You are on page 1of 29

Structures 56 (2023) 104975

Contents lists available at ScienceDirect

Structures
journal homepage: www.elsevier.com/locate/structures

Aerodynamic loads of tapered tall buildings: Insights from wind tunnel test
and CFD
Yi Li a, b, c, *, Yi Zhu b, Fu-bin Chen a, c, *, Qiu-Sheng Li d
a
School of Civil Engineering, Changsha University of Science and Technology, Changsha 410114, Hunan, China
b
School of Civil Engineering, Hunan University of Science and Technology, Xiangtan 411201, Hunan, China
c
Key Laboratory of Bridge Engineering Safety Control by Department of Education, Changsha University of Science and Technology, Changsha 410114, Hunan, China
d
Department of Architecture and Civil Engineering, City University of Hong Kong, Kowloon, Hong Kong

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

Keywords: Tapered modifications have been widely utilized for tall buildings to reduce their wind effects in practical en­
Tall building gineering. However, the flow mechanisms caused by tapering and relevant empirical models for wind loads
Tapered estimations of tapered tall buildings are still unknown. In this study, both wind tunnel test and CFD were used to
Wind load
study the aerodynamic performance of tapered tall buildings. Taken the CAARC standard tall building model as
Wind pressure
Wind tunnel test
the benchmark model, another four modified models with taper ratio of 0.05, 0.10, 0.15 and 0.20 were designed
Large eddy simulation and tested by pressure measurements and large eddy simulation (LES). The variations of wind pressure co­
efficients, local wind force coefficients, base moment coefficients, wind load power spectrum and vertical
coherence function of five models with different taper ratios were discussed and compared. Also, the streamline
of mean and fluctuating velocity around tapered tall building models are provided for better understanding of the
flow mechanisms. The results showed that aerodynamic performance of the tapered tall buildings has been
improved effectively, especial for across-wind effects. This study aims to provide reference for aerodynamic
optimization design of tapered tall buildings.

1. Introduction with tapered shape along the height have been built in some metro­
politan areas, as shown in Fig. 1.
Safety and comfort of tall buildings located in strong wind region Several scholars have carried out relevant researches on the topic of
depend most on wind excitations. How to evaluate and reduce wind effects of tapered modification on wind loads of high-rise buildings.
effects on tall buildings has always been an important issue for wind Cooper et al. [5] conducted wind tunnel tests on tapered tall buildings
engineers [12,15,30,17,20]. Due to the operability and economical ef­ model and pointed out that tapered modifications could reduce aero­
ficiency, aerodynamic optimization methods have been widely used in dynamic damping. Kim and You [14] indicated that the tapered tall
the process of wind-resistant design from fundamental research to en­ building may extend the vortex shedding to wider frequency range, thus
gineering practices for tall buildings [30,32,24,16,35]. Among these could effectively reduce the across-wind response. You et al. [36] found
aerodynamic modifications, tapering the section of tall building along that shedding position of the main vortex moved to the top when the
with its height, so called tapered tall building, is a common strategy for width of the tapered building structure decreased along the height. Xie
mitigation of wind effects of tall buildings. In recent years, tall buildings et al. [33] studied the wind effects of tapered tall building with square

* Corresponding authors at: School of Civil Engineering, Changsha University of Science and Technology, Changsha 410114, Hunan, China.
E-mail addresses: engineerliyi@163.com (Y. Li), fbchen88@126.com (F.-b. Chen).

https://doi.org/10.1016/j.istruc.2023.104975
Received 29 April 2023; Received in revised form 24 July 2023; Accepted 26 July 2023
Available online 3 August 2023
2352-0124/© 2023 Institution of Structural Engineers. Published by Elsevier Ltd. All rights reserved.
Y. Li et al. Structures 56 (2023) 104975

tapered ratios. The results of wind pressure distribution, local wind force
coefficients, base moment coefficients, power spectral densities and
vertical coherence functions of each tapered tall building model were
compared by two methods. Variations of wind loads characteristics of
tall buildings with different taper ratios were discussed in detail.
Moreover, the streamline of mean and instantaneous velocity around
tapered tall building models are provided by LES for analysis of flow
mechanism. The findings of this study are expected to provide useful
information for aerodynamic optimization design of similar tapered tall
buildings.

2. Objective and methodology

2.1. Study objective

The well-recognized CAARC standard tall building, which dimension


is 30.48 m (Depth) × 45.72 m (Breadth) × 183.88 m (Height), has been
selected as the benchmark tall building for tapered modification. Four
other tapered buildings with different tapered ratios (5%, 10%, 15%,
and 20%) are designed accordingly based on the taper ratio determined
in previous study [19]. For the application of wind tunnel test and saving
computation time of LES, all the five building models tested are scaled to
1/300 in both wind tunnel and CFD. In order to measure or monitor
wind pressure distribution on the surfaces of tapered tall building
models, measurement taps are arranged equally on four walls of the
building models for both wind tunnel test and numerical simulation. For
each model, there are seven measurement layers along with the model
height. Each layer has twenty measurement taps. The layouts of the
measurement taps of the tapered building model are shown in Fig. 2.

2.2. Wind tunnel test


Fig. 1. A typical tapered tall building in Shenzhen, China.

