Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

ARTICLE IN PRESS

Journal of Wind Engineering


and Industrial Aerodynamics 94 (2006) 365–376
www.elsevier.com/locate/jweia

On the Reynolds number sensitivity of the


aerodynamics of bluff bodies with sharp edges
G.L. Larosea,, A. D’Auteuilb
a
Aerodynamics Laboratory, National Research Council Canada, 1200 Montreal Road, Building M2,
Ottawa, Ont., Canada K1A 0R6
b
Department of Civil Engineering, University of Ottawa, Ottawa, Canada
Available online 3 March 2006

Abstract

A review of recent comparisons between the aerodynamic characteristics of bluff bodies with sharp
edges at low and high Reynolds number is presented. It is concluded that the relaxation of the
Reynolds number similitude requirement in the study of the aerodynamics of bluff body such as long
span bridge decks can lead to systematic errors. It is observed also that some bluff geometries appear
to be more sensitive than others to Reynolds number effects. Parameters that could link the geometry
of a bluff body with the sensitivity of its aerodynamics to Reynolds number have to be clearly
defined. An ongoing study of these issues based on experiments in a pressurized wind tunnel at the
National Research Council Canada is introduced and preliminary results are discussed. For
rectangular prisms, the fineness ratio or width-to-depth ratio is identified as one of the parameters.
Crown Copyright r 2006 Published by Elsevier Ltd. All rights reserved.

Keywords: Wind–structure interaction; Reynolds number; Bluff body; Bridge deck; Rectangular prism

1. Introduction

To describe the forces experienced by a body in a moving fluid, it is reasonable to


assume that the forces will depend at least on the density r and viscosity m of the fluid, the
relative velocity between the body and the fluid V, and a characteristic dimension of the
body L. Based on this assumption and dimensional analysis (e.g. [1]), a dimensionless
combination of the fundamental units: mass, length and time, can be defined to form
Corresponding author.
E-mail addresses: guy.larose@nrc-cnrc.gc.ca (G.L. Larose), adaut022@uottawa.ca (A. D’Auteuil).

0167-6105/$ - see front matter Crown Copyright r 2006 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2006.01.011
ARTICLE IN PRESS
366 G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376

naturally a parameter that has an influence on the majority of the fluid dynamics problem:
rVL
. (1)
m
This dimensionless P product of physical quantities shown by Eq. (1) is known as the
Reynolds number, Re, with reference to the British engineer Osborne Reynolds
(1842–1912), who was the first to demonstrate that this combination of variables had an
influence on the flow classification of a fluid in a pipe as being laminar, transitional or
turbulent (1883). It expresses the ratio of the inertia forces to the viscous forces of a
particle of fluid on an element. At high Reynolds number, inertia forces dominate over the
fluid viscous forces. At low Reynolds number, viscous forces play a determinant role on
the aerodynamics.
The Reynolds number is thus a similitude parameter that generally needs to be respected
for a model-scale laboratory experiment involving fluid to be representative of full-scale
conditions. However, the general belief is that bluff bodies with sharp edges such as bridge
decks and towers, buildings and many structural members can have aerodynamic
characteristics almost insensitive to Reynolds number as long as a Reynolds number of
10,000 is reached.
In his thorough study of the Tacoma Narrows Bridge collapse, Farquarson [2] quoted
the following statement by Prandtl and Tietjens [3] to justify the choice of the geometrical
scale for his experiments and the more or less imposed relaxation of the Reynolds number
similitude when working with aeroelastic models:

The simple relations just outlined (invariability of drag coefficients with change in
Reynolds’ number) are practically true for bodies where the total drag consists almost
exclusively of pressure drag and where the geometry of the flow, in particular the
breaking away of the fluid, is determined by sharp edges as, for instance, with a plate
perpendicular to the direction of the flow...

