Applsci 14 00988
Applsci 14 00988
Applsci 14 00988
sciences
Article
Computational Fluid Dynamics-Aided Simulation of Twisted
Wind Flows in Boundary Layer Wind Tunnel
Zijing Yi 1 , Lingjun Wang 2 , Xiao Li 2, * , Zhigang Zhang 2, * , Xu Zhou 2 and Bowen Yan 2
Abstract: The twisted wind flow (TWF), referring to the phenomenon of wind direction varying with
height, is a common feature of atmospheric boundary layer (ABL) winds, noticeably affecting the
wind-resistant structural design and the wind environment assessment. The TWF can be effectively
simulated by a guide vane system in wind tunnel tests, but the proper design and configuration of
the guide vanes pose a major challenge as practical experience in using such devices is still limited
in the literature. To address this issue, this study aims to propose an approach to determining the
optimal wind tunnel setup for TWF simulations using a numerical wind tunnel, which is a replica of
its physical counterpart, using computational fluid dynamics (CFD) techniques. By analyzing the
mechanisms behind guide vanes for generating TWF based on CFD results, it was found that the
design must take into account three key parameters, namely, (1) the distance from the vane system to
the side wall, (2) the distance from the vane system to the model test region, and (3) the separation
between the vanes. Following the optimal setup obtained from the numerical wind tunnel, TWF
profiles matching both the power-law and Ekman spiral models, which, respectively, reflect the ABL
and wind twist characteristics, were successfully generated in the actual wind tunnel. The findings of
this study provide useful information for wind tunnel tests as well as for wind-resistant structural
designs and wind environment assessment.
Keywords: twisted wind flow; wind tunnel testing; CFD; numerical wind tunnel
Citation: Yi, Z.; Wang, L.; Li, X.;
Zhang, Z.; Zhou, X.; Yan, B.
Computational Fluid Dynamics-Aided
Simulation of Twisted Wind Flows in
1. Introduction
Boundary Layer Wind Tunnel. Appl.
Sci. 2024, 14, 988. https://doi.org/
The wind field within the atmospheric boundary layer (ABL) is a crucial research
10.3390/app14030988
topic because it plays an important role in the wind-resistant design of structures as
well as the wind environment assessment. In most wind codes, the vertical wind profile
Academic Editor: Giuseppe Lacidogna
below the gradient height, typically 200–500 m, is stipulated as follows: the wind speed
Received: 23 November 2023 increases with height following the (semi-)empirical logarithmic or power law, whilst the
Revised: 4 January 2024 wind direction remains constant at all heights. However, field measurement studies have
Accepted: 15 January 2024 pointed out that the direction of natural wind flow within ABL may not be constant as it
Published: 24 January 2024 varies noticeably with height, which is attributed to various factors such as the Ekman
effect (i.e., a combined effect of Earth’s rotation, surface friction, and pressure gradient
force), atmospheric stability, terrain roughness, and baroclinity. This type of winds is
referred to as the twisted wind flow (TWF, also known as the turning of winds or veering
Copyright: © 2024 by the authors. winds), and the angular difference between the wind direction at a given height and that
Licensee MDPI, Basel, Switzerland. at the gradient height is referred to as the twist angle. For instance, Mendenhall et al. [1]
This article is an open access article
investigated the dependencies of the twist angle on the horizontal temperature gradient
distributed under the terms and
and the atmospheric stability and then proposed prediction models for the variation in the
conditions of the Creative Commons
twist angle with height over land and ocean terrains. According to the long-term SODAR
Attribution (CC BY) license (https://
observations of wind characteristics in a coastal area of Japan, Tamura et al. [2] reported
creativecommons.org/licenses/by/
that the twist angle may reach about 20◦ within the height range of 50 to 350 m above
4.0/).
ground. Liu et al. [3] analyzed 4766 30-min-long data segments of synoptic winds collected
by a wind profiler installed in a coastal area of China, concluding that the twist angle can
reach 5◦ to 40◦ over the first 1000 m above ground. Based on the field measurement data of
multiple tropical cyclones collected by Doppler radar wind profilers and anemometers in
Hong Kong, He et al. [4] and Shu et al. [5] suggested that wind direction generally varies
20◦ to 40◦ within the first 1000 m in height, and the twist angle is highly dependent on the
terrain condition, the wind speed, and the storm type (monsoons or typhoons). Based on
these field observations, it can be readily understood that the TWF may impose significant
asymmetrical loads on structures, and therefore the neglect of the TWF effect may lead
to inaccurate calculations of the wind loads acting on super-tall structures, including
skyscrapers and gigantic wind turbines [6,7]. For instance, the variation in wind yaw
angle may affect the drag and lift forces of the wind turbine blades, whilst such a TWF
effect is usually neglected in turbine designs. Moreover, ignoring the TWF may also result
in unreliable wind environment assessments, such as for pedestrian-level winds over
complex urban and/or mountainous terrain [8].
As a widely accepted experimental approach, boundary layer wind tunnel testing is
usually recommended or even mandatory for wind-resistant structural designs and wind
environment assessments. To artificially replicate the TWF in conventional wind tunnels
that employ passive devices (e.g., grids, spires, and ground roughness) to generate ABL
winds, research efforts have been devoted over the past few decades. One of the most
common approaches is to set up a guide vane system consisting of multiple vertical vanes
with a curved surface to redirect the flow at the outlet of the wind profile development
region (i.e., where the passive devices are installed). Given a properly designed guide
vane system, the wind flow can be twisted to achieve the desired orientations at various
heights, so that the flow in the downstream model test region follows a profile that has
both the speed and direction varying with height in the required manner. One of the
first attempts was made by Prof. Flay and his research group from the University of
Auckland, who evaluated the aerodynamic performance of downwind yacht sails by
installing plastic vanes [9]. Despite the drawbacks, such as high turbulence intensities and
dips in mean wind speeds at regular heights, this guide vane system delivered promising
test results that well agreed with the data obtained by numerical simulations and field
measurements. The successful application of this guide vane system has not only spurred
studies in sail aerodynamics [10,11], but has also garnered attention from the research
field of civil engineering. In recent years, a few wind tunnel tests have been conducted to
investigate the TWF effect on structures and the wind environment [12,13] In particular,
Tse et al. [8] fabricated five wooden vanes to simulate the TWF at the pedestrian-level height
caused by mountainous and densely built city terrain, indicating that the TWF noticeably
modifies flow features such as asymmetric wind speed distributions and reduced wind
speed near an isolated building or building groups. Liu et al. [14] generated two TWFs
with the characteristics of thousand-meter-high ABL winds using four vanes and pointed
out that the value of wind load applied to a super-tall building under the TWF effect highly
depends on the total twist angle instead of being a specific value. Yan et al. [15] developed
two one-piece molded-fiberglass vanes to investigate the TWF effect on a typical super-
tall building, highlighting that the TWF significantly affects the distribution of extreme
cladding pressures and local wind loads. Nevertheless, the wind tunnel investigation of
the TWF is still lacking, and one of the major difficulties it encounters is the design of a
proper guide vane system.
