You are on page 1of 9

Experimental Investigation of Effects of Wind Yaw Angles

and Turbulence on Postflutter Behaviors of a Suspension


Bridge Model
Zhixiong Wang1; Zhitian Zhang2; and Kai Qie3
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Nonlinear postflutter behaviors of a suspension bridge model with a π-shaped deck section are investigated, focusing on
influences of wind yaw angles and low-intensity turbulence. Structural damping properties are first identified by free decaying vibration
in still air. The results show that modal damping ratios increase significantly with motion amplitudes. Wind flows with yaw angles are de-
composed into components perpendicular and parallel to the bridge deck axis. The flutter instability under wind yaw angles is examined in
terms of the perpendicular component. The experimental results show that while small wind yaw angle (10° and 20°) only slightly improve
the critical wind speed U⊥, the largest wind yaw angle (30°) substantially improves the U⊥ value. Oncoming turbulence of low intensity is
found to stabilize the system to some extent, increasing the flutter threshold by about 13% without blurring the demarcation between stability
and instability. DOI: 10.1061/(ASCE)BE.1943-5592.0001832. © 2022 American Society of Civil Engineers.
Author keywords: Flutter; Suspension bridge; Wind tunnel test; Wind yaw angle; Turbulence; Damping.

Introduction of a π-shaped section. Zhu et al. (2020) examined nonlinear flutter


characteristics of several bridge sections by spring-suspended sec-
Flutter stability is a major issue in the wind-resistant design of long- tional model tests.
span bridge structures, which in general necessitates physical A full bridge aeroelastic model has a number of advantages over
model tests in wind tunnels due to bluff properties of deck config- a sectional model, including at least the ability to inflect effects re-
urations. Spring-suspended sectional models have been used sulting from wind yaw angles, multimodes coupling, and wind tur-
historically to perform such type of investigation (Scanlan and bulence. Aeroelastic performances of bridge structures are
Tomko 1971). In general, this methodology is thought to be able determined generally without consideration of wind yaw angles.
to determine flutter thresholds only. In recent years, however, it In practical engineering, however, bridge structures are rarely sub-
has also been adopted to investigate nonlinear flutter properties. ject to wind flows exactly orthonormal to the deck axes. Wind yaw
Daito et al. (2002) used a spring-supported one-degree-of-freedom angle effects generally cannot be determined by sectional models
(DOF) (torsional) system to investigate flutter properties of due to very limited length-to-width ratios (normally below 3.0).
two-edge girder sections. Náprstek et al. (2007) and Náprstek As far as the turbulence is concerned, a coherent recognition has
and Pospíšil (2011) designed a device allowing for large-amplitude not been reached in terms of its effects on the flutter instability.
coupled free vibrations in bending and rotation and investigated Huston et al. (1988) investigated turbulence effects on the flutter
postflutter behaviors of several typical bridge deck and rectangular stability of the original Tacoma Narrows Bridge, and the effects
sections. Amandolese et al. (2013) used a 2-DOF microsection on the flutter derivative A*2 were examined. However, the results in-
model system and investigated the postflutter characteristics of a dicated insignificant influence from the turbulence. Matsumoto
flat plate. Pigolotti et al. (2017) studied the influence of dynamic et al. (1992) investigated turbulence effects on a series of H-shaped
parameters (stiffness, mass, and damping) on the flutter instability sections. The results were both stabilizing and destabilizing, de-
of rectangular cross sections by using a sectional model. Tang and pending on a section’s slenderness ratio B/D. With a taut strip
Hua (2019) investigated the influence of initial wind attack angles model, Diana et al. (1993) investigated the flutter stability of the
and nonlinear mechanical damping ratios on postflutter responses original Tacoma Narrows Bridge. All the results showed no evi-
dence of destabilizing the flutter stability from the turbulence,
1 and, in some cases, it enhanced the flutter threshold remarkably.
Ph.D. Candidate, Wind Engineering Research Center, Hunan Univ.,
Changsha 410082, China; Provincial Key Laboratory of Wind and Bridge Irwin et al. (1995) investigated the flutter stability of the Golden
Engineering, Hunan Province, Changsha, China. ORCID: https://orcid.org Gate Bridge and found that turbulence blurs the wind speed, demar-
/0000-0001-8815-5653. Email: doctorwzx@hnu.edu.cn cating the flutter instability, which is otherwise clear in laminar
2
Professor, College of Civil Engineering and Architecture, Hainan flow. Later, Scanlan (1997) explained the turbulence effects on flut-
Univ., Haikou 570228, China (corresponding author). ORCID: https:// ter in terms of coherence effects of the flutter derivatives in the
orcid.org/0000-0002-5878-2530. Email: zhangzhitian@hnu.edu.cn spanwise direction. Recently, Lam et al. (2017) tested a truss girder
3
Ph.D. Candidate, Wind Engineering Research Center, Hunan Univ., section and found that turbulence enhanced the flutter stability
Changsha 410082, China; Provincial Key Laboratory of Wind and Bridge substantially.
Engineering, Hunan Province, Changsha, China. Email: qiekai@hnu.edu.cn
Note. This manuscript was submitted on May 10, 2021; approved on
The preceding issues can be better examined by using a full
November 14, 2021; published online on February 9, 2022. Discussion pe- bridge aeroelastic model than a sectional one. Traditionally,
riod open until July 9, 2022; separate discussions must be submitted for in- the bridge deck in an aeroelastic model consisted of a core
dividual papers. This paper is part of the Journal of Bridge Engineering, beam, modeling the structural stiffness, and separated deck seg-
© ASCE, ISSN 1084-0702. ments attached to the core beam, modeling the aerodynamic

