Sunteți pe pagina 1din 163

Wind-Induced Dynamic Instabilities

of Flexible Bridges
Nikolaos Nikitas
Department of Civil Engineering
University of Bristol
A dissertation submitted to the University of Bristol in
accordance with the requirements for the degree of
Doctor of Philosophy in the Faculty of Engineering.
April 2011
Abstract
The wind-induced vibrations of exible bridges and their components have long been
a major concern. Although a great level of sophistication has been reached in wind-
resistant design, there is still a signicant threat from the wind. Most intriguingly, often
when one problem is solved a new one seems to arise in its place. This study examines
a selection of such peculiar aerodynamic issues involved in the routine life of a bridge.
All of them share in common the need for further explanations to address previous
modelling omissions and weaknesses and oer new understanding of the underlying
phenomena.
Starting with utter, an inverse scheme was employed to identify the utter deriva-
tive description of aeroelastic loading using actual response measurements of a full-scale
suspension bridge. As expected for ambient data, the identication produced noisy es-
timates of the parameter values, yet clear trends could be distinguished. Encouragingly
the trends were in reasonably good agreement with results from wind tunnel tests on
similar cross-sections. Evidence of aeroelastic coupling between vertical and torsional
vibrations was identied from the recorded bridge data and the utter wind speed for
the single-degree-of-freedom torsional instability case was estimated. It is believed this
is the rst time this has been achieved based solely on full-scale data. The study shows
the viability of the method to identify the utter derivatives from full-scale data, which
has rarely been attempted previously and never with such clarity of the results. This is
potentially useful for identifying the actual aeroelastic behaviour and safety of bridges
in service, particularly as uncertainties of wind tunnel tests, such as Reynolds number
dependence are overcome.
Next, quasi-steady galloping theory was revisited aiming to address previous incon-
sistencies and clarify the correct generic equations and their implications. The case of a
section free to vibrate in two orthogonal directions was considered, subject to ow at an
arbitrary angle of attack to the principal structural axes. Putting forward the correct
galloping criterion for this non-classical case, it was possible to quantify dierences from
previous incorrect analysis of the galloping condition and structural damping demand.
The eects of the structural parameters on galloping thresholds were addressed, again
overcoming former shortcomings.
Finally there was an attempt to elucidate the inuence of critical Reynolds number
on the apparent galloping instability of dry circular cables. This instability per se
cannot fully t the earlier classical galloping description. Through a series of wind
tunnel tests ow conditions responsible for excessive motion were identied. These
conditions were more perplexing than the single-sided laminar separation bubble and
2
the associated steady lift that is generally believed to dominate the critical Reynolds
number range. For cylinders normal to the ow, discontinuities in the aerodynamic
loading appear to act like a quenching intermittency, eectively inhibiting response. For
other cable inclinations the unusual ow structures that emerge seem to be related to
the observed dynamic instability of the cable. According to this thesis, the transitional
behaviour of the boundary layer is entirely responsible for the large amplitude vibration
events of bridge cables in dry conditions attributed to so-called dry galloping.
Acknowledgements
I owe a great debt of gratitude to my supervisor John Macdonald. His support, his
trust in me, his advice and his critical views have been more than inspiring for this
work. I am sure that our endless, always unscheduled, discussions can be heard on
every page of this thesis. He introduced me to the world of bridge monitoring and he
made me piece of the legacy of the Clifton Suspension Bridge (CSB). Without him this
thesis would have never been this thesis.
I would also like to thank Jasna Jakobsen for all background information and support
provided when performing the identication analysis for the CSB. It was a pleasure to
work with her and learn from her.
Further all the team that made possible the wind tunnel tests at the National Research
Council (NRC) of Canada deserve a special thank. Terje Andersen, Mike Savage, Brian
McAulie, Guy Larose, along with the technical sta at NRC have been more than
helpful partners.
I gratefully acknowledge the support of The Clifton Suspension Bridge Trust during the
CSB site tests and the nancial support from EPSRC during my PhD course (under
John Macdonalds Advanced Research Fellowship).
Last but not least, I want to thank my family in Drama for their understanding and love
throughout the period of my research at Bristol. Especially my father, a real modern
Ulysses in my mind, has a unique contribution to my work for he has never stopped
advising me on how to defeat my strongest enemy; myself.
From being to becoming
Authors Declaration
I declare that the work in this dissertation was carried out in accordance with the
regulations of the University of Bristol. The work is original except where indicated by
special reference in the text and no part of the dissertation has been submitted for any
other degree.
Any views expressed in the dissertation are those of the author and in no way
represent those of the University of Bristol.
The dissertation has not been presented to any other University for examination
either in the United Kingdom or overseas.
Signed:
Dated:
Contents
1 Introduction 1
2 The Aeroelasticity Framework 5
2.1 Wind-induced structural loading . . . . . . . . . . . . . . . . . . . . . . 5
2.1.1 Static loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.2 Wind bueting . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.3 Vortex shedding . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.4 Galloping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.5 Flutter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.6 Wake-induced loading . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.7 Rain-wind Instabilities . . . . . . . . . . . . . . . . . . . . . . . . 22
2.2 Circular galloping: myth or true? . . . . . . . . . . . . . . . . . . . . . . 25
2.2.1 Reynolds number eects . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.2 Inclination eects . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2.3 Instability mechanisms . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3 State of the art in bridge wind design . . . . . . . . . . . . . . . . . . . 35
2.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3 Identication of utter derivatives from full-scale data 43
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2 The case study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.3 Wind characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4 Response and modal parameters . . . . . . . . . . . . . . . . . . . . . . 50
3.4.1 Response Characteristics . . . . . . . . . . . . . . . . . . . . . . 50
3.4.2 Modal Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5 Flutter derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.5.1 Flutter Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.5.2 Identication Method . . . . . . . . . . . . . . . . . . . . . . . . 55
3.5.3 Application to the Clifton Suspension Bridge . . . . . . . . . . . 57
3.6 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
i
ii CONTENTS
4 Quasi-steady galloping analysis revisited 67
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.2 Quasi-steady derivations . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.2.1 Relevance to uniform continuous systems . . . . . . . . . . . . . 74
4.3 Application: quantifying dierences . . . . . . . . . . . . . . . . . . . . . 75
4.4 The detuning eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.5 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5 Experiments on galloping vibration of a circular cylinder 89
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.2 Wind tunnel tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.2.1 Preliminary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.2.2 Setup details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.3.1 Overview and large responses . . . . . . . . . . . . . . . . . . . . 99
5.3.2 Pressure data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.4.1 Symmetry considerations . . . . . . . . . . . . . . . . . . . . . . 106
5.4.2 Mechanism implications . . . . . . . . . . . . . . . . . . . . . . . 111
5.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6 Conclusion and outlook 115
Publications 121
References 123
List of Tables
5.1 Orientation angles for studied cases. . . . . . . . . . . . . . . . . . . . . 93
5.2 Position details for the model. For rings and lowest cable end distance
from oor refers to stagnation points, while for cobra probes distance
from model refers to along-wind distance. . . . . . . . . . . . . . . . . . 98
iii
List of Figures
2.1 Mapping of Strouhal number against Reynolds number in the subcritical
range. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Free vibration tests for a circular cylinder. . . . . . . . . . . . . . . . . . 10
2.3 Circular cylinder vibration phenomena past the lock-in range U=5m/s.
Top: vertical response record with f
c
=9Hz. Bottom: surface pressure at
transverse tap, f
v
=13Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Storeblt Bridge vortex-induced vertical motion at max (left) and min
(right) of the amplitude cycle. Encircled is a parked van with its view
distorted due to motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.5 Typical galloping response curve for a rectangular prism. Inset the clas-
sical galloping mechanism illustrated. . . . . . . . . . . . . . . . . . . . . 15
2.6 Response of a rectangular prism with side ratio 2/1 against reduced
velocity for varying critical structural damping ratio. The prism has
Sr=0.081 that sets the relevant U
r
for vortex resonance at 12.34. . . . 16
2.7 (a) Displacements and aeroelastic forces on a thin airfoil; (b) Displace-
ments and aeroelastic forces for a bridge section. . . . . . . . . . . . . . 17
2.8 (a) Quantitative dierence of response characteristics for a full bridge
under dierent ow conditions. (b) A

2
from wind tunnel tests for a tor-
sionally unstable bridge section under laminar and turbulent ow condi-
tions. Changes appear minimal to sustain any substantial modication
in the utter behaviour. . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.9 (a) Rivulet formation on the circular cable section; (b) Inclination ge-
ometry of the inclined and yawed to the ow cable. . . . . . . . . . . . . 22
2.10 Upper water rivulet mean angular position during rain-wind vibrations
and lift force, displacement time series for a dierent large response
conguration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.11 Resume of wind tunnel test results on the proposed cables in the Higashi-
Kobe Bridge. Improvement in aerodynamic performance is apparent for
the solution with protuberances under all conditions. . . . . . . . . . . . 25
v
vi LIST OF FIGURES
2.12 Mean drag coecient versus Reynolds number. On top, transitions (Tr)
from laminar (L) to turbulent (T) ow are presented in relation to sep-
aration points (S) and boundary layers (BL). . . . . . . . . . . . . . . . 28
2.13 (a) The axial ow, evidenced by light ags positioned inside the wake,
act towards inhibiting communication between shear layers and promot-
ing a secondary circulatory ow. The function described, simulates the
galloping of a circular cylinder equipped with a long splitter plate. (b)
Enhanced vortices are produced when axial vortices from the inclined
cable, mix and interact with ordinary K arm an vortices. . . . . . . . . . 31
2.14 Dry cable instability design criteria together with real-bridge unstable
records. Dotted lines are due to the uncertainty in dening structural
damping values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1 Clifton Suspension Bridge (CSB) elevation showing monitoring instru-
ment locations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2 Sketch of the CSB cross-section . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Vertical and torsional modes of CSB . . . . . . . . . . . . . . . . . . . . 47
3.4 (a) Histogram of wind speeds during the 2003-04 recording period. (b)
Polar plots of 1h mean wind velocities from both anemometers. . . . . . 49
3.5 (a) 1h average wind speed over the monitoring period. (b) 1h RMS
vertical accelerations at the reference location over the monitoring period. 51
3.6 (a) RMS vertical accelerations
v
, in relation to wind speed for all 1h
records. (b) Same as (a) for 1h records dominated by wind loading,
with RMS vertical accelerations now divided by the vertical turbulence
intensity. The power-law approximating the obtained trend is also plotted. 52
3.7 PSDs for dierent loading conditions for (a) vertical (b) torsional and
(c) lateral accelerations. . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.8 (a) PSDs of ltered data for rst vertical and torsional modes for the
maximum wind speed record. (b) The evolution of the coupling action
is evident in the vertical PSD for wind speeds above 11m/s. . . . . . . . 53
3.9 Decolouring process. In the 1DOF system, lter application will produce
corrected spectra, see dashed line, simulating white noise loading. For
the 2DOF system, ltering with F
L
will erroneously modify the true
aeroelastic coupling, see dashed line versus greyed area. . . . . . . . . . 57
3.10 Example covariance functions (scaled with variance) for the combined
two degrees-of-freedom plotted against time lag. . . . . . . . . . . . . . . 58
3.11 Flutter derivatives of Clifton Suspension Bridge from full-scale data,
compared with wind tunnel extracted utter derivatives for various cross-
sections. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
LIST OF FIGURES vii
3.12 Identied H