The spires and roughness elements were combined to simulate a


section by pressure measurements and presented that the tapering could typical suburban terrain (exposure category B) in the Chinese design
effectively reduce the across-wind load and wind-induced response to a code [7]. This terrain specifies a mean wind speed profile with a power
certain extent. Kim et al. [13] also obtained a similar conclusion that the law exponent of α = 0.15 and a gradient height of 350 m. Since the top
position of the vortex shedding acting on the tapered tall buildings height of all the scaled tapered building models is 0.609 m, hereby this
would be forced to move upward. Chen et al. [4] used the forced vi­ height is chosen as the reference height for analysis. The turbulence
bration test technique to study the unsteady aerodynamic forces of a integral length scale at the reference height is 0.56 m, which is equiv­
tapered model with the taper of 5.6 % under different wind speeds. alent to 138 m in prototype. Fig. 3 illustrates the profiles of mean wind
Obviously, the previous studies on wind effects of tapered tall buildings speed and turbulence intensity at various heights over the turntable in
were originated from wind tunnel tests. The flow mechanism of tapered the wind tunnel. Both the mean wind speed profile and turbulence
tall buildings is still unknown. The numerical simulation method has profile agree well with those profiles stipulated in the design code. The
great advantages in revealing the flow mechanism of wind effects on tall longitudinal wind velocity spectrum at the reference height is shown in
buildings. Huang et al. [10] conducted numerical simulation of wind Fig. 4. It can be observed that the simulated wind velocity spectrum is in
effects on a tall steel building. Elshaer et al. (2016) conducted a LES good agreement with the von Karman spectrum. Due to the space lim­
technical evaluation of wind-induced responses for an isolated and a itation, more details about the test models and test setup can be referred
surrounding CAARC tall building model. Asghari and Kargarmoakhar to previous literature [19].
[3] discussed the applicability and challenging of CFD numerical
simulation technology in aerodynamic design of tall buildings. Almin­
hana et al. [2] adopted LES technology to investigate on the aero­ 2.3. Numerical simulation
dynamic performance of CAARC standard tall buildings with chamfer
and recession modifications. Shang et al. [29] studied the flow over two The numerical simulation adopted in this study was operated by the
tandem square cylinders by large eddy simulation. In general, there are commercial CFD software Fluent 18.1. A LES study was carried out for
few studies on wind effects of tapered tall buildings evaluated by nu­ the wind effects on the benchmark model firstly. The simulation results
merical simulations. Therefore, it is necessary to conduct numerical were compared with the wind tunnel test results to verify the reliability
simulation to investigate the flow mechanism of tapered tall buildings. and accuracy of the numerical simulation method. Furthermore, the
In this study, both wind tunnel test and numerical simulation were other four tapered models were simulated by LES to investigate the flow
mechanisms change caused by tapering. The Reynolds number (Re) was

used to evaluate aerodynamic loads of tall buildings with various determined by Eq. (1). So, the maximum Reynolds number of wind

2
Y. Li et al. Structures 56 (2023) 104975

Fig. 2. Layouts of measurement taps of tapered building models.

3
Y. Li et al. Structures 56 (2023) 104975

Fig. 3. Mean wind speed profile and turbulent intensity profile.

Fig. 5. Computational domain size and boundary conditions.


Fig. 4. Longitudinal velocity spectrum.

trum. These turbulence structures cannot be simulated in the RANS


tunnel test and LES for the tapered tall building models is about 9.7 × simulation. In order to obtain the turbulent flow field similar to wind
104. tunnel test, the DSRFG technique [9,34] is utilized hereby to generate
ρuB the turbulent flow field for large eddy simulation. The fluctuating wind
Re = (1) velocity field in × space is expressed as follow:
μ
kmax ∑
∑ ( m,n ) ( m,n )
where ρ is the air density of 1.225 kg/m3, u represents the incoming
N
u(x, t) = [Pm,n cos ̃
k ⋅̃x + ωm,n t + qm,n sin ̃
k ⋅̃x + ωm,n t ] (2)
wind velocity, μ represents dynamic air viscosity and B means the m=k0 n=1
characteristic length, which is equal to the breadth of the tapered model.
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
ϛ × km,n 4E(km )
2.3.1. Inflow turbulence generation for LES Pm,n = a (3)
|ϛ × km,n | N
The inflow turbulent flow field needs certain turbulence structures
for the purpose of meeting the requirements of the target power spec­

4
Y. Li et al. Structures 56 (2023) 104975

Fig. 6. Mesh arrangements of five tapered tall building models.

12

0
0 1 2 3

Fig. 7. Distribution of Y + values on the four walls of the benchmark tall building model.

√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
ξ × km,n 4E(km )
(4)
m,n
qm,n = (1 − a) ̃m,n k
|ξ × km,n | N k =
k0
x
x=
̃ (5) |km,n | = km (6)
LS
( )
ωm,n ∈ N 0, 2πkm Uavg (7)

5
Y. Li et al. Structures 56 (2023) 104975

Fig. 9. Comparisons of mean wind pressure coefficients obtained by wind


Fig. 8. Comparison of wind pressure coefficients between LES and wind tunnel tunnel test and LES.
tests for two grids.
characteristics simulated by the DSRFG are in good agreement with the
In which, x is the coordinate vector and t is the time. Uavg is the mean target value, demonstrating that this technique can effectively generate
wind speed. ϛ and ξ are vector form of ϛni and ξni , respectively, and ϛni ,ξni ∈ the turbulent flow filed for LES.
N(0, 1); a is a random number evenly distributed between 0 and 1; km is
the wave number; N is the number of samples per wave number, 2.3.2. Computational domain, meshing and boundary conditions
generally 100–200, and N in this simulation process is 200; LS represents The dimension of the computational domain is 20H(x) × 6.5H(y) ×
the turbulence integral length scale, which is an important parameter 5H(z) in the numerical simulation (H is the scaled height of building
used as a spatially correlated scale factor; E(km ) is the discrete single model). The origin of Cartesian coordinate in the numerical simulation
pieces spectrum. is located at the bottom center of the tall building model, as shown in
The DSRFG technique can generate the fluctuating wind velocity Fig. 5. The center of the bottom of the model is 5H away from the inlet of
flow field according to the specified power spectrum and ensure the the computational domain and 3.25H away from side boundary. The
continuity of the inlet turbulent flow. Fig. 3 and Fig. 4 shows the cross-sectional size of the computational domain is 4 m (width) × 3 m
generated mean wind speed profile, turbulence intensity profile and the (height), which is the same as the cross-sectional size of wind tunnel.
spectrum and comparison with the wind tunnel test results as well as the The maximum blockage rate of streamwise direction is 0.77%, which is
target results. It can be seen from the figures that all the wind field less than 5% and is generally accepted in wind engineering. [31,8,18].