© ASCE 04022012-1 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


(a) (b) (c)
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

(d)

Fig. 1. Dimensions of the suspension bridge model (units: mm): (a) deck section; (b) tower section; (c) transverse beam (deck); and (d) elevation
view.

configuration. This methodology gives rise to issues such as dis-


continuous deck configuration in the longitudinal direction, incor-
rect higher-order modal properties, and uncontrollable structural
damping. In this work, the authors made a core-beam-free aero-
elastic suspension bridge model with a continuous deck system
to investigate nonlinear flutter behaviors, with specific concerns
on the influences of wind yaw angles and low-intensity
turbulence.

Aeroelastic Bridge Model

Model Design
As a general principle, the model is designed in such a way as to
reduce various uncertainties as much as possible. To this end, a
continuous bridge deck is adopted instead of a combination of a Fig. 2. Design drawing of the suspension bridge model.
core-beam and deck segments. A π-shaped deck section is used
for the bridge model, as shown in Fig. 1(a). Rectangular sections
are employed for towers and deck transverse beams [Figs. 1(b manufactured by computer numerical control machining. The
and c)], which are located at positions where the deck is con- counterweight under each pair of hangers is 100 g [Fig. 3(g)].
nected to hangers. The main span of the model is set to 7.2 m Some major natural frequencies of the model obtained by fast
according to the width of the wind tunnel, the side span is Fourier transform (FFT) of response signals from free decaying
0.4 m long, and the main towers are 1.37 m in height. The sag are listed in Table 2. It is noted that the frequencies of the first
ratio of the main cables is set to 1:10 [Fig. 1(d)], and the lateral and second vertical modes are quite close and so are those of the
distance of two cables is 0.2 m. The design of the suspension torsional modes.
bridge model is shown in Fig. 2, where an additional cable is in-
stalled at the side span to compensate for the insufficient stiffness
Experimental Setup
provided by the side span main cables. The geometric scale of
the model designed in this paper is supposed to be somewhere The tests were carried out in the HD-2 boundary layer wind tun-
between 1:100 and 1:200, according to deck width in practical nel at Hunan University, China. It is 15.0 (length) × 8.5 (width) ×
engineering. 2 m (height) in dimension, and the wind speed is adjustable be-
The bridge deck is made of bended steel plates of 0.5 mm in tween 0 and 14 m/s. The oncoming turbulence intensity mea-
thickness, as shown in Fig. 3(a). Transverse beams of the deck sured from uniform flows in the wind tunnel is less than 2%
are carved from ABS plates of 5 mm thickness, and they are fas- when the wind speed is higher than 3 m/s. Four laser displace-
tened to the main girder by ten M 2.0 high-strength bolts, as ment sensors (optoNCDT Model 2300 with sampling frequency
shown in Fig. 3(b). The main cables and hangers are made of being 2,000 Hz) with an operating range of 100–300 mm are
high-strength steel wires and constantan wires of 0.5 mm diam- used to measure vertical motions, two being placed under the
eter, respectively. Members of the main towers are made of rec- middle span, and the other two placed under the quarter span.
tangular iron bars. Sectional properties of major members of the The deck’s rotation is calculated from vertical motions and cor-
model are listed in Table 1. The bearings are made of hard alu- responding sensor-to-sensor distances. The vertical displacement
minum, designed with two degrees of freedom [Figs. 3(c and d)]. is set to be positive downward and the torsional displacement
The anchorages [Fig. 3(e)] and saddle supports [Fig. 3(f)] are positive nose-up.