1
utter derivative from each 15-minute record. . . . . . . . 61
3.13 Flutter derivative A

2
with additional points from A

4
. . . . . . . . . . . . 63
3.14 H

1
for dierent H-sections and a possible Reynolds number based ex-
planation for the H

1
inversion. Crossovers can initiate when Reynolds
number changes alter the multiplier of U
r
in Eq.(3.6). . . . . . . . . . . 64
4.1 Geometry of a blu section indicating lift and drag forces (L, D), relative
angle of attack () and principal structural axes (x, y). (a) The general
case with
0
= 0 and the 2DOF motion potential. (b) The special case
for 1DOF across-wind oscillations. . . . . . . . . . . . . . . . . . . . . . 70
4.2 Sections used in the galloping analysis. . . . . . . . . . . . . . . . . . . . 76
4.3 Non-dimensional aerodynamic damping coecients (S
DH
, min(S
xx
, S
yy
),
S
2D
) as a function of angle of attack. Negative values indicate unstable
behaviour (in the absence of structural damping). . . . . . . . . . . . . . 78
4.3 (continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.4 Comparison between the erroneous S
sc
and the correct value for the 1D
rotated y-axis case, S
yy
, for (a) the square in Fig.4.2(d) and (b) the
triangle in Fig.4.2(f). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.5 Evolution of the aerodynamic damping solution for dierent values of
detuning, , for (a) the section in Fig.4.2(j) and = 123

and (b) the


section in Fig.4.2(k) and = 30

. The lower branch is the important


one. In (a) for perfect tuning the solution is unstable (negative aerody-
namic damping) and for detuning above about 7% it approaches the 1D
solution, which in this case is stable. In (b), on the other hand, for per-
fect tuning the solution is stable and for detuning above 1% it becomes
unstable while moving towards the 1D solution. . . . . . . . . . . . . . . 83
4.6 Modal trajectories corresponding to Fig.4.5(a) for (a) = 1, (b) =
1.005, (c) = 1.05 and (d) = 1.1. The applicable S
detuned
value is also
indicated. Unstable modes are plotted with solid lines while stable ones
are dotted. Note for comparison that S
xx
= 0.45, S
yy
= 0.06. For all
plots the structural damping value was c = 0. . . . . . . . . . . . . . . . 85
5.1 Transformation from real cable to wind tunnel model. . . . . . . . . . . 92
5.2 View of the NRC wind tunnel facility and its test section with the model
in place. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.3 Turbulence intensity and mean velocity across the wind tunnel section. . 95
5.4 Elevation of cable model showing instrumentation arrangement. . . . . . 97
5.5 Typical frequency response curves for three pressure taps. . . . . . . . . 98
viii LIST OF FIGURES
5.6 Three examples of motion traces during records of instability; Setup 2A
and 2C. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.7 Proportion of total variance from 20 POMs (from all pressure tap data)
against Reynolds number . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.8 First Proper Orthogonal modeshapes for a set of dynamic and static tests103
5.9 Spectra of the lift coecient on setup 2A, averaged over all four pressure
rings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.10 Correlation functions of lift coecients between pressure rings. . . . . . 105
5.11 Mean pressure coecient distribution around cylinder for a large re-
sponse case. Model setup 2A. . . . . . . . . . . . . . . . . . . . . . . . . 107
5.12 Drag evolution of the half-section with Reynolds number. Model setup
=60

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.13 Drag evolution of the half-section with Reynolds number. Model setup
=90

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.14 Ring 1 C
L
and C
D
transitional avalanche-like behaviour. Setup =90

. 110
5.15 Pressure distributions during a transitional state succession. Model
setup =90

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Nomenclature
The precise interpretation of notation and abbreviations must be obtained from the
local context in which it is used and in which it will be explained. As a further guide,
the following is a non-exhaustive list of commonly used terms
Notation
A

i
, H

i
, P

i
Flutter derivatives (i=1-6)
a
1
, , b
1
Wake oscillator tted parameters
B Representative sectional dimension (chord length)
b Half chord length
C(k) Theodorsen circulation function
C
D,L,M
Static force coecients for D, L, M
C
D
1/2
Static drag coecient from half the cylinder section
C
ij
Covariance estimate between i, j series
C
p
Static pressure coecient
c Structural damping
D Static drag force
D
b
Bueting drag force
D
se
Self-excited drag force
d Across-wind dimension
F
L
(f) Filter function for lift force
F
x,y
Force along direction x or y
f Frequency
f
c
Structural frequency
f
v
Vortex shedding frequency
G
ij
Modal integral of modes i, j
H
j
(f) Frequency Response Function of mode j
h(s, t), p(s, t), (s, t) Vertical, lateral and torsional displacement
h
i
(s), p
i
(s),
i
(s) Vertical, lateral and torsional i
th
mode shape
I
i
Turbulence intensity along component i
ix
x NOMENCLATURE
I
i
Generalised inertia of mode i
Imaginary unit
|J
j
(f)|
2
Joint acceptance function of mode j
K Reduced cyclic frequency based on the full chord length
k Reduced cyclic frequency based on the half chord length
L Static lift force
L
se
Self-excited lift force
L
b
Bueting lift force
x,y,z
L
i
Turbulence length scale parameters
l Characteristic length
Maximum lags for covariance estimates
M Static overturning moment
M
b
Bueting overturning moment
M
se
Self-excited overturning moment
m Mass
N Number of samples
n Number of lags in correlation function
Q
i
Generalised aerodynamic loading of mode i
q
i
, q
i
, q
i
, Generalised displacement, velocity and acceleration of mode i
R
ij
Correlation function
Re Reynolds number
r
g
Radius of gyration

S Power Spectral Density


S
c
Scruton number
S
2D
Generalised two tuned degree-of-freedom galloping criterion
S
DH
Den Hartog galloping criterion
S
sc
Non-generalised galloping criterion
S
ij
Dimensionless aerodynamic damping along i due to motion along j
Sr Strouhal number
s Space variable
U Wind velocity
U
n
Normal wind component
U
r
Reduced wind velocity
U
rel
Relative wind velocity
u, v, w Longitudinal, transverse and vertical directions
x, y, z Directional motion variables
Angle of attack
Orientation angle

Wind-cable related angle


NOMENCLATURE xi

ij
Kronecker delta
Roughness characteristic diameter

ij
Kronecker delta
(f) Sears function
Torsional motion
Cable inclination angle

r
Rivulet angular position
Critical damping ratio
Detuning ratio
Kinematic viscosity
Density of air

i
Standar deviation of i
Non-dimensional time variable
Mode shape function
Cable-wind angle

