You are on page 1of 14

2D CFD Simulations of a Static Twin Box Deck: Effects of

Circular Cylinder and Guide Vane on the Aerodynamics


Properties
Olivia Oey, Kenny C.S. Kwok, Félix Nieto, Robert Ong
1 ABSTRACT
This work investigates the aerodynamic interference of bare twin-deck configuration caused by a
circular cylinder fitted with and without guide vanes through the computational dynamics (CFD) technique.
Three configurations (twin-deck only, twin-deck with circular cylinder, twin-deck with circular cylinder
and guide vanes) were modelled in 2D using the Unsteady Reynolds-Averaged Navier-Stoke (URANS)
under k-ω SST turbulence model. The resultant mean and standard deviation pressure distribution along
with their corresponding force and moment coefficients were presented. The flow patterns in the wake
regime found to noticeably reduce the drag and lift coefficients as well as altering the pressure distribution
for each. The isosurfaces of spanwise vorticity for the three scenarios were presented, and found that
vortices in the small cylinder wakes distort and stretch to maintain connections until the mid-section of the
lower section of upstream deck. We also observed a delay in flow separation at the downstream deck which
was a direct consequence due to the absence in flow separation at the upstream deck. Furthermore, the
vortex shedding mechanisms was noticeably altered due to the presence of guide vanes, thereby mitigating
the vortex-induced vibrations (VIV).
2 BACKGROUND
Long-span cable bridges have opted twin-box girder configurations due to its high critical flutter
speed. Such configurations have been successfully implemented in the construction of the Stonecutters
Bridge in Hong Kong and the Xihoumen suspension bridge in China. Previous studies (conducted by Kwok
et al. (2012) [1]; and Laima et at. (2015) [2]) have shown improved flutter stability of twin-box girder;
however, the gap introduced between the upstream and downstream decks generates flow-induced
vibrations, known as vortex-induced vibration (VIV).
The application of Computational Fluid Dynamics (CFDs) in civil engineering projects continues
to be more prominent over the years. This research project provides as an extension to the former 2D
URANS simulations conducted by Nieto et al. (2017) [3]. The flow around the decks simulated by the study
mentioned in the latter was conducted under smooth, transient flow; however, in reality, long-span cable
bridges experience turbulence intensities as great as 20% and integral scales as large as ten times their deck
width. Therefore, this study focuses on the aerodynamic performance of twin-deck girders under rod-
generated turbulent flow.
Due to the complications presented by utilizing turbulent inflow in CFD, a rod-generated turbulent
flow was opted in this project instead under the limited timeframe. The placement of a thin circular rod
upstream of the deck instigates fine-scale turbulence to the decks in the computational domain [4]. The
addition of the guide vanes, which are parallel plates running along the bottom side of the decks, aims to
reduce turbulence associated with vortex shedding. This therefore, reduces the effects of vortex-induced
vibrations (VIV) mentioned above, producing a relatively laminar flow downstream the leeward deck. The
aim of this research project is to examine the effect of guide vanes installations on the force coefficients,
pitching moment, Strouhal number and the relative vortex shedding mechanism under a rod-induced
turbulent flow.
3 METHODS
The CFD model implemented in this project was adopted from the 2D URANS simulation proposed
by Nieto et al. (2017) [3]. The flow around the bodies were modelled by the Unsteady Reynolds-Averaged
Navier-Stokes (URANS) equations, particularly the k-ω SST turbulence model. The pimpleFoam solver in
OpenFOAM was utilized to simulate and evaluate incompressible, transient flow around the decks.
The geometry of the deck was modelled similarly to that of the Stonecutters Bridge in Hong Kong,
with chord length (C) of 19.5 m, deck height (D) of 3.5 m and a gap-width of (G) of 14.3 m. A geometric
scale of 1:80 is implemented to the computational domain. A gap-width (G) to total chord (B) ratio of 27%
was kept constant for all the simulations being carried out. The angle of wind incidence is kept at 0°, with
a corresponding wind speed of 15 m/s.
Three case files were generated with reference to the aims of the study. The first setup involves the
deck only, with no cylinder and guide vanes, acting as a control. The second configuration involves the
addition of a 10.2 mm (0.816 m) diameter cylinder at 20 mm (1.6 m) upstream from the far end corner of
the left deck, to instigate turbulent flow downstream the deck. Displayed in Fig. 1 is the final setup, which
includes the implementation of both the cylinder and the fitted guide vanes. The guide vanes are position
at 5.56 mm underneath the bottom surface of both the decks facing the gap.
The computational domain for deck-fitted vanes under rod-generated turbulent flow, shown in Fig.
1 below is divided into four different mesh regions. This includes the boundary layers of the decks, cylinder
and the guide vanes, which is not shown in the diagram below. A non-structured triangular mesh was
utilized for the entire computational domain, except for the boundary layer. A high-density mesh is
implemented in the region closest to the decks or the buffer zone. Especially in the regions between the
cylinder and the upstream deck as well as the gap between the two decks, special consideration was
implemented to ensure that a finer mesh is imposed. The wake region downstream of the buffer zone has a
moderate density mesh, while the coarsest tri-mesh is located away from the deck. The boundary layer
surrounding the decks, guide vanes and cylinder are made up of a quadrangular structured mesh consisting
of 10 layers with a scale factor of 4.
y1/C x1/y1 r lbl ybl/C Total Quadrilateral Triangular
cells cells cells
Bare 1.6432 · 10-4 4 1.167 10 3.5035 · 10-3 653998 66820 587178
Deck
Deck + 1.6432 · 10-4 4 1.167 10 3.5035 · 10-3 694374 70820 623554
Cylinder
Deck + 1.6432 · 10-4 4 1.167 10 3.5035 · 10-3 753756 77940 675816
Cylinder
+ Guide
vanes
Table 2. Mesh properties. The parameter y1 is the total height of the first element of the boundary layer (bl), x1 is the length of
first element in the bl, r is the growth ratio of the elements in the bl, l bl is the number of layers forming the bl and ybl is the total
height of the bl.
Fig. 1. Computational domain of deck-fitted guide vanes under rod-generated turbulent flow.