© ASCE 04022012-2 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


(a) (b) (c) (d)
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

(e) (f) (g)

Fig. 3. Details of the suspension bridge model: (a) deck; (b) transverse beam (deck); (c) bridge bearing; (d) supporting system; (e) anchorages; (f) sad-
dle supports; and (g) counterweight.

Table 1. Sectional properties of the model

Member E (GPa) A (mm2) Iy (mm4) Iz (mm4) Jd (mm4) w (N/m)


Deck section 210.0 149.500 1,032,312.1 4,428.9 12.5 11.4139
Transverse beam section (deck) 2.1 125.000 260.4 6,510.4 944.6 1.3632
Tower section 210.0 200.000 1,666.7 6,666.7 4,664.3 14.9513
Main cable section 210.0 0.196 — — — 0.0151
Hanger section 90.0 0.196 — — — 0.0164

Table 2. Tested modal frequencies of the model


Mode Mode shape Frequency (Hz)
1 First symmetric vertical 2.258
2 Second antisymmetric vertical 2.441
5 Third symmetric vertical 4.761
3 First symmetric torsional 3.021
4 Second antisymmetric torsional 3.448

For reference, the steady aerodynamic force coefficients tested


using a balance are given in Fig. 4, where the lift force coefficient
(CL) is positive downward, and the moment coefficient (CM) is pos-
itive clockwise.
Fig. 4. Aerostatic forces coefficients.

Wind Tunnel Tests


1 d ln [A(t)]
ξ=− (1)
Amplitude-Dependent Structural Damping 2πf dt
It has been reported that sectional models have obvious nonlinear
where f = vibration frequency and can be supposed as time-
characteristics and are dependent on motion amplitudes (Zhang
invariant. To obtain the structural damping ratio ξ that varies
et al. 2017; Tang and Hua 2019; Gao et al. 2020; Wu et al.
smoothly with the time, it is favorable to fit first the transient am-
2020). At present, only some literature has investigated damping
plitude obtained from HT with
nonlinearities of full bridge aeroelastic models. In this study, free
decaying histories of the first-order vertical and torsional modes
of the model were obtained by artificial excitations in still air, as A(t) = C1 + C2 e−C3 t + C4 e−C5 t (2)
shown in Fig. 5. Transient amplitudes are identified through Hilbert
transform (HT). The transient amplitudes A(t) corresponding to where Ci (i = 1–5) are constants. The fitting results are plotted in
Fig. 5 are given in Fig. 6. Fig. 6 against those obtained directly from HT.
Amplitude-dependent structural damping ratios are able to be Damping ratios identified from motions of the middle and
obtained by quarter span in still air are shown in Fig. 7. The results indicate

© ASCE 04022012-3 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


(a) (b)

Fig. 5. Time history of the single first-order vertical and torsional modes of free decaying vibration: (a) torsional Aα; and (b) vertical Ah.
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 6. Transient amplitudes in still air: (a) first-order torsional mode Aα; and (b) first-order vertical mode Ah.

(a) (b)

Fig. 7. Nonlinear modal damping ratios: (a) torsional modal damping ratio ξα; and (b) vertical modal damping ratio ξh.