Farquarson did not stop there though and actually conducted experiments at several
Reynolds numbers using either different flow speeds or geometrical scales to verify this
assumption for the deck of the Tacoma Narrows Bridge [2].
The quote above provides also the basis for a general definition of a bluff body: a body
for which the major contribution to the drag force is due to the pressure forces arising from
the occurrence of boundary layer separation and the formation of a recirculating flow
region aft of the body. In contrast, the major contributor to the drag on a streamlined
body is the viscous action of the fluid on the body.
For a bluff body with sharp edges, it is believed that the onset of the flow separation is
defined by the location of the edges, and flow re-attachment may or may not occur but if it
does, it might not be affected by Reynolds number. The Reynolds number similitude
parameter is thus generally relaxed in wind engineering studies where the effects of wind on
buildings and bridges are investigated on models at Re one or two orders of magnitude
lower than in full scale.
However, recent experiments at high Reynolds number and comparisons between full-
scale and model-scale experiments have indicated that there can be several limitations to
this belief (for bridge decks see [4–6], and for a low-rise building, [7]). A review of these
observations is presented in the next section.
ARTICLE IN PRESS
G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376 367

The aerodynamic characteristics of a body tested at low Reynolds numbers such as its
drag and lift coefficient or its Strouhal number have been shown to be different from the
high Reynolds number values defining what is called Reynolds number effects. It is argued
that the presence or the extent of the re-attachment region is a function of Reynolds
number as well as a function of the turbulence in the flow. Depending on the Reynolds
number regime, the static force coefficients, the characteristic shedding frequency and the
patterns of unsteady surface pressures could be affected.
As a consequence, the issue of Re sensitivity is now often raised in the planning of the
aerodynamic studies carried out to support the design of bridge decks and other components
of cable-supported structures, see for example Larose et al. [8] for the deck of the proposed
Stonecutters Bridge. Of particular interest is the accurate prediction of the Strouhal number,
St, for the cross-section investigated. This is important since a vortex-shedding mitigation
strategy could consist to push, using aerodynamic appendages, the critical velocity for
potential lock-in to higher wind speed having lower probability of occurrence. The possible
influence of Reynolds number on the amplitude of vibrations during lock-in is also of interest.
Experiments specifically aimed at studying Reynolds number effects for bluff bodies
with sharp edges have been carried out at the National Research Council Canada, NRCC.
They have been done on large sectional models in a large wind tunnel or on smaller models
in a pressurised wind tunnel. One of the latter studies [9] focussed on the depiction of
Reynolds number effects for cross-sectional shapes used in civil engineering structures and
for which Reynolds number similitude would normally be relaxed. Preliminary results of
this study are presented in Section 3 of this paper.

2. Lessons learned

In this section, a review of recent cases where Reynolds number effects have been
depicted for three bridge decks and a building is presented. With two exceptions, the
observations are based on full-scale measurements compared to wind-tunnel tests.

2.1. The Storebælt East Bridge

The Storebælt East Bridge in Denmark is a remarkable structure with a 6784 m long
closed-box girder steel deck. The centre portion of the structure is a 2716 m long
suspension bridge with a streamlined deck, having a width B to depth D ratio of
approximately 7.7 (see Fig. 1). The suspension bridge is flanked by two approach bridges,
1538 and 2530 m long. The steel girder of the approach bridges is also a closed box but is
bluffer than the suspension bridge deck, with a B=D of 3.7 (Fig. 1).

Fig. 1. Deck cross-sections for the Storebælt East Bridge, Approach Bridge (left), Suspension Bridge (right).
ARTICLE IN PRESS
368 G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376

During construction of the approach bridges and later on of the suspension bridge,
vortex-shedding induced oscillations of the respective bridge deck were observed. While they
did not come as a surprise since they had been predicted based on sectional model, taut strip
model and aeroelastic model tests, some lessons were learnt from these events. For the
approach bridge, it appeared that the vibrations occurred at a lower wind speed than
anticipated based on the Strouhal number obtained from tests conducted at Re ¼ 40; 000. A
Strouhal number of 0.17–0.18 was predicted. The oscillations occurred in full-scale at a
windspeed corresponding to a Strouhal number of 0.22. This triggered a thorough study of
this phenomenon in a pressurised wind tunnel by Schewe and Larsen [4]. It was concluded
that the difference in Strouhal number was attributed to Reynolds number effects.
However, for the deck of the cable-supported portion of the bridge, the wind velocities
at which the vortex-shedding induced oscillations occurred in full-scale appeared to be in
good agreement with the Strouhal number predictions based on more or less all scale
model tests for this cross-section. Vortex-shedding induced oscillations were observed at a
Strouhal number of 0.11 on a 1:300 scale taut strip model of the deck at Re ¼ 28; 000 [10],
at Re ¼ 70; 000 and 150,000 on sectional models and in full scale at Re ¼ 19; 000; 000 [11].
So, same bridge, two deck cross-sections, two completely different sensitivities to Reynolds
number. Part of the answer to the Re sensitivity issue lays likely within these observations.

