Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter analysis
Mannini, Bartoli and Borri
Proceedings of the Institution of Civil Engineers
Structures and Buildings 165 March 2012 Issue SB3
Pages 139–159 http://dx.doi.org/10.1680/stbu.2012.165.3.139
Paper 900008
Received 20/01/2009
Accepted 01/04/2011
Keywords: bridges/dynamics/wind loading & aerodynamics
ICE Publishing: All rights reserved
New developments in bridge
flutter analysis
1
j
Claudio Mannini PhD
Post-Doctoral Research Fellow, Inter-university Research Centre on
Building Aerodynamics and Wind Engineering, Department of Civil
and Environmental Engineering, University of Florence, Italy
2
Gianni Bartoli PhD
j
Associate Professor and Deputy Head of Inter-university Research
Centre on Building Aerodynamics and Wind Engineering,
1
j
2
j
Department of Civil and Environmental Engineering, University of
Florence, Italy
3
Claudio Borri Prof. Ing. Dr-Ing. H.c. Mult
j
Full Professor and Head of Inter-university Research Centre on
Building Aerodynamics and Wind Engineering, Department of Civil
and Environmental Engineering, University of Florence, Italy
3
j
The first part of this paper is devoted to an approximate approach to flutter, which is attained through simplification
of the flutter equations. The critical wind speed and the flutter frequency can be calculated with the proposed
formulas by employing only three flutter derivatives instead of the usual eight coefficients. This approach may be
seen as an easy engineering tool for a better tailoring of bridge structures at early design stages. In addition, the
simplicity of the equations allows better understanding of the mechanism of flutter instability and the role played by
structural parameters such as damping. In particular, an explanation is provided for soft- and hard-type flutter. The
second part of the paper outlines a model to take into account the uncertainty in the measurement of self-excited
forces in a flutter analysis. Ad hoc wind tunnel tests allowed determination of the statistical properties of the
measured flutter derivatives. These coefficients are treated as independent normally distributed random variables,
and Monte Carlo simulations are performed in order to determine the probability distribution function of the critical
wind speed. The paper concludes with an example of application of the proposed probabilistic flutter assessment
method.
1.
Introduction
Modern long-span bridges are more and more sensitive to wind
loads and aeroelastic phenomena, due to challenging designs
involving high-performance materials leading to lighter structures
and lower vibration frequencies. In particular, an adequate and
reliable safety margin with respect to the critical wind speed
leading to the aeroelastic instability known as flutter, which
induces catastrophic oscillations and even collapse of the structure, has to be guaranteed.
Classical flutter is a self-excited phenomenon due to the aeroelastic coupling of vertical bending and torsional modes that
introduces energy into the system, leading to divergent or largeamplitude limit-cycle oscillations. Torsional flutter (or torsional
galloping) is also relevant to bridge structures, wherein negative
damping in a torsional mode can be attained without any
coupling with other modes. Both phenomena are usually approached using semi-empirical models (Caracoglia and Jones,
2003; Costa and Borri, 2006; Scanlan, 1978a, 1978b; Scanlan
and Tomko, 1971; Simiu and Scanlan, 1996) in which some
aerodynamic coefficients (the so-called flutter derivatives) have to
be experimentally determined in a wind tunnel. Given the strong
dependence of the calculated flutter wind speed on these coefficients, the experimental phase and the identification procedure
are extremely important (Bartoli et al., 2009). The flutter
derivatives are, among other functions and coefficients, also
necessary for accurate estimation of the structural response to
turbulent wind (Scanlan, 1978b).
Two different aspects related to the analysis of the flutter
phenomenon are considered in this paper. Firstly, although it is
important to compute in the most accurate way possible the
flutter critical wind speed and the response to turbulent wind, the
authors have attempted to simplify the flutter problem, searching
for approximate formulas for flutter prediction to be employed at
early design stages and implemented for the improvement of
codes and standards (Mannini et al., 2007). Approximate approaches also help towards better understanding of the mechanism leading to flutter instability and consequently allow better
tailoring of the structural design in order to avoid flutter.
Several attempts to obtain simplified models of flutter assessment
139
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
have been reported (Bartoli and Mannini, 2008; Bartoli and Righi,
2006; Chen, 2007; Dyrbye and Hansen, 1997; Frandsen, 1966;
Nakamura, 1978; Mannini, 2006; Øiseth et al., 2010; Vairo, 2010),
demonstrating the relevance of this issue. A comparison of the
present approach and the methods proposed by Nakamura (1978)
and Chen (2007) will be discussed in the next section. Some
researchers (Bartoli and Righi, 2006; Dyrbye and Hansen, 1997;
Frandsen, 1966) tried to condense the aeroelastic performance of a
bridge deck section in one empirical coefficient, the aerodynamic
stability performance index. In fact, the flutter critical wind speed
can be calculated first for a dynamically equivalent theoretical flat
plate and then corrected through this coefficient, which is usually
less than one. This method is appealing for its simplicity but some
drawbacks have already been highlighted (Bartoli and Mannini,
2008; Mannini, 2006). Vairo (2010) set up an analytical approach
based on a simplified variational formulation for the dynamic
problem of the wind–structure interaction mechanism leading to
flutter instability in the case of cable-stayed bridges. However, the
most common simplified method of flutter assessment is the quasisteady theory. The limits of this approach have been discussed by
Bisplinghoff et al. (1996) and Bartoli and Mannini (2008), while
its absolute inadequacy to predict torsional flutter instability was
highlighted by Nakamura and Mizota (1975) and Nakamura
(1979). Øiseth et al. (2010) proposed a modified quasi-steady
approach, based on the linear or quadratic fit of experimentally
measured flutter derivatives, as an engineering tool to estimate the
flutter critical condition.
coefficient of variation of 0.20 was assumed. The same simple
model for the uncertainty affecting flutter derivatives was
assumed by Cheng et al. (2005) who, in a reliability analysis of
the Jing Yin Bridge, China, included a large number of random
variables and then performed a sensitivity study of the reliability
index with respect to the mean and standard deviation values of
these variables. However, in no research work has the actual
stochastic nature of the flutter derivatives been studied and no
probabilistic model has been proposed to determine the probability distribution of the critical flutter wind speed, given the
uncertainty in wind tunnel measurements. Therefore, in the third
section of this paper, a method is proposed in order to account
for aerodynamic input uncertainty in flutter analysis.
A second issue dealt with in this paper is related to the
experimental evidence that flutter derivatives involve uncertainties
(i.e. their measurement presents a significant dispersion), in
particular in the case of free-vibration setups (Mannini, 2006;
Righi, 2003). The literature indicates several attempts to account
for the stochastic effect on flutter boundaries due to oncoming
flow turbulence (Bartoli and Righi, 2006; Lin, 1996), but in only
a few cases and in a very approximate way has the uncertainty in
flutter derivatives been considered in the reliability analysis of
bridge structures.
Ostenfeld-Rosenthal et al. (1992) considered a log-normally
distributed random variable to account for uncertainty in structural damping and a Gaussian random variable for the uncertainty
factor related to the conversion from model to full scale. In
addition, the critical wind speed directly measured in a wind
tunnel was assumed to be a Gaussian random variable. Ge et al.
(2000) used an empirical formula to calculate the flutter critical
wind speed, considering uncertainty in the bridge deck mass,
mass moment of inertia and structural damping. A normally
distributed conversion factor from model to prototype was also
assumed. Pourzeynali and Datta (2002) considered the resistance
of a bridge structure as the product of the flutter critical wind
speed, log-normally distributed, and three independent log-normally distributed random variables accounting for the uncertainty
in structural damping, in mathematical modelling and in the
flutter derivatives. In particular, for the latter an almost arbitrary
140
2.
Simplified approach to flutter instability
2.1 Mechanical model
In flutter analysis, normally only the onset instability wind speed
is sought for the design of bridge structures. Therefore, under the
assumption of small oscillations perturbing the flow, the structure
can be modelled as a two degrees of freedom (dof) linear
oscillator:
1:
2:
h
i
_ þ ø2 h(t)
m h€(t) þ 2æ h ø h h(t)
h
_
Æ(t), Æ(t),
_
K
¼ Lse h(t), h(t),
€ (t) þ 2æÆ øÆ Æ(t)
I Æ
_ þ ø2Æ Æ(t)
_
¼ M se h(t), h(t),
Æ(t), Æ(t),
_
K
where h and Æ are the heaving displacement and the pitching
rotation (Figure 1), m and I the mass and the mass moment of
B
U
h
M
α
L
Figure 1. Reference scheme for displacements and self-excited
forces
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
inertia per unit length, ø h and øÆ are the circular eigenfrequencies, æ h and æÆ represent structural damping in heaving and
pitching modes respectively and Lse and Mse are the self-excited
lift and moment per unit length. K ¼ Bø/U is the reduced
frequency of oscillation, where B is the deck width, ø the circular
frequency of oscillation at flutter and U the undisturbed mean
flow speed; the dots denote derivatives with respect to time t.
Classically, the self-excited forces can be assumed to be linear
functions of structural displacements and velocities, parametrically dependent on the non-dimensional reduced frequency of
oscillation (Scanlan, 1978a, 1978b; Scanlan and Tomko, 1971;
Simiu and Scanlan, 1996):
3 and 4 has been found to be questionable in some instances
(Diana et al., 2008; Mannini et al., 2010a; Noda et al., 2003),
since a non-negligible dependence of flutter derivatives on the
amplitude of oscillation was observed.