(a) (b) (c)

Fig. 8. Model positioned at different wind yaw angles: (a) 10°; (b) 20°; and (c) 30°.

that both the vertical and the torsional damping are nonlinear, et al. 2007), which is partly accounted for by the fact that the aero-
showing strong dependence on motion amplitudes. dynamic configuration of the bottom part of a π-shaped deck girder
varies more significantly at wind yaw angles. Three wind yaw an-
gles are considered in this work, 10°, 20°, and 30° as opposed to 0°,
Effects of Wind Yaw Angles as shown in Fig. 8. Effects of the wind yaw angles on the nonlinear
flutter are investigated with smooth oncoming wind flows.
Aerodynamic properties of long-span bridges stiffened with The flutter thresholds under wind yaw angles are presented in
π-shaped decks are known to be more susceptible to wind yaw an- Table 3. If aerodynamic properties of a bridge deck are supposed
gles than those with enclosed box girders (Pindado et al. 2005; Zhu to be unaffected by wind yaw angles, the threshold of flutter

© ASCE 04022012-4 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


instability would be determined by a fixed wind velocity compo- deviation between the expected and tested values (as large as
nent perpendicular to the bridge axis, which is irrelevant to wind 39%), showing the aeroelastic stability being altered significantly
yaw angles. If that is the case, expected critical wind speeds at var- by the wind yaw angle. It is worthy of mention that, for the case
ious yaw angles are able to be obtained by simple vector decompo- concerned in this work, wind yaw angles play the role of stabilizing
sition according to specific wind yaw angles, as shown in Fig. 9. the bridge deck aerodynamically.
Therefore, the expected critical wind speeds are calculated by The postflutter limit cycle oscillation (LCO) amplitude at the
quarter- and midspan are plotted against the oncoming wind velocity
Ucr αyaw =0◦ in Fig. 10, while in Fig. 11, the same results are plotted against the
Ucr αyaw = (3)
cos (αyaw ) wind velocity component orthonormal to the bridge axis. Scrutiny of
the results leads to the following observations: first, the flutter insta-
where Ucr αyaw =0◦ = flutter critical wind speed for αyaw = 0°; bility exhibits DOF coupling (phenomenally) between the vertical
namely, the wind direction is perpendicular to the bridge axis. and torsional motions, despite the fact that the deck section is
The expected and tested flutter onset speeds are provided in bluff and hence a mechanism of single-DOF torsional flutter instabil-
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

Table 3. The results imply an important property. For cases of ity is expected. The observed phenomenon also agrees with findings
wind yaw angles less than 20°, the critical wind speed can be ob- by Náprstek and Pospíšil (2012) and Náprstek et al. (2015). The cou-
tained according to vector decomposition without much loss of ac- pling here ought to be viewed more as a concomitant phenomenon of
curacy, indicating that aeroelastic properties of the deck are not the instability than as an underlying mechanism. Second, the post-
being affected significantly by the wind yaw angles. When the flutter amplitude evolves quite linearly with respect to the wind
wind yaw angle increases to 30°, however, there is a substantial speed, with torsional LCO amplitude developing from less than 1°
at the onsets to about 7° after wind speeds being increased about
Table 3. Flutter critical wind speed of the 2D analysis and tests results 50%. Third, comparison between Figs. 10 and 11 indicates that,
Ucr (m/s)
while small yaw angles have very limited effects on the instability,
the largest yaw angle (30°) not only stabilizes significantly the struc-
αyaw (°) Expected Tested Discrepancy (%) ture (an increase in U⊥) but also obviously changes the shape of the
0 2.51 2.51 0 evolution path of postflutter LCO. In general, results with wind yaw
10 2.55 2.56 0.4 angles indicate that the axial wind flow (along the bridge deck axis)
20 2.67 2.82 5.6 plays an important role in stabilizing the structure aeroelastically.
30 2.90 4.03 38.9 The ratios of the midspan motion amplitude to the quarter-span
motion amplitude are shown in Fig. 12. It is noticed that ratios of
the torsional motion for all wind yaw angles are almost identical
(varies slightly around 1.8), indicating a single symmetric torsional
mode and a fixed modal shape [Fig. 12(b)]. For the vertical motion,
however, the ratios differ substantially among different wind yaw
angles. For example, the ratios of wind yaw angle 20° remains ap-
proximately a constant being greater than 1.0, indicating a stable,
symmetric vertical motion. For the other three cases of wind yaw
angle, the ratios are all below 1.0 and vary drastically [sometimes
very close to zero; see Fig. 12(a), yaw angle 10°]. The ratio being
less than 1.0 indicates that an asymmetric mode, which does not
Fig. 9. Speed decomposition diagram.
match the torsional motion, is dominant in the vertical motion.