As an alternative approach to investigating the wind effect, numerical simulation
using computational fluid dynamics (CFD) techniques has also been widely adopted in
numerous studies. Compared with typical wind tunnel testing, numerical simulations have
prominent advantages, such as the detailed illustration of flow fields, which is especially
important to the investigation of complex flow structures, e.g., the TWF. To numerically
simulate the pedestrian-level wind field under the TWF effect, Weerasuriya et al. [16]
introduced new inflow boundary conditions to model the wind speed profiles of the TWF.
Appl. Sci. 2024, 14, 988 3 of 23
Feng et al. [17] and Lu and Li [18] incorporated the Coriolis force term in the momentum
equations to numerically generate the TWF profiles, so that the influence of the Coriolis
force on the TWF could be clearly demonstrated by comparing the TWF profiles with the
reference profiles generated without this term. However, most of these studies generated
the TWF profiles directly as the inlet boundary condition, and thus they hardly provided
useful advice for the design of guide vane systems in wind tunnel tests. On the other hand,
by combining advanced CFD techniques with the skills and experience in wind tunnel
testing accumulated over the decades, the concept of a numerical wind tunnel has emerged.
A numerical wind tunnel mimics its physical counterpart by setting the inlet boundary
condition as a uniform flow and creating models of passive devices (e.g., grids, spires, and
ground roughness) in the computational domain, allowing the desired ABL wind profile in
the downstream area to be generated in a similar manner to that in an actual wind tunnel.
Aly and Bitsuamlak [19] established a numerical wind tunnel using models of spires and
ground roughness, which allowed them to test solar panels at various geometric scales
with ease. Phuc et al. [20] modeled spires and ground roughness to generate ABL winds,
suggesting that the simulation results of wind speed and turbulence intensity well agree
with those obtained from wind tunnel tests. To investigate the performance of a novel
device for the wind tunnel simulation of transient gust-front outflow, Le et al. [21] created a
numerical wind tunnel and analyzed the flow field generated by this device in detail. Given
the successful applications of numerical wind tunnels in the literature, it is reasonable
to assume that one can explore various experimental setups in a numerical environment
through CFD techniques to identify the optimal one, which can then be implemented in
an actual (i.e., physical) wind tunnel test. Such an approach, known as CFD-aided wind
tunnel testing, is highly beneficial for testing novel experimental devices for which prior
experience is limited, enabling researchers to gain insights into the performance of different
designs of this device and to select the optimal design for the tests in the actual wind tunnel.
In this study, a procedural framework will be proposed to effectively optimize the
wind tunnel experimental setup for TWF simulations utilizing CFD techniques, with a
particular focus on the design of a guide vane system. The present paper contributes to the
improvement in experimental and numerical simulation techniques for the TWF, promoting
the application of the CFD-aided wind tunnel testing approach in the investigation of
non-conventional wind flow. This paper is organized as follows: Section 2 proposes the
procedural framework for the optimization of a wind tunnel setup; Section 3 presents the
preparation stage of the framework, including the selection of target wind profiles and
the basic information of the wind tunnel setup; Section 4 presents the development stage,
including the establishment of a numerical wind tunnel using CFD techniques and the
optimization of the design of a guide vane system; Section 5 describes the closeout stage,
which finalizes the optimal experimental setup in the actual wind tunnel and validates the
results, followed by the concluding remarks in Section 6. A nomenclature for the main
parameters involved in this study is presented below for ease of reading.
2. Procedural Framework
To generate the desired TWF in the boundary layer wind tunnel efficiently and prop-
erly, this study proposes a procedural framework for the configuration of the wind tunnel
experimental setup using CFD techniques. The general approach is to first preliminarily
determine the experimental setup for the generation of the TWF in the numerical wind
tunnel established using CFD techniques and, subsequently, optimize the experimental
setup by conducting tests in the actual (i.e., physical) wind tunnel. Figure 1. illustrates
the procedural framework of the CFD-aided wind tunnel simulation of the TWF in the
form of a flowchart, demonstrating that the framework consists of three stages, namely, the
preparation stage, the development stage, and the closeout stage. The steps involved in
each stage are described as follows:
Appl. Sci. 2024, 14, x FOR PEER REVIEW
Appl. Sci. 2024, 14, 988 4 of 23
Figure 1. Procedural framework for CFD-aided wind tunnel simulation of twisted wind flow.
Figure 1. Procedural framework for CFD-aided wind tunnel simulation of twisted
2.1. Preparation Stage
2.1.Step A determines the target wind profiles of the TWF as well as the passive devices
Preparation Stage
needed to generate these profiles, such as grids, spires, ground roughness, and guide
vanes. Step A determines
To describe the target wind
the TWF comprehensively, this profiles of the
study proposes TWF
three as profiles:
target well as the p
(1) the longitudinal mean wind speed profile, U (z), where z denotes the height; (2) the
needed to generate these profiles, such as grids, spires, ground roughne
longitudinal turbulence intensity profile, Iu (z); and (3) the twist angle profile, θT (z).
vanes. To describe the TWF comprehensively, this study proposes three tar
2.2.
theDevelopment Stage mean wind speed profile, U(z), where z denotes the heig
longitudinal
Step B validates
gitudinal the CFD
turbulence setup for profile,
intensity the TWF simulations
Iu (z); andby
(3)comparing
the twisttheangle
resultsprofile,
obtained from the numerical wind tunnel with those from an actual one. The validation
involves the selection of the turbulence model, the computational domain, the grid
2.2. Development
scheme, and the solutionStage
strategy in CFD. Only the CFD setup that leads to a simulation
error, er WT,CFD , below the preset threshold, er lim , will be adopted for further steps
Step B validates the CFD setup for the TWF simulations by compar
(i.e., er WT,CFD < er lim ). To ensure sufficient accuracy of the CFD simulations, this study
obtained
sets fromerthe
the threshold lim at numerical
20%. Notably, wind tunnel
this step aims with
only tothose from
validate an actual
the CFD setup, one.
and thereforethe
involves the TWF generated
selection of atthe
this turbulence
step is generated withoutthe
model, the characteristics
computational of dom
ABL winds.
scheme, and the solution strategy in CFD. Only the CFD setup that leads t
Step C optimizes the design of the guide vane system to generate the desired twisted
error,
wind erWT,CFD
field. , below
The design the
includes two preset
aspects,threshold,
including (1)er
the , will be shape
limgeometrical adopted
of thefor furt
erWT,CFD < erlim ). To ensure sufficient accuracy of the CFD simulations, this
threshold erlim at 20%. Notably, this step aims only to validate the CFD se
fore the TWF generated at this step is generated without the characteristics
Step C optimizes the design of the guide vane system to generate the d
Appl. Sci. 2024, 14, 988 5 of 23
vane and (2) the spatial configuration of the vanes. The shape of a guide vane is quantified
by two terms, namely, the variation in the twist angle with height, θvane (z), and the total
height of the guide vane, Hvane . As for the spatial configuration, the location of the i-th
guide vane is denoted by ( xi , yi )vane using a Cartesian coordinate system. By adjusting the
above parameters of the guide vanes in the numerical wind tunnel, a TWF over the entire
model test region, typically the turntable area, can be generated.
Step D incorporates the ABL wind characteristics into the TWF generated at Step C.
This is achieved by adding the conventional passive devices (e.g., grids, spires, and ground
roughness) into the numerical wind tunnel. Multiple parameters, such as the amount,
dimensions, and locations of these conventional devices, are carefully adjusted to generate
the desired TWF. The setting up of the conventional devices can usually be accomplished
with ease based on the experience gained from the actual wind tunnel.
