science, IOP Publishing, 2022.
[5] P.A. Irwin, Bluff body aerodynamics in wind engineering, J. Wind Eng. Ind. Aerod. 96 (6–7) (2008) 701–712.
[6] G. Hu, et al., Wind-induced responses of a tall building with a double-skin façade system, J. Wind Eng. Ind. Aerod. 168 (2017) 91–100.
[7] K. Shahab, H. Irtaza, A. Agarwal, Comparative study of aerodynamic coefficients of prismatic and twisted tall buildings with various cross sections using CFD,
J. Inst. Eng.: Series C 102 (3) (2021) 635–650.
[8] A. Elshaer, G. Bitsuamlak, A. El Damatty, Enhancing wind performance of tall buildings using corner aerodynamic optimization, Eng. Struct. 136 (2017)
133–148.
[9] Y. Tamura, et al., Aerodynamic control of wind-induced vibrations and flow around supertall buildings, in: Proceedings of the 6th International Conference Adv.
Exp. Strutural Eng. International Work. Adv. Smart Mater, Smart Strutures Technol, 2015. Urbana-Champaign.
[10] Y. Tamura, et al., Aerodynamic and response characteristics of supertall buildings with various configurations, in: Proceedings of the 8th Asia-Pacific Conference
on Wind Engineering, 2013. Chennai, India.
[11] F.A. Johann, M.E. Carlos, F.L. Ricardo, Wind-induced motion on tall buildings: a comfort criteria overview, J. Wind Eng. Ind. Aerod. 142 (2015) 26–42.
[12] Y. Kim, J. Kanda, Effects of taper and set-back on wind force and wind-induced response of tall buildings, Wind Struct. 13 (6) (2010) 499–517.
[13] T. Deng, X. Yu, Z. Xie, Aerodynamic measurements of across-wind loads and responses of tapered super high-rise buildings, Wind Struct. 21 (3) (2015) 331–352.
[14] Y. Li, et al., Investigation of wind effect reduction on square high-rise buildings by corner modification, Adv. Struct. Eng. 22 (6) (2019) 1488–1500.
[15] K. Kwok, Effect of building shape on wind-induced response of tall building, J. Wind Eng. Ind. Aerod. 28 (1–3) (1988) 381–390.
[16] H. Hayashida, Y. Iwasa, Aerodynamic shape effects of tall building for vortex induced vibration, J. Wind Eng. Ind. Aerod. 33 (1–2) (1990) 237–242.
[17] T. Tamura, T. Miyagi, T. Kitagishi, Numerical prediction of unsteady pressures on a square cylinder with various corner shapes, J. Wind Eng. Ind. Aerod. 74
(1998) 531–542.
[18] Z. Zhengwei, et al., Effects of corner recession modification on aerodynamic coefficients of square tall buildings, in: The Seventh International Colloquium on
Bluff Body Aerodynamics and Applications, 2012.
[19] K.C. Kwok, P.A. Bailey, Aerodynamic devices for tall buildings and structures, J. Eng. Mech. 113 (3) (1987) 349–365.
[20] P. Sanyal, S.K. Dalui, Forecasting of aerodynamic coefficients of tri-axially symmetrical Y plan shaped tall building based on CFD data trained ANN, J. Build.
Eng. 47 (2022), 103889.
[21] S. Dai, et al., Impact of corner modification on wind characteristics and wind energy potential over flat roofs of tall buildings, Energy 241 (2022), 122920.
[22] N. Gaur, R. Raj, Aerodynamic mitigation by corner modification on square model under wind loads employing CFD and wind tunnel, Ain Shams Eng. J. 13 (1)
(2022), 101521.
[23] P. Sanyal, S.K. Dalui, Comparison of aerodynamic coefficients of various types of Y-plan-shaped tall buildings, Asian J. Civil Eng. 21 (7) (2020) 1109–1127.
[24] F. Alkhatib, et al., Wind-resistant structural optimization of irregular tall building using CFD and improved genetic algorithm for sustainable and cost-effective
design, Front. Energy Res. 10 (2022).
[25] A.H. Al-Masoodi, et al., A review of the sustainability and parametric design approach of complex tall buildings at an early design stage, in: 2021 Third
International Sustainability and Resilience Conference: Climate Change, IEEE, 2021.
[26] F. Alkhatib, et al., Performance-driven evaluation and parametrical design approach for sustainable complex-tall building design at conceptual stage, in: IOP
Conference Series: Earth and Environmental Science, IOP Publishing, 2022.
[27] K. Miyashita, et al., Wind-induced response of high-rise buildings effects of corner cuts or openings in square buildings, J. Wind Eng. Ind. Aerod. 50 (1993)
319–328.
[28] K.-T. Tse, et al., Economic perspectives of aerodynamic treatments of square tall buildings, J. Wind Eng. Ind. Aerod. 97 (9–10) (2009) 455–467.
[29] M. Gu, Y. Quan, Across-wind loads and effects of supertall buildings and structures, Sci. China Technol. Sci. 54 (10) (2011) 2531–2541.

14
A.H.H. Al-Masoodi et al. Journal of Building Engineering 76 (2023) 107017

[30] T. Deng, et al., An experimental study on the wind pressure distribution of tapered super high-rise buildings, Struct. Des. Tall Special Build. 27 (13) (2018)
e1483.
[31] K.T. Tse, et al., Effect of aerodynamic modifications on the surface pressure patterns of buildings using proper orthogonal decomposition, Wind Struct. 32 (3)
(2021) 227–238.
[32] F.-Q. Meng, et al., Sensitivity analysis of wind pressure coefficients on CAARC standard tall buildings in CFD simulations, J. Build. Eng. 16 (2018) 146–158.
[33] W. Melbourne, Comparison of measurements on the CAARC standard tall building model in simulated model wind flows, J. Wind Eng. Ind. Aerod. 6 (1–2)
(1980) 73–88.
[34] P. Luo, Wind Tunnel Test on Standard CAARC Tall Building Model, Tongji University, Shanghai, China, 2004.
[35] B. Mou, et al., Numerical simulation of the effects of building dimensional variation on wind pressure distribution, Eng. Appl. Comput. Fluid Mech. 11 (1)
(2017) 293–309.
[36] D.-X. Zhao, B.-J. He, Effects of architectural shapes on surface wind pressure distribution: case studies of oval-shaped tall buildings, J. Build. Eng. 12 (2017)
219–228.
[37] J. Franke, et al., Best practice guideline for the CFD simulation of flows in the urban environment-a summary, in: 11th Conference on Harmonisation within
Atmospheric Dispersion Modelling for Regulatory Purposes, Cambridge, UK, July 2007, Cambridge Environmental Research Consultants, 2007.

15

You might also like