6
Y. Li et al. Structures 56 (2023) 104975

Fig. 10. Scatter diagrams of mean wind pressure coefficients.

The inlet condition and outlet condition of the computational domain ratio is 1.2. Two grid sizes were utilized in this study to test the sensi­
select velocity inlet and outflow, respectively. Both side boundary and tivity of the grid. The analysis content is in the following Section 3.1.1.
top boundary of the computational domain adopt symmetrical boundary After weighing the mesh error and computational efficiency, hexcore
condition. The surface of the ground and building model chooses the grid of 0.05D (Depth) size was chosen for the other four tapered building
non-slip wall condition to assume the zero velocity on boundary models. The mesh arrangements of five tapered tall building models are
surfaces. illustrated in Fig. 6. The total numbers of grid cells of the tapered tall
Hexcore is used to mesh the whole computational domain with building models are about 2.2 × 106. The distribution of dimensionless
structured and unstructured grids [28]. The surface of the tall building height Y + value at the near-wall of the benchmark model grid is shown
models and the ground are meshed by unstructured grids [6] and the in Fig. 7. It is noted that the dimensionless height Y + value is less than 9
stretching ratio is set to 1.05 for reduction of the difference in cut-off on the four walls of the building model, which indicated that the wall
wave numbers between adjacent grids in the LES [10]. The other re­ treatment can be used.
gions are meshed by structured grids and the corresponding stretching

7
Y. Li et al. Structures 56 (2023) 104975

where pi(t) is time history of the measured wind pressure; p0 and p∞


stands for the total pressure and the static pressure at the reference
height, respectively.
The mean wind pressure coefficient Cpmean and the RMS (root-mean
square) wind pressure coefficient Cprms are respectively determined as
follows:

1 T
Cpmean = Cpi (t)dt (9)
T 0
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
/
√ N
√∑
Cprms =√ (Cpik − Cpmean )2 (N − 1) (10)
K=1

in which T is the sampling time, N is the sampling length.

3.1.1. Grid sensitive analysis


The grid sensitivity analysis is necessary for meshing quality in nu­
merical simulation. So, two different grid sizes including Grid 1 and Grid
2 have been used to assess the grid sensitivity. The grid lengths of Grid 1
were set as 0.05D (D is depth of the building model) for the building
surface and 0.005 D for height of the first boundary layer, respectively.
Fig. 10. (continued). For Grid 2, the grid lengths were set as 0.03D for the building surface
and 0.003 D for height of the first boundary layer, respectively. Due to
2.3.3. Solution settings space limitation, only the benchmark model was selected to assess the
In this study, the Wall-Adaptive Local Eddy-viscosity (WALE) model grid sensitivity in this study. In order to ensure the accuracy of the
[26] was chosen for the LES subgrid-scale stress model. The Pressure- assessment, the reference pressure in the LES is set at a height of 0.6 m at
Implicit with Splitting of Operators (PISO) algorithm was adopted for the inlet boundary, which is equivalent to the height of the top of the
pressure and velocity coupling[11]. PISO has better performance in model. Fig. 8 compares the mean wind pressure coefficients and the RMS
unsteady simulation than the Semi-Implicit Method for Pressure-Linked wind pressure coefficients on the surface of the benchmark model. It can
Equations (SIMPLE) [9]. The second-order implicit scheme is applied for be found that wind pressure distributions resulted from the two grids are
time discretization. The momentum discretization uses the bounded consistent with those obtained from wind tunnel test. After weighing the
central difference (BCD) method. In order to meet the Courant- mesh error and computational cost, the Grid 1 was finally chosen to
Friedrichs-Lewy (CFL) condition, the time step is chose to be 5 × 10-4. simulate the flow field for the five tapered building models in the study.
The results of LES are obtained by calculating 20,000 time steps after the
flow field is stable, and then sampling. A total of 40,000 time steps are 3.1.2. Wind pressure distribution for different tapered models
calculated, representing 20 s of time on the model scale. The results of Fig. 9 shows the mean wind pressure coefficients at 2/3 height of the
LES are sampled by the subsequent 20,000 time steps after the flow field tall building models obtained by wind tunnel test and LES. It can be
becomes stable in the first 20,000 time steps. All numerical simulation observed that the mean wind pressure coefficients distribution regu­
works were operated on a 64-core Linux system workstation. The larities obtained by the LES and wind tunnel test are basically the same.
computational expense for each time step is about 71.6 h. As the taper ratio increases, the mean wind pressure coefficients Cpmean
gradually decrease on the windward. The values of Cpmean on the
3. Results and discussions windward for 5% and 10% tapered models decrease rapidly, but the
values of Cpmean on the windward for 15% and 20% tapered models
Since the unfavorable wind direction of rectangular tall buildings decrease gently. This is caused by the reduced width of the windward
affected by wind is usually perpendicular to the surface of the building and thus the incoming flow is quickly separated from the edge of
[21], the following content only presents and discusses the results of windward, which resulting in a slight drop in the mean wind pressure on
0◦ wind direction. Monitoring points with the same configuration as the windward. For the leeward and side walls of the model, the tapered
wind tunnel test were set on the building surface in numerical simula­ modifications have a remarkable effect on the values of Cpmean . As the
tion to extract the time histories of the instantaneous pressure. taper ratio increases, the absolute values of Cpmean on the leeward and
side walls of the model gradually decrease, especially for the rear edge
(pressure taps 9–10 and 16–17).
3.1. Wind pressure coefficients
Also, the comparisons of the Cpmean for five models obtained from
wind tunnel test and LES are illustrated in Fig. 10. This part compares
The wind pressure coefficient Cpi(t) at pressure tap i on the surface of
the mean wind pressure coefficients obtained from all the 140 pressure
a model is obtained by normalizing the measured pressures as follows:
taps of the models. It can be found that the LES technology has a good
pi (t) − p∞ simulation on the windward and sidewalls of the five models. The
Cpi (t) = (8)
p0 − p∞ maximum prediction error generally does not exceed 20%. Although