_
BÆ(t)
_
h(t)
þ KH
2 (K)
U
U
h(t)
2
(K)Æ(t)
þ
K
H
(K)
þ K2 H
3
4
B
Lse ¼ qB KH
1 (K)
3:
By assuming coupled harmonic oscillations at frequency ø and
imposing the system complex determinant to vanish, one obtains
fourth-order and third-order polynomial equations with respect to
the non-dimensional frequency of oscillation Y ¼ ø/ø h (the socalled real and imaginary flutter equations (e.g. Dyrbye and
Hansen, 1997)). The polynomial coefficients are non-linear functions of the reduced frequency of oscillation K. The couple (Y,
Kc ), which identifies the critical condition, can be obtained as the
intersection of the solutions of the real and imaginary flutter
equations. Then the critical wind speed can be easily calculated
from:
5:
_
BÆ(t)
_
h(t)
þ KA
2 (K)
U
U
h(t)
2
2
þ K A3 (K)Æ(t) þ K A4 (K)
B
M se ¼ qB2 KA
1 (K)
4:
where q ¼ 0.5rU 2 is the mean dynamic pressure, r is the air
density and the coefficients H i and A i are the flutter derivatives,
which are functions of the reduced frequency of oscillation K as
well as the mean angle of attack (Diana et al., 2004; Mannini,
2006; Mannini and Bartoli, 2008a). This simple 2-dof model can
be extended to multi-mode analysis (Chen and Kareem, 2006;
Chen et al., 2000; Jain et al., 1996; Katsuchi et al., 1999;
Scanlan, 1978a, 1978b) of the bridge structure. Nevertheless, in
most cases the flutter critical wind speed can be accurately
calculated by considering two modes only (Bartoli and Mannini,
2005a; Chen et al., 2000). A complete flutter approach also
requires inclusion of the self-excited drag force and the contribution of the along-wind (sway) motion (Chen et al., 2000; Jain et
al., 1996; Katsuchi et al., 1999; Scanlan, 1978a, 1978b), so that
the flutter derivatives to be determined in the wind tunnel total 18
instead of eight. In most cases, the contribution of drag and sway
motion were found to be negligible (Bartoli and Mannini, 2005a;
Chen and Kareem, 2006; Chen et al., 2000; Øiseth et al., 2010),
although in a few instances a significant effect on the flutter
critical condition was encountered (Katsuchi et al., 1999; Vairo,
2010). Chen and Kareem (2006) explained that the additional
flutter derivatives are non-negligible only if the drag is large and
the self-excited lift and moment are small, so that the contribution of the latter to the aerodynamic damping builds up slowly
with increasing wind speed.
It is worth remarking that the classical linear model of Equations
U c ¼ Bø h
Y
Kc
Alternatively, it is possible to define the dissipative forces in the
left-hand side of Equations 1 and 2 as imaginary stiffness terms
(Fung, 1993) by introducing the rate-independent damping coefficients g h and gÆ :
h
i
m €h(t) þ (1 þ ig h )ø2h h(t)
6:
7:
_
Æ(t), ia(t), K
ih ¼ Lse h(t), h(t),
€ (t) þ (1 þ ig Æ )ø2Æ Æ(t)
I Æ
_
i ¼ M se h(t), iah(t),
Æ(t), Æ(t),
_
K
where i ¼ (1)1=2 denotes the imaginary unit. The rate-independent damping coefficients can be related to the ratio-to-critical
damping coefficients by the following expressions:
8:
g h ¼ 2æ h
ø
øh
9:
g Æ ¼ 2æÆ
ø
øÆ
The previously mentioned flutter equations can now be written
as:
141
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
ì2 2 2
1
2 2
ð1 g h g Æ Þ 2 r Æ X r Æ ì 1 þ 2
ªø
ªø
þ
12:
ì
þ r 2 ì H g H X
A
g
A
Æ
h
4
1
3
2
Æ
ª2ø
þ r2Æ ì2 þ ìA3 þ r2Æ ì H 4 þ H
4 A3
10:
H
1 A2 H 3 A4 þ H 2 A 1 ¼ 0
ð g h þ gÆ Þ
þ
ì2 2 2
gh
2 2
r
X
r
ì
g
þ
Æ
Æ
ª2ø
ª2ø Æ
#
ì
2
X
A2 þ g h A
3 þ rÆ ì H 1 þ g Æ H 4
2
ªø
H
3 A1 H 2 A4 ¼ 0
where
1=2
1 I
rÆ ¼
B m
is the non-dimensional radius of gyration,
ì¼
BøÆ
K c X 1=2
2.2 Approximate formulas
It is possible to remark that by assuming 1 g h g Æ ffi 1, Equation
10 can be written as:
2
þ ìA
2 þ r Æ ì H 1 þ H 4 A2 þ H 1 A3
11:
Uc ¼
13:
X
2 2
ìA3 rÆ ì (X 1) 1 2
ªø
X
r2Æ ì H 4 (X 1) þ g h ìA2 þ gÆ ª2ø r2Æ ì H
1
ª2ø
þ H 4 A3 H 1 A2 H 3 A4 þ H 2 A1
X
ffi ìA3 r2Æ ì2 (X 1) 1 2 ¼ 0
ªø
The approximation in Equation 13 holds for all the dynamic and
aerodynamic data of real bridge structures collected by the
authors in a database (Bartoli and Mannini, 2005b, 2008;
Mannini, 2006), unless X ffi ª2ø (i.e. ø ffi ø h ), a condition which
occurs only if the still air frequency ratio is very close to one.
A simplified equation is therefore obtained for the frequency
parameter:
2m
rB2
14:
X ¼1þ
A3
r2Æ ì
is the mass ratio of the deck section,
øÆ
ªø ¼
øh
is the still-air frequency ratio and
X ¼ ø2Æ =ø2
Furthermore, Equations 10 and 11 can be linearly combined in
order to eliminate the term proportional to X 2 and all the terms
that contain a second power of the damping coefficients can be
neglected. Several terms in the resulting equation are found to be
negligible with respect to the others, once the range of variability
of the dynamic and aerodynamic parameters has been identified
with reference to the previously mentioned bridge database
(Bartoli and Mannini, 2005b, 2008; Mannini, 2006). Therefore,
the following simplified equation can be obtained:
is a non-dimensional frequency parameter.
The advantage of this approach is that the flutter equations
simplify to second-order polynomial equations of the frequency
parameter X and hence manipulations are easier. Given the couple
(Xc , Kc ) which satisfies Equations 10 and 11, one can calculate
the flutter critical wind speed as:
142
15:
X ¼ ª2ø
r2Æ ì( g h þ gÆ ) þ A3 ( g h þ gÆ ) A2 r2Æ H 1
r2Æ ì(ª2ø g h þ gÆ ) þ gÆ A3 A2 ª2ø r2Æ H
1
By comparison with Equation 14, one obtains:
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
2
2
2
gÆ (A
3 ) þ g Æ r Æ ìA3 (2 ªø ) A2 A3
2
2
ª2ø r2Æ H
1 A3 þ r Æ ìA2 (ªø 1)
r4Æ ì2 gÆ (ª2ø 1) ¼ 0
16:
Finally, observing that the first two terms of Equation 16 are
negligible with respect to the others, the following simplified
equations can be obtained:
2 2
2
2
A
2 A3 þ ªø rÆ H 1 A3 r Æ ìA2 (ªø 1)
17:
þ r4Æ ì2 gÆ (ª2ø 1) ¼ 0
Equation 17 allows determination of the flutter critical reduced
frequency of oscillation, which appears as the argument of the
flutter derivatives. Then, through Equation 14, the non-dimensional frequency parameter X can be calculated and, through
Equation 12, the critical wind speed is finally obtained.
More details about the simplification procedure can be found in
previous publications (Bartoli and Mannini, 2005b, 2008; Mannini,
2006) in which Equations 15 and 16 had been taken as final
approximate formulas. Compared with these previous formulations, the first two terms of Equation 16 were discovered to be
always negligible in practice, while the equation for the frequency
parameter (Equation 15) was substituted by a simpler and more
meaningful relation (Equation 14) that involves only the flutter
derivative A3 instead of H 1 , A2 and A3 :
Apart from the convenient analytical form of the proposed equations, it is important to stress the fact that only three out of eight
flutter derivatives are retained. These functions are the so-called
‘uncoupled’ flutter derivatives, which can be measured in a wind
tunnel with 1-dof experimental setups. Clearly, Equations 14 and
17 represent a remarkable simplification of the flutter problem.
From analysis of Equation 14 one can notice that the mechanism
with which the torsional frequency decreases up to the flutter
frequency is very simple and analogous to the case of 1-dof
systems, that is, depending only on the flutter derivative A3 , and
is unaffected by structural damping. In fact, from Equation 14 it
is possible to write:
18:
ø ¼ øÆ
rB4
1þ
A
2I 3
1=2
Conversely, the analysis of Equation 17 is less immediate but it
can be remarked that structural damping can play a significant
role in the equation for the critical reduced frequency of
oscillation through one term depending only on the structural
properties of the bridge. These observations are in agreement
with the results of Chen and Kareem (2003) who found that
structural damping does not significantly influence frequency,
aerodynamic damping or complex mode shape and the aeroelastic
modal damping can be simply estimated as the sum of structural
modal damping and aerodynamic damping estimated with zero
structural damping. Moreover, Equation 17 shows that only
structural damping of the torsional mode has an influence on the
flutter mechanism. This is reasonable as, usually, it is the
evolution under wind of the torsional mode (torsional branch)
that becomes unstable; in other words, at the critical wind speed
it is the eigenvalue relative to torsion that exhibits a positive real
part. The assumption that the heaving branch is stable was also
stated by Nakamura (1978) in obtaining a simplified formula of
flutter assessment. Nevertheless, Matsumoto et al. (1999, 2002,
2008) observed that, in a few particular cases, the flutter
derivative A2 assumes sufficiently large negative values to
stabilise the torsional branch. In this instance it is the heaving
branch that gives rise to the flutter instability. This issue will be
discussed later on in the paper.
As is clear from the procedure of simplification of the flutter
equations, the approximate Equations 14 and 17 do not hold for
frequency ratios very close to unity, in the same way as Selberg’s
(Selberg, 1961) and Rocard’s (Frandsen, 1966) formulas for a flat
plate. Nevertheless, it seems that a frequency ratio of about 1.3,
or sometimes even less, is sufficient to obtain an acceptable
degree of approximation. Frequency ratios very close to unity are
fairly uncommon and usually characterise either super-long-span
bridges or very unconventional structures that are not expected to
be analysed with simplified methods, requiring instead careful
experimental campaigns from the very beginning of the design
procedure.
If damping is neglected, the proposed model reduces to that
proposed by Nakamura (1978), although it was obtained in a
completely different way (Bartoli and Mannini, 2008; Mannini,
2006). The simplified closed-form solution of the flutter equations
proposed by Chen (2007) accounts for the structural damping
contribution and is also based on the assumption of wellseparated frequencies. Chen’s approach retains only four flutter
derivatives, namely H
3 , A1 , A2 and A3 : The fact that the flutter
derivatives considered here are different from those emphasised
in that work should not be seen as a contradiction in view of
existing flutter derivative inter-relations (Matsumoto, 1996; Scanlan et al., 1997). In particular, the fact that the proposed formulas
depend only on the coefficients H
1 , A2 and A3 does not mean
that these are the most important derivatives for the instability
mechanism and that the others can be neglected. On the contrary,
the apparent redundancy of the classical model of self-excited
forces (Equations 3 and 4) just allows expression of the critical
condition with three aerodynamic parameters instead of eight.