4 RESULTS
4.1 Sectional static aerodynamic force coefficients
Sectional aerodynamic forces of the twin-box girder include drag force (Fd), lift force (Fl) and
pitching moment (M), which are calculated by integrating the surface pressure around twin-box girder.
These aerodynamic forces and moment are then normalized to form:
𝐹𝑑 𝐹𝑙 𝑀
𝐶𝑑 = , 𝐶𝑙 = , 𝐶𝑚 =
1 2 1 2 1 2 2
2 𝜌𝑈 𝐷 𝑑𝑧 2 𝜌𝑈 𝐶 𝑑𝑧 2 𝜌𝑈 𝐶 𝑑𝑧
where, Fd is drag force, Fl is lift force, M is pitching moment, ρ is air density, U is mean wind speed, C is
chord length, D is deck height and dz is the length of the cross-section of the deck.
Throughout this paper, positive lift and moment are considered in the upward direction and
clockwise rotation respectively. In addition, positive drag is taken in the downstream wind direction, which
is along the direction of the x-axis. This is shown clearly in Fig. 2 below.
Fig. 2. Sign convention.

Fig. 3a, 3b and 3c below displays the mean pressure distribution for a deck configuration only, a
deck under rod-induced turbulence and a deck installed with guide vanes under rod-induced turbulence
respectively. Regions of flow separation and reattachment can be distinguished by fluctuations in pressure
distribution.