Step E evaluates the performance of the experimental setup established at the previous
step. If the error for numerical simulation (i.e., the difference between the simulation results
and the target values), denoted by erobj,CFD , is lower than the preset threshold of erlim = 30%
(proposed based on practical experience), the current experimental setup will be adopted
for the next step in the closeout stage. Otherwise, the procedures in Steps C and D will be
carried out again until the requirement erobj,CFD < erlim is met.
3. Preparation Stage
3.1. Target Wind Profiles
As pointed out by previous full-scale field measurement studies [3,5,8,22], the profiles
of the TWF are highly dependent on the meteorological background and the local terrain
effect. A variety of wind profile models have been proposed to quantitatively describe the
TWF caused by different mechanisms. This study selects four widely recognized models
of twisted wind angle for discussion, namely, (1) the Ekman spiral model for the TWF
induced by Coriolis force [3,23], (2) the inverse proportion model [3] and the exponential
law model [8] for the TWF induced by mountainous terrain, and (3) the logarithmic law
model [4] for the TWF induced by tropical cyclones over coastal areas.
In particular, for the Ekman spiral model, the profiles of the wind speed in the longitu-
dinal and lateral directions, denoted by u and v, respectively, and the twist angle, θ, are
expressed by:
− hz z − hz z
u(z) = u g 1 − e E cos − vg e E sin (1)
hE hE
− z z − z z
v(z) = v g 1 − e hE cos − u g e hE sin (2)
hE hE
v(z)
θ (z)= tan−1 (3)
u(z)
where u g and v g are the gradient wind speeds in the longitudinal and lateral directions,
respectively, and h E is the Ekman height determined by the Coriolis parameter and the
turbulence viscosity coefficient.
By setting the maximum twist angle at the ground level at 30◦ and that at a height of
500 m at 0◦ , Figure 2 depicts the twist angle profiles obtained from the above four models.
It is readily observed that the twist angle following the inverse proportion model rapidly
Appl. Sci. 2024, 14, 988 6 of 23
drops to a value below 5◦ , which is a very small angle corresponding to negligible TWF
effects, as the height reaches 100 to 200 m, and those following the exponential law
Appl. Sci. 2024, 14, x FOR PEER REVIEW 6 ofand
24
the logarithm law models decrease to approximately 5◦ at a height of 200 m. In contrast,
the twist angle expressed by the Ekman spiral model remains higher than 5◦ until the
height reaches 400 m, indicating that the twisted wind effect represented by such a model
is much more prominent than those represented by the three other models. Therefore, to
is much more prominent than those represented by the three other models. Therefore, to
consider the most prominent twisted wind effect for conservative purposes, this study
consider the most prominent twisted wind effect for conservative purposes, this study
selects the twist angle profile expressed by the Ekman spiral model as the target profile
selects the twist angle profile expressed by the Ekman spiral model as the target profile for
for
thethe generation
generation of the
of the TWFTWF in the
in the wind
wind tunnel.
tunnel.
Figure
Figure 2.
2. Comparison
Comparison of
of common
common twist
twist angle
angle curve
curve models.
models.
3.2. Wind
3.2. Wind Tunnel
Tunnel Setup
Setup
AA series
series of
of tests
tests were
were conducted
conducted in in the
the Wind
Wind Tunnel
Tunnel Laboratory
Laboratory at
at the
the School
School of
of Civil
Civil
Engineering, Chongqing
Engineering, Chongqing University.
University. The
The test
testsection
sectionisis2.4
2.4mm×× 1.8
1.8 m (breadth ×
m (breadth height) in
× height) in
size, and the turntable is 2.0 m in diameter. Multiple passive devices were
size, and the turntable is 2.0 m in diameter. Multiple passive devices were employed in employed in
the wind tunnel to generate the target TWF profile that follows the Ekman spiral model.
the wind tunnel to generate the target TWF profile that follows the Ekman spiral model.
Different combinations of spires and ground roughness elements, with the dimensions
Different combinations of spires and ground roughness elements, with the dimensions
indicated in Figure 3, were available to generate the ABL wind profiles. In this study, the
indicated in Figure 3, were available to generate the ABL wind profiles. In this study, the
target profiles of wind speed and turbulence intensity are expressed as follows:
target profiles of wind speed and turbulence intensity are expressed as follows:
α𝛼
u
z) = = ( 𝑢 )
u(𝑢(𝑧) (4)
(4)
u𝑢g
𝑔
Iu −−𝛼−0.1
α−0.1
Iu (z) = 𝐼𝑢 (5)
𝐼𝑢 (𝑧) = (Iu,g ) (5)
𝐼𝑢,𝑔
where u g and Iu,g denote the longitudinal wind speed and turbulence intensity at the
where 𝑢𝑔 and 𝐼𝑢,𝑔 denote the longitudinal wind speed and turbulence intensity at the
gradient height, respectively, and α denotes the power-law exponent, which is taken as 0.12
gradient height, respectively, and 𝛼 denotes the power-law exponent, which is taken as
in this study.
0.12 inOn thestudy.
this other hand, the TWF was generated using the guide vane system shown in
Figure 4. This system comprises two identical vanes made of 10 mm thick fiberglass
molded into a single piece. Each guide vane is of a total height of 1.5 m and consists of
two sections, as plotted in Figure 5, namely, a twist section ranging from ground level to
1.0 m in height and a transition section from 1.0 m to 1.5 m. The leading edge of the twist
section is straight, whilst the trailing edge is curved to form a guide angle varying with
height to redirect the approach flow. Since the Ekman spiral model has been selected in
Appl. Sci. 2024, 14, 988 7 of 23
this study as discussed above, the trailing edge is curved following Equation (3) with a
twist angle of 30◦ at ground level and 0◦ at a height of 1.0 m. The transition section is
in the shape of a straight board to avoid altering the flow direction and to prevent the
unfavorable eddies generated at its leading edge from reaching the downstream area.
At the preparation stage, the two vanes were positioned as shown in Figure 6a, i.e., one
vane was installed on the right side of the wind tunnel (as observed in Figure 4)7 with
Appl. Sci. 2024, 14, x FOR PEER REVIEW of 24
its trailing edge 850 mm from the side wall, and the other vane was installed on the left
side with a separation of 900 mm between the vanes.
On the other hand, the TWF was generated using the guide vane system shown in
Figure 4. This system comprises two identical vanes made of 10 mm thick fiberglass
molded into a single piece. Each guide vane is of a total height of 1.5 m and consists of
two sections, as plotted in Figure 5, namely, a twist section ranging from ground level to
1.0 m in height and a transition section from 1.0 m to 1.5 m. The leading edge of the twist
section is straight, whilst the trailing edge is curved to form a guide angle varying with
height to redirect the approach flow. Since the Ekman spiral model has been selected in
this study as discussed above, the trailing edge is curved following Equation (3) with a
twist angle of 30° at ground level and 0° at a height of 1.0 m. The transition section is in
the shape of a straight board to avoid altering the flow direction and to prevent the unfa-
vorable eddies generated at its leading edge from reaching the downstream area. At the
preparation stage, the two vanes were positioned as shown in Figure 6a, i.e., one vane was
installed on the right side of the wind tunnel (as observed in Figure 4) with its trailing
edge 850 mm from the side wall, and the other vane was installed on the left side with a
separation of 900 mm between the vanes.