2
(f) Aerodynamic admittance function
Cyclic frequency
Abbreviations
1DOF One Degree of Freedom
2D Two-dimensional
2DOF Two Degree of Freedom
3D Three-dimensional
3DOF Three Degree of Freedom
CEV Complex Eigenvalues Analysis
CHBM Covariance Block Hankel Matrix
CSB Clifton Suspension Bridge
DOF Degree of Freedom
ERA Eigenvalue Realization Algorithm
FHWA Federal Highway Administration
FRF Frequency Response Function
HDPE High Density Poly-Ethylene
IRF Impulse Response Function
IWCM Iterative Windowed Curve tting Method
MIV Motion Induced Vortices
NRC National Research Council
xii NOMENCLATURE
PSD Power Spectral Density
POM Proper Orthogonal Mode
RMS Root Mean Square
SDOF Single Degree of Freedom
SVD Singular Value Decomposition
WIV Wake Induced Vibrations
Chapter 1
Introduction
Structural disasters have always served engineering as a trigger for moving forward. The
image of the Tacoma Narrows Bridge (TNB) collapsing back in 1940, under the action
of only moderate winds, has been one of the most striking failures ever recorded and
markedly inuenced design practice. Since then bridge spans have expanded greatly
with no catastrophic event similar to the Tacoma Narrows failure recurring. This should
be attributed to the great wealth of knowledge that was developed in addressing this
notorious incident. Yet there are still remaining doubts about the exact mechanism
that instigated the dynamic instability. Most wind-structure interaction phenomena,
especially where blu bodies are concerned, possess an intriguing duality. Explaining
them in a certain way does not always rule out alternate interpretations. Cooperative
phenomena seem to form and it is exactly when such hybrid cases are encountered
that our abilities of prediction break down. In any case the design framework currently
in use for aerodynamic eects on bridges has generally proved successful, combining
theory and wind tunnel experimentation. The latter acts as an eective safety net
that can warn us of potential aerodynamic problems. But apart from that, many in-
novative solutions have been conceived in wind tunnels. The form of many modern
bridges are actually owed to them. Through the wind tunnel testing for the Severn
Bridge the streamlined bridge section was invented, which allowed more ecient aero-
dynamically stable bridges. Nevertheless in a wind tunnel not all parameters of the
full-scale conditions can be faithfully reproduced, while size-eects could yield any links
between prototype and model inaccurate. Ultimately denitive information about the
wind resistance of a bridge can only be taken from the actual bridge. In the current
work a combination of full-scale monitoring, theory and wind tunnel tests is utilised in
investigating dynamic instabilities that threaten a bridges structural integrity.
The research begins by considering the old question of utter (utter being the
widely accepted dynamic instability blamed for the TNB collapse) on an existing long
1
2 Chapter 1. Introduction
suspension bridge - the historic Clifton Suspension Bridge (CSB). As a matter of fact
the CSB belongs to the era of the rst European steps in building suspension bridges.
Many examples of the time (e.g. Menai Straits Bridge, Brighton Chain Pier Bridge,
and Roche-Benard Bridge) had severe wind-induced vibrations leading to destruction
or serious damage. Still the issue addressed is not only the specic CSB application but
a more generic approach to answering challenges such as Can we predict instabilities
for bridges that were not designed against them? and Can we predict the true safety
margins to compare with our design idealisations?. Field monitoring of dynamic re-
sponses combined with identication methods on the ambient data were performed in
order to estimate the vulnerability of the bridge to wind action. Exploring the potential
to identify actual full-scale behaviour rather than relying on wind tunnel tests is the
the main target of this section of the thesis.
Subsequently the interest shifts to another bridge-related issue; the dynamic in-
stability of bridge cables and the explanation of their excessive motion rst witnessed
and reported on Japanese bridges in the early 1980s. To justify this research further,
the TNB collapse is reconsidered. A recent explanation of the collapse says that just
before switching into an uncontrollable twisting motion, a pair of the middle cross-tie
cables on the bridge had snapped. Such an explanation could well mean that even
local parameters can have a major eect on the overall response, so before addressing
the overall design of the bridge the response characteristics of the discrete bridge parts
should be fully recognised rst.
It has long been thought that a circular cable, due to perfect symmetry, cannot
gallop (as in classical Den Hartog galloping of iced transmission lines). This view
has recently been challenged and this study aims to uncover the details characterising
the instability generation mechanism known as dry galloping. The specic objective
adopted is to analyse the Reynolds number eects, which were previously overlooked.
For this purpose a series of large-scale wind tunnel tests was carried out with an aeroe-
lastic cable model inclined at dierent angles and equipped with pressure measuring
taps. There have been very few similar large-scale tests in the past and none of them
had equipment to measure the aerodynamic forces on a moving model.
The layout of the thesis is as follows:
Chapter 2 introduces the aeroelasticity framework necessary to follow all concepts
discussed in this work. The literature is critically reviewed while information on
both loading characteristics and modelling practices are provided. The descrip-
tion of conventional well-known self-excitation phenomena such as galloping and
utter is followed by a review of the more sparse research on dry galloping vibra-
3
tions. A series of controversial points are highlighted and additional justication
is given for the main study that follows.
Chapter 3 presents the utter derivative identication scheme performed on the
CSB full-scale ambient vibration data . The chapter commences by introduc-
ing background information on the monitoring procedure and the wind condi-
tions encountered on the CSB. Subsequently the stochastic subspace identica-
tion method used is explained. Flutter derivative results are compared with other
similar sections and eventually an estimate for the utter critical speed is made.
Chapter 4 is devoted to addressing shortcomings of the current literature regard-
ing galloping, quantifying the dierences in the results that can be observed with
dierent conditions or assumptions. The correct two-degree-of-freedom quasi-
steady galloping analysis is put forward for the rst time, identifying the true
eect of the structural parameters on the dynamic instability. The analysis is
useful in understanding the characteristics of galloping and in interpreting the
observed response traces in the following chapter.
Chapter 5 addresses the dry-galloping vibrations of circular cables that dier from
the instability considered in the previous chapter. It starts with describing the
experimental setup used and the large responses encountered. Next, a discussion
of the results regarding the inuence of Reynolds number on the dry galloping
mechanism is made. The inclination angle appears to have key role in the oc-
curence of unstable motion. Discontinuities in the aerodynamic forces originating
from intermittent jumps between dierent ow states along with unusual ow
structures are unique features found due to the transitional behaviour.
Chapter 6 concludes the thesis summarising the main ndings and contributions
of the present research. It also includes suggestions for further work to advance
the present investigation.
Chapter 2
The Aeroelasticity Framework
Long-span bridges are exposed to severe wind action and the main target of this chapter,
before getting into the detailed analysis of this work, is to present the main attributes
of wind-structure interaction that are essential for the aerodynamic bridge design and
highlight yet unresolved issues that are later confronted. Background information at-
tempts to present both the loading characteristics as well as modelling practices that
have been employed in constructing wind-resistant structures. The interest centres on
self-excitation phenomena, which inherently introduce a feedback eect since the actual
aerodynamic forcing becomes a function of the ow-induced motion. Such phenomena,
where cause and eect are interlinked, naturally pose additional hurdles in their treat-
ment. Formulations and theoretical knowledge used in the later chapters are mostly
provided here.
A key feature is to also present the inuence of Reynolds number on ows past cir-
cular cylinders, which later is proved to be an essential piece in the puzzle of excessive
bridge cable vibrations. For cable vibrations specically, a number of inconsisten-
cies underlying the proposed explanation mechanisms are illustrated for the rst time.
Throughout the chapter, real bridge examples are given in relation to the instability
issues discussed to further justify the needs for and benets of the current research.
2.1 Wind-induced structural loading
In general, the wind approaching a structure can be decomposed into a steady and
a uctuating vector component. In exible bridges the gust loading falls well in the
range of their many low natural frequencies and on top of that even the uniform ow is
capable of imposing unsteady pressures, which can dynamically interact with structural
motion causing either a detrimental or a benecial eect. The two mechanisms interact
5
6 Chapter 2. The Aeroelasticity Framework
with the net eect being in some cases a nonlinear one, which cannot be fully perceived
in the realm of simple linear superposition. There is also the static eect from the
ow, which is capable too of leading to failures, as in the case of torsional divergence.
What follows should be treated as a synopsis of the basic aerodynamic and aeroelastic
theory as included in the classical textbooks of Blevins [1], Simiu and Scanlan [2],
Zdravkovich [3] and Dowel et al. [4].
2.1.1 Static loads
The mean wind around a structure engenders static lift (L) and drag (D) forces together
with an overturning moment (M). The L, D, M expressions per unit length are given
by
L =
BU
2
2
C
L
,
D =
BU
2
2
C
D
, (2.1)
M =
B
2
U
2
2
C
M
,
where is the air density, B is a representative sectional dimension, U is the mean
wind velocity and C
L,D,M
are the static lift, drag and moment coecients respectively.
C
L,D,M
are in probably all shapes functions of the angle of attack . The scenario of
torsional divergence earlier referenced is brought about when C
M
has the tendency to
increase with increasing . Thus for a structure with nite torsional stiness, above a
critical value, wind will cause an ever-increasing ultimately leading to destruction.
For dynamic loading a quasi-steady formulation considers that Eqs.(2.1) are still valid
with U replaced by the relative wind velocity U
rel
, which accounts for the inuence
of either uctuating wind components or structural motion. Complications with this
idealisation do exist, since evidently the introduction of rotational velocity waives the
uniqueness of such a U
rel
.
2.1.2 Wind bueting
The unsteady loading imposed by wind turbulence is termed wind bueting. Wind
turbulence, although chaotic and complex in its nature, is briey characterised by a
number of parameters. These are:
The turbulence intensity I
i
=
i
/U where
i
is the Root Mean Square (RMS)
of the uctuating wind component along i = u, v, w the longitudinal, transverse
and vertical directions relative to the wind.
2.1. Wind-induced structural loading 7
The relevant auto-Power Spectral Densities (PSDs)

S
i
(f) and cross-PSDs

S
i
1
i
2
(f),
i
1
= i
2
, where f is the frequency variable and it holds
2
i
=
_

S
i
(f)df. Various
PSDs have been proposed in the literature for design purposes [57], wherein most
cases an ecient description of the uctuations of strong winds is accomplished
through the von K arm an model of isotropic turbulence [8].
The length scale parameters
x,y,z
L
i
, nine in total, which designate a directional
(x, y, z) average size of turbulent eddies in each wind component i.
The classical treatise for bueting loading, utilising the statistical concepts from ran-
dom vibration theory was rst applied by Liepmann [9] to aircraft wings and later by
Davenport to long-span bridges [10, 11]. The formulation, expressing the force com-
ponents in quasi-steady terms, attempts to rst establish the sectional loading PSDs
from the sectional (or point) directional wind PSDs. Further the generalised loading of
each mode is recovered by integration along the length, utilising the modal shape and
wind correlation information. In the frequency domain the PSD expression for vertical
response in each mode j is given by [11]
j

S
z
(f) =
(BU)
2
4
_
_
C
D
+
dC
L
d
_
2

S
w
(f) + 4C
2
L

S
u
(f)
_

2
(f)|H
j
(f)|
2
|J
j
(f)|
2
, (2.2)
where H
j
(f) is the Frequency Response Function (FRF) of mode j, |J
j
(f)|
2
is the
joint acceptance of mode j and
2
(f) is the admittance function. The joint acceptance
function is the means of translating the point-like load to a span-wise load, while the
admittance function should be deemed to be a correction factor to compensate for
the frequency dependence of the instantaneous aerodynamic load [12, 13] and for its
sectional variation. The rst to theoretically evaluate such a function was Sears [14].
For a vertical gust w that is a sinusoidal time function of the form w = w
0
e
2ft
(note
here is the imaginary unit), Sears derived the corresponding oscillatory lift on an airfoil
as
L(t) =
BU
2
2w
0
(f)e
2ft
, (2.3)
where (f) is called the Sears function. Restating Eq.(2.3) in the frequency domain it
becomes

S
L
= (BU)
2

2
(f)

S
w
, (2.4)
yielding quite easily that |(f)|
2
=
2
(f). Many other empirical options exist for
envisaging
2
(f), and there remains an ambiguity over which is the most appropriate
to use. In studies with long bridges with very low fundamental natural frequencies
the contribution from aerodynamic admittance is usually conservatively neglected [15],
though in some cases it is essential for the reliable response estimation [16]. As presented
8 Chapter 2. The Aeroelasticity Framework
in Hays [17] analysis for the Wye and Erskine Bridges, during narrow band vertical
response for the vertical displacement RMS values
z
, the bueting action should set

z
IU
2.83
.
1
(2.5)
The recovered relations from full-scale monitoring of the actual bridge responses were
actually containing a somewhat lower exponent than 2.83. In any case Eq.(2.5) shows
that response is proportional to turbulence intensity and is only asymptotically di-
verging. A nal factor that is worth mentioning is the eect of signature turbulence,
which consists of wind uctuations imposed not by the approach natural ow but by
the actual submerged structure or elements ahead of it. Any shape, unless very well
streamlined, will produce signature turbulence; however its eect in standard bueting
analysis is generally ignored.
2.1.3 Vortex shedding
Vortices trailing behind blu bodies are a very common picture in nature. Of course
the picture is not alone but is accompanied by forces with well dened dynamic char-
acteristics. Strouhal [18] was the rst to observe and suggest that the frequency of the
wake oscillations follow
Sr =
f
v
d
U
, (2.6)
where Sr is the shape-dependent Strouhal constant, f
v
the vortex shedding frequency
and d the across-wind dimension. An across-wind force with frequency f
v
is exerted on
the submerged blu body and as logically expected if the body is allowed to move across-
wind with a structural frequency f
c
, then the two frequencies coalescence will result in
resonance. Actually the classical conception of the wake structure (two counter-rotating
vortices shed alternately from each side during one cycle) implies that an along-wind
force also exists with twice the shedding frequency and is able too of causing resonance.
Unfortunately there are slight inconsistencies in this idealisation. The example of the
circular cylinder will allow to eectively present them. For a circular cylinder Sr is
strongly a function of Reynolds number (Re)
Re =
Ud