An important feature of the model is that the present equation for
the critical reduced frequency of oscillation does account for the
143
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
contribution of structural damping is an important feature of the
model. In fact, as noted by Chen and Kareem (2003), while the
role of structural damping on the onset of instability is negligible
in the case of ‘hard-type flutter’, it is significant in the case of
‘soft-type flutter’. In the latter instance, negative damping builds
up slowly with increasing wind speed, so that a small translation
of the curves of total damping due to an increase of the structural
damping is able to induce a significant increment of the critical
wind speed. By contrast, the change of sign of the total damping
is abrupt for hard-type flutter and therefore the instability
condition is only weakly sensitive to small translations of its
curve. Only in the case of soft-type flutter devices such as tuned
mass dampers can be effective in controlling flutter instability
(Chen and Kareem, 2003) and the effect of large self-excited drag
forces can be non-negligible (Chen and Kareem, 2006). This
issue will be further analysed later in the paper.
tests were performed considering the flutter derivatives of a
theoretical flat plate (Fung, 1993; Theodorsen, 1934), which is
always an important benchmark for bridge aeroelasticity, and a
rectangular cylinder with a chord to thickness ratio B/H ¼ 12.5
(specimen R12.5, the bluffer one among those studied by
Matsumoto (1996) not prone to torsional flutter).
Finally, if damping is provided as a ratio-to-critical instead of a
rate-independent coefficient, the procedure of calculation of the
critical reduced frequency of oscillation through Equation 17
becomes iterative. Equation 9 with ø ¼ øÆ is used to obtain a
first estimate of g Æ and then, once the critical flutter frequency
has been calculated through Equation 14 or 18 , the rateindependent damping coefficient is updated and the flutter
calculation is repeated. One iteration is normally enough to
obtain convergence.
2.3 Model validation
In order to validate Equations 14 and 17, the following procedure
was adopted. A large number of structural parameters for various
typologies of existing bridges (suspension, cable-stayed and
footbridges) were collected and compared (Bartoli and Mannini,
2005b, 2008; Mannini, 2006). Since mass and moment of inertia
enter in the flutter equations (Equations 10 and 11) in the form of
the non-dimensional parameters ì, r2Æ ì and r2Æ ì2 , two bridge
structures were identified as representative of opposite extreme
dynamics and used as references in the following analyses:
Tsurumi Fairway Bridge in Japan (ì ¼ 35:7 and rÆ ¼ 0:249) and
Rio Guamà Bridge in Brazil (ì ¼ 178:6 and rÆ ¼ 0:353). Both
are cable-stayed bridges but Rio Guamà Bridge is characterised
by a concrete deck and consequently by a remarkable mass with
respect to the chord B. In contrast, Tsurumi Fairway Bridge has a
relatively light steel deck with a large chord length. In all the
flutter calculations presented in the following, air density was
always assumed to be r ¼ 1.25 kg/m3 :
In order to draw as general as possible conclusions, it is
important to consider a large number of dynamic and aerodynamic data. Assuming that flutter derivatives depend on reduced
frequency of oscillation and cross-section geometry only, it is
possible to combine the aerodynamics of a bridge with the
dynamic properties of a completely different one, performing
calculations on idealised structures. This is important: it allows
one to employ all the reliable aerodynamic and dynamic data
available for the calculations. In particular, the first validation
144
Figure 2 combines the previously discussed reference dynamics
and aerodynamics, and compares the rigorous solution of the
flutter equations (Equations 10 and 11) with those given by the
approximate formulas (Equations 14 and 17) in terms of critical
reduced wind speed URc ¼ 2ð/Kc for different values of the
frequency ratio ªø and structural damping. It is clear that the
approximation offered by the simplified method is good unless
the frequency ratio is very close to unity. It is also apparent in
Figure 2(d) that when the effect of damping is significant, the
simplified equations are able to correctly account for it.
A similar analysis was performed for the case of a deck crosssection prone to torsional flutter. In particular, the flutter derivatives measured by Matsumoto (1996) for a rectangular cylinder
with a chord to thickness ratio B/H ¼ 5.0 (specimen R5) were
considered, as this geometry is considered a benchmark test case
for bridge aerodynamics and aeroelasticity (Bartoli et al., 2008).
The results are reported in Figure 3, along with the solution of the
1-dof problem. It is worth noting that the 2-dof and 1-dof
approaches usually give close results, unless the frequency separation is small. Nevertheless, Chen et al. (2000) and Chen and
Kareem (2006) observed that the coupling of heaving and pitching
motions generally reduces the critical wind speed in the case of
torsional flutter. The agreement between exact and approximate
solutions is again very good and the simplified equations are able
to take into account the unfavourable effect of coupling between
bending and torsion and the role played by damping (Figure 3(b)).
As further validation, various case studies were taken into
account for classical and torsional flutter instabilities (Tables 1–
4). In particular, cases 5 and 6 in Tables 1 and 2 refer respectively
to Akashi Kaikyo Bridge (Katsuchi et al., 1999) in Japan and
Tsurumi Fairway Bridge in Japan (Singh et al., 1995); cases 1
and 9 refer to the Original Tacoma Narrows Bridge, USA
(Scanlan and Tomko, 1971) and Golden Gate Bridge, USA
(Simiu and Scanlan, 1996) for both dynamic and aerodynamic
properties. All the other case studies are idealised. The flutter
derivatives of rectangular cylinders with chord to thickness ratios
of 20 (R20 (Matsumoto, 1996)) and 10 (R10 (Matsumoto, 1996))
and a rectangular cylinder with semi-circular fairings and a chord
to thickness ratio of 14.3 (R14.3F (Chowdhuri and Sarkar, 2004))
were also studied. It is worth noting in Tables 3 and 4 that,
according to the flutter classification outlined by Matsumoto et
al. (2002), R5, the Tacoma and the Golden Gate sections are
prone to the so-called low-speed torsional flutter, while R10 is
susceptible to high-speed torsional flutter.
Concerning structural properties, case 2 of Tables 1 and 2 and
Structures and Buildings
Volume 165 Issue SB3
45
Flat plate – µ ⫽ 35·7, rα ⫽ 0·249
16
40
14
35
12
30
10
URc
URc
18
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
8
ζh ⫽ ζα ⫽ 0%; exact solution
25
20
6
ζh ⫽ ζα ⫽ 1%; exact solution
15
4
ζh ⫽ ζα ⫽ 0%; approximate solution
10
2
ζh ⫽ ζα ⫽ 1%; approximate solution
5
0
0
1
14
2
γω
3
4
5
1
2
γω
3
4
(a)
(b)
R12.5 – µ ⫽ 35·7, rα ⫽ 0·249
R12.5 – µ ⫽ 178·6, rα ⫽ 0·353
5
20
12
10
15
8
URc
URc
Flat plate – µ ⫽ 178·6, rα ⫽ 0·353
6
10
4
5
2
0
1
2
γω
3
4
5
(c)
0
1
2
γω
3
4
5
(d)
Figure 2. Comparison of approximate and exact solutions of the
flutter equations (URc ¼ 2ð/Kc is the flutter critical reduced wind
speed) for different values of the frequency ratio ªø and two
values of the ratio-to-critical damping coefficients in the case of
classical flutter instability
case 5 of Tables 3 and 4 refer to the 1992-proposed design of
Messina Strait Bridge, Italy; case studies 4 and 9 of Tables 1 and
2 refer to Bosporus I Strait Bridge, Turkey; case study 7 of
Tables 1 and 2 refers to Indiano Bridge, Italy; case study 10 of
Tables 1 and 2 refers to Normandy Bridge, France. The interested
reader can find more details about the dynamic properties of
these bridges in the literature (Bartoli and Mannini, 2005b, 2008;
Mannini, 2006).
It is apparent in the tables that the degree of approximation
offered by the simplified formulas is good, often limited to a
few percent and in most cases below 10%. Only in cases 2, 6
and 7 (Tables 1 and 2) are the errors larger, but still well below
20%. For case 2, the discrepancy was expected as the frequency
ratio is small, namely ªø ¼ 1:32 (the dynamic parameters of
the 1992-proposed design of Messina Strait Bridge are considered in this case). In contrast, test cases 6 and 7 refer to the
flutter derivatives of Tsurumi Fairway Bridge. As shown in
Figure 4, for this fairly streamlined cross-section, the values
assumed by the coefficient H
4 are surprisingly high, thus
rendering the proposed formulas slightly less accurate. In fact,
the absolute value of the coefficient is in this case much larger
not only than the one predicted by Theodorsen’s theory for a
flat plate (Fung, 1993; Theodorsen, 1934) but also for most
bridge deck sections, which usually show values close to zero.
This is the reason why this coefficient was neglected along with
A
4 in several formulations (e.g. Scanlan, 1978a, 1978b; Scanlan
and Tomko, 1971).
Tables 2 and 4 show that the approximate formula for the critical
frequency is very accurate. Also, agreement with the rigorous 2dof results is extremely good in the case of structures prone to
torsional flutter (both for low-speed and high-speed torsional
flutter; Table 4) and generally significantly better than the results
offered by the 1-dof approach. Moreover, comparison with the
results reported by Bartoli and Mannini (2008) demonstrates that
145
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
R5 – µ ⫽ 35·7, rα ⫽ 0·249
the additional simplification of the formulas introduced here does
not imply any significant loss of accuracy.
5
URc
4
Figure 5 shows the pattern of the function on the right-hand side
of Equation 17 and its zero crossing as compared with the exact
solution of the flutter equations in the case of Akashi Kaikyo
Bridge (case 5 in Tables 1 and 2) and the Original Tacoma
Narrows Bridge (case 1 in Tables 3 and 4).