Figure
Figure 4. 4. Guide
Guide vanes
vanes for tunnel
for wind wind simulation
tunnel simulation of twisted
of twisted wind flow. wind flow.
Appl. Sci. 2024, 14, 988 8 of 23
Figure 4. Guide vanes for wind tunnel simulation of twisted wind flow.
Figure 5. Dimensions of guide vanes: (a) front view; (b) top view.
Figure 5. Dimensions of guide vanes: (a) front view; (b) top view.
To comprehensively measure the flow field generated in the wind tunnel, a Cobra
probe (Turbulent Flow Instrumentation Pty. Ltd., Victoria, Australia) was deployed to
measure the winds over five different locations within the model test region (i.e., turnta
ble), indicated as Points A to E in Figure 6a. The winds were measured at 13 differen
heights ranging from 50 to 1050 mm, as shown in Figure 6b, i.e., the heights from 50 to 250
mm with intervals of 50 mm and from 250 to 1050 with intervals of 100 mm. By setting th
geometric scale as 1:500, the above height range in the model scale corresponds to 25 to
525 m in full scale. The measurements were performed for 30 s at each height. The sam
pling frequency was set at 625 Hz, which is adequately high to capture the turbulenc
characteristics of the flow. To obtain accurate mean wind speed and direction profiles, th
measurements were repeated three times at each height, and the average values wer
adopted.
Figure 6. Positions
Figure of of
6. Positions guide
guidevanes andmeasurement
vanes and measurement points:
points: (a)view;
(a) plan plan(b)
view; (b) side view.
side view.
To comprehensively measure the flow field generated in the wind tunnel, a Cobra
Figure 7 depicts the twist angle profiles measured over Points A to E. These profi
probe (Turbulent Flow Instrumentation Pty. Ltd., Victoria, Australia) was deployed to
generally match
measure eachover
the winds other, suggesting
five different thatwithin
locations the TWF generated
the model by (i.e.,
test region the turntable),
vanes was distr
uted indicated
over theasmodel
Points test
A to region in a6a.
E in Figure uniform manner
The winds (i.e., flow
were measured speed
at 13 andheights
different direction at
same height are similar within the region). It is also observed that the measured maximu
twist angles were slightly below 30°, which is attributed to the inevitable attenuation
the twisting characteristic along the flow direction [8]. Overall, the variations in the me
ured twist angles decreased slowly with height following the Ekman spiral model, in
Appl. Sci. 2024, 14, 988 9 of 23
ranging from 50 to 1050 mm, as shown in Figure 6b, i.e., the heights from 50 to 250 mm with
intervals of 50 mm and from 250 to 1050 with intervals of 100 mm. By setting the geometric
scale as 1:500, the above height range in the model scale corresponds to 25 to 525 m in full
scale. The measurements were performed for 30 s at each height. The sampling frequency
was set at 625 Hz, which is adequately high to capture the turbulence characteristics of the
flow. To obtain accurate mean wind speed and direction profiles, the measurements were
repeated three times at each height, and the average values were adopted.
Figure 7 depicts the twist angle profiles measured over Points A to E. These profiles
generally match each other, suggesting that the TWF generated by the vanes was distributed
over the model test region in a uniform manner (i.e., flow speed and direction at the same
height are similar within the region). It is also observed that the measured maximum twist
angles were slightly below 30◦ , which is attributed to the inevitable attenuation of the
twisting characteristic along the flow direction [8]. Overall, the variations in the measured 10 of 24
Appl. Sci. 2024, 14, x FOR PEER REVIEW
twist angles decreased slowly with height following the Ekman spiral model, indicating
that the selection of the shape of the guide vanes used in this study was reasonable.
7. Twist
Figure 7.
Figure Twistangle
angleprofiles measured
profiles within
measured model
within test region.
model test region.
4. Development Stage
4. Development
4.1. Establishment ofStage
Numerical Wind Tunnel
4.1.
4.1.1.Establishment
CFD Setup forofNumerical
NumericalWind
WindTunnel
Tunnel
4.1.1.InCFD
the Setup
CFD simulation, the complex
for Numerical fluid dynamics issues can be simplified for
Wind Tunnel
numerical resolution through the Navier–Stokes (N-S) equations. By neglecting the
In the CFD simulation, the complex fluid dynamics issues can be simplified for nu-
temperature effect, a closed set of governing equations is formed by the continuity
merical resolution through the Navier–Stokes (N-S) equations. By neglecting the temper-
equation and the momentum conservation equation. In this study, both the Reynolds-
ature effect, a closed set(RANS)
Averaged Navier–Stokes of governing equations
model and is formed
the Large-Eddy by the continuity
Simulation equation
(LES) model are and
the momentum
utilized conservation
for the CFD equation. In this study, both the Reynolds-Averaged Navier–
simulations.
Stokes (RANS) model and the Large-Eddy Simulation (LES) model are utilized for the
CFD simulations.
Vortex filtration is involved in the LES model by considering the feature grid size.
Large-scale vortices are solved directly by the N-S equation, and small-scale ones are
solved by the sub-grid scale model. The governing equations in the instantaneous state
after filtration can be expressed as:
Appl. Sci. 2024, 14, 988 10 of 23
Vortex filtration is involved in the LES model by considering the feature grid size.
Large-scale vortices are solved directly by the N-S equation, and small-scale ones are solved
by the sub-grid scale model. The governing equations in the instantaneous state after
filtration can be expressed as:
∼
∂ui
=0 (6)
∂xi
∼ ∼ ∼ ∼
∂ui ∼ ∂ui ∂p 1 ∂2 u i ∂τij
+ uj =− + 2
− (7)
∂t ∂x j ∂xi Re ∂x j ∂x j
∼ ∼∼
where τij is the sub-grid scale stress, given by τij = ρui u j − ρ u i u j to make the equations closed.
The standard Smagorinsky sub-grid model [24] is adopted in this study as the LES
turbulence model, expressed as:
1 ∼
τij − τkk δij = −2µt S ij (8)
3
→
µt = (Cs ∆)2 | S | (9)
∼ ∼ ∼
S ij = 1/2 ∂ u i /∂x j + ∂ u j /∂x I (10)
r
→ ∼ ∼
|S| = 2 S ij S ij (11)
1
∆ = ∆ x ∆y ∆z 3 (12)
→
where µt is the dynamic viscosity in the sub-grid scale; | S | is the velocity scale; ∆ x , ∆y , and
∆z are the grid size in the x, y, and z directions, respectively; and Cs is the Smagorinsky
constant taken as 0.1 in the present study.
On the other hand, statistics are used in the RANS approach to represent turbulence
flow as time-averaged and fluctuating parts. When the N-S equations are averaged, the
instantaneous velocity is regarded as the sum of the average velocity and the fluctuation
velocity. The equations are expressed below:
∂ui
=0 (13)
∂xi
" ! #
∂ui ∂ui 1 ∂p ∂ ∂ui ∂u j ′ ′
+ uj =− + υ + − ui u j (14)
∂t ∂x j ρ ∂xi ∂x j ∂x j ∂xi
where ρ is the air density; I and u j are the wind velocity components; υ is the coefficient
of kinetic viscosity; and Ii and x j are the coordinates vector of corresponding velocity
components. An additional term, namely the Reynolds stress ρui′ u′j , is contained in the
incompressible N-S equations for the RANS simulation. To make the equations closed,
several turbulence models were proposed based on different assumptions about this term,
including the Standard k-ε model [25], the RNG k-ε model [26], the Realizable k-ε model [27],
the Standard k-ω model [28], and the SST-k-ω model [29].