, (2.7)
where is the kinematic viscosity. For subcritical values (i.e Re< 10
5
) customarily Sr
is considered to have a value of 0.2. Still this is not strictly true, the actual behaviour
is illustrated in Fig. 2.1 and is governed by discontinuity intervals as well as a strange
1
Hay [17] does not explicitly state which component the turbulence intensity I refers to.
2.1. Wind-induced structural loading 9
inversion of the monotonicity. The critical range behaviour is of far greater interest
(and variation) in large-scale engineering applications and will be discussed in detail
later.
0.00
0.17
0.19
0.21
0.23
S
r
0.02 0.04

Re 1/
2E4 2E3 1000 500 300
Re
retarded
transition
parallel
shedding
5000
1300
240
230
180
360
0.06
Figure 2.1. Mapping of Strouhal number against Reynolds number in the subcritical range.
Adapted from Fey et al. [19].
A great number of experiments have revealed unique features of the aeroelastic char-
acter of the vortex shedding loading and response. The tests by Feng [20] were among
the rst to demonstrate a series of intriguing nonlinear features, such as the capture of
f
v
from the structural vibration frequency in an extended range near resonance during
the phenomenon called lock-in. Seminal reviews on vortex phenomena refer to this
study [2124]. Feng performed his wind tunnel tests with freely oscillating (restrictively
across-wind) cylinders for varying structural damping values in the subcritical range.
He measured the vibration amplitude

Y (scaled with cylinders diameter d), f
v
, f
c
, the
phase dierence of the two throughout the lock-in region and for limited runs the
across-wind force coecient dynamic amplitude

C
Y
. Part of his results are presented
in Fig. 2.2, where all recorded variables are plotted as functions of the reduced velocity
U
r
=U/f
c0
d, with f
c0
being the still air structural frequency. As shown in the gure
the lock-in behaviour establishes itself when approximately reaching the Strouhal res-
onance indicated earlier. From then on, f
v
remains equal to f
c
until the point where it
jumps back to f
v
1.4f
c
following naturally Eq.(2.6). For other shapes the lock-in re-
gion is double-sided extending also to frequencies <f
c
. It is noteworthy that hysteresis
exists in the self-limiting amplitude response of Fig(2.2), with magnitude depending on
whether the wind velocity is increasing or decreasing. The upper branch attained for
increasing wind cannot be reached from rest displaying the interesting feature of depen-
dence on the loading history. For the few runs that

C
Y
is recovered it is obvious that
10 Chapter 2. The Aeroelasticity Framework
its value is strongly amplied. Although not shown here, similar amplication holds
also for the along-wind dynamic forcing. It should be pointed out that the extreme re-
sponse and forcing do not match the initiation of the lock-in zone but they are situated
close to its middle. Finally the phase dierence values along with their discontinuous
jump right in the heart of the synchronised regime, have become a great matter of
controversy and added trouble in the modelling task. In any case phase values indicate
both in phase and in quadrature forcing components. The detailed reasoning on phase
jumps was provided by Williamson and Roshko [22], who identied dierent modes
of vortex shedding and abrupt transitions between them in the amplitude-structural
frequency-wind velocity parameter space. Later work of Williamson with Jauvtis [25]
and Morse [26] completed, for the time being, the shedding mode characterisation.
Sr =0.198
Y
-
U
5 6 7 8 4
Y
-
U
r
=U/f
c0
d
0.5
1.0
1.5
0.4
0.2
0.6
C
Y
1
2
0
0
50
100

f
v
f
c

Y
-
Y
-
from rest
-
d
C
Y
-
f
v,c
/f
c0
Figure 2.2. Free vibration tests for a circular cylinder. Adapted from Feng [20].
2.1. Wind-induced structural loading 11
Outside the 1:1 (f
c
:f
v
) synchronisation range there is also a multitude of interesting
phenomena. Williamson and Roshko [22] identied a region of 1:3 subharmonic reso-
nance and reasoned that wake stability considerations prevent 1:2 similar occurrences.
As a matter of fact they point that subharmonic resonance should be possible for any
1:n ratio, with n being an odd number. Experiments from Durgin et al. [27] partly con-
rm this view by nding large response of free vibrating cylinders in such 1:3 regimes
but not in 1:2 or elsewhere. Still it is worth noting that such phenomena are very rare
and according to classical synchronisation theory should readily vanish when n>3 [28].
A nal unresolved feature picked for this synopsis consists in the force-displacement
relation outside lock-in. As presented in Fig. 2.3 the displacement, top signal, is virtu-
ally a pure sinusoid (at f
c
=9Hz) of very stable magnitude. On the other hand the force
causing this displacement, illustrated in terms of the representative transverse pres-
sure tap trace in the lower signal, shows a strong modulation and a dierent frequency
(i.e. f
v
=13Hz). Thus a question surfaces on how these two signals can combine in a
cause-eect relation. According to Minorsky [29] this seems to be a classical example
of asynchronous excitation. A system possessing a stable focus point (i.e. initial equi-
librium point) followed by two adjacent limit-cycles (such can be a system expressed
by a polynomial of at least fth order) could bifurcate and rest to the outer limit-cycle
when a suitably sized periodic action of random frequency is applied.
Figure 2.3. Circular cylinder vibration phenomena past the lock-in range after Ferguson and
Parkinson [30]. U=5m/s. Top: vertical response record with f
c
=9Hz. Bottom: surface pressure
at transverse tap, f
v
=13Hz.
Modelling of vortex wake phenomena is far from complete. Leaving aside the purely
computational treatises of discrete vortex potential ow models and the numerically
12 Chapter 2. The Aeroelasticity Framework
solved Navier Stokes equations, what remains to be the most ecient attacking tool
for our descriptive low-dimensional studies are the so-called wake-oscillator models.
Their conception belongs to Bishop and Hashan [31, 32] who were the rst to suggest
that a cylinders wake behaves like a mechanical oscillator. Taking into account the
non-linear behaviour proven above, and particularly the lock-in and limit-cycle (the
stationary cylinder vortex shedding is what is idealised as limit-cycle) attributes, a
wise compatible modelling choice from the world of mechanics would be a Van der
Poll oscillator [33, 34]. Such a model was rst implemented by Hartlen and Currie [35]
and acquires the form
d
2
Y
d
2
+ 2
dY
d
+Y =
d
2