Fig. 3. Mean pressure distribution: (a) bare deck; (b) deck + cylinder; (c) deck + cylinder + guide vanes
a

Fig. 4. Pressure standard deviation distribution: (a) deck; (b) deck + cylinder; (c) deck + cylinder + guide vanes

For the twin-deck only scenario, the flow separation at the leading edge of the upstream deck is
highlighted by the large negative pressures, which in turn will induce a positive mean pitching moment.
The presence of cylindrical rod causes the growing region of constant pressure underneath that would
generate a negative pitching moment. Furthermore, the presence of a small circular cylinder placed
upstream of the deck resulted in the alteration of flow fields at the wake region elongated in the direction
of the flow. Consequently, the velocity outside the boundary layer at the separation point is reduced, which
would in turn affecting the pressure distribution at the leading edge of the upstream deck and to some extent
the downstream deck. Indeed, highly concentrated vorticity structures imply the presence of high
perturbation energy in the wake, which is directly proportional with both the instantaneous and mean drag
forces. The flow field around the deck with due to the presence of upstream cylinder is noticeably different
from that of without cylinder. Hence, the pressure distribution at the leading edge of the upstream deck
(forebody) is slightly different, and the pressure returns to values not very far from those of deck-only in
the afterbody. Note that there is a slight reduction in drag forces due to the presence of upstream flow
disturbance which would introduce different distribution, thereby the introduction of cylindrical rod acts as
a buffer which would reduce pressure drag on the deck.
Conversely, the force component of the deck due to circular cylinder in the across-wind direction
(lift force) is substantially quite different than for the case without. This can be explained due to a non-
symmetrical potential flow induced by the cylinder, which present a rapid and impulsive vortex
development. The positive vorticity induces a stronger vortex passing over its lower surface, which in turn
would cause the flow to accelerate and the resulting velocity difference produces a pressure difference and
a resultant negative downward lift force.
Guide vanes act as effective vortex mitigation devices, which are installed along the underside of
the twin-deck. Mean pressure distribution and its standard deviation depicted in Fig. 3 and 4 for both cases
with and without the vanes are closely interlinked with flow separation and reattachment. When examined
further, the negative pressure distribution of the upstream deck fitted with guide vanes is fairly constant and
low in magnitude. The vanes ensure that flow remains attached to the bottom non-streamlined surface of
the upstream deck and hence, lift forces is expected to be positive while moment coefficient remains more
negative. In addition, drag force is seen to increase for decks fitted with vanes, which can be explained by
the larger positive pressures at the downside of the downstream deck. However, the upstream deck without
fitted vanes has large negative pressures peaks at the upper corner of its streamlined face, explaining for
the greater negative lift. The significant decrement in standard deviation of the drag and lift coefficient for
deck-fitted vanes is explained by the suppression of vortices and vortex-induced oscillations. Supporting
this statement is the lower magnitudes of standard deviation pressure in Fig. 4 for decks fitted with guide
vanes and those without.
Following this, standard deviations pressure distribution of the deck under laminar flow, deck under
rod-generated turbulent flow and deck-fitted vanes are also displayed in Fig. 4a, 4b and 4c accordingly.
Standard deviation of the aerodynamic forces (Cʹd and Cʹl) and moment coefficient (Cʹm) are also recorded
to examine fluctuations on the forces and moment of the twin-deck girder. These values are recorded in
Table 2 below.
Cm Cd Cl Cʹm Cʹd Cʹl St
Bare deck 0.028 0.146 -0.234 0.109 0.058 0.524 0.301
Deck + cylinder -0.031 0.116 -0.304 0.104 0.053 0.479 0.291
Deck + cylinder -0.106 0.138 0.073 0.133 0.033 0.214 0.134
+ guide vanes
Table 2. Drag, lift and moment coefficient and its standard deviation for deck, deck + cylinder and deck + cylinder + guide
vanes.

4.1.1 Effect of rod-generated turbulence on aerodynamic forces

ca

Fig. 5. Pressure around computation domain: (a) deck; (b) deck + cylinder; (c) deck + cylinder + guide vanes.