8
Y. Li et al. Structures 56 (2023) 104975

Fig. 11. Comparisons of RMS wind pressure coefficients obtained by wind tunnel test and LES.

some simulation results for wind pressures on leeward are greater than The distribution regularities of the RMS wind pressure coefficients ob­
20%, it is still can be accepted for engineering application. tained by LES and wind tunnel test are almost the same. As the taper
Fig. 11 shows the RMS wind pressure coefficients (Cprms ) at 2/3 ratio increases, the RMS wind pressure coefficients gradually decrease.
height of five model obtained by wind tunnel test and LES respectively. The change of taper ratio has little effect on the Cprms of the windward of

9
Y. Li et al. Structures 56 (2023) 104975

Fig. 12. Scatter diagrams of RMS wind pressure coefficients.

the model. Obviously, the RMS wind pressure coefficients of the leeward pressure taps for the five models are presented in Fig. 12. As can be seen,
and the side walls are more sensitive to the taper modification. This is the wind pressure coefficient errors between the simulation results and
because the Cprms of the side wall is related to the separation of the wind tunnel test are within 20% for most cases, which indicating that
incoming flow and the Cprms of leeward is related to the vortex structure the LES technology can be used to predict the RMS wind pressure co­
in the wake flow. The RMS wind pressure coefficients of all the 140 efficients of tapered tall building models properly in this study.

10
Y. Li et al. Structures 56 (2023) 104975

Fig. 12. (continued).

In order to illustrate the distribution of wind pressure on the facade taper ratio. For Cprms on the leeward, the minimum value of 0.18
of the tapered tall building model, contour distribution of the wind benchmark model. After the taper ratio increases to 20%, the minimum
pressure coefficients are provided. The mean wind pressure coefficients value has been reduced to 0.14.
contours of the five models obtained by LES under wind direction of
0◦ are shown in Fig. 13. From left to right, the figures are mean wind 3.2. Local wind force
pressure coefficients contours on windward, side wall and leeward,
respectively. By comparing the mean wind pressure contours of five Time series of wind forces at each measurement layer on a tapered
models, it can be seen that the distributions of the mean wind pressure model can be obtained by integrating the simultaneously measured wind
coefficients of each tapered model are basically similar. The maximum pressures. Non-dimensional coefficients of the local wind forces are
value of Cpmean on the windward is 0.85 for benchmark model (tapered defined as follows:
ratio of 0%), and the height of the maximum zone is about 0.8H. As the
taper ratio increases from 0% to 20%, the maximum value of Cpmean on FD (zi ) ′ σD (zi )
CD (zi ) = , C (zi ) = (11)
the windward decreases from 0.85 to 0.68, and the corresponding height A(zi )qH D A(zi )qH
decreases from 0.8H to 0.75H. The maximum negative pressure of
Cpmean occurs at the leading edge of the sidewall for benchmark model. C′L (zi ) =
σ L (zi )
(12)
As the taper ratio increases, the range of the zone of maximum negative A(zi )qH
pressure gradually shrinks and the height of the zone of maximum
negative pressure moves down. After the taper ratio increases to 20%, where, CD (zi ) is the mean drag coefficient; C′D (zi ) and C′L (zi ) are the RMS
the absolute value of the negative pressure decreases from 0.72 to 0.68. drag coefficient and RMS lift coefficient, respectively; FD (zi ) is the mean
For the negative pressure zone on the leeward, the minimum absolute drag force; σD (zi ) and σL (zi ) are the RMS drag force and RMS lift force,
value of the Cpmean of is 0.57 for benchmark model and the minimum respectively; qH is the incoming wind pressure at the height of the top of
zone is happened at 0.5 H of the model. When the taper ratio increases to the model. A(zi ) is the wind area of each measurement layer. Since mean
20%, the minimum absolute value of Cpmean on the leeward decreases lift coefficient, mean and RMS torque are close to zero under wind di­
from 0.57 to 0.51, and the position corresponding to the minimum zone rection of 0◦ , they will not be discussed in this study.
drops from 0.5 H to 0.35H.
Fig. 14 shows the contour distributions of RMS wind pressure co­ 3.2.1. Mean and RMS local wind force coefficients
efficients (Cprms ) of the five models obtained by the LES under wind The local wind force acting on the bluff body mainly depends on the
direction of 0◦ . It can be found that the Cprms contour distributions of the characteristics of the incoming flow, the geometric parameters of the
five models are quite different. For benchmark model, the maximum building and the relative height [20]. This section mainly studies the
value of Cprms on the windward is 0.26. After the taper ratio increases, effect of the tapered modifications on the local wind force coefficients of
the maximum value is reduced from 0.26 to 0.19 on the windward. The the tall building models, and compares the results of the wind tunnel test
Cprms of the sidewall decreases significantly with the increasing of the with those obtained from the LES. Fig. 15 shows the variation the mean
and RMS local wind force coefficients along the height of five models

11
Y. Li et al. Structures 56 (2023) 104975

Fig. 13. Mean wind pressure coefficients contours of five models.