2
0
8
2
Sr
2
m
C
Y
,
d
2
C
Y
d
2
a
1

0
dC
Y
d
+

2
0
(
dC
Y
d
)
3
+
2
0
C
Y
= b
1
dY
d
, (2.8)
where Y is the across-wind motion scaled with d, C
Y
is the instantaneous across-
wind force coecient, =2f
c
t is the non-dimensional time variable, is the critical
structural damping ratio in still air, m is the mass per unit length,
0
is the ratio of
shedding to structural frequency f
v
/f
c
and a
1
, , b
1
are empirical constants to be tted
from experiments. This archetypal form of the model assumes nonlinear phenomena
originating from the uidic Van der Poll oscillator and being driven, in the case of
motion, by a linear coupling motion-dependent term. Hartlen and Currie employed
the rst approximation solution of Krylo and Bogoliubo [33] and recovered parts
of the behaviour in Fengs [20] experiments. Amplitude or phase hysteresis was not
apparent in analytical results but this is only due to the solution method employed.
Later variants of the model altered the uid stiness term, the uid damping term, the
structural damping term or the form of the coupling-forcing term in Eqs.(2.8) improving
each time the match to experiments,(for a review see [3639]). The same model was
also used on a less phenomenological basis, being derived from rst principles, having
though C
Y
substituted by a hidden-ow variable [1]. The sophistication of the model
although extensive has been little concerned with a great branch of phenomena, such
as chaotic and quasi-periodic oscillations.
The preceding discussion tacitly assumed smooth ow conditions. The introduction
of turbulence would make vortices lose coherence along the body length and resonant
peaks in the C
Y
spectrum degenerate, broaden, and evidently waive their ecacy in
setting up motion. Somewhat similar eects can be brought upon by surface protu-
berances, surface roughness, long splitter plates or more complex additions such as
wavy separation lines, or spirally arranged bumps [40, 41]. Earlier it was seen that
the enhancement of vortex shedding during large response leads to amplication of
2.1. Wind-induced structural loading 13
the dynamic loading. Similarly when the vortices shed by a body start losing their
strength, the mean pressure drag C
D
is expected to decrease. The most prominent
example of this rule is the circular cylinder and its behaviour along the Re range.
For bridges, although the British revolution of streamlining sections was thought
to adequately handle vortex phenomena [42], this was not actually the case. Vortex-
induced vibrations have been quite systematic in bridges. The Wye Bridge had such
occurrences, but only of small amplitude, while the very similar Erskine Bridge did
not [17]. The discrepancy was reasoned in view of the higher turbulence intensities
measured at site for the latter, which as referenced earlier could disorganise the vortex
formation and propagation processes
2
. The Kessock Bridge [44] had similar observa-
tions with winds of only quite low turbulence intensities exciting moderate amplitude
bending oscillations. The extremely wind-prone Deer Isle Bridge [45, 46] sustained vor-
tex oscillations in probably all of its congurations (being retrotted or repaired in
many instances). For the Storeblt Bridge Larsen et al. [47] presented the excessive
form the phenomenon acquired, with Fig. 2.4 being indicative of amplitudes observed.
The Shanghai Lupu Bridge [48] is a unique reference, being an arch bridge with vortex
issues while on operation.
Figure 2.4. Storeblt Bridge vortex-induced vertical motion at max (left) and min (right) of
the amplitude cycle, after Larsen et al. [47]. Encircled is a parked van with its view distorted
due to motion.
The presentation in this section was mainly based on the circular cylinder paradigm
and one should expect that exible bridges possessing cables, that can most of the times
fall into the circular cylinder category (disregarding only for the time being any incli-
nation), should have issues regarding them. Fortunately as presented by Virlogeux [49]
2
This is not to be confused with the case of turbulence strengthening vortices as Matsumoto et
al. [43] observed on some bridge section types. Their analysis also includes non-classical Motion-
Induced-Vortices (MIV) making up for the seeming inconsistency.
14 Chapter 2. The Aeroelasticity Framework
it can be easily shown, by considering typical values, that large amplitude vortex shed-
ding should not be a concern for bridge cables, since lower mode excitation is restricted
to quite low wind speeds of consequently limited energy content. Nevertheless, higher-
modes would receive large vortex forcing but due to their higher damping would not
produce excess motion. But the stresses exerted in the cables apart from being func-
tions of amplitude, they are also depending on curvature, which evidently increases
in higher modes. Thus even for low-amplitudes but higher mode persistent response,
large stress loading cycles will result, raising fatigue concerns. The record of a specic
cable in the Saint-Nazaire Bridge that initially sustained fatigue damage, was replaced
and later was found to be driven in large higher mode vortex oscillations, should be
deemed to be an indicative example.
2.1.4 Galloping
Classical galloping refers to across-wind motion arising due to the so-called incidence
eect, which translates to a wind forcing contribution originating from variations in
the eective ow incidence angle. It is considered equivalent to the condition
dC
L
d
+C
D
< 0 . (2.9)
Rotational asymmetry is evidently a basic requirement in the operating mechanism. A
simplistic view of galloping is provided in the inset sketch of Fig. 2.5, where a body
in a ow of velocity U is moving downward with velocity y, thus altering according to
denition, the eective ow incidence. The shear layer on the lower side moves closer to
the body, therefore getting more curved, while the upper side shear layer moves away
from the body and becomes less curved. As a result a net downward pressure force is
acted across the side faces (in contrast to an upward force that a streamlined airfoil
would experience) further assisting motion.
It will be shown in Chapter 4, Eqs.(4.7&4.8), that in this classical scenario the
threshold reduced wind velocity U
0
for setting o galloping, calculated by means of
linear theory, is proportional to structural damping. Thereof a diverging response
should result. In reality a limiting mechanism operates to set response into a steady
state. Parkinson et al. [50, 51] were among the rst to implement nonlinear concepts to
evaluate steady galloping amplitudes that well match experimental observations. Their
analysis recovers that galloping increases roughly proportional to reduced velocity U
r
,
as shown in Fig. 2.5, with also hysteresis eects and jump behaviour emerging in
the range noted by a dashed line and bounded between arrows. Later Novak [5254]
proposed the notion of universal response curves after rescaling axes in Fig. 2.5 with
2.1. Wind-induced structural loading 15
Figure 2.5. Typical galloping response curve for a rectangular prism. Inset the classical
galloping mechanism illustrated.
S
c
=m/d
2
. This dimensionless factor, named after Scruton, was rst proposed by
Scruton [55] as the inuencing parameter against most aeroelastic instabilities. Novak
extended on the distinction between hard and soft galloping oscillators, soft being
the ones able to gallop from rest while hard the ones in need of an initial hard push
to get into motion, presenting also a number of special cases where stable sections can
turn into weak hard ones.
Galloping forces in general terms should be considered as the product of ow-
afterbody interaction. Parameters altering any of the two constituents inevitably will
aect the instability characteristics. Introduction of turbulence for instance can turn
an unstable section to stable or the opposite, with any changes being strongly shape-
dependent. Similarly insertion of a splitter plate in the near wake can be quite dramatic
concerning the separation process [56, 57] and induce galloping in cases where it would
not be expected, as in the examples of a rectangular section with along to across-
wind dimension ratio higher than 3 and the perfectly symmetric circular cylinder. The
splitter plate inuence on galloping is the opposite of the inuence earlier quoted on
vortex shedding, thus a question is raised on the link between galloping and vortex
shedding. In most practical applications in wind, the regions where the two phenomena
become dominant are well separated, with vortex shedding being conned in relatively
low reduced velocities and galloping appearing later only for much higher U
r
values. A
typical example of aeroelastic response would be expected in the form of Fig. 2.6, with
vortex shedding giving a response hump near U
r
=5 and after a quiescent transition
period galloping taking over.
16 Chapter 2. The Aeroelasticity Framework
0 5 10 15 20
0.1
0.2
U
r
Damping ratio :
0.37 %
0.76 %
1.40 %
2.12 %
4.40 %
U
1
Y
-
Y
-
2
Figure 2.6. Response of a rectangular prism with side ratio 2/1 against reduced velocity for
varying critical structural damping ratio. The prism has Sr=0.081 that sets the relevant U
r
for vortex resonance at 12.34. Adapted from [53].
Actually the specic example of Fig. 2.6 although explained in this expected way
even in well respected textbooks, see [1], it contains some paradoxical features. Con-
trary to the aforesaid descriptions what seems like galloping here starts abruptly at
U
r
11 for all dierent values of structural damping, while in the =4.40% case it
attains a rather unexpected decay. Taking into account that the rectangular prism
in hand has Sr=0.081, setting the relevant U
r
for vortex resonance at 12.34, this
paradigm is probably a good indication of the rather complicated form that hybrid
vortex-galloping oscillations may obtain. For =0.37% the galloping threshold is eval-
uated at U
0
=5.2 [53] but oddly is inhibited up to where vortex shedding is regularly
located. On the other hand for =4.40% it follows U
0
=61, which explains the non-
increasing response character. The interaction range includes many more possibilities
to be pursued in later chapters. In the context of the present work the term galloping
will be used to characterise any motion-triggered translational aeroelastic instabilities
regardless of orientation and extending to also cover the combined participation of
orthogonal motion.
Galloping tends to be considered of less interest in cable-supported bridges, still
there were a number of instances in design or construction where it did show up. Ac-
cording to Virlogeux [58], the hexagonal bunches of strands to be installed on the
Normandy Bridge were abandoned due to galloping concerns that arose while in wind
tunnel testing. Instead a circular strand distribution was promoted, which was also cov-
ered by a high density polyethylene (HDPE) sheath with helical llets. Wardlaw [59]
2.1. Wind-induced structural loading 17
reports the wind tunnel investigation of the single tower of the Aratsu-Ohashi cable-
stayed Bridge where an apparent galloping susceptibility was documented throughout
the construction stages. A more controversial appearance should be the one docu-
mented in Deer Isle Bridge with deck motion amplitudes reaching up to 6100mm [46].
Although also other loading mechanisms could be held responsible for this incident the
characteristics of the bridge as presented by Cai et al. [60] readily support the galloping
scenario. Galloping events and the countermeasures for their treatment, for a series of
arch and truss bridges e.g. Burton Bridge, Bras dOr Bridge and Commodore Bary
Bridge are presented by Wardlaw [59, 61]. Finally the case with excessive vibrations
recorded on the circular iced hangers of Storeblt Bridge [62], reminiscent of iced trans-
mission line galloping, is another phenomenon worthy of engineering attention along
with other less evident similar instances.
2.1.5 Flutter
Flutter is the aeroelastic instability that naturally follows galloping. It is nominally
of divergent character, rapidly building up, and has provided engineering history with
some of the most spectacular failure pictures. In dierent forms it is thought to have
been met rst in slender bridges, dating as long as two centuries ago, and much later on
the torsionally weak wings (and tails) of World War I ghter aircraft, while nowadays
it is still one of the most serious concerns of aerodynamic design. Although many
classications have been used by dierent authors, here the broadest categorisation
into coupled (or classical) utter and Single-Degree-of-Freedom (SDOF) utter will be
introduced.
Figure 2.7. (a) Displacements and aeroelastic forces on a thin airfoil; (b) Displacements and
aeroelastic forces for a bridge section
Coupled utter was initially used to name the combined torsional-bending insta-
bility of airfoils. The phenomenon necessitates for a torsional and a bending mode
to oscillate at the same frequency but with a decisive phase dierence between them,
that allows their cooperation to extract energy from the wind. Theodorsen [63] laid
18 Chapter 2. The Aeroelasticity Framework
the basis for ensuing utter analysis by estimating the self-excited lift force (L
se
) and
pitching moment (M
se
) of a thin airfoil section immersed in an incompressible uid
ow, while performing small amplitude harmonic vibrations of cyclic frequency . The
model problem is illustrated in Fig. 2.7(a), where typically the drag force and displace-
ment are unimportant and additionally it is assumed that the shear centre and chord
centre coincide. Theodorsens derivation established L
se
, M
se
as linear functions of
translation h, rotation and of their rst and second time derivatives,

h,

h, , ,
where each overdot denotes one dierentiation with respect to time. His well-known
solution is given as follows [64, 65]
L
se
= b
2
[U +

h] 2bUC(k)[U +

h +
b
2
] ,
M
se
= b
2
[U
b
2
+
b
2
8
] +b
2
UC(k)[U +

h +
b
2
] , (2.10)
where k=b/U is the reduced cyclic frequency, based on the half chord length b, and
C(k) the complex Theodorsen circulation function that is assigned to time delays,
and is analytically expressible in terms of Bessel functions. C(k) becomes 1 for static
conditions, i.e. k 0, reverting to the quasi-steady formulation. It was early realised
[6668] that direct application of Eqs.(2.10) to the analysis of bridges, owing to their
generally blu sections, will yield inaccurate results. Hence the rationale of expressing
aeroelastic forces as linear k-dependent motion functions was preserved but C(k) was
replaced by a more reliable empirically determined substitute. Many options [6971]
toward dening a reliable such substitute were formulated but the one originating from
Scanlan and co-workers [72, 73] became dominant. The Scanlan approach in its latest
amended form [74], expresses the self-excited force and moment components for the
bridge section in Fig. 2.7(b) as
L
se
=
1
2
UB
_
KH

h
U
+KH

2
B
U
+K
2
H

3
+K
2
H

4
h
B
+KH

5
p
U
+K
2
H

6
p
B
_
,
D
se
=
1
2
UB
_
KP

1
p
U
+KP

2
B
U
+K
2
P

3
+K
2
P

4
p
B
+KP

h
U
+K
2
P

6
h
B
_
,
M
se
=
1
2
UB
2
_
KA

h
U
+KA

2
B
U
+K
2
A

3
+K
2
A

4
h
B
+KA

5
p
U
+K
2
A

6
p
B
_
,
(2.11)
where K=B/U is the reduced cyclic frequency, based on the full deck length B this
time, and H