The deck in smooth, laminar flow serves as a control for deck under rod-generated turbulent flow.
The mean pressure distribution for the bare deck and deck and cylinder are displayed in Fig. 2a and Fig.
2b, has several noticeable differences amongst one another which can be explained by closer examination
of Fig. 7 and 8.
Under smooth laminar flow, streamlines in Fig. 8a shows flow separation occurring at the top
surface of the streamlined section of the upstream deck. This phenomenon is resulted by the large adverse
pressure gradient recorded at the specified location, as shown in Fig 3a. Consequently, the increase in the
fluid pressure downstream has resulted in decreasing velocity (i.e. kinetic energy) being recorded in Fig.
6a and a zero-velocity region starting at the far-left corner of the upstream deck, suggesting flow separation.
On the other hand, no flow separation is observed upon addition of the rod. Turbulence formed by the
laminar separation of flow in the wake region of the cylinder, possesses increased transport of momentum
and the turbulent boundary layer which is formed has an increased wall shear stress. Separation is therefore,
prevented as shown in Fig. 8b, when compared to the bare deck in Fig. 8a. Pressure differences between
the leading and trailing edge of the upstream deck caused by flow separation explains the higher drag (Cd)
coefficient compared to deck under rod-generated turbulent flow recorded in Table 2. Due to the increase
drag force, the boundary layer observed in Fig. 8a is thicker than that of Fig. 8b.
Reduced drag and lift forces indicated in Table 2 can also be deduced by examining the mean
pressure distribution around the deck under smooth, laminar flow and rod-induced turbulent flow.
Differences in pressure distribution is caused by differences in flow separation as mentioned above. Under
rod-induced turbulent flow, magnitudes of the negative pressure on the bottom surface of the upstream deck
is greater compared to that of its top surface. As a result, flow is more accelerated over the more streamlined
bottom surface of the deck, leading to a lower drag force and an increased negative lift, in turn inducing a
negative pitching moment.
Further analysis of the vorticity field of the two scenarios depicted in Fig. 7a and 7b provides the
same conclusion. Under rod-generated turbulent flow, eddies are shed as soon as the flow hits the stagnation
point of the cylinder. As the flow progresses downstream, fewer eddies are continuously shed at the top
surface of the upstream deck, while eddies are no longer generated as the streamline from the cylinder
reattaches to the bottom surface of the upstream deck (Fig. 8b) after circulation. Fewer eddy formation
emphasizes the notion that flow from the cylinder remains continuous and that separation does not take
place upstream deck. Unlike the deck under turbulent flow, the formation of eddies for the bare deck
occurred immediately at the leading edge of the upstream deck as the smooth, laminar flow strikes the sharp
corner of the deck.
The placement of the rod causes flow to strike the cylinder at its stagnation point directly, inducing
high pressures on its free side, as shown in Fig. 5b. Since the inflow is smooth and laminar with low
Reynolds number (Re), vortices are shed alternatively from the lower and upper surface of the cylinder and
into the wake region of the cylinder, forming what is known as the von Kármán vortex street. Shown in
Fig. 7b, the upstream deck is immersed in these unsteady vortices. Hence, being sheltered by the cylinder
forms a region of low pressure and velocity between the cylinder and upstream deck as shown in Figure
Fig. 5b and Fig. 6b.
By closer examination of the streamlines around the downstream deck in Fig. 8b, a delay in flow
separation is observed, when compared to Fig. 8a. Again, this might be a result of the turbulent flow formed
by the cylinder, which has a Reynolds number (Re) greater than the critical Re. This causes an earlier
transition from laminar to turbulent boundary layer flow in the downstream deck. Thus, flow separation is
delayed with the addition of the cylinder. Moreover, a shorter separation bubble is observed for the deck
under smooth flow compared to that under turbulent flow. Streamlines are seen to reattach downstream the
leeward deck in Fig. 8b, while this phenomenon is not identified in Fig. 8a. No flow reattachment is
identified for the deck under smooth flow, since pressure distribution at the top surface of the downstream
deck remains negative and large in magnitude.
Results obtained in for the downstream deck is in agreement to the conclusions drawn by Kwok et
al. (1986) [4], except for the upstream deck. The study concluded that an increase in free-stream turbulence
caused by the placement of the cylinder promotes an earlier transition of laminar to turbulent boundary
layer flow; and hence, a delayed flow separation. For the upstream deck, no flow separation is inferred from
Fig. 8a, which is slightly different to the results obtained by previously mentioned study.
The deck without the cylinder has large negative pressures, which are induced by flow separation
are located further away from the rotational center, particularly near the leading edge of the upstream deck.
Hence, a large positive mean pitching moment is expected. Nevertheless, since no flow separation is
identified in Fig. 8b, negative pressures are not distinguished at the leading edge of the upstream deck of
Fig. 5b. This is demonstrated in the positive Cm for the configuration under smooth flow and a negative Cm
for the other under turbulent flow.
No significant differences were identified in the mean pressure distribution and the standard
pressure distribution of the downstream deck, except for their magnitudes. As shown in Fig. 4a and 4b, the
magnitudes of the standard deviation pressure distribution of the rod-induced turbulent flow is examined to
be much higher than that of the smooth, laminar flow. Vortex shedding impeding the downstream deck in
the bare deck is mostly sourced from the upstream deck and the gap, however with the implementation of
the cylinder, the additional vortex shedding caused by the placement of the cylinder accounts for the higher
magnitudes of standard deviation pressure recorded in Fig. 5b. Further, since both cases have the
downstream deck being immersed in the wake of the vortices shed by the upstream deck, positive pressures
are recorded at the non-streamlined windward surface of the downstream deck.
Periodic shedding of vortices is associated with oscillating lift and drag forces, and is represented
by Cʹl and Cʹd. The relatively unchanged Strouhal number, St and the pattern of vortices being shed, shown
in Fig. 7a and 7b suggest that the frequency of vortices being shed remains approximately the same in both
cases. Hence, it is expected that standard deviations of lift, drag and moment coefficients remain relatively
constant.