Using CFD techniques, a numerical wind tunnel was established as a replica of the
wind tunnel at Chongqing University introduced in Section 3. As shown in Figure 8, the
dimensions of the numerical wind tunnel were identical to those of its physical counterpart,
and all passive devices available for this study, as shown in Figure 3, were also modeled
numerically. As for the boundary conditions, the computational domain employed a
velocity inlet and a pressure outlet, the floor was set as a rough wall, and all other surfaces,
including the surfaces of the experimental devices, were set as non-slip walls. A mixed-grid
scheme was adopted for both the RANS and LES simulations. The structured grids were
generated near the inlet and outlet, while unstructured grids were employed in the region
Appl. Sci. 2024, 14, 988 11 of 23
Figure 8. Basic
Figure 8. Basic information information
of numerical of numerical
wind tunnel. wind tunnel.
Figure 10.
Figure Comparisons of
10. Comparisons of mean
mean wind
wind speeds
speeds obtained
obtained by
by CFD
CFD with
with different
different RANS
RANS turbulence
turbulence
models and by actual wind tunnel tests.
Figure 10. Comparisons of mean wind speeds obtained by CFD with different RANS turb
models and by actual wind tunnel tests.
Appl. Sci. 2024, 14, x FOR PEER REVIEW 14
Appl. Sci. 2024, 14, 988 4.2. Optimization of Guide Vane Configuration 13 of 23
Figure 11. Distribution of lateral mean wind speed at z = 0.3 m: (a) d vane,wall = 450 mm;
Figure
(b) 11. =Distribution
d vane,wall 600 mm; (c) dof lateral= mean
vane,wall wind speed at z = 0.3 m: (a) 𝑑𝑣𝑎𝑛𝑒,𝑤𝑎𝑙𝑙 = 450 mm; (b)
750 mm.
𝑑𝑣𝑎𝑛𝑒,𝑤𝑎𝑙𝑙 = 600 mm; (c) 𝑑𝑣𝑎𝑛𝑒,𝑤𝑎𝑙𝑙 = 750 mm.
4.2.2. Spacing between Guide Vanes
After determining d = 750 mm, the next step is to identify the optimal
4.2.2. Spacing betweenvane,wallGuide Vanes
spacing between the two vanes. Figure 12 shows the distribution of the lateral wind
speed After determining
at z = 0.3 𝑑𝑣𝑎𝑛𝑒,𝑤𝑎𝑙𝑙
m with d vane,wall = 750 mm= 750
andmm, the nextbetween
the spacing step isthe
to identify the optimal s
vanes, denoted
ingd vane,vane
by between herein, ranging
the two vanes.fromFigure
700 to 12
1000 mm with
shows the intervals of 100ofmm.
distribution the The figure
lateral wind spe
shows that a so-called “low-speed area”, where the mean lateral
z = 0.3 m with 𝑑𝑣𝑎𝑛𝑒,𝑤𝑎𝑙𝑙 = 750 mm and the spacing between the vanes, denotewind speed is below
0.7 m/s (colored in green), is formed in the downstream region of the vanes. The length
𝑑 herein, ranging from 700 to 1000 mm with intervals of 100 mm. The fi
of 𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒
the low-speed area increases with d vane,vane and eventually extends to the turntable
shows
after that areaches
d vane,vane so-called
900“low-speed
mm. Obviously, area”, where
the the mean
low-speed lateral
area is wind speed
unfavorable to the is below
m/s (colored
generation in green),
of evident TWF is formed
within in thetest
the model downstream
region, and,region of the
therefore, theconfiguration
vanes. The length o
low-speed
with d vane,vanearea
= 700increases
mm, which with 𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒
corresponds and
to the eventually
smallest extends
low-speed area, to the turntable
is selected
for the following simulations.
𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒 reaches 900 mm. Obviously, the low-speed area is unfavorable to the gen
tion of evident TWF within the model test region, and, therefore, the configuration
𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒 = 700 mm, which corresponds to the smallest low-speed area, is selected fo
following simulations.
Appl. Sci.
Appl. Sci. 2024,
2024, 14, x FOR PEER REVIEW
14, 988 1514of
of 24
23
Figure 12. Distribution of lateral mean wind speed at z = 0.3 m: (a) d vane,vane = 700 mm;
(b) dvane,vane
Figure = 800 mm; of
12. Distribution dvane,vane
(c)lateral mean= 900 mm;
wind (d) dat
speed z = 0.3 =m:1000
vane,vane 𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒 = 700 mm; (b)
(a) mm.
𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒 = 800 mm; (c) 𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒 = 900 mm; (d) 𝑑𝑣𝑎𝑛𝑒,𝑣𝑎𝑛𝑒 = 1000 mm.
4.2.3. Distance between Trailing Edge of Guide Vane and Center of Turntable
4.2.3.As can be readily
Distance betweenobserved
Trailingfrom EdgeFigures 11 and
of Guide Vane12, theCenter
and flow significantly
of Turntableaccelerates in
the lateral direction after it passes the leading edges of the vanes,
As can be readily observed from Figures 11 and 12, the flow significantly forming a long and narrow
accelerates
area with a lateral wind speed over 1.4 m/s (colored in orange in the figure), referred to as
in the lateral direction after it passes the leading edges of the vanes, forming a long and
the “high-speed area” herein, at the downstream region of each vane. These two so-called
narrow area with a lateral wind speed over 1.4 m/s (colored in orange in the figure), re-
high-speed areas begin to merge after a certain distance, and the lateral wind speed becomes
ferred to as the “high-speed area” herein, at the downstream region of each vane. These
more uniformly distributed over the downstream region. Therefore, the distance from
two so-called high-speed areas begin to merge after a certain distance, and the lateral wind
the center of the model test region to the trailing edge of the vanes, denoted by dvane,model
speed becomes more uniformly distributed over the downstream region. Therefore, the
herein, needs to be sufficiently long so that the high-speed areas merge before they reach
distance from the center of the model test region to the trailing edge of the vanes, denoted
the model test region. Meanwhile, if dvane,model is too great, it may cause the attenuation of
by 𝑑𝑣𝑎𝑛𝑒,𝑚𝑜𝑑𝑒𝑙 herein, needs to be sufficiently long so that the high-speed areas merge
the twist characteristics, as pointed out by Tse et al. [8]. Therefore, it is necessary to balance
before they reach the model test region. Meanwhile, if 𝑑𝑣𝑎𝑛𝑒,𝑚𝑜𝑑𝑒𝑙 is too great, it may
dvane,model to ensure a sufficient development of the twist characteristics while avoiding
cause the attenuation of the twist characteristics, as pointed out by Tse et al. [8]. Therefore,
excessive attenuation of these characteristics over distance.
it is necessary to balance 𝑑𝑣𝑎𝑛𝑒,𝑚𝑜𝑑𝑒𝑙 to ensure a sufficient development of the twist char-
Based on the numerical simulation results, Figure 13 illustrates the profiles of the
acteristics
longitudinalwhilewindavoiding
speed,excessive
the lateral attenuation
wind speed,of these
andcharacteristics
the twist angle. overThedistance.