2.2. The Ikara Bridge

The Ikara Bridge in Japan was also the object of several model scale experiments at 1:20,
1:30 and 1:121 scale. Full-scale measurements during construction were also conducted,
including surface pressure measurements that provided an evaluation of the Strouhal
number [5]. The bridge has a closed-box girder deck, with a B=D of 5.5 and is relatively
streamlined (see Fig. 2). Kubo et al. [5] addressed the issue of Reynolds number in their
study of the bridge and concluded that for this type of deck cross-section, a minimum
Reynolds number of Re ¼ 110; 000 (based on the deck width, B) should be respected in
model scale experiments. This minimum was defined by inspection of the relationship
between Strouhal number and Reynolds number, a Strouhal number of 0.16 having been
measured at a Reynolds number of 20,000 compared to St ¼ 0:20 in full-scale. Note that
for the deck cross-section of the Approach Bridges of the Storebælt East Bridge, a
minimum Re of 500,000 can be defined from Schewe and Larsen’s (1998) study [4].

2.3. The IHI Bridge

A tremendous effort was put forward to depict Reynolds number effects for a twin-box
streamlined bridge deck referred to here as the IHI Bridge. Tests at several geometrical
scales were conducted in Japan in IHI wind tunnel and a very large model at 1:10 scale was

Fig. 2. Deck cross-section for the Ikara Bridge, after Kubo et al. [5] (dimensions: m).
ARTICLE IN PRESS
G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376 369

Fig. 3. Deck cross-sections for the IHI Bridge, 1:10 scale model, after Matsuda et al. [6].

Fig. 4. Deck cross-section for the proposed Stonecutters Bridge.

tested in NRCC’s 9 m  9 m Wind Tunnel [6]. Four decks cross-section configuration were
investigated but only one of them (80 gratings on Fig. 3) exhibited important Reynolds
number sensitivity, when the gaps between the boxes were covered with porous grating. Re
effects were seen only for angle of wind incidence larger than 3 . Not only the Strouhal
number was affected but also the drag coefficient as for the Storebælt East Bridge and also
the aeroelastic characteristics such as the pitch aerodynamic damping. For this case a
minimum Re of 1,000,000 can be extracted from the study.

2.4. The Stonecutters Bridge

The proposed Stonecutters Bridge, a 1018 m span cable-stayed bridge, has also a twin-box
deck cross-section, relatively streamlined with curved bottom panels (see Fig. 4). Reynolds
number sensitivity of the deck was suspected, given the curved surface. Wind-tunnel tests
were conducted at geometrical scale of 1:80 and also at 1:20 to depict Re effects, in particular
on the vortex-shedding induced oscillations. A clear sensitivity to Reynolds number was
depicted. It affected the drag and pitching moment static coefficients, the vortex-shedding
amplitude and the character of the lock-in phenomenon [8]. More importantly, it was
observed that aerodynamic appendages devised to mitigate vortex excitation (guide vanes as
shown in Fig. 4 at the bottom plate knuckles) were not effective at lower Reynolds number
but worked perfectly during the higher Re test campaign. It was concluded that flow passing
through the gap between the appendage and the deck was insufficient to disturb the vortex
formation and the source of the excitation. A minimum local Reynolds number based on the
gap size of 40,000 was thus defined which also alerted the wind engineering community
about this important physical modelling issue [12].
ARTICLE IN PRESS
370 G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376

Fig. 5. Difference between full scale (high Re) and model scale (low Re) pressure coefficients for the Silsoe
Building; top: sharp edges; bottom: curved eaves, after Hoxey et al. [7]. The flow is from the left.