ζh ⫽ ζα ⫽ 0%; exact solution
3
ζh ⫽ ζα ⫽ 1%; exact solution
ζh ⫽ ζα ⫽ 0%; approximate solution
2
ζh ⫽ ζα ⫽ 1%; approximate solution
ζh ⫽ ζα ⫽ 0%; 1-dof solution
1
ζh ⫽ ζα ⫽ 1%; 1-dof solution
0
1
2
4
3
γω
(a)
5
R5 – µ ⫽ 178·6, rα ⫽ 0·353
6
5
URc
4
3
2
1
0
1
2
γω
(b)
4
3
5
Figure 3. Comparison of approximate and exact 1-dof and 2-dof
solutions of the flutter equations (URc ¼ 2ð/Kc is the flutter critical
reduced wind speed) for different values of the frequency ratio ªø
and two values of the ratio-to-critical damping coefficients in the
case of torsional flutter instability
Case
Aerodynamics
1
2
3
4
5
6
7
8
9
10
Flat plate
Flat plate
R12.5
R12.5
Akashi Kaikyo Bridge
Tsurumi Fairway Bridge
Tsurumi Fairway Bridge
R14.3F
R14.3F
R20
Table 1. Case studies (classical flutter)
146
ì
rÆ
ªø
35.7
24.1
178.6
27.7
55.6
35.7
55.5
47.2
27.7
38.7
0.249
0.374
0.353
0.357
0.422
0.249
0.250
0.539
0.357
0.286
2.38
1.32
1.96
2.29
2.34
2.38
2.06
1.54
2.29
2.27
The case of wind tunnel section models, to which Equations 14
and 17 also apply, is very particular. As an example, a common
trapezoidal box girder with lateral cantilevers (Figure 6), similar
to Sunshine Skyway Bridge in Florida, tested in the CRIACIV
(Inter-university Research Centre on Buildings Aerodynamics and
Wind Engineering) wind tunnel, was considered (Mannini, 2006;
Mannini et al., 2010b). Two different configurations were
selected: the second one was simply obtained from the first by
adding eccentric masses to the suspension system (for configuration Dyn. 1 B ¼ 0:45 m, rÆ ¼ 0:294 and ì ¼ 45:0; for Dyn. 2
B ¼ 0:45 m, rÆ ¼ 0:426 and ì ¼ 52:0). Flutter derivatives were
measured on configuration Dyn. 1. It is worth noting that, from a
structural dynamic point of view, these two test cases are within
the range defined by the previously mentioned reference dynamics (i.e. those of Tsurumi Fairway Bridge and Rio Guamà
Bridge) in spite of the fact that the section model was not
dynamically scaled with respect to any real bridge but just
designed to measure flutter derivatives (Mannini, 2006). Figure 7
shows a comparison of the approximate and exact solutions in
terms of the dimensional critical wind speed for several values of
the frequency ratio and two ratio-to-critical damping coefficients.
The agreement between the different sets of results is good up to
small frequency separations.
In some particular cases, it is not the torsional branch that is
unstable, but coupled flutter arises in the heaving-branch (Matsumoto et al., 1999, 2002, 2008). For instance, this is the case for a
rectangular cylinder with chord to thickness ratio B/H ¼ 20 and
vertical plates at mid-chord (R20-VP (Matsumoto et al., 1999,
æ h : % æÆ : %
0.5
0.6
0.8
0.5
0.5
0.5
0.2
0.5
0.5
0.2
0.5
0.7
0.8
0.5
0.3
0.5
0.2
0.5
0.5
0.5
Structures and Buildings
Volume 165 Issue SB3
Case
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
Exact solution
1
2
3
4
5
6
7
8
9
10
Approximate solution
fc : Hz
URc
Uc : m/s
fc : Hz
˜fc : %
URc
˜URc : %
Uc : m/s
˜Uc : %
0.315
0.071
0.562
0.308
0.138
0.377
0.880
0.180
0.284
0.338
10.75
6.72
16.80
7.64
16.14
8.34
11.04
11.41
10.16
11.21
128.7
28.9
134.2
66.0
79.2
119.5
217.6
24.6
80.9
90.2
0.315
0.071
0.563
0.314
0.138
0.352
0.836
0.179
0.283
0.340
+0.1
+0.5
+0.1
+1.7
0.3
6.7
5.0
0.1
0.6
+0.7
10.43
5.51
16.83
7.01
15.91
9.89
12.72
10.59
9.35
10.79
3.0
18.0
+0.2
8.2
1.5
+18.5
+15.3
7.2
8.0
3.7
124.9
23.8
134.5
61.6
77.2
132.2
238.3
22.8
74.0
87.4
2.9
17.6
+0.2
6.6
1.7
+10.6
+9.5
7.3
8.5
3.1
Table 2. Results for the case studies in Table 1; fc is the coupling
frequency, URc is reduced flutter wind speed and Uc is flutter
wind speed
Case
Aerodynamics
1
2
3
4
5
6
7
8
9
10
Tacoma Narrows Bridge
R5
R5
R5
R5
R10
R10
R10
Golden Gate Bridge
Golden Gate Bridge
ì
rÆ
ªø
47.2
47.2
35.7
178.6
24.1
47.2
35.7
178.6
94.2
35.7
0.539
0.539
0.249
0.353
0.374
0.539
0.249
0.353
0.369
0.249
1.54
1.54
2.38
1.96
1.32
1.54
2.38
1.96
2.20
2.38
æ h : % æÆ : %
0.5
0.5
0.5
0.8
0.6
0.5
0.5
0.8
0.5
0.5
0.5
0.5
0.5
0.8
0.7
0.5
0.5
0.8
0.5
0.5
Table 3. Case studies (torsional flutter)
Case
1-dof solution
fc : Hz
1
2
3
4
5
6
7
8
9
10
URc
0.200 4.81
0.189 5.05
0.376 4.64
0.626 5.73
0.066 4.74
0.182 9.89
0.334 9.15
0.602 11.21
0.188 4.31
0.448 3.46
2-dof exact solution
Approximate solution
Uc : m/s
fc : Hz
URc
Uc : m/s
fc : Hz
˜fc : %
URc
˜URc : %
Uc : m/s
˜Uc : %
11.5
11.5
66.3
50.9
19.0
21.6
116.3
95.9
22.2
58.9
0.200
0.190
0.377
0.626
0.067
0.188
0.372
0.610
0.188
0.452
4.81
4.41
4.08
5.56
2.76
7.65
7.21
10.16
4.23
3.28
11.5
10.0
58.5
49.4
11.2
17.3
101.9
87.9
21.8
56.4
0.200
0.190
0.385
0.626
0.070
0.189
0.375
0.611
0.188
0.453
0.0
+0.4
+2.0
+0.1
+3.3
+0.1
+0.9
+0.2
+0.0
+0.1
4.81
4.35
3.94
5.52
2.90
7.46
6.99
10.00
4.10
3.10
0.0
1.4
3.2
0.7
+5.0
2.5
3.0
1.6
3.1
5.4
11.5
9.9
57.7
49.1
12.2
16.9
99.8
86.7
21.2
53.4
0.0
1.0
1.3
0.6
+8.5
2.3
2.1
1.4
3.0
5.3
Table 4. Results for the case studies in Table 3 (differences are
calculated with respect to the 2-dof exact solution)
147
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
10
2
Exact
Approximate
8
0
6
4
f (UR)
H *4
⫺2
⫺4
2
0
⫺6
Flatplate theory
Tsurumi Fairway Bridge
⫺8
⫺2
⫺4
⫺10
0
5
10
15
⫺6
UR
Figure 4. Comparison of theoretical values of the flutter derivative
H4 for a flat plate (Fung, 1993; Theodorsen, 1934) and measured
values for the Tsurumi Fairway Bridge deck section (Singh et al.,
1995)
5
0
15
10
UR
20
(a)
4
3
2
The case of vertical galloping is not taken into account here as
this instability is rarely a problem for bridge designs. Moreover, a
simple quasi-steady approach seems to be fairly adequate to
estimate critical wind speed.
2.4 Analysis of the mechanism of instability
The simplified formulas of Equations 14 and 17 can also be used
to better understand the mechanism of flutter instability. For this
purpose, all the terms in Equation 17 containing the aerodynamic
stiffness coefficient A3 can be moved to the right-hand side of
the equation:
19:
2
A
2 rÆ ì g Æ ¼
1
A
3
A2 þ ª2ø r2Æ H
1
2
2
ªø 1 rÆ ì
1
f (UR)
2002)), which shows large negative values of A2 able to stabilise
the torsional branch. A few case studies for this type of
instability are presented in Tables 5 and 6. Despite the fact that
they were considered for the torsional branch instability, the
simplified formulas give reasonable results also in these cases.
Nevertheless, by reducing the frequency ratio, the flutter equations do not allow any solution while the approximate approach
does. The minimum frequency ratio for which a flutter solution is
attained strongly depends on the bridge dynamic properties as is
evident in Tables 5 and 6.
0
⫺1
⫺2
⫺3
⫺4
0
1
2
4
3
5
6
Figure 5. Graphic solution of the approximate Equation 17 for
(a) Akashi Kaikyo Bridge and (b) Tacoma Narrows Bridge.
2
2 2
2
4 2
2
f (UR ) ¼ A
2 A3 + ªø r Æ H1 A3 r Æ ìA2 (ªø 1) + r Æ ì gÆ (ªø 1).
The exact solution of the flutter equations is also reported
equation can be interpreted as follows. The progressive reduction
of the torsional frequency X 1 due to the aerodynamic coefficient A
3 is transformed into negative aerodynamic damping
through a weighting factor depending on the flutter derivatives
H
1 and A2 , the frequency ratio and the non-dimensional radius
of gyration. Instability is reached when this negative damping
equals the positive structural damping and the direct aerodynamic
damping in torsion that depends on A
2:
Substituting Equation 14 into Equation 19 yields:
Equation 19 can also be rearranged in the following manner:
20:
X 1
2
A2 þ ª2ø r2Æ H
A
1
2 rÆ ì g Æ ¼ 2
ªø 1
In the case of sections not prone to torsional flutter (A2 , 0), this
148
7
UR
(b)
21:
gÆ
A2
¼0
r2Æ ì
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
450
150
150
150
2%
Aluminium rib s ⫽ 8 mm
Stiffners
70
65·5
4·6
4
2%
8
8
10
10
Sheet aluminium s ⫽ 0·5 mm
76·2
71·1
155·4
76·2
71·1
Figure 6. Section-model geometry (dimensions in millimetres)
70
ζh ⫽ ζα ⫽ 0%; exact solution
Case
Aerodynamics
40
1
2
3
4
R20-VP
R20-VP
R20-VP
R20-VP
30
Table 5. Case studies (heaving-branch classical flutter)
ζh ⫽ ζα ⫽ 1%; exact solution
60
ζh ⫽ ζα ⫽ 0%; approximate solution
Uc: m/s
50
ζh ⫽ ζα ⫽ 1%; approximate solution
ì
rÆ
ªø
35.7
35.7
178.6
178.6
0.249
0.249
0.353
0.353
2.38
1.65
1.96
1.20
æ h : % æÆ : %
0.5
0.5
0.8
0.8
0.5
0.5
0.8
0.8
20
where
10
0
0
1
2
70
γω
(a)
3
4
5
22:
60
Uc: m/s
50
40
30
20
10
0
0
1
2
γω
(b)
3
4
5
Figure 7. Comparison of approximate and exact solutions of the
flutter equations (Uc is the dimensional critical wind speed) for
different values of the frequency ratio ªø and two values of the
ratio-to-critical damping coefficients in the case of the wind
tunnel section model of Figure 6 (Mannini, 2006; Mannini et al.,
2010b). (a) Dyn. 1 (B ¼ 0:45 m, rÆ ¼ 0:294, ì ¼ 45:0); (b) Dyn. 2
(B ¼ 0:45 m, rÆ ¼ 0:426, ì ¼ 52:0)
A2 ¼ A2
1 A2 A3
ª2ø H 1 A3
ª2ø 1 r2Æ ì
ª2ø 1 ì
Then the 2-dof flutter critical condition is written as for the 1-dof
torsional flutter instability, by substituting the flutter derivative A
2
with the modified function A
2 : It is clear from Equation 21 that
2
A
2 =rÆ ì can be interpreted as aerodynamic damping of the
unstable branch. In the case of a cross-section prone to torsional
flutter, the difference between A
2 and A2 tends to be small
(Mannini and Bartoli, 2008b). The term proportional to H 1 A
3,
which is equivalent to A1 H
3 , gives the main destabilising
contribution, in agreement with the conclusions of others (Chen,
2007; Matsumoto, 1996; Øiseth et al., 2010). Nevertheless, unless
the non-dimensional mass moment of inertia r2Æ ì ¼ 2I=rB4 and
the frequency ratio are very large, the term proportional to A
2 A3
can also play a non-negligible destabilising role.