12
Y. Li et al. Structures 56 (2023) 104975

Fig. 13. (continued).

obtained by both wind tunnel test and LES under wind direction of 0◦ . In coefficients at the height of 2/3H on the five models under wind di­
Fig. 15(a), the mean drag force coefficients of the LES gradually decrease rection of 0◦ are studied. Fig. 16 shows the power spectral densities of
with the increasing of the taper ratio, and they are very close to the wind the local wind force coefficients at along-wind, across-wind and
tunnel test results. The mean drag force coefficients of five models torsional direction, respectively. It can be seen from Fig. 16(a) that the
decrease slowly along the height within the range of heights below along-wind force power spectrum results of LES are in good agreement
0.35H. When the height is larger than 0.35H, the mean drag force co­ with the results of wind tunnel test. The variation trends of the power
efficients of model gradually increase along the height and reach to the spectral density of the along-wind force with the reduced frequency are
maximum value at the height of 0.85H. basically the same, showing obvious broadband characteristics as a
Fig. 15(b) shows the comparison of the RMS drag force coefficients of whole. From Fig. 16(b), it can be observed that there are obvious peaks
the five models. As the taper ratio increases, the RMS drag force co­ emerged at reduced frequency of 0.1 of the power spectral density of the
efficients decrease gradually. The RMS drag force coefficients obtained across-wind force coefficients for five models at both wind tunnel and
by LES are close to the wind tunnel test results, although the values are LES. As the taper ratio increases, the peak value of the power spectral
slightly larger at the bottom of models, which may be caused by the density of the across-wind force decreases and the position of peak
difference between the ground boundary set by the numerical simula­ moves to the lower reduced frequency. The bandwidths of the power
tion and the actual ground in the wind tunnel test. It can be seen from spectrum of the 15% and 20% models are enlarged and the peaks are no
Fig. 15(c) that the LES results of RMS lift force coefficients are generally longer obvious, which demonstrates that the engery of vortex shedding
consistent with the results of wind tunnel test. By comparing the model is suppressed by the tapered modification, thus indicating that the
before and after the tapered modifications, it is found that the maximum tapered modifications can effectively reduce the dynamic across-wind
of RMS lift force coefficients for models with tapered ratios of 0% and forces. In Fig. 16(c), the power spectral densities of torque from wind
5% were appeared at 0.47H, while the maximum of RMS lift force co­ tunnel tests are also in good agreement with those obtained from LES.
efficients for models with tapered ratios of 10%, 15% and 20% were The benchmark model has a narrow-band spectrum peak near the
emerged at 0.65H. This is because the tapered modifications increase the reduced frequency 0.1. As the taper ratio increases, the power spectral
relative height of the main body vortex shedding phenomenon, which density of torque force exhibits broadband characteristic. When the
leads to rasing height of the maximum value of RMS lift force taper ratio is greater than 10%, the power spectral density of torque
coefficients. force shows a bimodal characteristic, which may be caused by the
reattachment of the separated flow on the side walls of the model.
3.2.2. Power spectral densities of local wind force coefficient
In this section, the power spectrum densities of the local wind force

13
Y. Li et al. Structures 56 (2023) 104975

Fig. 14. RMS wind pressure coefficients contours of five models.

14
Y. Li et al. Structures 56 (2023) 104975

Fig. 14. (continued).

3.3. Wind flow field analysis increasing of taper ratio, the vortex scale in the wake region decreases
and the width of the wake flow gradually become narrow, indicating
Using streamline to describe the flow of fluid around the building can that the RMS drag force of the models are reduced. It is noted that the
clearly show the movement process of the fluid hindered by the build­ small vortices on the side walls of the 15% and 20% tapered model are
ing, including the formation of vortex, the shedding of vortex and the gradually attached to the sidewalls of the model. This existence of vortex
process of small vortex absorbing energy and developing into large reattachment can suppresses the vortex shedding strength [23] and thus
vortex. In this section, the X-Y plane with the height of model z = 0.406 can significantly reduce the RMS lift force coefficients.
m (2/3H) was chosen to study the effect of the tapered modifications on
the fluid movement around the model. The mean velocity contours with
streamline of the five models under wind direction of 0◦ are plotted in 3.4. Base moment coefficients
Fig. 17. From the distribution of the entire flow field, the mean flow
fields of the five models are basically the same, with slight differences The base moment can be used to calculate the first-order response
around the building. For the benchmark model, the incoming flow and equivalent static wind load of the structure. The non-dimension
separates from the front edges of both side walls for the building model mean and RMS base moment coefficients are expressed as follows:
after it hits the windward of the building model, and develops two
FMD ′ σ MD
symmetrical horseshoe-shaped vortices in the wake region on the CMD = C = (13)
BH 2 qH MD BH 2 qH
leeward of the building model. As the taper ratio increases, the width of
the low-velocity region on the windward of the model is gradually σ ML
C′ML = (14)
reduced, which means that the mean wind pressure on the windward of BH 2 qH
the model decreases. The dimension of the horseshoe-shaped vortex in
the wake region also gradually decreases, resulting in the reduction of where, CMD is mean along-wind base moment coefficient; C′MD and C′ML
the along-wind forces of the tapered model. This phenomenon provides are the RMS base moment coefficients of along-wind and across-wind
explanation for the fact that the tapered modifications can reduce the directions, respectively; FMD , σ MD and σML are the corresponding mean
mean drag force coefficients of the tall building models effectively. and RMS base moments. H and B are the height and width of the model,
Fig. 18 shows the instantaneous velocity contour with streamline of respectively; qH is the incoming wind pressure at the height of the top of
five models at 2/3 height obtained by LES. Compared to the mean flow the model.
field, there are many small vortices on the side walls of the building
models in the instantaneous flow field. The small vortices that shed from 3.4.1. Mean and RMS base moment coefficients
the trailing edge of the side walls for the building model enter into the Fig. 19 shows the base moment coefficients of the five tapered
wake region and form large-scale vortex on the leeward. With the models obtained from the wind tunnel test and LES. Comparisons with

15
Y. Li et al. Structures 56 (2023) 104975

Fig. 15. Comparison of local wind force coefficients between wind tunnel test and LES.