2.5. The Silsoe Building

There exist also examples of Reynolds number sensitivity of building with sharp
edges. The most striking example is the observation by Hoxey et al. [7] that the Silsoe
Building, a low-rise structure fully exposed to atmospheric boundary layer flow could be
sensitive to Reynolds number, when comparing model scale experiments to field
measurements (see Fig. 5). An in-depth analysis of the full-scale data also showed that the
configuration of the Silsoe Building with sharp eaves was more sensitive to Re effects than
the same building with curve eaves. This is a contradiction with the general belief that curved
corners are more Re sensitive than sharp edges. It was observed that the character of the
separated flow region of the windward roof slope was varying with Reynolds number as
shown in Fig. 5.

2.6. Summary and discussions

By definition, the aerodynamics of bluff body with sharp edges such as bridge decks and
buildings involve flow separation and reattachment phenomena, laminar or turbulent
boundary layer, laminar or turbulent shear layer, separation bubbles and laminar-to-
turbulent transition in the boundary layer or shear layer. Experience has shown that many
of these phenomena can be affected by Reynolds number as illustrated by the cases
summarized above. These cases show that some bluff bodies with sharp edges are more
ARTICLE IN PRESS
G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376 371

sensitive to Re effects than others and that it is hard at a first glance to define a general
minimum Re above which these effects will be limited. Schewe [13] argues that most global
Re effects seen on bluff bodies can be attributed to fundamental variations with Re in the
topological structure of the wake. It is suspected that the location of the laminar-to-
turbulent transition is at the root of these structural changes.
The case of the Storebælt East Bridge where the more streamlined box girder of the
suspension bridge appeared to be less sensitive to Reynolds number than the bluffer box
girder of the approach bridges could be a turning point. This is corroborated by
observations made by researchers at Politecnico di Milano for the deck of the streamlined
closed-box girder of the Humber Bridge. Comparisons between model scale tests at low Re
and exhaustive field measurements have shown a low level of sensitivity to Re for this deck
cross-section [14].
While these conclusions can be disheartening for wind engineers entrusted with the
design of small scale full-bridge aeroelastic models, there are some possible compromises.
The most common approach is to define the aerodynamic characteristics of the deck
studied at the highest Re possible using a large sectional model at high wind speeds for
example and to adjust the geometry of a smaller scale sectional model to match the
aerodynamic characteristics obtained at high Re. This new geometry with equivalent
aerodynamics is then used for the construction of the aeroelastic model. This approach was
used for example by the first author for the aeroelastic model study of the 1992-design of
the Messina Straits Bridge (1:250 scale aeroelastic model equivalent to a 1:30 scale
sectional model) and the Pont de Normandie during construction (1:150 scale aeroelastic
model equivalent to a 1:10 scale and 1:60 scale sectional models). Nevertheless, the need
for guidelines as to when Re effects can be expected persists.

3. Work in progress: rectangular prisms

In a recent study conducted in the NRCC’s Blowdown Trisonic Wind Tunnel,


rectangular prisms with fineness ratio (width-to-depth ratio) of 2–4 as shown on Fig. 6
were studied in smooth and turbulent flow for Reynolds number ranging from 0:15  106
to 4  106 and for a Mach number, Ma, not exceeding 0.3 (see [9] for details of the

Fig. 6. Exploded view of the model (left side) and cross-section tested: square edges, small chamfer ð4:1%DÞ and
large chamfer ð12:5%DÞ (right side).
ARTICLE IN PRESS
372 G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376

experiments). Rectangular prisms were chosen as cross-sectional shapes to study since they
are generally found in civil engineering structures and since Re similitude would normally
be relaxed for such bodies. The ultimate goal is to identify the parameters that could help
characterize the sensitivity of bluff bodies with sharp edges to Re effects. Three main
parameters were varied during the study: fineness ratio of the rectangular prisms, edge
treatments (see Fig. 6) and the turbulence intensity.