This simple flutter formulation can give some indications to tailor
a structure at early design stages and increase the flutter critical
wind speed by modifying the structural dynamic properties of a
bridge. In particular, it can help to understand to what extent the
classical strategies are really efficient. Obviously an increase in
torsional frequency (especially by increasing the torsional stiffness of the structure) has a positive effect because the dimensional critical wind speed increases once the reduced critical
149
Structures and Buildings
Volume 165 Issue SB3
Case
1
2
3
4
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
Exact solution
Approximate solution
fc : Hz
URc
Uc : m/s
fc : Hz
˜fc : %
URc
˜URc : %
Uc : m/s
˜Uc : %
0.249
—
0.377
—
14.13
—
27.84
—
133.8
—
149.0
—
0.242
0.220
0.403
0.340
3.1
—
+7.0
—
12.77
9.56
24.06
13.66
9.7
—
13.6
—
117.2
79.9
137.8
66.0
12.4
—
7.5
—
Table 6. Results for the case studies in Table 5; fc is the coupling
frequency, URc is reduced flutter wind speed and Uc is flutter
wind speed
The simplified formulas can also help to explain the reasons for
the different types of flutter instability highlighted by Chen and
Kareem (2003) and discussed in Section 2.2. It is worth noting
that, once the bridge structural properties have been fixed, one can
observe hard- or soft-type flutter depending on the aerodynamic
properties, as clearly shown in Figure 8. In the same way, given
the flutter derivatives, the type of instability can change depending on the structural properties of the bridge. From the analysis of
Equations 21 and 22 it can be seen that a reduction in the absolute
value of the flutter derivative, A2 , or in the frequency ratio ªø ,
determines a decrease of the slope of A2 near the critical
condition, thus fostering soft-type flutter instability. Also, a
diminution of the non-dimensional mass moment of inertia, r2Æ ìj
has a similar effect. Instead, a variation of the non-dimensional
mass, ì ¼ 2m=rB2 does not significantly change this slope. It is
also worth noting that the structural damping coefficient g Æ in
Equations 19–21 is directly scaled by r2Æ ì, so that the importance
of damping is amplified if the mass moment of inertia of the deck
section is large, in agreement with the results of Figures 2 and 3
(consider that r2Æ ì differs by a factor of about 10 between the Rio
150
0·4
R12.5, heaving branch
R12.5, torsional branch
Flat plate, heaving branch
Flat plate, torsional branch
0·3
Total damping
wind speed has been fixed (see Equation 12). However, particularly interesting is to understand how the reduced critical wind
speed can be increased. From Equation 22 it is apparent that in
the case of classical flutter instability an increase in deck mass
per unit length m always has a favourable effect. Conversely, if
A
2 takes small values and the frequency ratio is already fairly
large, a further increment of the frequency separation does not
produce any appreciable increase of the reduced critical wind
speed (see Figures 2(c) and 2(d). In the same conditions, an
increase of the mass moment of inertia I tends not to be effective.
However, the reduced critical wind speed might be more sensitive
to variations of the mass moment of inertia in the case of large
values of the structural damping in torsion, as shown by
Equation 21. In fact the discussed parameters are generally
related (for instance, the frequency ratio depends on the ratio of
the mass moment of inertia to the mass of the deck) and it is not
possible to change them in a completely independent way. A
detailed discussion of the effects of different structural dynamic
parameters on the flutter instability mechanism can be found in
Mannini and Bartoli (2008b).
0·2
Hard-type flutter
0·1
Soft-type flutter
0
⫺0·1
⫺0·2
0
5
15
10
20
25
U/(Bfα)
Figure 8. Example of soft-type and hard-type flutter for the same
structure (ì ¼ 178:6, rÆ ¼ 0:353, ªø ¼ 1:96, æ h ¼ æÆ ¼ 0) with
different aerodynamic properties. The total modal damping
(structural plus aerodynamic) is plotted against the normalised
wind speed ( fÆ is the still-air torsional frequency)
Gramà and Tsurumi Fairway bridges). Therefore, if the slope of
the curve of the aerodynamic damping near the critical condition
is small enough (soft-type flutter) and the non-dimensional mass
moment of inertia is anyway large, then structural damping plays
a non-negligible role on the mechanism of instability. It is clear
from what has been said so far that a key parameter that
contributes to trigger the type of instability is the flutter
derivative, A
2 . In fact, Figure 9 shows that the rectangular
cylinder R12.5 with respect to the thin flat plate presents similar
values of H 1 and A
3 but much smaller absolute values of A2 ;
this is why the test case of Figure 8 shows soft-type flutter for the
former and hard-type flutter for the latter. In conclusion, soft- and
hard-type flutters have a complicated combined dynamic and
aerodynamic origin and the simplified equations proposed are
clearly able to distinguish between the two.
Finally, it should be noted that when a 2-dof flutter calculation is
performed, either rigorous or simplified, the choice of the critical
modes susceptible to couple is crucial. This selection must be
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
0
fairly complicated and this choice is not trivial. Sometimes even
a multi-mode approach is necessary to properly account for the
actual mechanism of instability (Øiseth et al., 2010).
⫺2
⫺4
H*1
3.
3.1 Uncertainty in the flutter derivative measurements
Flutter derivatives are usually employed in flutter and buffeting
analyses as deterministic coefficients. Nevertheless, several studies
have demonstrated that these type of wind tunnel measurements are
characterised by significant dispersion (Mannini, 2006; Righi,
2003). This is particularly true for free-vibration setups while the
problem is expected to be reduced for the forced-vibration technique where it is easier to control a few important test conditions
(e.g. amplitude of vibration or length of measurements). Nevertheless, in the second case, uncertainty is not eliminated as the
measurement of small aerodynamic quantities is obtained through
the difference of large forces (due to the presence of inertial forces)
under wind and in still-air conditions. If computational fluid
dynamics simulations are employed to determine the flutter
derivatives, random uncertainties are substituted by the concern of
systematic errors due to the accuracy of the governing equations (in
particular the turbulence model equations) or to the numerical
methods used to solve them. In addition, if advanced strategies of
turbulence modelling are adopted (such as three-dimensional large
eddy simulation or detached eddy simulation), solutions show
significant stochastic components and statistical convergence is
obtained only through long and computationally expensive simulations. Therefore, if short computations are performed, the results
again involve random uncertainties.
⫺6
Flatplate theory
⫺8
⫺10
R12.5
0
2
4
6
UR
(a)
8
10
0
2
4
6
UR
(b)
8
10
12
0
⫺0·2
A*2
⫺0·4
⫺0·6
⫺0·8
⫺1·0
⫺1·2
⫺1·4
12
5
4
A*3
3
2
1
0
0
2
4
6
UR
(c)
8
10
Probabilistic approach to flutter
12
Figure 9. Comparison of flutter derivatives for a thin flat plate
(Fung, 1993; Theodorsen, 1934) and a 12.5:1 rectangular cylinder
(R12.5 (Matsumoto, 1996))
based on the mode coupling coefficient (Bartoli and Mannini,
2005a; Dyrbye and Hansen, 1997), which accounts for imperfect
mode shape similarity, on the frequency ratio and on the absolute
value of the torsional frequency. In many cases the choice is
obvious in practice but sometimes more than one calculation is
necessary to find the smallest flutter critical wind speed. In a few
cases, especially at bridge erection stages, the mode shapes are
As an example of the possible scatter in wind tunnel results,
2
Figure 10 shows the functions KH 1 , K 2 H
3 , KA2 and K A3
measured in the CRIACIV wind tunnel for a bridge deck section
model (Figure 6) characterised by a geometry similar to the
cross-section of Sunshine Skyway Bridge, Florida. The tests were
performed according to the free-vibration technique (Mannini,
2006). The modified unifying least squares (Bartoli et al., 2009)
method was used to identify the flutter derivatives and it was
shown to perform very well (Bartoli et al., 2009; Mannini, 2006).