X-axis in
Based on the numerical simulation results, Figure 13 illustrates
Figure 13 originates from the inlet boundary of the numerical wind tunnel illustrated the profiles of the
in
longitudinal
Figure 9, andwind speed,
the vanes the set
were lateral windtheir
up with speed, and the
leading twist
edges at angle.
X=6m The X-axis
and in Figure
trailing edges
13
at Xoriginates
= 7 m. The from
windtheprofiles
inlet boundary of the numerical
in the downstream region of wind
the tunnel
vanes atillustrated
X = 8 to 12inm,Figure
with
9, and the vanes were set up with their leading edges at X = 6 m and trailing
intervals of 1 m, are plotted in Figure 13 to demonstrate the flow propagation over distance. edges at X =
7Inm. The wind
addition, theprofiles in the downstream
wind profiles regionregion
in the upstream of theatvanes
X = at5m X =are
8 to 12 plotted
also m, with ininter-
the
vals of as
figure 1 m, are plottedItin
a reference. is Figure
observed 13 to demonstrate
that the flow
the longitudinal windpropagation
speeds at over distance.
all heights In
were
addition, the wind
approximately profiles
12 m/s while inthe
thelateral
upstreamwindregion
speeds at and
X = 5twist
m are also plotted
angles in the
were zero, figure
which is
as a reference. It is observed that the longitudinal wind speeds at all heights
desired because the inflow was set to be uniform in the simulation. As shown in Figure 13a, were approx-
imately
the wind12speeds
m/s while the lateral
at heights below wind
0.5 mspeeds and due
dropped twisttoangles were zero,
the blockage which
caused is desired
by the vanes,
because the inflow was set to be uniform in the simulation.
whilst those above 0.5 m maintained the same amplitude as in the upstream region.As shown in Figure 13a,The
the
wind speeds at heights below 0.5 m dropped due to the blockage caused by the vanes,
Appl. Sci. 2024, 14, x FOR PEER REVIEW 16 of 24
whilst those above 0.5 m maintained the same amplitude as in the upstream region. The
lateral wind speeds and the twist angles, as observed from Figure 13b,c, became evident
lateral wind speeds and the twist angles, as observed from Figure 13b,c, became evident
at heights below around 0.6 m due to the guide vanes. Since the profiles at X = 11 m are
at heights below around 0.6 m due to the guide vanes. Since the profiles at X = 11 m are
similar tothose
similar to thoseatatXX= =1212 m,m, this
this indicates
indicates thatthat the high-speed
the high-speed areas areas generated
generated by the two
by the two
vanes have been merged and created a stable TWF field uniformly
vanes have been merged and created a stable TWF field uniformly distributed over these distributed over these
distances. Hence,
distances. Hence, it isitconcluded
is concluded that X from
that X ranging ranging
11 to from 11 equivalently,
12 m, or, to 12 m, or, equivalently,
dvane,model
𝑑
ranging from
𝑣𝑎𝑛𝑒,𝑚𝑜𝑑𝑒𝑙 ranging
4 to 5 m, from
is 4
the to 5
optimalm, is the
value optimal
for the value
present for
wind the present
tunnel wind
setup. tunnel setup.
Figure 13. Wind profiles at different locations on the X-axis: (a) longitudinal mean wind speed;
Figure 13. Wind profiles at different locations on the X-axis: (a) longitudinal mean wind speed; (b)
(b) longitudinal turbulence intensity; (c) twist angle.
longitudinal turbulence intensity; (c) twist angle.
Appl. Sci. 2024, 14, x FOR PEER REVIEW 1
Appl. Sci. 2024, 14, 988 16 of 23
Figure 15. Time variations in the instantaneous wind speeds: (a) longitudinal speed; (b) lateral speed.
Figure 15. Time variations in the instantaneous wind speeds: (a) longitudinal speed; (b)
speed.
Figure 16 plots the profiles of TWF at the center of the model test region. As shown
in Figure 16a, the longitudinal wind speed and turbulence intensity follow the power
Figure 16 plots the profiles of TWF at the center of the model test region. As s
law expressed by Equation (4) and Equation (5), respectively. This indicates that the
in Figure 16a, the longitudinal wind speed and turbulence intensity follow the powe
objective of incorporating ABL characteristics is generally satisfied. On the other hand,
expressed by Equation (4) and Equation (5), respectively. This indicates that the ob
Appl. Sci. 2024, 14, 988 of incorporating ABL characteristics is generally satisfied. On the other18hand, of 23 Figu
shows that the twist angle profile is generally in good agreement with the target
(i.e., theFigure
Ekman spiralthat
16b shows model) at heights
the twist above
angle profile 0.4 m. However,
is generally the twist
in good agreement with angles
the be
m weretargetconsiderably lower
profile (i.e., the Ekmanthan themodel)
spiral desired values,
at heights which
above 0.4 m.isHowever,
believedtheto be attrib
twist
angles below 0.4 m were considerably lower than the desired values, which is believed
the attenuation over distance and will be further improved at the subsequent c
to be attributed to the attenuation over distance and will be further improved at the
stage. subsequent closeout stage.
5. Closeout Stage
5. CloseoutFollowing
Stage the setup in the numerical wind tunnel as discussed in Section 4.3, the
passive devices, including the vane system, were implemented in the actual wind tunnel
Following the setup in the numerical wind tunnel as discussed in Section 4
as shown in Figure 17. Considering that the twist angle values at heights below 0.4 m
passiveweredevices, including
lower than the desired thevalues,
vanethe system, were
guide vanes implemented
were moved closer toin thethe actual
model test wind
as shown regionin by
Figure
200 mm 17. Considering
(i.e., dvane,model = 3.8that
m) tothe twist
alleviate theangle values
attenuation at heights
of twist below 0.4
characteristics.
Table 2 summarizes the final experimental setup of the guide vanes at the closeout stage,
lower than the desired values, the guide vanes were moved closer to the model test
i.e., the optimal guide vane setup obtained following the procedure framework proposed
by 200 by mm study. 𝑑
this(i.e., It𝑣𝑎𝑛𝑒,𝑚𝑜𝑑𝑒𝑙
is noteworthy = that
3.8 the
m)experimental
to alleviate the are
setups attenuation of twist
highly dependent on thecharacte
Table 2specifications
summarizes the
of the windfinal experimental
tunnel, and, therefore,setup of theofguide
modifications vanes atmay
setup parameters thebecloseou
needed when generating the TWF wind profiles in another wind tunnel.
i.e., the optimal guide vane setup obtained following the procedure framework pr
by thisTable
study. It is noteworthy
2. Summary that the
of optimal experimental experimental
setup of guide vanes. setups are highly dependent
specifications of the wind tunnel, and, therefore, modifications of setup paramete
Parameter Value
be needed when generating No. of vanes
the TWF wind profiles in another
2
wind tunnel.
θvane (z) model Ekman spiral
θvane (0) experimental setup of guide vanes.