(a) (b)

(c) (d)

Fig. 10. LCO amplitudes plotted versus the wind speed: (a) vertical, middle span; (b) vertical, quarter span; (c) torsional, middle span; and (d) tor-
sional, quarter span.

© ASCE 04022012-5 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


(a) (b)
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

(c) (d)

Fig. 11. LCO amplitudes versus the wind component perpendicular to the bridge axis: (a) vertical, middle span; (b) vertical, quarter span; (c) tor-
sional, middle span; and (d) torsional, quarter span.

(a) (b)

Fig. 12. Amplitude ratio under different wind yaw angles: (a) vertical amplitude ratio; and (b) torsional amplitude ratio.

combination with the unstable ratio values and unmatched vertical


motions shown in Fig. 13, imply the insignificance of the partic-
ipation of vertical motions in the underlying mechanism of aero-
dynamic instability.

Oncoming Turbulence Effects

Wind Field Properties


Cubic roughness elements are used to generate the desired turbulent
wind field, as shown in Fig. 14. Main parameters of the arrays of
Fig. 13. Intensity of the vertical motion relative to the torsional motion. cubic roughness elements include lateral and longitudinal spacing,
number of rows, and distance from the bridge model. They are de-
termined in a trial-and-error way to result in target wind properties
The intensity of the vertical motion relative to the torsional at the location where the model is positioned. The installation and
motion is able to be reflected by comparing it with the girder- setting of the displacement lasers are the same as in Fig. 3, and the
edge’s motion caused by rotation, namely Av compared with wind angle of attack is set as 0°.
πbAα /180, where b is the half-width of the bridge deck, and tor- Wind properties are examined at five positions as shown in
sional motion Aα is denoted in degrees. The results, which differ Fig. 15, including the girder center, girder ends, and quarter
obviously among different wind yaw angles, are shown in Fig. 13. spans (about 0.6 m above the floor). Turbulence intensities at the
It can be seen that the relative motion intensity of the vertical mo- midspan position are shown in Fig. 16. The horizontal turbulence
tion varies as the wind speed increases. Smallest values are very profiles obtained at different wind speeds are plotted in Fig. 17,
close to 0, implying a state of nonparticipation of the symmetric where it can be seen that the variations of turbulence intensity
vertical motion in the flutter instability. These phenomena, in among five positions are satisfactory. Fig. 18 shows power spectra

© ASCE 04022012-6 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


Fig. 14. Suspension bridge model in turbulence wind field.
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 15. Monitoring site of turbulence profile: (a) plan view; and (b) elevation view.

(a) (b) (c)

Fig. 16. Turbulence intensity versus wind speed: (a) Iuu; (b) Ivv; and (c) Iww.

(a) (b) (c)

Fig. 17. Horizontal turbulence profile of turbulence wind field: (a) Iuu; (b) Ivv; and (c) Iww.

(a) (b)

Fig. 18. Power spectrum profile at a wind speed of U = 2.80 m/s: (a) nSu/(u*)2; and (b) nSw/(u*)2.

© ASCE 04022012-7 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


(a) (b)
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

(c) (d)

Fig. 19. Motion amplitudes: (a) vertical, middle span; (b) vertical, quarter span; (c) torsional, middle span; and (d) torsional, quarter span.

of wind fluctuations at different positions with a mean wind speed deck axis, indicating that the axial wind flow stabilizes the struc-
of U = 2.80 m/s. ture aeroelastically.
3. Low-intensity turbulence postpones the flutter instability.