3.1. Fineness ratio parameter

The aerodynamic characteristics of a square prism (fineness ratio of 1) are known to be


independent of Reynolds number. However, for prism with fineness ratio of 2–4, there are
uncertainties about their aerodynamic coefficients such as the drag coefficient as shown in
Fig. 7, after Simiu and Scanlan [15]. This figure shows distinct values of drag coefficient for
rectangular prisms with fineness ratio smaller than  1 or larger than 4. On the other hand,
a grey zone exists for the fineness ratio between 1 and 4 indicating some dispersion in the
published drag coefficients for these cases. It is speculated that this dispersion is related to
the wind-tunnel test conditions, likely the Reynolds number of the experiments and the
turbulence intensity.
Results from tests done at NRCC [9] showed that for a prism with square edges, there is
a Re effect on the lift coefficient, C L and on the drag coefficient, C D . The lift coefficient
appeared to be the most sensitive. In smooth flow, with a Mach number of 0.15 and a
pitching angle of 2 , there is a difference for the lift coefficient of 20% for a 2:1, 11% for
a 3:1 and 10% for a 4:1 fineness ratio. The drag coefficient variations are approximately
6.5% for 2:1, 8% for 3:1 and 9% for 4:1 (see Fig. 8). The level of sensitivity of C L and C D
with Re is similar for the other pitching angles tested (0 , 2 , 4 , 6 , 8 , 10 ).
The after-body length has a considerable impact on the sensitivity to Re of the lift
coefficient of the prisms. In fact, C L is more affected for a 2:1 and this impact is reduced
with increasing fineness ratio. However, the 4:1 fineness ratio showed the largest variation
of C D versus Re with a difference of 9%. Looking at the drag coefficient (curves of Fig. 8),
the higher value of C D for a 2:1 prism could be related to the impossibility of the flow to
reattach to the prism sides that appeared too short. However, for the 3:1 and 4:1 fineness
ratio prisms, the lower C D value tends to express that a reattachment took place and then,

Fig. 7. Variation of drag coefficient C D with fineness ratio ðb=hÞ for prisms with square edges, after [15].
ARTICLE IN PRESS
G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376 373

0.3 0.8
0.7
0.2
0.6
0.1 0.5

CD
CL

0 0.4
0.3
-0.1
0.2
-0.2 0.1
-0.3 0
0 1 2 3 0 1 2 3
Re (106) Re (10 6)

Fig. 8. Variation of aerodynamic coefficients with Reynolds number (based on D) for a square edge prism, in
smooth flow, Mach: 0.15 and for a pitching angle of 2 . E: 2:1, m: 3:1, ’: 4:1. C L and C D are normalized by the
width of the body ðBÞ.

the wake width behind the prism reduced considerably. Also, the non-constant C D shows
that the boundary layer could be dependent on Re. In fact, the state of the boundary layer
(laminar or turbulent) that develops along the body before separation at the windward
square edges could affect the flow structure around the body and in particular
the reattachment point. In other words, the 2:1, 3:1 and 4:1 fineness ratio prisms could
be affected by Re because the airflow might not separate at the sharp edges with the
same state. This is especially true for cases where the angle of wind incidence is different
from 0 .

3.2. Edge treatment parameter

The study done by Delany and Sorensen [16] showed that for a rectangular prism with
small round corners (radius of curvature of 4% of the depth, D), there was a sudden drop
of C D around Re of 1  106 . For the present study, it was decided to study the effect of
square edges, small chamfer and large chamfer on aerodynamic coefficient for Re ranging
from 0.3 to 2:5  106 . Such prisms with sharp edge corners are of most interest in civil
engineering since they represent the basic shapes for bridge deck, bridge pylon or
buildings. The results showed that for a 2:1 prism, in smooth flow, at Ma:0.15 and for a
pitching angle of 2 , the variation of C L with Re is 20% for square edges, 55% for small
chamfer and 136% for large chamfer (see Fig. 9).
However, the differences for the drag coefficient were, respectively 6.5%, 10% and
6.3%. Again, the effect of Re on C L is important and it is not negligible for C D . Bluff
bodies with sharp edges are affected by Re and the largest difference was for the large
chamfer configuration. In fact, the state of the boundary layer could be again the reason
for this variation of C L with Re. The development of the boundary layer on the surface of
the prism is affected by the presence of the chamfer that could allow the flow to stay
attached longer to the body and to separate at the second edge of the chamfer for a state
that influence its reattachment further along the side of the body. The first edge of the
chamfer can also energize the boundary layer to pass to a turbulent state. In light of these
results, it is concluded that bluff bodies with sharp edges are not always independent of Re
ARTICLE IN PRESS
374 G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376