16
Y. Li et al. Structures 56 (2023) 104975

101

100

10-1

10-2 101

10-3 100
10-3 10 -2
10 -1
10 0

10-1

10-2
WT 0% LES 0%
WT 5% LES 5%
WT 10% LES 10%
WT 15%
WT 20%
LES 15%
LES 20%
10-3
10-3 10-2 10-1 100

Fig. 16. Comparison of power spectrum of local wind force coefficients at 2/3H of building models between wind tunnel test and LES.

17
Y. Li et al. Structures 56 (2023) 104975

Fig. 17. Streamline diagram of mean velocity at 2/3H height of models.

18
Y. Li et al. Structures 56 (2023) 104975

Fig. 17. (continued).

other researches are also illustrated. Regarding to the RMS along-wind Laboratory at the University of Notre Dame (UND) provides an online
and across-wind base moment coefficients, the NatHaz Modeling aerodynamic load database for tall buildings [25] (http://www.nd.

19
Y. Li et al. Structures 56 (2023) 104975

Fig. 17. (continued).

edu/nathaz/2003). Fig. 19(a) and Fig. 19(b) present the mean and RMS direction of 0◦ . The results show that the power spectral densities of base
along-wind base moment coefficients, respectively. The values of mean moment coefficients obtained from the wind tunnel test are in good
along-wind base moment coefficients obtained from both the wind agreement with the LES results. Variations of the power spectral den­
tunnel test and the LES gradually decrease with the increasing of the sities of along-wind base moment coefficients with the reduced fre­
taper ratio. Compared to the benchmark model, the mean along-wind quency are almost the same with those of the local along-wind force
base moment coefficients of the 20% tapered model are decreased by coefficients, basically showing broadband characteristics. For the power
35.3% and 37.8% for the wind tunnel test and LES, respectively. The spectral densities of the across-wind base moment coefficients, the
RMS along-wind base moment coefficients for all tapered models are power spectral densities always have peaks near the reduced frequency
between open terrain (UND-A) and urban terrain (UND-C) terrain. The of 0.1 and change little with the increasing of the taper ratio, which is
values of the RMS along-wind base moment coefficients also show different from the power spectral densities of the across-wind lift force.
decreasing trend with the increasing of taper ratio. Compared to the The torsional power spectral densities of the benchmark model show an
benchmark model, the RMS along-wind base moment coefficients of the obvious peak near the reduced frequency of 0.1. As the taper ratio in­
20% tapered model are decreased by 39.7% and 37.4% for the wind creases, the peak of the spectrum tends to become gentle. The LES results
tunnel test and LES, respectively. As shown in Fig. 19(c), the RMS across- show the same tendency. Therefore, it can be concluded that the tapered
wind base moment coefficients decrease as the taper ratio increases. The modifications can also reduce the torsional base moments of the model.
RMS across-wind base moment coefficient of the benchmark model are
close to the values of AIJ Recommendation for Loads on Buildings [1] 3.5. Vertical coherence function
and Quan [27], which is slightly smaller than the values calculated by
UND-A and UND-C. However, the RMS across-wind base moment co­ For the tall building with tapered modifications, vertical coherence
efficients of the five models are much less than the values suggested by of wind loads is important for wind load calculation in frequency
Liang et al. [22] and the Chinese design code [7]. For the 20% tapered domain. Hereby, this study selects the vertical coherence between the
model, the value of the RMS across-wind base moment coefficient is the measurement layer C (1/2H) and measurement layer D (2/3H) of each
smallest, and the minimum values for the wind tunnel test and LES re­ tapered building model. The vertical coherence function of wind loads at
sults are 0.074 and 0.083, respectively, which are 40.1% and 36.2% the height of any two measurement layers Z1 and Z2 can be determined
smaller than the benchmark model. as:
⃒ ⃒
3.4.2. Power spectral densities of base moment coefficients ⃒SZ ,Z (f ) ⃒2
2
Coh (Z1 , Z2 , f ) = 1 2
(15)
Fig. 20 presents the power spectral densities of base moment co­ SZ1 (f )SZ2 (f )
efficients in the along-wind, across-wind and torsional directions of the
five models obtained from the wind tunnel test and LES under wind in which, SZ1 ,Z2 (f) is the cross power spectral density of the local wind

20
Y. Li et al. Structures 56 (2023) 104975

Fig. 18. Instantaneous velocity streamline diagram at 2/3H height of models.