Results and Discussion


The results in turbulent flows are plotted in Fig. 19, where those in Data Availability Statement
smooth flows are provided for comparison. It is clear that existence
of the turbulence does not change the “soft” nature of flutter insta- All data, models, and code that support the findings of this study are
bility, and it does not blur the demarcation between stability and in- available from the corresponding author upon reasonable request.
stability, which differs from the work of Irwin et al. (1995) and
Scanlan (1997) but agrees with that of Diana et al. (1993). It is
also noticed that the flutter threshold has increased to 2.84 m/s, a Acknowledgments
13.2% promotion based on 2.51 m/s in smooth flow. This is in
agreement with main findings by Diana et al. (1993) and Lam The authors would like to express their gratitude for the financial
et al. (2017). The motion amplitudes are also affected by turbu- support from the Hainan Provincial Natural Science Foundation of
lence. An interesting phenomenon is that while the transverse mo- China (Grant No. 520CXTD433). The authors are also indebted to
tion amplitude in turbulent flow remains significantly lower than the National Natural Science Foundation of China for research pro-
that in smooth flow throughout the range of wind speeds covered, jects granted in recent years (Grant Nos. 51578233 and 51938012).
the rotation in turbulent flow catches up progressively with its
counterpart in smooth flow as U increases. Although the mecha-
nism of this phenomenon is not clear, the stabilization effect of References
low-intensity turbulence on flutter instability is demonstrated to
be positive. Amandolese, X., S. Michelin, and M. Choquel. 2013. “Low speed flutter
and limit cycle oscillations of a two-degree-of-freedom flat plate in a
wind tunnel.” J. Fluids Struct. 43: 244–255. https://doi.org/10.1016/j
.jfluidstructs.2013.09.002.
Conclusions Daito, Y., M. Matsumoto, and K. Araki. 2002. “Torsional flutter mecha-
nism of two-edge girders for long-span cable-stayed bridge.” J. Wind
In this study, the effects of wind yaw angles and low-intensity tur- Eng. Ind. Aerodyn. 90 (12–15): 2127–2141. https://doi.org/10.1016
bulence on the flutter behaviors of a suspension bridge model with /S0167-6105(02)00329-X.
a π-shaped deck section are investigated by wind tunnel tests. The Diana, G., S. Bruni, A. Cigada, and A. Collina. 1993. “Turbulence effect on
following conclusions are drawn based on the discussions flutter velocity in long span suspended bridges.” J. Wind Eng. Ind.
presented: Aerodyn. 48 (2–3): 329–342. https://doi.org/10.1016/0167-6105(93)
1. The structural damping ratios are found to increase nonlinearly 90144-D.
and drastically with respect to the motion amplitudes. The struc- Gao, G., L. Zhu, J. Li, W. Han, and B. Yao. 2020. “A novel
two-degree-of-freedom model of nonlinear self-excited force for cou-
tural damping ratio of the vertical motion identified at the mid-
pled flutter instability of bridge decks.” J. Sound Vib. 480: 115406.
span position deviate progressively from that at the quarter span, https://doi.org/10.1016/j.jsv.2020.115406.
indicating evolution of the modal shape with the motion Huston, D. R., H. R. Bosch, and R. H. Scanlan. 1988. “The effects of fair-
amplitude. ings and of turbulence on the flutter derivatives of a notably unstable
2. Wind yaw angles increase significantly the flutter thresholds in bridge deck.” J. Wind Eng. Ind. Aerodyn. 29 (1–3): 339–349. https://
terms of the wind speed component perpendicular to the bridge doi.org/10.1016/0167-6105(88)90172-9.