16
, A

16
, P

16
are the so-called utter derivatives, which are functions
of K and are derived by means of sectional free or forced wind tunnel tests. No
inertial contributions are considered, since for heavy bridge decks in air they should
2.1. Wind-induced structural loading 19
be minimal. P

derivatives linked with D


se
were a later addition to the formulation,
which actually proved critical for explaining the aeroelastic behaviour of the Akashi
Kaikyo full bridge model [75]. The linearisation concept utilised in deriving Eqs.(2.11)
should be applicable for small structural deections, corresponding to only incipient
instability action, and in the absence of concerted vortex shedding with its subsequent
strong nonlinear characteristics.
The main attribute enabled by coupled utter is having motions (one of them
being necessarily torsional) which when autonomously considered would behave stably,
i.e. have positive eective damping, but in common operation would exchange energy
between them through coupling terms H

2,3,5,6
, A

1,4,5,6
, P

2,3,5,6
with a total positive
net energy eect for their system. Well streamlined bridges such as Severn, Lilleblt,
Burrard Inlet, Humber and Bosporus should exhibit catastrophic coupled utter at very
high wind speeds, when interestingly part of their direct aeroelastic forces (velocity
products with H

1
, A

2
, P

1
) contribute positive aerodynamic damping that adds to
the structural damping [76]. Note that these outcomes were derived only in wind
tunnel tests, and design wind speeds nowadays by far exceed the recorded operational
envelopes of modern real long span bridges, making the coupled utter phenomenon a
quite improbable event. Still, complying with the strict utter design guidelines has
imposed shape modications on major modern bridge prototypes, referring in short
to the slot and stabiliser additions in the Akashi Kaikyo Bridge, the slot addition in
Zhejiang Xihoumen and Tsing Ma bridges [77], and the fairing addition to the Ting
Kau Bridge [78].
SDOF utter should more accurately for the onomatology adopted in this work
be termed as torsional utter, since the galloping convention introduced earlier should
cover any exclusively translational self-excited response. The mechanism operation re-
lies, as in incipient galloping, on an aerodynamic negative damping eect which reduces
structural damping and ultimately turns their sums sign to negative. Apparently from
Eqs.(2.11) A

2
becoming positive serves this goal. This function which is a typical
characteristic of many blu sections previously used in bridges (e.g. H-sections [73]),
constitutes the most striking dierence between airfoil and bridge behaviour. Thus
historically SDOF utter was realised as a separated-ow phenomenon. Model stud-
ies have shown that there could be cases where a strongly positive A

2
coexists with a
tendency toward torsional-bending coupling. Then the intrinsic proclivity for torsional
utter may drive the participation of vertical motion too, establishing optical (but not
functional) resemblance with coupled utter.
Further on bridge utter, its interaction with turbulence remains mystifying. Name-
ly a retardation of divergence of the full bridge response was early witnessed [12, 76] as
qualitatively illustrated in Fig. 2.8(a). The phenomenon was initially accredited to the
20 Chapter 2. The Aeroelasticity Framework
R
M
S

d
i
s
p
l
a
c
e
m
e
n
t
Wind velocity U
Laminar
flow
Turbulent
flow
(a)
0.2
-0.2
0
2 4
Turbulent flow
Laminar flow
fU/B
A
2
*
(b)
Figure 2.8. (a) Quantitative dierence of response characteristics for a full bridge under
dierent ow conditions. Adapted from [59, 76]; (b) A

2
from wind tunnel tests for a torsionally
unstable bridge section under laminar and turbulent ow conditions. Changes appear minimal
to sustain any substantial modication in the utter behaviour. Data after [79].
eect of turbulence on utter derivatives themselves, but a series of later studies [74,79]
proved that any alterations were minimal. A specic example for a torsionally weak
bridge section exposed to both laminar and turbulent ow conditions, is presented in
Fig. 2.8(b). Shown changes in A

2
are negligible, with even a minutely earlier negative
crossing for the turbulent scenario that should contradict experience. Other forms
of reasoning the utter-turbulence interaction consist of multi-modal behaviour that
spreads the uttering modes energy into other modes with various damping and
frequency characteristics, essentially setting up a complex tuned damping system [12,
76]. Finally explanations were acquired by means of the distracting action of turbulence
on the span-wise coherence of aeroelastic loading [80], although Scanlan also pointed out
that there is a possibility for such inhomogeneity to act detrimentally if local extreme
excitation regions, e.g. of very positive A

2
, are introduced.
2.1.6 Wake-induced loading
Wake-induced loading refers to wind forces exerted on a body when this is situated in
the unsteady wake of another upstream blu body. Meeting a structured disturbed
stream of diused and convected vorticity, when not far downstream of the source,
seems quite dierent than facing directly a fully developed turbulent ow, which also
has unsteady components but any localised imprints in it have been smeared out.
Notwithstanding the intuitive discrepancy, the forcing mechanism has been historically
perceived as only a special case of the previously presented instabilities and have been
explained on the same grounds. The term wake galloping is customarily employed to
2.1. Wind-induced structural loading 21
characterise any excessive response events where wake loading seems to occur, still Assi
et al. [81] recently, revisiting an old problem set forward by Zdravkovich [82], exempli-
ed the case where a tandem pair of circular cylinders, not far apart (distance/diameter,
l/d=4), performs vibrations inherently dierent from common galloping. Galloping
should be successfully attained by quasi-steady theory, which strongly relies on the
mean wind velocity distribution. In the Wake-Induced-Vibrations (WIV) of Assi et al.
the unsteadiness brought upon by the vortices of the upstream obstacle are the critical
driving parameter and not the mean wind velocity distribution, while on the other hand
the phenomenon is non-resonant and induces a wake stiness (or frequency) attribute
that will not match vortex shedding as presented earlier. Such phenomena, and any
interaction phenomena in general, will not be pursued further, but it is of relevance for
this work to shortly present them.
A very interesting feature in modelling wake interference eects is that ordinary
modelling tools can capture their non-conservative nature once endowed with a mem-
ory eect, which assumes the form of an articially induced time-lag between motion
and aeroelastic force [83, 84]. This is striking when considering the discontinuous, or
even bistable [85] character observed. Bridge cables in parallel arrangements and on the
lee side of bridge towers, for certain wind directions, could evidently fall in the realm
of wake-induced response. Preventing such events consists of spacing cables far apart.
A distance of more than ve or six diameters is thought to be an adequate remedy ac-
cording to Matsumoto et al. [86] and Tokoro et al. [87]. Records of large responses that
should be possibly accredited to wake eects can be found in many modern bridges.
In the Akashi Kaikyo Bridge, during the construction stage, hangers spaced at l/d=9
vibrated violently [88]. In another case of inclined stays this time, on the Second Severn
Crossing with even greater inter-cable distances (approximately 4m), relatively large
cable vibrations were witnessed in February 1999. Wind was blowing almost in parallel
to the cable fans, exciting all cables into rst mode large transverse motions. The
phenomenon looks similar to tube arrays response, as in breathing [49] for instance,
though the large spacing discourage engineers of linking it to a wake-related source. On
the resundsbron Bridge where inclined twin cables of d=250mm spaced 670mm apart
were used as stays, many extreme cable vibrations occurred [89]. The small spacing,
determined through a series of scaled wind tunnel tests though, is thought to be of some
connection to the events. Yet it is quite dicult to conclude, when even crude details
(e.g. of the vortex shedding process) due to the individual moving cables inclination
remain in ambiguity.
22 Chapter 2. The Aeroelasticity Framework
2.1.7 Rain-wind Instabilities
The cooperative action of rain and wind can produce forcing, which exceeds the previous
categorisation. The vibrations of yawed cable-stays in the Meiko-Nishi Bridge during
erection, reported by Hikami and Shiraishi [90], were the rst to be attributed the
designation rain-wind instabilities. The name derives from the initial observation that
as soon as either rain or wind ceases, forcing dies down. Actually the contribution of
rain was even earlier postulated to be strongly inuential in aeroelastic loading owing to
the monitoring work of Hardy and Bourdon [91] on transmission lines. Nowadays most
of the vibration recordings in many cable-stayed bridges concern the cables and are
classied as such phenomena, so evidently they deserve dedicated space in this study.
Figure 2.9. (a) Rivulet formation on the circular cable section; (b) Inclination geometry of
the inclined and yawed to the ow cable.
What seems to accompany the instability is the formation of water rivulets along
the cable length, as in Fig. 2.9. The formation is a process governed by many param-
eters such as the rainfall intensity, the cable inclination, the wind yaw, the treatment
and material of the cable surface and naturally the wind velocity and motion frequency.
There still remains some abstruseness on the subtle characteristics of these vibrations,
while in some cases contradicting results have been brought forward. Hikami and Shi-
raishi [90] performing wind tunnel tests on cable geometries with =45

, =45

,
according to Fig. 2.9(b), observed that it is the upper rivulet motion that contributes
the excitation force with a lower rivulet only adding positive damping. Similarly Fla-
mand [92] for =25

, =20

50

attained large vibrations when an upper rivulet (with


no lower one though) formed and oscillated in the circumferential direction. Further
attaching false xed rivulets at the place where the real ones were previously seen, no
2.1. Wind-induced structural loading 23
instability was found, thus yielding that it is the synchronised motion of the rivulet
that causes vibrations and not simply its appearance. Oppositely Bosdogiani and Oli-
vari [93] adopted the view that the motion of the liquid rivulets is not indispensable,
and by attaching rigid bars, of realistic size and shape, where rivulets would normally
form proved that classical galloping could emerge. On the same grounds Matsumoto
et al. [94] using a horizontal (=0

) circular cylinder, generically yawed (=0

45

) to
the ow, performed an extensive parametric study varying the angular position of an
articial glued upper rivulet, in order to assess the eectiveness of the rivulet place-
ment. It is interesting to note that they obtained an unstable response even for the
typical non-yawed, non-inclined cross-ow scenario. Their results do not agree with the
rain-wind tests of Cosentino et al. [95] on a geometrically similar set-up. Still the impli-
cation of rain-wind vibrations on non-inclined cables was also made by Verwiebe [96],
who quotes the examples of hangers in two arch bridges that sustained such events just
before being commissioned. Verwiebe for cable orientations with =30

, =0

90

presents three dierent vibration mechanisms underlying the rain-wind phenomenon.