Fig. 6. Velocity around computational domain: (a) deck; (b) deck + cylinder; (c) deck + cylinder + guide vanes.

4.1.2 Effect of the guide vanes on aerodynamic forces


In this section, the performance of the twin-deck configuration fitted with and without guide vanes
will be analyzed under rod-generated turbulent flow. From Fig. 7b, it is easily determined the twin-box
arrangement without the vanes is prone to vortex shedding. When no guide vanes are installed downside of
the deck, vortex-induced vibrations results in positive pressures at the leading edge of the upstream deck
shown in Fig. 3b. The suppression of vortex shedding (discussed more in Section 4.2.2) by the guide vanes
has resulted in mild negative pressures being recorded at the leading edge of the upstream deck. Since
pressure for both cases remains positive and mildly negatively, flow separation is not observed at the
leading edge of the upstream deck, depicted by the streamlines plotted in Fig. 8b and 8c.
The shape of the pressure distribution of the downstream deck for both cases remains unchanged,
except for its magnitudes. The non-streamlined bottom surface of the downstream deck fitted with guide
vanes experiences significant positive pressures, which can be explained by the concentrated streamline in
Fig. 8c hitting the bottom, windward surface of the downstream deck. This accounts for the higher drag
force recorded in Table 2. Since vortex shedding is suppressed by the installation of the guide vanes,
fluctuations in pressure distribution in Fig. 4c is lower in magnitude.
A higher positive pressure at the bottom surface of the upstream deck and the low negative
pressures surrounding the deck-fitted guide vanes, explains for its overall positive lift force. Since the
positive pressure inducing a positive lift force is located near the leading edge of the upstream deck and
further away from the rotational center, a larger overall negative pitching moment is induced.
The implementation of guide vanes has altered aerodynamic forces significantly, which concludes
that flow features around the deck has changed since the addition of the guide vanes. In both cases, no flow
separation is observed and that flow separated by the cylinder reattaches to the top and bottom surfaces of
the upstream deck respectively. Due to imminent vortex shedding in the deck without the vanes, flow
circulation is greatly emphasized in Fig. 8b compared to that of Fig. 8c. In the downstream deck, flow
separation takes place at roughly the same location, explaining the similar pressure distribution diagram in
Fig. 3b and 3c of the downstream deck. One difference to take note is the double separation bubble observed
in Fig. 8b.
Fluctuations in lift, drag and moment coefficient is associated with vortex induced vibrations of the
deck. By looking at the values recorded in Table 2 under Section 4.1, fluctuations in lift and drag
coefficients reduces dramatically in deck-fitted vanes, demonstrating the suppression of vortex shedding
by the guide vanes.