© ASCE 04022012-8 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012


Irwin, P. A., J. Xie, and G. Dunn. 1995. Wind tunnel studies for the Golden Pindado, S., J. Meseguer, and S. Franchini. 2005. “The influence of the sec-
Gate Bridge. Final Rep. 93-144F-4. Prepared for T. Y. Lin International. tion shape of box-girder decks on the steady aerodynamic yawing mo-
Guelph, ON, Canada: Rowan, Williams, Davies and Irwin. ment of double cantilever bridges under construction.” J. Wind Eng.
Lam, H. T., H. Katsuchi, and H. Yamada. 2017. “Investigation of turbulence Ind. Aerodyn. 93 (7): 547–555. https://doi.org/10.1016/j.jweia.2005
effects on the aeroelastic properties of a truss bridge deck section.” .05.005.
Engineering 3 (6): 845–853. https://doi.org/10.1016/j.eng.2017.10.001. Scanlan, R. 1997. “Amplitude and turbulence effects on bridge flutter
Matsumoto, M., H. Shirato, and S. Hirai. 1992. “Torsional flutter mecha- derivatives.” J. Struct. Eng. 123 (2): 232–236.
nism of 2-D H-shaped cylinders and effect of flow turbulence.” Scanlan, R., and J. Tomko. 1971. “Airfoil and bridge deck flutter deriva-
J. Wind Eng. Ind. Aerodyn. 41 (1–3): 687–698. https://doi.org/10 tives.” J. Eng. Mech. 97 (6): 1717–1737.
.1016/0167-6105(92)90480-X. Tang, Y., and X. Hua. 2019. “Experimental investigation of flutter charac-
Náprstek, J., and S. Pospíšil. 2011. “Post-critical behavior of a simple non- teristics of shallow Π section at post-critical regime.” J. Fluids Struct.
linear system in a cross-wind.” Eng. Mech. 18: 193–201. 88: 275–291. https://doi.org/10.1016/j.jfluidstructs.2019.05.010.
Náprstek, J., and S. Pospíšil. 2012. “Response types and general stability Wu, B., X. Chen, Q. Wang, H. Liao, and J. Dong. 2020. “Characterization
conditions of linear aero-elastic system with two degrees-of-freedom.” of vibration amplitude of nonlinear bridge flutter from section model
Downloaded from ascelibrary.org by University of Central Florida on 09/22/22. Copyright ASCE. For personal use only; all rights reserved.

J. Wind Eng. Ind. Aerodyn. 111: 1–13. https://doi.org/10.1016/j.jweia test to full bridge estimation.” J. Wind Eng. Ind. Aerodyn. 197:
.2012.08.002. 104048. https://doi.org/10.1016/j.jweia.2019.104048.
Náprstek, J., S. Pospíšil, and S. Hrač ov. 2007. “Analytical and experimen- Zhang, M., F. Xu, and X. Ying. 2017. “Experimental investigations on the
tal modelling of non-linear aeroelastic effects on prismatic bodies.” nonlinear torsional flutter of a bridge deck.” J. Bridge Eng. 22 (8):
J. Wind Eng. Ind. Aerodyn. 95 (9–11): 1315–1328. https://doi.org/10 04017048. https://doi.org/10.1061/(ASCE)BE.1943-5592.0001082.
.1016/j.jweia.2007.02.022. Zhu, L., M. Wang, D. Wang, Z. Guo, and F. Cao. 2007. “Flutter and buf-
Náprstek, J., S. Pospíšil, and J. Yau. 2015. “Stability of two-degrees- feting performances of third Nanjing bridge over Yangtze river under
of-freedom aero-elastic models with frequency and time variable para- yaw wind via aeroelastic model test.” J. Wind Eng. Ind. Aerodyn.
metric self-induced forces.” J. Fluids Struct. 57: 91–107. https://doi.org 95 (9–11): 1579–1606. https://doi.org/10.1016/j.jweia.2007.02.019.
/10.1016/j.jfluidstructs.2015.05.010. Zhu, L., G. Gao, and Q. Zhu. 2020. “Recent advances, future application
Pigolotti, L., C. Mannini, and G. Bartoli. 2017. “Experimental study on the and challenges in nonlinear flutter theory of long span bridges.”
flutter-induced motion of two-degree-of-freedom plates.” J. Fluids J. Wind Eng. Ind. Aerodyn. 206: 104307. https://doi.org/10.1016/j
Struct. 75: 77–98. https://doi.org/10.1016/j.jfluidstructs.2017.07.014. .jweia.2020.104307.

© ASCE 04022012-9 J. Bridge Eng.

J. Bridge Eng., 2022, 27(4): 04022012

You might also like