21
Y. Li et al. Structures 56 (2023) 104975

Fig. 18. (continued).

force at the heights of Z1 and Z2; SZ1 (f) and SZ2 (f) are the self-power the frequency; Z1 and Z2 are the heights of the two layers; U is the mean
spectral density of local wind force, respectively. The least square value of the wind speed between the heights of Z1 and Z2.
method was adopted to fit the coherence function results obtained from Fig. 21 shows the coherence function of drag force coefficients for
wind tunnel test according to Eq. (16): five models by wind tunnel test and LES under wind direction of 0◦ .
( ) Since the coherence function of the drag force coefficients obtained by
f |Z2 − Z1 |
Coh(Z1 , Z2 , f ) = exp − CZ (16) LES is similar to the wind tunnel test result, only the coherence function
U
curve of the drag force coefficients obtained from the experiment result
is fitted here. The coherence function of the drag force coefficients
in which, CZ is the coherence coefficient of the coherence function; f is
basically decreases with increasing frequency, and the coherence

22
Y. Li et al. Structures 56 (2023) 104975

Fig. 19. Base moment coefficients.

23
Y. Li et al. Structures 56 (2023) 104975

Fig. 20. Comparison of wind tunnel test and LES for power spectral densities of base moment coefficients.

24
Y. Li et al. Structures 56 (2023) 104975

Fig. 20. (continued).

function curve tends to be gentle in high frequency region. As the taper pressure coefficients and RMS wind pressure coefficients
ratio increases, the coherence coefficient (CZ ) of the drag force co­ decrease with the increasing of taper ratio. Tapered modifications
efficients increases from 3.27 to 4.13, which shows that larger taper have greater effect on wind pressure distribution on the wind­
ratio can lower the coherence of drag force coefficients for the building ward and leeward sides rather than the side wall.
model. (2) With increasing of tapered ratio, the drag force coefficients and
Fig. 22 shows the coherence function of lift force coefficients for the lift force coefficients of high-rise building models are effectively
five models under wind direction of 0◦ . It can be seen that the coherence reduced. Moreover, the bandwidth for power spectral of across-
functions of lift force coefficients obtained by LES are in good agreement wind force expands and the power spectrum peak value de­
with those obtained by wind tunnel test, with slight differences at high creases. The reduced frequency of the peak value moves towards
frequencies. As the taper ratio increases, the peak value around fre­ the low frequency.
quency 10 Hz gradually decreases. The peak of the coherence function of (3) As the taper ratio increases, the horseshoe vortex gradually
the lift force coefficients nearly disappears for the 15% and 20% tapered became smaller and the width of the wake decreased gradually,
models. This phenomenon may be caused by the disturbed vortex leading to the reduction in the along-wind forces of the tapered
shedding along with the model height. model. The tapered modifications suppressed the intensity of
vortex shedding, thus resulting in the reduction of the across-
4. Conclusions wind forces.
(4) Base moment coefficients of high-rise buildings have been
This paper investigated the aerodynamic loads of tapered high-rise reduced by tapered modification. The maximum reductions were
buildings by both wind tunnel test and numerical simulation. The observed at tall building models with 20% taper ratio. Compared
wind pressure coefficient distribution, local wind coefficients, base to the benchmark model, the maximum mean along-wind base
moment coefficients, power spectrum and coherence function of the moment coefficients, RMS along-wind base moment coefficients
tapered building models were discussed. The main conclusions are and RMS across-wind base moment coefficients were reduced by
drawn as follows: 35.3%, 39.7% and 40.1%, respectively.
(5) As the taper ratio increases, the coherence coefficient of the
(1) The tapered modifications could effectively reduce wind pressure coherence function of drag force coefficients increases from 3.27
coefficients on the surface of tall buildings. Both mean wind to 4.13, indicating larger taper ratio could weaken the coherence

25
Y. Li et al. Structures 56 (2023) 104975

Fig. 21. Vertical coherence function of drag force coefficients.

26

Fig. 21. (continued).


Y. Li et al. Structures 56 (2023) 104975

Fig. 22. Vertical coherence function of lift force coefficients..

27
Y. Li et al. Structures 56 (2023) 104975

Fig. 22. (continued).

of the drag force coefficients of the high-rise building models. For Declaration of Competing Interest
lift force coefficients, the peak value around frequency 10 Hz
gradually decreases as the taper ratio increases. The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
CRediT authorship contribution statement the work reported in this paper.

Yi Li: Conceptualization. Fu-Bin Chen: Funding acquisition.