The originally measured coefficients KH 1 , K 2 H
3 , KA2 and
2
K A3 are plotted instead of H 1 , H 3 , A2 and A3 in order not to
alter the experimental scatter by dividing by the reduced
frequency K or K2 : It is evident that the dispersion of the data
tends to increase with the reduced wind speed UR ¼ 2ð/K and
becomes definitely important near the flutter boundary. This is
probably due to the fact that the total damping (structural plus
aerodynamic) becomes very small near flutter and therefore any
disturbance has a large effect on the mechanical system. In
addition, as is typical in free-vibration setups, the mean angle of
attack tends to increase with increasing wind speed due to the
static aerodynamic moment and is no longer negligible near
flutter (Mannini, 2006; Mannini and Bartoli, 2008a). This effect
implies a change in the aerodynamic properties of the crosssection and enhanced vortex shedding may be partly responsible
151
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
0
⫺2·5
⫺1
⫺3·0
⫺2
⫺3·5
K 2H*3
KH*1
⫺3
⫺4
⫺5
⫺4·0
⫺4·5
⫺6
⫺5·0
⫺7
⫺8
5
0
⫺5·5
15
10
UR
(a)
1·4
⫺0·1
1·2
⫺0·2
1·0
KA*2
K 2A*3
0
⫺0·3
0·8
⫺0·4
0·6
⫺0·5
0
2
4
UR
(c)
6
8
10
0·4
0
0
2
2
4
4
UR
(b)
UR
(d)
6
8
10
6
8
10
Figure 10. Wind tunnel measurements of the flutter derivatives
H1 , H3 , A2 and A3 for the common bridge deck geometry of
Figure 6 (Mannini, 2006; Mannini et al., 2010b)
for the increased dispersion of the results. However, the reasons
for this large scatter are not yet fully understood and, for most
derivatives, this is not ascribable to the identification procedure;
for instance, it could be easy to show that different values of A
2
actually correspond to signals with appreciable difference in
damping in the pitching mode. The undesirable, but to a certain
extent unavoidable, rolling motion of the section model (rotation
about a horizontal axis parallel to the flow direction) surely
represents a disturbance in the tests. Nevertheless, from analysis
of the experimental data this third degree of freedom does not
seem to be responsible for such a large scatter of the measurements. Other possible causes are wind tunnel free-stream turbulence, interference with the vortex-shedding excitation and the
partial inadequacy of the mathematical model to explain the
physical phenomenon. In particular, the previously mentioned
possible non-linear dependence of the self-excited forces on the
amplitude of vibration could be a source of dispersion, but only
to a limited extent as a big effort was made to keep the initial
condition constant during the tests and a system of electromagnets
152
was employed for this purpose (Mannini, 2006). It is worth noting
that the least uncertain coefficients are H 3 and A
3 , which are
related to the aerodynamic stiffness in the pitching mode.
In order to characterise the flutter derivatives from a statistical
point of view, relatively large samples of measurements (N ¼ 30)
were obtained at three different wind speeds: a low wind speed
(U ¼ 4.0 m/s); a fairly high wind speed far from flutter
(U ¼ 15.0 m/s); a wind speed immediately before the instability
onset (U ¼ 19.2 m/s). The first interesting point is whether the
flutter derivatives are normally distributed or not. The Gaussian
approximation is acceptable in most cases, although some
distributions seem to be affected by a certain skewness. In
particular, the normal hypothesis does not seem completely
adequate for the coefficients H
3 and A3 at low wind speed.
Figure 11 shows some examples of normal probability plots for
the aerodynamic coefficients A2 and A
3 at low wind speed.
Another important issue is the correlation between different
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
0·99
0·98
no significant correlation was observed between H
1 and H 4 and
between H 2 and H 3 for the lift force, and between A1 and A
4
and between A
2 and A3 for the moment, which are proportional
to the in-phase and in-quadrature parts of the forces respectively
due to the heaving and pitching motion. Correlation was found
neither between H
1 and H 3 nor A1 and A3 , despite the well
known interrelations H 1 ¼ KH 3 and A1 ¼ KA3 discussed
by Matsumoto (1996), Scanlan et al. (1997) and Bartoli and
Mannini (2008). Finally, it seems that lift and moment flutter
derivatives are mostly uncorrelated.
Probability
0·95
0·90
0·75
0·50
0·25
0·10
0·05
0·02
0·01
⫺0·054
⫺0·053
⫺0·052
A*2
(a)
⫺0·051
0·99
0·98
0·95
0·90
Probability
0·75
0·50
0·25
0·10
0·05
0·02
0·01
0·071
0·072
0·073
0·074
0·075
A*3
(b)
Figure 11. Cumulative probability distribution for flutter
derivatives A2 and A3 at low wind speed (U ¼ 4.0 m/s) for the
bridge section of Figure 6. Comparison with the corresponding
normal distributions
flutter derivatives. For this purpose, the correlation coefficient R
was calculated for all possible couples of aerodynamic coefficients. The probability p of getting a correlation as large as the
observed value by chance, when the true correlation is zero, was
also determined. At low wind speed all the random variables
associated with the aerodynamic derivatives seem to be uncorrelated. At U ¼ 15.0 m/s, appreciable correlation can be observed
between H 1 and H 2 (R ¼ 0.647, p ¼ 0.0001), H
3 and A3
.
.
(R ¼ 0 389, p ¼ 0 0336) and, above all, A1 and A2
(R ¼ 0.802, p ¼ 9.7 3 108 ). Near flutter (U ¼ 19.2 m/s), there
is a strong correlation between H 1 and H 2 (R ¼ 0.967,
p ¼ 3.1 3 1018 ), H 3 and H 4 (R ¼ 0.939, p ¼ 1.8 3 1014 ),
A1 and A2 (R ¼ 0.968, p ¼ 2.1 3 1018 ) and, to a lesser extent,
between A3 and A4 (R ¼ 0.751, p ¼ 1.7 3 106 ) and A
2 and
A4 (R ¼ 0.425, p ¼ 0.0193). Visual examples of correlation
between flutter derivatives are shown in Figure 12. Interestingly,
3.2 Probabilistic model
The main problem from an engineering point of view is to
understand how the uncertainty in the input of a flutter calculation is transferred into the output (i.e. the critical wind speed).
This is the classical problem of error propagation (see for
instance the clear overview given by Näther (2009)). A deterministic calculation, which employs mean values of the flutter
derivatives (or in some cases only one measured value per
reduced wind speed) and structural dynamic parameters, risks
becoming meaningless if a perturbation of the input is strongly
amplified by the non-linear flutter equations or if the flutter
critical wind speed presents a bimodal distribution (Bartoli and
Mannini, 2005c). In other words, it is extremely important to
estimate the variance of the critical wind speed and not just a
representative deterministic value. Even better is to obtain the
probability distribution of the output, which allows one to
determine quantiles and confidence intervals. In addition, if X is
the vector of input random variables and f is the function that
relates the input and the output, the relation
23:
E[ f (X )] ¼ f (E[X ])
where E[ . ] denotes the expectation operator, is in general not
valid unless f is linear and it can lead to large errors for strongly
non-linear problems (such as the flutter equations).
As discussed in the previous section, the flutter critical wind
speed depends on several dynamic and aerodynamic quantities.
Nevertheless, mass, mass moment of inertia, natural frequencies
and deck dimensions are known with good accuracy and Cheng
et al. (2005) showed that the flutter reliability index is only
slightly sensitive to the uncertainty in these random variables. By
contrast, the uncertainty in the structural damping is quite high,
but this parameter has a non-negligible influence on the flutter
critical wind speed only in the particular case of soft-type flutter.
Consequently, in this work only the flutter derivatives are considered as random variables while all the other parameters are
treated as deterministic quantities. Cheng et al. (2005) showed
that the flutter reliability index is very sensitive to uncertainty in
the flutter derivatives.
Bartoli and Mannini (2005c) considered the values of the flutter
derivatives for two rectangular cylinders measured in smooth and
153
Structures and Buildings
Volume 165 Issue SB3
1·2
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
⫺0·4
R ⫽ ⫺0·967, p ⫽ 3·1 ⫻ 10⫺18
1·0
⫺0·5
0·8
⫺0·6
A*2
H*2
0·6
0·4
⫺0·7
0·2
⫺0·8
0
⫺0·2
⫺13
R ⫽ ⫺0·968, p ⫽ 2·1 ⫻ 10⫺18
⫺12
⫺11
⫺10
⫺9
⫺8
⫺0·9
1·5
⫺7
2·0
2·5
H*1
(a)
2·45
3·0
A*1
(b)
3·5
4·0
4·5
⫺9
⫺8
⫺7
⫺0·4
R ⫽ ⫺0·023, p ⫽ 0·902
⫺0·5
2·40
A*3
A*2
⫺0·6
⫺0·7
2·35
⫺0·8
R ⫽ ⫺0·328, p ⫽ 0·076
2·30
⫺0·9
⫺0·8
⫺0·7
⫺0·6
⫺0·5
⫺0·4
⫺0·9
⫺13
⫺12
⫺11
A*2
⫺10
H*1
(c)
(d)
Figure 12. Examples of correlation between flutter derivatives at
high wind speed (U ¼ 19.2 m/s) for the bridge section of Figure
6. R denotes the correlation coefficient, while p is the probability
of obtaining a correlation as large as the observed value by
chance when the true correlation is zero
turbulent flow (Righi, 2003) as independent normal random
variables. Monte Carlo realisations were generated in correspondence of the reduced wind speeds where measurements were
available. Sets of flutter derivative functions were then obtained
by interpolation. A deterministic flutter calculation was performed for each set of functions. By contrast, Mannini (2006)
and Mannini and Bartoli (2009) tried to maximise the statistical
perturbation of aeroelastic functions by first interpolating the
mean and the variance of the measured coefficients and then
generating Monte Carlo realisations of the flutter derivatives.
Flutter calculations were then performed for the previously
mentioned rectangular cylinders and for the realistic bridge
section discussed hereafter.
In this work a more rigorous probabilistic model for flutter is
proposed, based on the assumption that the mechanical model of
Equations 1–4 applies and the flutter derivatives are independent,
154
normally distributed random variables. As previously discussed,
experimental evidence (confirmed by the results obtained for two
rectangular cylinders in smooth and turbulent flow (Righi, 2003))
suggests that the flutter derivatives can be considered as normally
distributed. More debatable is their statistical independency but
since the correlation seems to be much lower than expected it is
reasonable to waive this complication in this first application of
the model.
The aim is to determine the probability that a given value of the
reduced wind speed UR ¼ 2ð=K is larger than the critical reduced
wind speed, p(UR > URc ); that is, flutter occurs at that reduced
wind speed. From a theoretical point of view, it could be possible
to solve the probabilistic flutter problem by considering the flutter
equations and calculating the probability that the solution of the
so-called imaginary equation (real values of Y ¼ ø=øh , such that
1 < Y < ªø , for which the imaginary part of the determinant of
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
the system of equations obtained imposing coupled harmonic
oscillations vanishes) is larger than (or equal to) that of the real
equation (values of Y for which the real part of the determinant
vanishes); this corresponds to an unstable condition (Mannini,
2006) if the common form of fourth- and third-order equations (as
reported for instance in Dyrbye and Hansen (1997)) is used (see
Figure 13). Otherwise, if the form of Equations 10 and 11 is
employed, instability is reached when the solution of the imaginary equation (Equation 11) is smaller than that of the real equation
(Equation 10). This probability could be calculated through the
formulas for probability density functions (PDFs) of the product,
the sum and the square root of continuous random variables.