4.2 Vortex shedding characteristics of twin-box girder


When steady, uniform flow separates as the fluid travels over bluff bodies, alternating vortices are
periodically shed into the wake of a bluff body. This phenomenon is also known as vortex shedding. This
periodically shedding of vortices exerts periodic forces by inducing fluctuating pressures; hence resulting
in the undesired vortex-induced vibration (VIV) of the system. It is therefore, critical to ensure that the
VIV of the system does not equate to the natural frequency of the structure or commonly known as lock-in
instability.
Vortex shedding is generally measured by the dimensionless Strouhal number, listed below:
𝑓𝑠 𝐻
𝑆𝑡 =
𝑈
where, fs is the vortex shedding frequency, H is the height of the deck and U is the mean speed of the
wind from the inlet.

Fig. 7. Vorticity field around computational domain: (a) deck; (b) deck + cylinder; (c) deck + cylinder + guide vanes.

4.2.2 Effect of rod-generated turbulence on vortex shedding


The vorticity fields for the deck configuration only and the rod-induced turbulence deck is shown
in Fig. 6a and 6b respectively. By inspection, both configurations display a relatively similar vorticity field.
This similarity in the displayed vorticity fields can account for the low 3.3% difference in Strouhal number.
In the upstream deck, vortices are continuously shed at the trailing edge of the upstream decks, travelling
downstream through the gap spacing and striking the downstream deck. The difference between the two
tests can be explained by the vortices generated by the placement of the cylinder upstream. Between the
cylinder and the upstream deck, a region of negative vorticity at the top and positive vorticity at the bottom
are formed. This leads to a delay in positive vorticity in the bottom surface of the upstream deck under
smooth, laminar flow.
Vortex shedding is also observed at the trailing edge of the downstream deck, especially for the
rod-induced turbulence configuration. A negative vorticity can be seen to detach from the top surface of
the downstream deck into the wake region for both tests.
4.2.2 Effect of guide vanes on vortex shedding
Fig. 6b and 6c respectively displays the instantaneous vorticity fields of the bare deck and deck-
fitted vanes under rod-generated turbulent flow. Evident differences in the formation of vortices being shed
for both cases can be observed. When no guide vanes are installed, shear layers from the upstream deck
rolls and forms alternating von Kármán type vortices at regular intervals into the wake of the upstream deck
which impinges the downstream deck. Streamlines from the upstream deck in Fig 8b is observed to
continuously recirculate after the trailing edge of the upstream deck before striking the downstream deck.
However, when vanes are installed downside of the decks, Fig. 7c shows that shear layers from the top
surface of the upstream deck flows directly towards the top surface of the downstream deck, with no
formation of vortices in the wake region of the upstream deck. In Fig. 8c, streamlines from the upstream
deck reattaches downstream into the leeward deck. When free stream flow through the gap spacing strikes
the leeward deck, vortices of lower vorticity are shed at the leading edge of the downstream deck. The
formation of vortices along the downstream deck for both cases justifies the similar shape yet lower
magnitudes of the fluctuating pressure of deck-fitted vanes.
Fig. 5b and 5c shows the velocity distribution for both respective scenarios. The introduction of the
guide vanes generated a low velocity region in the wake zone of the upstream deck. The formation of this
low velocity region might explain the absence of vortex formation in the wake region of the upstream deck.
It is also examined that the Strouhal number of the deck under rod-generated turbulent flow both
arrangements undergoes significant reduction with the installation of guide vanes at the inner corner of the
twin-deck. Therefore, guide vanes act as effective aerodynamic appendages, which are capable of
mitigating vortex-induced oscillations caused by unsteady separation of flow and vortex shedding. These
results are subsequently in conjunction to the previous studies conducted by Larsen et al. (2000) [5] and
Nieto et al. (2010) [6].
Fig. 8. Streamline around computational domain: (a) deck; (b) deck + cylinder; (c) deck + cylinder + guide vanes.