In two of them the participation of both upper and lower rivulets is mandatory, with
resulting trajectories being planar across or along wind, depending on the symmetry
of the rivulets movement. For his remaining third mechanism only the lower rivulet
drives response resulting in an elliptic motion close to across-wind that can ultimately
cross over to purely across-wind when an upper rivulet forms.
Notwithstanding the discrepancies there are features unanimously accepted. Inci-
dents show up in a velocity restricted range visually similar to the vortex shedding
response of Fig. 2.2. Yet unlike vortex shedding the bounds of this range are inde-
pendent of the motion frequency [90, 96] becoming functions of simply the wind speed.
Close to this observation lies that Reynolds numbers are of the order of 10
4
10
5
, which
is nominally subcritical or very early critical. Amplitudes can reach up to several
meters i.e. >10d. There is some consistent repeatability of vibrations for a range of
eective

(see Fig. 2.9 for the angle denition) between approximately 20

and 35

,
but separately and are also relevant due to the gravity force inuence on the water
rivulet motion. Flamand and co-workers [95, 97] recorded the water thickness and lift
force characteristics during large rain-wind induced excitation, as presented in Fig. 2.10.
Both plots are noisy and have intermittent ring intervals, raising questions as to how
synchronisation can be feasible in such erratic waveforms.
The self-exciting character of rain-wind vibrations renders them similar to galloping
and alike formulations have been used for their modelling. Yamaguchi [98] employing
quasi-steady theory, showed that a sliding upper rivulet can oscillate along the cir-
cumference with a frequency originating from aerodynamic stiness. As this frequency
measure varies with increasing wind speed, it approaches the cable frequency with their
24 Chapter 2. The Aeroelasticity Framework
12
7
2
-3
1 2 3 4
Lift force
Displacement
time (s)
c
m
,

N
/
m
10 12 14 16 18 20 22 24 26 28 30
17
19
21
23
time (s)
p
o
s
i
t
i
o
n

(

)
r
i
v
u
l
e
t
U=11.5m/s, =25, =50
=25, =30
Figure 2.10. Upper water rivulet mean angular position during rain-wind vibrations and
lift force, displacement time series for a dierent large response conguration. Adapted from
Flamand et al. [97].
coalescence setting o instability. This model along with its later successors does not
produce the random-like features shown in Fig. 2.10, however it has been deployed with
relative success in modern cable-stays (e.g. [99]) for treating a problem that is not yet
resolved in eld. A long list of large-scale real events can be presented. The examples
of Puente Real Bridge, Veterans Memorial Bridge, Fred Hartman Bridge, Erasmus
Bridge, Dubrovnik Bridge, Far and Higashi-Kobe Bridge are only indicative of how
widespread the phenomenon is. The actual case of Higashi-Kobe Bridge is of particular
interest. Fig. 2.11 from Kitazawa et al. [100] presents a summary of the preliminary
cable wind tunnel tests that were performed in order to mitigate the rain-wind response
of the proposed bridge design. Assuming that regular spaced protuberances along the
cable circumference will inhibit formation and motion of rivulets, it was found that in-
deed they are very eective in restraining large motions, keeping response low relative
to the plain smooth cylinder option. This was the verdict for all the tested wind speed
range and for both rainy and dry conditions. Thus such an aerodynamic measure was
applied to the newly erected bridge for rst time ever.
Still there is another feature worthy of note in Fig. 2.11. For the cable encased in a
smooth circular duct, large vibrations were attained not only under rain but also in dry
conditions. The frequencies (at around 1Hz) are far o to reason direct K arm an vortex
shedding resonance, and no water on the cable surface exists to form aerodynamically
unstable rivulets. Thus the origin of this new type of response sets forward a new puzzle
to later consider. Closing up this section it should be referred that large vibrations on
the protuberance-equipped cables of Higashi-Kobe Bridge did occur for an extreme
velocity outside the tested range of the preliminary tests (i.e. around 40m/s) [101].
It is the empiricism in the current state of knowledge that clearly necessitates for
additional testing and understanding of underlying mechanisms in order to get into a
state of successfully predicting problems.
2.2. Circular galloping: myth or true? 25
Figure 2.11. Resume of wind tunnel test results on the proposed cables in the Higashi-
Kobe Bridge. Improvement in aerodynamic performance is apparent for the solution with
protuberances under all conditions. Adapted from Kitazawa et al. [100].
2.2 Circular galloping: myth or true?
The unheralded diverging response illustrated for the plain cable in Fig. 2.11, seems
to resemble the classical galloping of Fig. 2.5. However a nominally perfect circular
body cannot t galloping per se. Due to perfect symmetry any incidence eect is
expected to cancel out. As Parkinson argues in his cogent galloping review [23] for
such shapes their afterbodies do not interfere with the separated shear layers and the
subsequent vortex formation, so that only vortex-induced vibration from rest will occur
for elastically-mounted cylinders, and galloping is not an issue. This statement puts
forward a fundamental challenge: Is galloping possible for a circular cylinder? and if
not what is this new phenomenon captured by Fig. 2.11?
Answering these questions is far from obvious and entails rst providing a short
background on aerodynamics specic to the circular cylinder. Actually Parkinsons
quote is stripped from inuences brought upon by Reynolds number transitions and
three-dimensionality of the ow, so it is natural to pursue discrepancies over these
parameters.
2.2.1 Reynolds number eects
For circular cylinders aerodynamic characteristics were early found to be decisively al-
tered by Reynolds number. Further to the subcritical Sr changes illustrated in Fig. 2.1,
the later critical and post-critical regions embody many interesting features (not only
26 Chapter 2. The Aeroelasticity Framework
in terms of Sr variation) that could well be held responsible for complex dynamic
behaviour of structures.
Bearman [102] was the rst to systematically map the evolution of all main aero-
dynamic features in the Re range 10
5
to 7.510
5
. He discovered that a discontinuous
jump in time-averaged base pressure, and concomitantly in time-averaged pressure
drag
3
force, takes place at Re 3.410
5
suggesting that this is brought forward by
the establishment of a laminar separation bubble only on one side over the complete
length of the cylinder. Interestingly the bubble formed consistently on the same side,
despite the very smooth cylinder nish and the absence of apparent asymmetries in ow
conditions. A large steady lift force (i.e. C
L
1.3) also resulted due to this one-sided
bubble. At the same time the frequency characteristics for lift, acquired through wake
velocity measurements, changed in an intermittent manner. For Re=3.5510
5
, when
the bubble was thought to be unstably bursting, Sr was transiting between two well
dened values at 0.23 and 0.32. Subsequently for a small Re increase the bubble sta-
bilised and acquired a single Sr 0.32. At larger Re a bubble similar to the rst formed
on the other side of the cylinder bringing an end to the asymmetry-induced lift and
the so-called critical Reynolds region. Sr after this was about 0.48 with intensity more
than an order of magnitude lower than in any previous state. Bearman nally noted
that contamination of the surface, e.g. by a dust particle, would trip the ow, alter the
ow uniformity over a considerable length, and transform the periodic vortex shedding
regardless of Sr to a wide-band process. An argument along the same lines, concerning
the high sensitivity while in near-critical Re, was earlier suggested by Humphreys [103],
who found that the addition of few light silk threads on the stagnation line can promote
regular 3D ow-patterning.
A substantial contribution to this insightful schema would come more than ten years
later. Kamiya et al. [104] performing similar experiments to Bearman, measured the
pressure distribution at a near-middle section and acquired plots translating meticu-
lously the laminar separation bubble notion to pressure proles. In addition to before,
they obtained one-bubble states in alternate sides, waiving any uncertainty left that the
lift appearance may be an artefact caused only by geometric imperfections. They also
captured a hysteresis eect on all their recorded transitions (i.e. zero to one bubble,
one to two bubbles, and vice versa). This designates that the history of the ow is
crucial in the critical Re region. The actual transitions illustrated a transient period,
where steady lift coecients would not jump directly to a large (or zero) lift value but
rst gradually increase (or decrease) and this way reduce the amplitude of the ensuing
discontinuity. In terms of the formation of bubbles this could probably be seen as a
3
Reference to drag throughout this study concerns the pressure part of drag, which is dominant for
moderate to high Re numbers.
2.2. Circular galloping: myth or true? 27
gradual growing with increasing growth rate. This conceptual view was also shared
by Almosnino and McAlister [105] who described the phenomenon as a supercritical
bifurcation.
Conversely, Schewe [106] in his seminal work, where he rst proposed the similar-
ity of ow transitions to bifurcations he recorded sudden abrupt ow-state jumps and
consequently termed the bifurcations subcritical. In any case the bifurcations implied
by both Schewe and Almosnino and McAlister require higher order polynomials (at
least fth, see [107]) for their accurate description. Schewe also suggested a number
of innovating ideas. Among them, he quotes that the observed transition phenomenon
is a hydrodynamic instability that should be treated in the framework of phase (or
critical) transitions. One particular contribution to the initial Bearman description
includes the recovery of the so-called critical uctuations. Following Schewe, before
any discontinuous drag or lift jump takes place, the periodic shedding moves away
from its well dened Sr value, and acquires low frequency components that designate
the so-called critical slowing inherent in the proximity of any critical point. Another
interesting feature Schewe nds is that in the one-bubble regime, lift and drag PSDs
show the same pronounced frequency (Sr=0.33), which does not match the classical
shedding mode, where they should have a 1:2 relation. He estimates that this origi-
nates from vortices being strong only from the one side where a bubble has not yet
formed. Further he completes the characterisation of ow for Re up to 7.110
6
, in
what should be deemed as the supercritical and transcritical Reynolds range. In the
course of increasing Re, a series of new overlapping transitions are postulated, where
bubbles progressively disappear towards reaching the fully turbulent state. In this nal
turbulent state ordinary shedding seems to have revived at Sr 0.27, very close to the
subcritical value, a nding which has been known for many decades due to the work
of Roshko [108]. Roshko also quoted that the obtained mean pressure distributions for
these high Reynolds numbers are insensitive to the addition of a splitter plate in the
wake.
A concise schematic description of all the above was devised by Zdravkovich [109],
and is illustrated in Fig. 2.12. As shown, the time-averaged drag coecient C
D
varies
as the position of the laminar-turbulent transition travels from the wake toward the
stagnation point. The one and two bubble states earlier dened, are mapped to the
designated TrBL1 and TrBL2-3 regimes. Most importantly Zdravkovich notes that
the whole classication should be valid only for a disturbance-free ow. The eects of
freestream turbulence or surface roughness could drastically alter the image. Although
most of the time both these inuences are modelled as a simple shift to the left of the C
D
curve in Fig. 2.12, the actual modications are more subtle. The changing operations
concern exclusively the states where turbulence intrudes the shear and boundary layers,
28 Chapter 2. The Aeroelasticity Framework
Figure 2.12. Mean drag coecient versus Reynolds number. On top, transitions (Tr) from
laminar (L) to turbulent (T) ow are presented in relation to separation points (S) and boundary
layers (BL). Adapted from Zravkovich [109].
termed TrSL and TrBL respectively. Large enough roughness values can for instance
completely preclude separation bubbles and thus sweep away the greatest part of TrBL,
whereas freestream turbulence would also very eciently relocate and contract TrSL2.
All this description is exclusive to a static cylinder. Motion similarly perturbs the ow
by bringing vortex formation nearer to the cylinder surface, altering many details of
the resultant aerodynamic force. According to Humphries [110], who performed water-
tunnel tests on a large-scale exible cylinder, motion can preserve vortex shedding
unaltered (i.e. without Sr transitions) throughout the critical Reynolds number range.
2.2.2 Inclination eects
Imagine a circular cylinder generically inclined to the ow. In this case, there is not only
the normal to the body wind component but also an axial wind contribution running
along the span-wise direction. This heuristic approach, where the ow is decomposed
into two independent orthogonal parts (i.e. independence principal), although appeal-
ing on its simplicity is not always accurate. The actual ow is far more complicated
and 3D features emerging from the complex uid-structure interaction can render such
a two-component ow partition invalid.
Bursnall and Loftin [111] performed one of the very few test studies on yawed cylin-
ders inside the critical Reynolds number range. Their results suggest that the wind
component normal to the cylinder U
n
, is not enough to adequately characterise the
mean drag evolution. For their tested cases with cylinders inclined at 90