Nevertheless, in the general case of 2-dof flutter, this analytical
procedure becomes too complicated and burdensome. It is applicable only in the case of 1-dof torsional flutter or if an approximate
model such as that of Equation 17 or Equation 21 is considered.
Based on Equation 21, the probability of the flutter condition
would read:
speed is characterised by the same probability function as the
reduced critical wind speed, the probability distribution and the
statistical moments of the former can be simply calculated
through the relation U ¼ Bøc UR =2ð, where the critical frequency øc is deterministically calculated.
24:
p(UR > URc ) ¼ p A
2 (U R ) >
2Ig Æ
rB4
25:
Much more practical for engineering applications is the use of
Monte Carlo strategies. At each reduced wind speed a set of
flutter derivatives is generated from the known PDFs. For every
realisation, the flutter condition is verified by comparing the
solutions of the real and imaginary equations and the probability
of encountering flutter is obtained. By spanning all the reduced
wind speeds in the range of interest, the cumulative probability
distribution function (CDF) p(URc < UR ) can be determined.
Furthermore, if one assumes that the dimensional critical wind
1·9
1·8
1·7
Y
Finally, it is worth noting that the CDF of the flutter critical wind
speed p(Uc < U ) expresses the probability of encountering
flutter given a certain wind speed. Therefore, such a function is
suitable to be used in the framework of the performance-based
design, accounting for the bridge collapse limit state due to
flutter. In fact, considering the uncertainty in hazard, vulnerability
and damage, a bridge structure can be conceived according to the
Pacific Earthquake Research Centre (PEER) equation (Ciampoli
et al., 2009; Porter, 2003). In particular, once a limit state related
to the structural response has been assumed as a decision
variable, this equation simplifies to (Mannini, 2006):
Pfail p(EDP) ¼
ð
p(EDP j IM) g(IM) d(IM)
where Pfail is the probability of failure, EDP is an engineering
demand parameter representing the structural response with
which the considered limit state is quantified, IM is the intensity
measure of the hazard, p( . ) denotes the probability of exceedance, p( . | . ) is a conditional probability of exceedance and
g(IM) is the probability density function of the hazard. The
integral is evaluated over the entire space of possible intensity
measures. Considering the ultimate limit state for flutter, assuming a probability of 1 for the collapse of the bridge structure
under flutter vibrations and taking the mean wind speed as the
intensity measure, the structural vulnerability term p(EDP|IM)
becomes the probability of encountering flutter given a certain
wind speed. Therefore p(EDP|IM) is the CDF of the flutter
critical wind speed. As a consequence, determining the probability function of the flutter critical wind speed is important not only
in order to estimate the variance of the results of a flutter
calculation but also to account for this aeroelastic instability
phenomenon in a risk analysis.
1·6
1·5
Real equation
Imaginary equation
1·4
1·3
0
2
4
6
8
10
UR
Figure 13. Deterministic solution of the real and imaginary flutter
equations. Y ¼ ø=ø h is the non-dimensional flutter frequency
and ªø ¼ øÆ =ø h is the still-air frequency ratio. Real and imaginary
equations are those reported in several texts (e.g. Dyrbye and
Hansen, 1997). ªø ¼ 1:948
3.3 Example of application
As an example of application of the outlined probabilistic flutter
approach, the wind tunnel section model depicted in Figure 6 is
taken as a case study (Mannini, 2006). In order to give a statistical
description of the flutter derivatives, measurements were repeated
ten times for each reduced wind speed while trying to keep the test
conditions unchanged. Four out of eight aerodynamic coefficients
are shown in Figure 10, even though the complete set of eight
flutter derivatives was used in both deterministic and probabilistic
calculations. Mean and standard deviation values of the flutter
derivatives are interpolated and a very limited extrapolation up to
UR ¼ 10 was necessary. The basic parameters of the mechanical
system are: B ¼ 0.45 m, m ¼ 5.45 kg/m, I ¼ 0.095 kg m2 /m,
ø h ¼ 19.20 rad/s, øÆ ¼ 37.40 rad/s, æ h ¼ 0.24% and æÆ ¼ 0.18%.
155
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
1·0
The graphical deterministic solution of the flutter equations,
obtained by considering the mean values of the flutter derivatives,
is shown in Figure 13. Table 7 shows that the calculated critical
value agrees very well with the experimentally measured value.
Calculated
Normal
0·8
0·7
0·6
CDF
The CDF of the critical wind speed was obtained through Monte
Carlo simulation with 500 000 realisations for each reduced wind
speed. The reasonable convergence of the results with respect to
the number of realisations was verified. The range of reduced
wind speed spanned was from 0 to 10 in increments of 0.02. The
PDF was obtained from the CDF by central differencing. Both
functions are reported in Figure 14. Comparison with the
corresponding normal distributions shows that the result is not
perfectly Gaussian, with a small positive skewness and a kurtosis
smaller than 3.0. Moreover, it is possible to remark that the
probability of encountering flutter is zero for reduced wind
speeds lower than 8.5, whilst it is 1.0 for reduced wind speeds
larger than 10.
0·9
0·5
0·4
0·3
0·2
0·1
0
8·0
8·5
9·0
9·5
10·0
9·5
10·0
UR
(a)
1·8
1·6
It is also worth noting that, due to the non-linearity of the flutter
equations, a large dispersion characterising several flutter derivatives is transformed in this case into a small dispersion of the
critical wind speed. In fact, in the range of interest of reduced
wind speed, coefficients of variation up to about 8% were
observed for H 1 and A1 , up to 3–4% for H 3 and A3 , up to
30% for A2 and even beyond 40, 50 and 100% respectively for
H
2 , A4 and H 4 : For the same reason, despite the normal
1·4
1·2
PDF
Table 7 lists the first four statistical central moments. It is evident
that the mean value of the reduced wind speed is in this case very
close to the one calculated in a deterministic way and to the wind
tunnel result. Nevertheless, the probabilistic approach also reveals
that the coefficient of variation is small (less than 3%); that is,
large deviations of the solution from the mean are not likely,
similarly to the previous predictions of error propagation (Bartoli
and Mannini, 2005c; Mannini and Bartoli, 2009). In these
conditions, the mean value of the distribution or the result of a
deterministic calculation is a reliable estimate of the flutter
critical wind speed, although the 95% confidence interval for the
reduced critical wind speed is (8.86, 9.86), the amplitude of
which is not negligible for engineering purposes.
1·0
0·8
0·6
0·4
0·2
0
8·0
8·5
9·0
UR
(b)
Figure 14. Cumulative probability distribution function (CDF) and
probability density function (PDF) of the reduced critical wind
speed for the case study considered. The corresponding normal
distributions are also reported for comparison
distribution of the input, the output of the calculation is not
perfectly Gaussian.
Experimental
9.31
Deterministic
9.42
Probabilistic
ì
CoV: %
9.36
2.69
skw
krt
0.22 2.64
Table 7. Comparison between the reduced critical wind speed
measured in the wind tunnel for the test case considered, the
result of the deterministic flutter calculation and those of the
probabilistic approach in terms of mean value ( ì), coefficient of
variation (CoV), skewness (skw) and kurtosis (krt)
156
Finally, it must be noted that the well-behaving results (mean
value of the critical wind speed close to the deterministic value,
small variance, mono-modal distribution, etc.) observed for the
particular case study considered can in no way be extended to
different test cases.
4.
Conclusion
This paper has shown that the problem of flutter stability can be
analytically simplified by manipulating the flutter equations on
the basis of experimental evidence. The procedure leads to two
approximate equations through which the critical reduced wind
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
speed and the flutter frequency can be calculated with just three
flutter derivatives instead of the usual eight coefficients. These
formulas give accurate results provided that the frequency separation of the critical modes is not too small. They represent a
considerable simplification and allow a deeper understanding of
the flutter mechanism and consequent better tailoring of a bridge
structure at early design stages. In particular, the proposed
formulas help to explain the role played by damping and other
structural parameters in the onset of instability and to give an
explanation for soft- and hard-type flutter.
Document. See http://www.aniv-iawe.org/barc for further
details (accessed 27/08/2011).
Bartoli G, Contri S, Mannini C and Righi M (2009) Towards an
improvement in the identification of bridge deck flutter
derivatives. ASCE Journal of Engineering Mechanics 135(8):
771–785.
Bisplinghoff RL, Ashley H and Halfman RL (1996) Aeroelasticity,
1st edn. Dover Publications, New York.
Caracoglia L and Jones NP (2003) Time domain vs. frequency
domain characterisation of aeroelastic forces for bridge deck
sections. Journal of Wind Engineering & Industrial
Aerodynamics 91(3): 371–402.
Chen X (2007) Improved understanding of bimodal coupled
bridge flutter based on closed-form solutions. ASCE Journal
of Structural Engineering 133(1): 22–31.
Chen X and Kareem A (2003) Efficacy of tuned mass dampers
for bridge flutter control. ASCE Journal of Structural
Engineering 129(10): 1291–1300.
Chen X and Kareem A (2006) Revisiting multimode coupled
bridge flutter: some new insights. ASCE Journal of
Engineering Mechanics 132(10): 1115–1123.
Chen X, Matsumoto M and Kareem A (2000) Aerodynamic
coupling effects on flutter and buffeting of bridges. ASCE
Journal of Engineering Mechanics 126(1): 17–26.
Cheng J, Cai CS, Xiao RC and Chen SR (2005) Flutter reliability
analysis of suspension bridges. Journal of Wind Engineering
& Industrial Aerodynamics 93(10): 757–775.
Chowdhuri AG and Sarkar PP (2004) Identification of eighteen
flutter derivatives of an airfoil and a bridge deck. Wind and
Structures 7(3): 187–202.
Ciampoli M, Petrini F and Augusti G (2009) A procedure for
performance-based wind engineering. Proceedings of 10th
International Conference on Structural Safety and Reliability,
Osaka. Taylor & Francis, London, pp. 1843–1850.
Costa C and Borri C (2006) Application of indicial functions in
bridge deck aeroelasticity. Journal of Wind Engineering &
Industrial Aerodynamics 94(11): 859–881.
Diana G, Resta F, Zasso A, Belloli M and Rocchi D (2004) Forced
motion and free motion aeroelastic tests on a new concept
dynamometric section model of the Messina suspension
bridge. Journal of Wind Engineering & Industrial
Aerodynamics 92(6): 441–462.
Diana G, Resta F and Rocchi D (2008) A new numerical approach
to reproduce bridge aerodynamic non-linearities in time
domain. Journal of Wind Engineering & Industrial
Aerodynamics 96(10–11): 1871–1884.