5 CONCLUSION
Computational Fluid Dynamics (CFD) was used in this project to analyze the effect of a rod-generated
turbulence and guide vanes on the static aerodynamic force coefficients and vortex shedding of a twin-deck
configuration bridge. Three tests setups, with constant gap-width to total chord ration were modelled and
meshed accordingly.
It was shown that mean pressure distribution around the decks under rod-induced turbulence is relatively
similar to that under smooth, laminar flow. This is perhaps due to the placement and size of the cylinder
being relatively inadequate in simulating high turbulent flow. However, under rod-induced turbulent flow,
a number of critical conclusions can be made, which includes no flow separation around the upstream deck
and a delayed flow separation for the downstream deck. Further, flow is observed to be accelerated along
the bottom surface of the streamlined upstream deck. Hence, under rod-induced turbulent flow, there is a
reduction in moment, drag and lift forces.
Following this, the implementation of the guide vanes in the deck mitigates vortex-induced vibrations
generated by upstream deck in the gap spacing between the two decks. They introduce a region of low
velocity fluid in the wake region of the upstream deck, suppressing the formation of vortices. However, this
paper fails to study the effect guide vanes has on the flutter performance of the twin-deck configuration,
which can be included in future pursuit of this research.
The study conducted in this paper lacks further examination on the aerodynamic performance of the twin-
deck girders at different turbulence characteristics. Therefore, future studies must include the installation
of different cylinders of various diameters placed at various distances upstream of the deck. This will allow
for the examination of the decks’ performance under different turbulence intensity and turbulence integral
scale. Furthermore, experimentally obtained data through wind tunnel testing is required to validate the
results produced by the CFD in this study, since the use of CFD in estimating forces of an oscillating bridge
remains limited, especially for turbulent flows.
6 REFERENCES
[1] KWOK, K. C. S., QIN, X. R., FOK, C. H. & HITCHCOCK, P. A. 2012. Wind-induced pressures around
a sectional twin-deck bridge model: Effects of gap-width on the aerodynamic forces and vortex
shedding mechanisms. Journal of Wind Engineering and Industrial Aerodynamics, 110, 50-61.
[2] LAIMA, S., LI, H., CHEN, W. & OU, J. 2018. Effects of attachments on aerodynamic characteristics
and vortex-induced vibration of twin-box girder. Journal of Fluids and Structures, 77, 115-133.
[3] SÁNCHEZ, R., NIETO, F., KWOK, K. C. & HERNÁNDEZ, S. CFD analysis of the aerodynamic
response of a twin-box deck considering different gap widths. Proceedings of the 2015 Congress
on Numerical Methods in Engineering (CMN 2015), Lisbon, Portugal, 29 June to 2 July 2015,
2015.
[4] KWOK, K. C. S. & KWOK, K. C. S. 1986. TURBULENCE EFFECT ON FLOW AROUND
CIRCULAR CYLINDER. Journal of Engineering Mechanics, 112, 1181-1197.

[5] NIETO, F., KUSANO, I., HERNÁNDEZ, S. & JURADO, J. Á. CFD analysis of the vortex-shedding
response of a twin-box deck cable-stayed bridge. Proceedings of the 5th International Symposium
on Computational Wind Engineering, Chapel Hill, NC, 2010.
[6] LARSEN, A., ESDAHL, S., ANDERSEN, J. E. & VEJRUM, T. 2000. Storebælt suspension bridge –
vortex shedding excitation and mitigation by guide vanes. Journal of Wind Engineering &
Industrial Aerodynamics, 88, 283-296.

You might also like