0.4 0.8
0.7
0.3
0.6
0.2 0.5

CD
CL

0.1 0.4
0.3
0
0.2
-0.1
0.1
-0.2 0
0 1 2 3 0 1 2 3
Re (106) Re (106)

Fig. 9. Variation of aerodynamic coefficients with Reynolds number for a square edges prism, in smooth flow,
Mach: 0.15 and for a pitching angle of 2 . E: Square edges, m: small chamfer, ’: large chamfer.

effects and other parameters are needed to characterize the behaviour of a given bluff body
with sharp edges and one of them could be the edge treatment.

3.3. Turbulence parameter

Turbulence is of great importance for the aerodynamics of bluff bodies with sharp edges.
At low Re, Laneville [17] showed that for a rectangular prism with fineness ratio of 2 and
square edges, the aerodynamic characteristics of the body in smooth flow showed the
potential for an unstable condition and by increasing the turbulence intensity up to 11%,
this potentially unstable state vanished. The stability is related to the variation of C L with
the pitch angle at 0 . In smooth flow, the C L slope was negative and in turbulent flow, the
slope was positive. Results of the present wind tunnel tests at high Reynolds number
showed that the turbulence intensity I u tends to increase the stability of a given prism (see
Fig. 10). For a turbulence I u : 5%, the 2:1 prism showed unstable aerodynamic
characteristics in smooth flow and turbulent flow. For a 3:1 prism, the 5% turbulence
was enough to bring the characteristics to a stable state and the 4:1 prism had already
stable features in smooth flow and without any surprise in turbulent flow. By increasing the
turbulence intensity, it gives a chance to the flow to have more energy before reaching the
separation point, increasing the entrainment rate, affecting the separation bubble and the
reattachment point, consequently affecting the lift and drag coefficients.

3.4. Future work

The main objectives of this research are to expand the definition of bluff body; to
establish the sensitivity limit to Re of rectangular prisms by making parallel with circular
cylinders that are highly sensitive to Re and square prisms that are not sensitive and to
establish which parameters could help define the sensitivity limit of a bluff body with sharp
edges. The other data taken during the investigation such as unsteady surface pressures
and fluctuations in the wake will help to understand the airflow development around the
ARTICLE IN PRESS
G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376 375

0.6
0.4
0.2
0

CL
-0.2
-0.4
-0.6
-0.8
-5 0 5 10 15
wind angle (deg)

Fig. 10. Variation of lift coefficient with angle of wind incidence for a square edges prism, Re: 1:97  106 , Mach:
0.15. Solid: turbulent, open: smooth. ’: 2:1, m: 3:1, : 4:1.

body and behind it. Also, data correction for blockage and wall interference, even if small,
will improve the quality, the richness and the relevance of the results.

4. Conclusions

It has been demonstrated that the aerodynamics of bluff bodies with sharp edges can be
sensitive to Reynolds number effects. Relaxing the Reynolds number similitude
requirements in wind engineering might thus be the cause of systematic errors. A
thorough analysis of the possible Re effects should be done almost case by case before the
similitude requirements be relaxed. Based on recent experiments, the relative fineness of the
body and the edge treatment can help indicate if Reynolds number effects should be
investigated. Flow turbulence is also an important parameter to consider when studying Re
effects for bluff bodies.

Acknowledgements

The financial support for the second author provided by the Natural Science and
Engineering Research Council of Canada through a Discovery Grant is gratefully
acknowledged.