Dyrbye C and Hansen SO (1997) Wind Loads on Structures, 1st
edn. Wiley, New York.
Frandsen AG (1966) Wind stability of suspension bridges.
Application of the theory of ‘thin airfoils’. Proceedings of
International Symposium on Suspension Bridges, Lisbon.
National Laboratory for Civil Engineering, Lisbon, pp. 609–
627.
Fung YC (1993) An Introduction to the Theory of Aeroelasticity.
Dover Publications, New York.
The problem of the uncertainty observed in wind tunnel measurements of flutter derivatives (especially in the case of freevibration setups) was discussed. A model to account for the nondeterministic nature of these coefficients in the calculation of the
flutter critical wind speed was presented. The main purpose of
this model is estimation of the variance of the calculated critical
wind speed, which is absolutely unknown by performing the usual
deterministic flutter analysis. An application in the case of a
bridge deck section model of common geometry was outlined,
employing wind tunnel data measured ad hoc for this purpose. As
a future development of this work it will be interesting to
consider the actual correlation observed between a few flutter
derivatives. Finally, it is worth remarking that the approximate
formulas for the critical wind speed can also be fruitfully
employed in order to solve analytically the problem of probabilistic flutter assessment.
Acknowledgements
This research work was partially performed in the framework of
the Italian National Research Contracts AER-BRIDGE (PRIN
2006) and Wi-POD (PRIN 2007), 2-year grants from the Italian
Ministry of Education, University and Research (MIUR).
REFERENCES
Bartoli G and Mannini C (2005a) From multimodal to bimodal
approach to flutter. Proceedings of 6th European Conference
on Structural Dynamics, Paris. Millpress, Rotterdam, pp.
349–354.
Bartoli G and Mannini C (2005b) A simplified method to evaluate
the critical flutter speed for flexible bridge decks.
Proceedings of 4th European and African Conference on
Wind Engineering, Prague. ITAM ASCR, Prague.
Bartoli G and Mannini C (2005c) Reliability of bridge deck
flutter derivative measurements in wind tunnel tests.
Proceedings of 9th International Conference on Structural
Safety and Reliability, Rome. Millpress, Rotterdam, pp.
1193–1200.
Bartoli G and Mannini C (2008) A simplified approach to bridge
deck flutter. Journal of Wind Engineering & Industrial
Aerodynamics 96(2): 229–256.
Bartoli G and Righi M (2006) Flutter mechanism for rectangular
prisms in smooth and turbulent flow. Journal of Wind
Engineering & Industrial Aerodynamics 94(5): 275–291.
Bartoli G, Bruno L, Buresti G et al. (2008) BARC Overview
157
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
Ge YJ, Xiang HF and Tanaka H (2000) Application of a reliability
The complex branch characteristics of coupled flutter. Journal
of Wind Engineering & Industrial Aerodynamics 96(10–11):
1843–1855.
Nakamura Y (1978) An analysis of binary flutter of bridge deck
sections. Journal of Sound and Vibration 57(4): 471–482.
Nakamura Y (1979) On the aerodynamic mechanism of torsional
flutter of bluff structures. Journal of Sound and Vibration
67(2): 163–177.
Nakamura Y and Mizota T (1975) Torsional flutter of rectangular
prisms. Journal of the Engineering Mechanics Division ASCE
101(2): 125–142.
Näther W (2009) Copulas and T-norms: Mathematical tools for
modeling propagation of errors and interactions. Proceedings
of 10th International Conference on Structural Safety and
Reliability, Osaka. Taylor & Francis, London, pp. 1238–
1245.
analysis model to bridge flutter under extreme winds. Journal
of Wind Engineering & Industrial Aerodynamics 86(2–3):
155–167.
Jain A, Jones NP and Scanlan RH (1996) Coupled flutter and
buffeting analysis of long-span bridges. ASCE Journal of
Structural Engineering 122(7): 716–725.
Katsuchi H, Jones NP and Scanlan RH (1999) Multimode coupled
flutter and buffeting analysis of the Akashi Kaikyo Bridge.
ASCE Journal of Structural Engineering 125(1): 60–70.
Lin YK (1996) Stochastic stability of wind-excited long-span
bridges. Probabilistic Engineering Mechanics 11(4): 257–
261.
Mannini C (2006) Flutter Vulnerability Assessment of Flexible
Bridges. PhD thesis, University of Florence–TU
Braunschweig.
Mannini C and Bartoli G (2008a) Investigation on the dependence
of bridge deck flutter derivatives on steady angle of attack.
Proceedings of 5th International Colloquium on Bluff Bodies
Aerodynamics and Applications, Milan. Politecnico di
Milano, Milan.
Mannini C and Bartoli G (2008b) Analisi del meccanismo del
flutter per impalcati da ponte: sensibilità rispetto ai parametri
dinamici della struttura. Proceedings of 10th Italian National
Conference on Wind Engineering, Cefalù, Italian National
Association of Wind Engineering, Rome (in Italian).
Mannini C and Bartoli G (2009) The problem of uncertainty in
the measurement of aerodynamic derivatives. Proceedings of
10th International Conference on Structural Safety and
Reliability, Osaka. Taylor & Francis, London, pp. 824–831.
Mannini C, Bartoli G, Flamand O and Grillaud G (2007) A
simplified method of flutter assessment: first steps towards
flutter derivative generalization. Proceedings of 12th
International Conference on Wind Engineering, Cairns,
Australasian Wind Engineering Society, pp. 2343–2350.
Mannini C, Ferrucci M and Bartoli G (2010a) Effetti non lineari
nelle derivate di flutter di un impalcato da ponte. Proceedings
of 11th Italian National Conference on Wind Engineering,
Spoleto, Morlacci, Perugia (in Italian).
Mannini C, Šoda A, Voâ R and Schewe G (2010b) Unsteady
RANS simulation of flow around a bridge section. Journal of
Wind Engineering & Industrial Aerodynamics 98(12): 742–
753.
Matsumoto M (1996) Aerodynamic damping of prisms. Journal
of Wind Engineering & Industrial Aerodynamics 59(2–3):
159–175.
Matsumoto M, Yoshizumi F, Yabutani T, Abe K and Nakajima N
(1999) Flutter stabilization and heaving-branch flutter.
Journal of Wind Engineering & Industrial Aerodynamics
83(1–3): 289–299.
Matsumoto M, Taniwaki Y and Shijo R (2002) Frequency
characteristics in various flutter instabilities of bridge girders.
Journal of Wind Engineering & Industrial Aerodynamics
90(12–15): 1973–1980.
Matsumoto M, Okubo K, Ito Y, Matsumiya H and Kim G (2008)
158
Noda M, Utsunomiya H, Nagao F, Kanda M and Shiraishi N
(2003) Effects of oscillation amplitude on aerodynamics
derivatives. Journal of Wind Engineering & Industrial
Aerodynamics 91(1–2): 101–111.
Øiseth O, Röonquist A and Sigbjörnsson R (2010)
Simplified prediction of wind-induced response and
stability limit of slender long-span suspension bridges, based
on modified quasi-steady theory: a case study. Journal of
Wind Engineering & Industrial Aerodynamics 98(12): 730–
741.
Ostenfeld-Rosenthal P, Madsen HO and Larsen A (1992)
Probabilistic flutter criteria for long span bridges. Journal of
Wind Engineering & Industrial Aerodynamics 42(1–3):
1265–1276.
Porter KA (2003) An overview of PEER’s performance-based
engineering methodology. Proceedings of 9th International
Conference on Applications of Statistics and Probability in
Civil Engineering, San Francisco. IOS Press, Amsterdam, pp.
973–980.
Pourzeynali S and Datta TK (2002) Reliability analysis of
suspension bridges against flutter. Journal of Sound and
Vibration 254(1): 143–162.
Righi M (2003) Aeroelastic Stability of Long Span Suspended
Bridges: Flutter Mechanism on Rectangular Cylinders in
Smooth and Turbulent Flow. PhD thesis, University of
Florence.
Scanlan RH (1978a) The action of flexible bridges under wind, I:
Flutter theory. Journal of Sound and Vibration 60(2): 187–
199.
Scanlan RH (1978b) The action of flexible bridges under wind, II:
Buffeting theory. Journal of Sound and Vibration 60(2): 201–
211.
Scanlan RH and Tomko JJ (1971) Airfoil and bridge deck flutter
derivatives. Journal of the Engineering Mechanics Division
ASCE 97(6): 1717–1737.
Scanlan RH, Jones NP and Singh L (1997) Inter-relations among
flutter derivatives. Journal of Wind Engineering & Industrial
Aerodynamics 69–71: 829–837.
Selberg A (1961) Oscillation and aerodynamic stability of
Structures and Buildings
Volume 165 Issue SB3
New developments in bridge flutter
analysis
Mannini, Bartoli and Borri
suspension bridges. Acta Polytechnica Scandinavica: Civil
Engineering and Building Construction Series, 13(308):
1–69.
Simiu E and Scanlan RH (1996) Wind Effects on Structures:
Fundamentals and Applications to Design, 3rd edn. Wiley,
New York.
Singh L, Jones NP, Scanlan RH and Lorendeaux O (1995)
Simultaneous identification of 3-dof aeroelastic parameters.
Proceedings of 9th International Conference on Wind
Engineering, New Delhi. Wiley Eastern, New Delhi, pp.
972–981.
Theodorsen T (1934) General Theory of Aerodynamic Instability
and the Mechanism of Flutter. NACA, Langley Research
Center, Hampton, Technical report 496, Annual report 20.
Vairo G (2010) A simple analytical approach to the aeroelastic
stability problem of long-span cable-stayed bridges.
International Journal for Computational Methods in
Engineering Science & Mechanics 11(1): 1–19.
WHAT DO YOU THINK?
To discuss this paper, please email up to 500 words to the
editor at
[email protected]. Your contribution will be
forwarded to the author(s) for a reply and, if considered
appropriate by the editorial panel, will be published as a
discussion in a future issue of the journal.
Proceedings journals rely entirely on contributions sent in
by civil engineering professionals, academics and students.
Papers should be 2000–5000 words long (briefing papers
should be 1000–2000 words long), with adequate illustrations and references. You can submit your paper online via
www.icevirtuallibrary.com/content/journals, where you
will also find detailed author guidelines.
159
Copyright of Structures & Buildings is the property of Thomas Telford Ltd and its content may not be copied or
emailed to multiple sites or posted to a listserv without the copyright holder's express written permission.
However, users may print, download, or email articles for individual use.