30 ◦
Table 2. Summary of optimal ◦
θvane (1) 0
θvane (1.5) 0◦
Parameter
Hvane 1.5 m Value
No. ofdvane,wall
vanes 750 mm 2
dvane,vane 700 mm
θvane (𝑧dvane,model
) model 3800 mmEkman spiral
θvane (0) 30°
θvane (1) 0°
θvane (1.5) 0°
Hvane 1.5 m
𝑑𝑣𝑎𝑛𝑒,𝑤𝑎𝑙𝑙 750 mm
Appl. Sci. 2024, 14, x FOR PEER REVIEW
Appl. Sci. 2024, 14, 988 19 of 23
Figure
Figure 17. 17.
FinalFinal
setup setup
in actualin actual
wind wind
tunnel. tunnel.
Figure 18 presents the variations in the profiles of longitudinal wind speed and
Figure
turbulence 18 obtained
intensity presents the
from thevariations in tunnel
numerical wind the profiles
over time,ofindicating
longitudinal
that the wind
bulence intensity obtained from the numerical wind tunnel over time, ind
flow simulated by the LES technique became steady within approximately 30 s. In addition,
the profiles obtained from the actual wind tunnel tests are also plotted in Figure 18 for
flow simulated by the LES technique became steady within approximatel
comparison purposes, and good agreement between the LES-simulated profiles and the
tion, the
measured profiles
ones obtained
can be readily from
observed. the
Such actual suggests
agreement wind tunnel tests arewind
that the numerical also plotte
for comparison
tunnel can serve as anpurposes,
efficient and and good
accurate agreement
approach between
to provide the LES-simulate
useful reference to the
actual wind tunnel tests.
the measured ones can be readily observed. Such agreement suggests that
Figure 19 plots the profiles of longitudinal speed, turbulence intensity, and twist an-
wind
gle of thetunnel
TWF. Ascan serveinasFigure
presented an efficient
19a, both and accurate
profiles approach
follow the power law to provide use
expressed
bythe actual (4)
Equations wind tunnel
and (5) with antests.
exponent of 0.12 and −0.22, respectively. Furthermore,
the profiles of conventional ABL wind flow (CWF) without twist were also generated
in the wind tunnel as a reference. Good agreement can be readily observed between
the profiles of the TWF and those of the CWF, indicating that the experimental setup
for the TWF has produced ABL characteristics that are satisfactory. On the other hand,
Figure 19b shows that the maximum twist angle of the TWF reached 27◦ , which is only
slightly below the 30◦ set for the Ekman spiral model, and the measured twist angles well
match the target values with a maximum discrepancy below 3◦ . Furthermore, Figure 20
shows that the power spectral densities of the longitudinal and lateral wind speeds,
obtained by the Welch and Yule–Walker methods, are in good agreement with the Von
Karman spectrum model. These results indicate that the guide vane system, configured
following the procedural framework proposed in this study, has generated the desired
TWF with satisfying accuracy.
tion, the profiles obtained from the actual wind tunnel tests are also plotted in Figure 18
for comparison purposes, and good agreement between the LES-simulated profiles and
the measured ones can be readily observed. Such agreement suggests that the numerica
Appl. Sci. 2024, 14, 988 wind tunnel can serve as an efficient and accurate approach to provide useful reference
20 of 23 to
the actual wind tunnel tests.
Figure 18. Variations in wind profiles at different locations on the X-axis: (a) longitudinal mean w
speed; (b) twist angle.
Figure 19 plots the profiles of longitudinal speed, turbulence intensity, and twist
gle of the TWF. As presented in Figure 19a, both profiles follow the power law expres
by Equations (4) and (5) with an exponent of 0.12 and −0.22, respectively. Furthermo
the profiles of conventional ABL wind flow (CWF) without twist were also generated
the wind tunnel as a reference. Good agreement can be readily observed between the p
files of the TWF and those of the CWF, indicating that the experimental setup for the TW
has produced ABL characteristics that are satisfactory. On the other hand, Figure
shows that the maximum twist angle of the TWF reached 27°, which is only slightly bel
the 30° set for the Ekman spiral model, and the measured twist angles well match
target values with a maximum discrepancy below 3°. Furthermore, Figure 20 shows t
the power spectral densities of the longitudinal and lateral wind speeds, obtained by
Welch and Yule–Walker methods, are in good agreement with the Von Karman spectr
model. These results indicate that the guide vane system, configured following the pro
dural framework proposed in this study, has generated the desired TWF with satisfy
Figure 18. Variations in wind profiles at different locations on the X-axis: (a) longitudinal mean wind
accuracy.
speed; (b) twist angle.
FigureFigure
20. Wind spectra
20. Wind of of
spectra TWFTWFat
at zz = 1.0 m:(a)
1.0 m: (a)longitudinal
longitudinal speed;
speed; (b) lateral
(b) lateral speed. speed.
6. Concluding Remarks
This paper has presented a procedural framework to achieve the optimal experimental
setup for the generation of TWF in a wind tunnel based on CFD simulation results. Follow-
ing the proposed procedures, the optimal configuration of the passive devices, especially
the design of the guide vane system, to generate the desired TWF can be efficiently obtained
by CFD simulation results in a numerical wind tunnel, enabling informed decisions for the
final experimental setup in the actual wind tunnel. Several concluding remarks are drawn
as follows:
(1) The numerical wind tunnel established using the RANS and LES techniques can serve
as an effective tool to examine the rationality of the experimental setup. The RNG k-ε
model for RANS delivered the best predictions of wind speed in a numerical wind
tunnel, according to the comparison with the results from the actual wind tunnel. The
LES model, the more computationally efficient approach, was also able to numerically
simulate the wind field with satisfying accuracy.
(2) Based on the numerical simulation results, the mechanism of a guide vane system
in a TWF simulation has been discussed in detail. Three parameters governing the
positions of the vanes have been emphasized in this study and need to be considered
carefully. First, dvane,wall is required to be sufficiently large so that the flow can
travel through the gap and form a desired high-speed area in the downstream region.
Second, dvane,vane needs to be sufficiently low to shorten the unfavorable length of the
low-speed area and to prevent this area from reaching the model test region. Third,
dvane,model must strike a balance so that it is not too large to cause the attenuation of
the twist characteristics and not too small to allow the low-speed region to reach the
model test region.
(3) The numerical wind tunnel testing has demonstrated its applicability in providing
valuable information for optimizing the experimental setups in the actual wind tunnel.
By implementing passive devices in the actual wind tunnel following the setup used
in its numerical counterpart, the desired TWF profiles have been generated with
satisfying accuracy, i.e., the profiles of wind speed and turbulence intensity follow
the power-law model for typical ABL winds, while the twist angle profile follows the
Ekman spiral model. These results underscore the importance of utilizing numerical
simulations to aid in the design of experimental setups, especially for novel devices
such as the guide vane system discussed in the present study.
Appl. Sci. 2024, 14, 988 22 of 23
Author Contributions: Conceptualization, Z.Y., X.L., Z.Z. and B.Y.; methodology, Z.Y., X.L. and Z.Z.;
software, L.W.; investigation, L.W., X.L. and X.Z.; resources, B.Y.; writing—original draft preparation,
L.W.; writing—review and editing, Z.Y. and X.L.; supervision, B.Y.; project administration, B.Y.;
funding acquisition, B.Y. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the National Natural Science Foundation of China
(No. 51878104 and No. 52278483), the Graduate Research and Innovation Foundation of Chongqing,
China (CYB21028), and the 111 Project of China (B18062).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author. The data are not publicly available due to privacy.