References

[1] A.M. Kuethe, J.D. Shetzer, Foundations of Aerodynamics, Wiley, New York, USA, 1959.
[2] F.B. Farquarson, Aerodynamic stability of suspension bridges with special reference to the Tacoma Narrows
Bridge, Part I, Investigations prior to October 1941, Structural Research Laboratory, University of
Washington, Seattle, USA, 1950.
[3] L. Prandtl, O.G. Tietjens, Applied Hydro and Aeromechanics, McGraw-Hill, New York, 1934.
[4] G. Schewe, A. Larsen, Reynolds number effects in the flow around a bluff bridge deck cross section, J. Wind
Eng. Ind. Aerodyn. 74–76 (1998) 829–838.
[5] Y. Kubo, C. Nogami, E. Yamaguchi, K. Kato, Y. Niihara, K. Hayashida, Study on Reynolds number effect
of a cable-stayed bridge girder, in: A. Larsen, G.L. Larose, F.M. Livesey (Eds.), Wind Engineering into the
21st Century, vol. 2, Balkema, Rotterdam, 1999, pp. 935–940.
ARTICLE IN PRESS
376 G.L. Larose, A. D’Auteuil / J. Wind Eng. Ind. Aerodyn. 94 (2006) 365–376

[6] K. Matsuda, K.R. Cooper, H. Tanaka, M. Tokushige, T. Iwasaki, An investigation of Reynolds number
effects on the steady and unsteady aerodynamic forces on a 1:10 scale bridge deck model, J. Wind Eng. Ind.
Aerodyn. 89 (2001) 619–632.
[7] R.P. Hoxey, A.M. Reynolds, G.M. Richardson, A.P. Robertson, J.L. Short, Observations of Reynolds
number sensitivity in the separated flow region on a bluff body, J. Wind Eng. Ind. Aerodyn. 73 (1998)
231–249.
[8] G.L. Larose, S.V. Larsen, A. Larsen, M. Hui, A.G. Jensen, Sectional model experiments at high Reynolds
number for the deck of a 1018 m span cable-stayed bridge, in: Proceedings of 11th International Conference
on Wind Engineering, Lubbock, TX, USA, 2003, pp. 373–380.
[9] G.L. Larose, A. D’Auteuil, Experiments on 2-D rectangular prisms at high Reynolds number in a pressurized
wind tunnel, in: Proceedings of Fifth International Symposium on Bluff Body Aerodynamics and
Applications, Ottawa, Canada, 2004, pp. 177–180.
[10] A.G. Davenport, J.P.C. King, G.L. Larose, Taut strip model tests, in: A. Larsen (Ed.), Proceedings of the
International Symposium on the Aerodynamics of Large Bridges, Copenhagen, Balkema, Rotterdam, 1992.
[11] A. Larsen, S. Esdahl, J.E. Andersen, T. Vejrum, Vortex shedding excitation of the Great Belt suspension
bridge, in: A. Larsen, G.L. Larose, F.M. Livesey (Eds.), Wind Engineering into the 21st Century, vol. 2,
Balkema, Rotterdam, 1999, pp. 947–962.
[12] A. Larsen, M.C.H. Hui, M.G. Savage, S.V. Larsen, Investigation of vortex response of a twin box bridge
section at high and low Reynolds number, in: Proceedings of Fifth International Symposium on Bluff Body
Aerodynamics and Applications, Ottawa, Canada, 2004, pp. 249–252.
[13] G. Schewe, Reynolds number effects in flow around a more-or-less bluff bodies, J. Wind Eng. Ind. Aerodyn.
89 (2001) 1267–1289.
[14] G. Diana, Personal communication, November 2004.
[15] E. Simiu, R.H. Scanlan, Wind Effects on Structures, third ed., Wiley, New York, USA, 1996.
[16] N.K. Delany, N.E. Sorensen. Low-speed drag of cylinders of various shapes, NACA Technical Note 3038,
Washington, USA, 1953.
[17] A. Laneville, I.S. Gartshore, G.V. Parkinson, An explanation of some effects of turbulence on bluff bodies,
in: Proceedings of the Fourth International Conference on Wind Effects on Buildings and Structures,
Heathrow, London, UK, 1975, pp. 333–342.

You might also like