28
Y. Li et al. Structures 56 (2023) 104975

Acknowledgments [18] Li Y, Li YG, Li QS, Tee KF. Investigation of wind effect reduction on square high-
rise buildings by corner modification. Adv Struct Eng 2019;22(6):1488–500.
[19] Li Yi, Song Q, Li C, Huang X, Zhang Y. Reduction of wind loads on rectangular tall
The work presented in this paper was fully supported by the grants buildings with different taper ratios. Journal of Building Engineering 2022;45:
from National Natural Science Foundation of China (Project no: 103588.
51708207, 52278479) and a grant from Civil Engineering Key Discipline [20] Li, Y., Tian, X., Tee, K.F., Li, Q.S. and Li, Y.G. (2018), “Aerodynamic treatments for
reduction of wind loads on high-rise buildings”, Journal of Wind Engineering and
of Changsha University of Science and Technology (23ZDXK11). Industrial Aerodynamics, 172, 107-115.
[21] Li Yi, Zhang JW, Li QS. Experimental investigation of characteristics of torsional
References wind loads on rectangular tall buildings. Struct Eng Mech 2014;49(1):129–45.
[22] Liang SG, Liu SC, Li QS, Zhang LL, Gu M. Mathematic model of across wind
dynamic loads on rectangular tall buildings. J Wind Eng Ind Aerodyn 2002;90:
[1] AIJ Recommendation for Loads on Buildings. Architectural Institute of Japan. 1757–70.
Japan: Tokyo; 2004. [23] Ma W, Zhang X, Macdonald J, Jin L, Li Y. Effects of reattachment and three-
[2] Alminhana GW, Braun AL, Loredo-Souza AM. A numerical-experimental dimensionality on the aerodynamics of a circular cylinder in the critical Reynolds
investigation on the aerodynamic performance of CAARC building models with number range. J Wind Eng Ind Aerodyn 2022;220:104839.
geometric modifications. J Wind Eng Ind Aerodyn 2018;180:34–48. [24] Mooneghi MA, Kargarmoakhar R. Aerodynamic Mitigation and Shape
[3] Asghari Mooneghi M, Kargarmoakhar R. Aerodynamic mitigation and shape Optimization of Buildings: Review. Journal of Building Engineering 2016;6:
optimization of buildings: review. Journal of Building Engineering 2016;6:225–35. 225–35.
[4] Chen Z, Huang H, Xu Y, Tse KT, Kim B, Wang Y. Unsteady aerodynamics on a [25] NatHaz Modeling Laboratory, University of Notre Dame. Aerodynamic loads
tapered prism under forced excitation. Eng Struct 2021;240:112387. database, http://www. nd. edu/nathaz/, 2003.
[5] Cooper KR, Nakayama M, Sasaki Y, Fediw AA, Resende-Ide S, Zan SJ. Unsteady [26] Nicoud F, Ducros F. Subgrid-scale stress modelling based on the square of the
aerodynamic force measurements on a super-tall building with a tapered cross velocity gradient tensor. Flow Turbul Combust 1999;62(3):183–200.
section. J Wind Eng Ind Aerodyn 1997;72:199–212. [27] Quan, Y., 2002. Across-wind Loads and Responses on Super High-Rise Buildings.
[6] Dai SF, Liu HJ, Chu YJ, Lam HF, Peng HY. Impact of corner modification on wind Ph. D. Dissertation. Tongji University, Shanghai.
characteristics and wind energy potential over flat roofs of tall buildings. Energy [28] Roache PJ. Quantification of uncertainty in computational fluid dynamics. Annu
2022;241:122920. Rev Fluid Mech 1997;29(1):123–60.
[7] GB 50009–2012. Load Code for the Design of Building Structures. Beijing: China [29] Shang J, Zhou Q, Alam MM, Liao H, Cao S. Numerical studies of the flow structure
Architecture & Building Press; 2012. and aerodynamic forces on two tandem square cylinders with different chamfered-
[8] Gousseau P, Blocken B, van Heijst GJF. Quality assessment of Large-Eddy corner ratios. Phys Fluids 2019;31(7):075102.
Simulation of wind flow around a high-rise building: Validation and solution [30] Tanaka H, Tamura Y, Ohtake K, Nakai M, Kim YC. Experimental investigation of
verification. Comput Fluids 2013;79:120–33. aerodynamic forces and wind pressures acting on tall buildings with various
[9] Huang SH, Li QS, Wu JR. A general inflow turbulence generator for large eddy unconventional configurations. J Wind Eng Ind Aerodyn 2012;107–108:179–91.
simulation. J Wind Eng Ind Aerodyn 2010;98(10-11):600–17. [31] Tominaga Y, Mochida A, Yoshie R, Kataoka H, Nozu T, Yoshikawa M, et al. AIJ
[10] Huang SH, Li QS, Xu SL. Numerical evaluation of wind effects on a tall steel guidelines for practical applications of CFD to pedestrian wind environment
building by CFD. J Constr Steel Res 2007;63(5):612–27. around buildings. J Wind Eng Ind Aerodyn 2008;96(10):1749–61.
[11] Issa RI. Solution of the implicitly discretised fluid flow equations by operator- [32] Xie JM. Aerodynamic optimization of super-tall buildings and its effectiveness
splitting. J Comput Phys 1986;62(1):66–82. assessment. J Wind Eng Ind Aerodyn 2014;130:88–98.
[12] Kareem A. Mitigation of wind induced motion of tall buildings. J Wind Eng Ind [33] Xie ZN, Li J. Experimental research on cross wind effect on tapered super-tall
Aerodyn 1983;11(1-3):273–84. buildings under action of strong wind. Journal of Building Structures 2011;32(12):
[13] Kim YC, Kanda J. Wind pressures on tapered and set-back tall buildings. J Fluids 118–26. In Chinese.
Struct 2013;39:306–21. [34] Yan BW, Li QS. Inflow turbulence generation methods with large eddy simulation
[14] Kim Y-M, You K-P. Dynamic responses of a tapered tall building to wind loads. for wind effects on tall buildings. Comput Fluids 2015;116:158–75.
J Wind Eng Ind Aerodyn 2002;90(12-15):1771–82. [35] Yan B, Ren H, Li D, Yuan Y, Li Ke, Yang Q, et al. Numerical Simulation for Vortex-
[15] Kwok KCS. Effect of building shape on wind-induced response of tall building. Induced Vibration (VIV) of a High-Rise Building Based on Two-Way Coupled Fluid-
Journal of Wind Engineering and Industrial Aerodynamic 1988;28(1–3):381–90. Structure Interaction Method. Int J Struct Stab Dyn 2022;22:2240010.
[16] Li YG, Li MG, Tan WJ, Li Y. Experimental investigation of blockage effects on [36] You, K.P., Kim, Y.M. and Ko, N.H. (2008) “The evaluation of wind-induced
fluctuating wind forces on standard tall building models. Journal of Building vibration responses to a tapered tall building”, Structural Design of Tall and Special
Structures 2019;40(7):70–8. In Chinese. Buildings, 17(3), 655-667.
[17] Li Yi, Li C, Li Q-S, Song Q, Huang X, Li Y-G. Aerodynamic performance of CAARC
standard tall building model by various corner chamfers. J Wind Eng Ind Aerodyn
2020;202:104197.

29

You might also like