Conflicts of Interest: The authors declare no conflicts of interest.
Nomenclature
References
1. Mendenhall, B.R. A Statistical Study of Frictional Wind Veering in the Planetary Boundary Layer. Atmospheric Science Paper.
Ph.D. Dissertation, Colorado State University, Fort Collins, CO, USA, 1967.
2. Tamura, Y.; Iwatani, Y.; Hibi, K.; Suda, K.; Nakamura, O.; Maruyama, T.; Ishibashi, R. Profiles of mean wind speeds and vertical
turbulence intensities measured at seashore and two inland sites using Doppler sodars. J. Wind Eng. Ind. Aerodyn. 2007, 95,
411–427. [CrossRef]
3. Liu, Z.; Zheng, C.; Wu, Y.; Song, Y. Investigation on characteristics of thousandmeter height wind profiles at non-tropical cyclone
prone areas based on field measurement. Build Environ. 2018, 130, 62–73. [CrossRef]
4. He, Y.C.; Chan, P.W.; Li, Q.S. Observations of vertical wind profiles of tropical cyclones at coastal areas. J. Wind Eng. Ind. Aerodyn.
2016, 152, 1–14. [CrossRef]
5. Shu, Z.R.; Li, Q.S.; He, Y.C.; Chan, P.W. Observational study of veering wind by Doppler wind profiler and surface weather
station. J. Wind Eng. Ind. Aerodyn. 2018, 178, 18–25. [CrossRef]
6. Peña, A.; Gryning, S.E.; Floors, R. The turning of the wind in the atmospheric boundary layer. J. Phys. Conf. Ser. 2014, 524,
012118. [CrossRef]
7. Yan, B.; Li, Y.; Li, X.; Zhou, X.; Wei, M.; Yang, Q.; Zhou, X. Wind tunnel investigation of twisted wind effect on a typical super-tall
building. Buildings 2022, 12, 2260. [CrossRef]
8. Tse, K.T.; Weerasuriya, A.U.; Kwok, K.C.S. Simulation of twisted wind flows in a boundary layer wind tunnel for pedestrian-level
wind tunnel tests. J. Wind Eng. Ind. Aerodyn. 2016, 159, 99–109. [CrossRef]
9. Flay, R.G. A twisted flow wind tunnel for testing yacht sails. J. Wind. Eng. Ind. Aerodyn. 1996, 63, 171–182. [CrossRef]
10. Viola, I.M.; Fossati, F. Downwind Sails Aerodynamic Analysis. In Proceedings of the 6th International Colloquium on Bluff
Bodies Aerodynamics and Applications, Milano, Italy, 20–24 July 2008.
11. Graf, K.; Müller, O. Photogrammetric investigation of the flying shape of spinnakers in a twisted flow wind tunnel. In Proceedings
of the 19th Chesapeake Sailing Yacht Symposium, Annapolis, MD, USA, 20 March 2009.
12. Weerasuriya, A.U.; Tse, K.T.; Zhang, X.L.; Li, S.W. A wind tunnel study of effects of twisted wind flows on the pedestrian-level
wind field in an urban environment. Build Environ. 2018, 128, 225–235. [CrossRef]
13. Zhou, L.; Tse, K.T.; Hu, G. Experimental investigation on the aerodynamic characteristics of a tall building subjected to twisted
wind. J. Wind Eng. Ind. Aerodyn. 2022, 224, 104976. [CrossRef]
Appl. Sci. 2024, 14, 988 23 of 23
14. Liu, Z.; Zheng, C.; Wu, Y.; Flay, R.G.J.; Zhang, K. Wind tunnel simulation of wind flows with the characteristics of thousand-meter
high ABL. Build Environ. 2019, 152, 74–86. [CrossRef]
15. Yan, B.W.; Li, Q.S.; He, Y.C.; Chan, P.W. RANS simulation of neutral atmospheric boundary layer flows over complex terrain by
proper imposition of boundary conditions and modification on the k-ε model. Environ. Fluid Mech. 2016, 16, 1–23. [CrossRef]
16. Weerasuriya, A.U.; Hu, Z.Z.; Zhang, X.L.; Tse, K.T.; Li, S.; Chan, P.W. New inflow boundary conditions for modeling twisted
wind profiles in CFD simulation for evaluating the pedestrian-level wind field near an isolated building. Build. Environ. 2018, 132,
303–318. [CrossRef] [PubMed]
17. Feng, C.; Gu, M.; Zheng, D. Numerical simulation of wind effects on super highrise buildings considering wind veering with
height based on CFD. J. Fluids Struct. 2019, 91, 102715. [CrossRef]
18. Lu, B.; Li, Q.S. Large eddy simulation of the atmospheric boundary layer to investigate the Coriolis effect on wind and turbulence
characteristics over different terrains. J. Wind Eng. Ind. Aerodyn. 2022, 220, 104845. [CrossRef]
19. Aly, A.M.; Bitsuamlak, G. Aerodynamics of ground-mounted solar panels: Test model scale effects. J. Wind. Eng. Ind. Aerodyn.
2013, 123, 250–260. [CrossRef]
20. Phuc, P.V.; Nozu, T.; Kikuchi, H.; Hibi, K.; Tamura, Y. Wind pressure distributions on buildings using the coherent structure
Smagorinsky model for LES. J. Comput. 2018, 6, 32. [CrossRef]
21. Le, V.; Caracoglia, L. Generation and characterization of a non-stationary flow field in a small-scale wind tunnel using a
multi-blade flow device. J. Wind. Eng. Ind. Aerodyn. 2019, 186, 1–16. [CrossRef]
22. Marks, F.D.; Dodge, P.; Sandin, C. WSR-88D observations of hurricane atmospheric boundary layer structure at landfall. In
Proceedings of the 23d Conference on Hurricanes and Tropical Meteorology (American Meteorology Society), Dallas, TX, USA,
10–15 January 1999; pp. 1051–1054.
23. Ekman, V.W. On the Influence of the Earth’s Rotation on Ocean Currents; Arkiv för Matematik, Astronomi Och Fysik; Almqvist &
Wiksells Boktryckeri, A.-B.: Uppsala, Sweden, 1905; Band 2; pp. 1–52.
24. Smagorinsky, J. General circulation experiments with the primitive equations. Mon. Weather. Rev. 1963, 91, 99–164. [CrossRef]
25. Launder, B.E.; Spalding, D.B. Lectures in Mathematical Model of Turbulence; Academic Press: London, UK, 1972.
26. Yakhot, V.; Orszag, S.A. Renormalization group analysis of turbulence: Basic theory. J. Sci. Comput. 1986, 1, 3–51. [CrossRef]
27. Shih, T.H.; Liou, W.W.; Shabbir, A.; Yang, Z.; Jiang, Z. A new k-ε eddy viscosity model eynolds reynolds number turbulent flows.
Comput. Fluids 1995, 24, 227–238. [CrossRef]
28. Wilcox, D.C. Reassessment of the scale-determining equation for advanced turbulence models. AIAA J. 1988, 26, 1299. [CrossRef]
29. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [CrossRef]
30. Roache, P.J.; Ghia, K.N.; White, F.M. Editorial policy statement on the control of numerical accuracy. J. Fluids Eng. 1986,
108, 2. [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.