, 75

, 60

, 45

2.2. Circular galloping: myth or true? 29


and 30

to the freestream, they produced maps of U


n
versus the normal to the cylinder-
axis mean drag force, and obtained ve dierent curves. They found that the lower
the inclination the lower the U
n
at which the critical drag-drop occurs. Namely for
the vertical cylinder the drag begins declining at 3.710
5
, whereas rotating 45

or 60

(transiting to more shallow congurations) this value becomes 2.310


5
and 1.0610
5
respectively. The situation is also preserved if the total wind is used instead of just
the normal wind component. Additionally there are vast discrepancies in the nal
supercritical mean drag coecients from dierent inclination set-ups. However, these
dierences would much reduce when the coecients are estimated based on the actual
wind speed. This sets a fundamental question on which would be the most appropriate
wind measure to use in our descriptions. Further, acquired pressure distributions re-
vealed that increasingly deviating from the vertical case, bubbles became less apparent
and less stable.
Ramberg [112] when treating inclined cylinders, focused on vortex shedding and
much lower Reynolds numbers. He found that for Strouhal number calculations, the
use of the normal wind component is a convincing approximation for a wide range of
inclinations. He also presented the increased geometric sensitivity that is inherent in
the inclined cylinder ow, by producing variously slanted, multistable wakes for only
minimal boundary alterations. These results were also corroborated by Shiraishi et
al. [113], who set o to test the validity of considering the yawed circular section as
analogous to an elliptical one. They disproved this view and with ow visualisations
illustrated that indeed streaklines bend and cross the static inclined cylinder at almost
right angles, which should readily support the use of simply the normal wind com-
ponent. Following Ramberg, they found that the addition of end-plates would alter
the shedding strength and uniformity. Resonant peaks in lift PSDs do not have the
sharpness seen in the non-inclined scenario. Yet, Ruscheweyh [114] testing a range of
inclined cantilevered (where probably also tip vortices contribute in loading) cylinders,
showed that the vortex shedding response would reduce due to inclination only for
higher Scruton numbers. Depending on the end-plate spacing, lift PSDs would also
attain a low frequency content, probably similar to what was previously referred as
critical transitional uctuations. Actually Ramberg captured dierent wake modes co-
existing during his tests. Hence, critical-like uctuations could result by transitions,
not on the boundary layers this time, but between the referenced wake modes. On an-
other aspect, exploiting the along-length inhomogeneity, Hayashi and Kawamura [115]
uncovered a pressure gradient on the lee side of their cylinder, regardless of boundary
conditions. Its sign promotes ow directed from downstream toward upstream, which
noticeably is opposite to the axial wind component. Detailed characteristics of this
combined axial ow are largely undened.
30 Chapter 2. The Aeroelasticity Framework
2.2.3 Instability mechanisms
So is the circular cylinder response exemplied in Fig. 2.11 galloping? In terms of de-
nitions it was actually named dry galloping, mainly due to its seeming divergence with
increasing wind speed, but whether it truly holds any resemblance to classical galloping
is a view seriously disputed. A series of recent studies [116118] assign the phenomenon
to ordinary vortex shedding origin, and refute the galloping characterisation. In any
case, the study of such phenomena has lately received a lot of research interest. Large
vibrations of bridge cable-stays that were initially suspected to be due to rain, are now
claried to have occurred under dry conditions. Although it is clear that to mitigate
the previously presented rain-wind vibrations one should inhibit the rivulet formation
and motion, the required countermeasures for dry wind vibrations are a disturbing
mystery. To answer it, it is essential to uncover the mechanisms stoking the instability.
Matsumoto and his co-workers with a series of seminal papers [94, 119, 120], were the
rst to attempt an explanation of such unexpected wind behaviour. Initially it was sug-
gested that three dierent types of response can occur. A galloping type, which could
be either diverging or velocity restricted, a vortex shedding type with long period, and
their mixed type. It was postulated that rain-wind phenomena would also fall into this
broad framework. Rivulets when formed would only amplify these identical instability
types. Amplication is usually envisaged in a quasi-steady manner, simply imposing
an added geometric asymmetry.
Galloping type response
The galloping type response emerges due to the axial ow that runs in the lee side of
the inclined circular body. The early suggestion [94, 119] was that this axial ow is a
non-vortex ow, close in value to the approaching wind axial component. Acting as an
air-curtain it would simulate a long-rigid splitter plate that as indicated earlier it will
induce classical galloping on a plain circular section. The intensity of the axial ow,
is the critical parameter that decides the vibration occurrence. Evidently, according
to this rule, non-yawed cables are not susceptible to dry galloping. As a matter of
fact, Matsumoto added an external articial uniform ow in the wake of a horizontal
cylinder, and observed large response similar to the yawed equivalent. This was consid-
ered a convincing proof of the suggested theory. The schematic representation of this
archetypal idea is given in Fig. 2.13(a). A splitter plate in principle has a dual oper-
ation. It reduces the mean drag force and inhibits vortex formation in the near wake.
A more accurate wording of the latter eect is that vortex formation is postponed to
far downstream, where it becomes ineective in feeding back substantial forcing [121].
It was experimentally observed by Matsumoto [101], that when dry galloping occurs,
2.2. Circular galloping: myth or true? 31
Figure 2.13. (a) The axial ow, evi-
denced by light ags positioned inside the
wake, act towards inhibiting communica-
tion between shear layers and promoting
a secondary circulatory ow. The func-
tion described, simulates the galloping of
a circular cylinder equipped with a long
splitter plate. (b) Enhanced vortices are
produced when axial vortices from the in-
clined cable, mix and interact with ordi-
nary K arm an vortices. Adapted from Mat-
sumoto et al. [94, 119, 123].
classical vortex shedding has become weak and intermittent. As a matter of fact it was
also stated that K arm an vortices are capable of suppressing large low frequency mo-
tion, but no supporting explanation was given on this. In the latest theory amendment,
Matsumoto et al. [122] consider vortex mitigation to be an indication of the axial ow
intensity. When axial ow is strong then it faithfully resembles a long rigid splitter
plate that will completely inhibit the vortex forcing and readily cause galloping. On
the other hand the vortex appearance would mean that the axial ow is closer to a
less ecient perforated splitter plate, which could even become unable to set o the
instability. Subsequently unstable and diverging galloping are distinguished in terms
of the ecacy of the splitter plate analogue.
Vortex type response
This type of response is more ambiguous and probably controversial. Its original con-
ception was based on the observation that for both wind tunnel tests and eld record-
ings, large events seemed to cluster at discrete reduced velocity ranges, at around
20, 40, 80, 120 etc. These gures look like following a certain pattern of multiples
of the reduced velocity that corresponds to ordinary shedding (i.e. 5). Still, no ev-
ident reason existed for this connection. The work of Bearman and Tombazis [124]
around the same period, provided a plausible explanation. They introduced a mild
three-dimensionality in the wake of an ellipse with a blunt trailing edge and acquired
32 Chapter 2. The Aeroelasticity Framework
wake velocity PSDs with span-wise distinct frequency peaks. The spatial transitions
between alternate shedding frequencies were accommodated by so-called vortex dis-
locations (or splittings). These dislocations were then associated a characteristic low
frequency of switching-states. Matsumoto et al. [120] suspected that a similar source
for low frequency loading could exist behind cables. Evidently the axial ow has to
become a vortex ow in this scenario. Acquiring wake velocity PSDs, Matsumoto et al.
found that even for a non-yawed cylinder there is a slight shedding variation along the
length. This among other justication, encouraged the belief that even normal to the
ow cylinders, latently have the ability of producing large response. In this explanation
attempt however, the frequency forcing characteristics in all cases could not be proved
to follow as submultiples of ordinary vortex shedding. Bearman and Tombazis with
their imposed three-dimensionality, could accurately control their dislocations posi-
tions but in Matsumotos et al. cables, dislocations, if any, are randomly distributed.
Thus although promising, this mechanism was abandoned for not tting the specic
details. Yet the idea of a vortex type response was not invalidated.
Matsumoto et al. [123] came back soon afterwards claiming that the mechanism
could be founded on subharmonic resonance. As earlier noted Durgin et al. [27] pre-
viously observed strong 1:3 resonant vortex shedding response for a vertical cylinder.
Shirakashi et al. [125] argued that this was only an end eect, but Matsumoto et al.
married the two views and showed their applicability on inclined cable test results.
Flow visualisations presented an axial vortex originating from the upstream end and
propagating towards the cables middle. Once shed it interacted with ordinary K arm an
vortices for enhancing every third produced vortex. Taking into account that for an
inclined cable Sr 0.15, the quoted timeliness results U
r
=20, giving a well match to
the earlier said reduced velocity ranges. An illustration of the above description is given
in Fig. 2.13(b). Subsequently this mechanism successfully entered design guidelines for
bridge cables [49]. Two marked features found, are that turbulence can enhance this
instability and that axial ow is again the regulating parameter. When the latter is
increased diverging galloping will emerge. This makes disputable the actual distinction
between unstable galloping and the vortex-response type, which seem to overlap (at
least for inclined sections). As a matter of fact in his later published work Matsumoto
et al. [122] seem to view this mechanism as redundant. Thus all comes down to the
simple rule that complete mitigation of vortex shedding designates diverging galloping
while intermittent and partial mitigation unstable galloping. Even responses found
inside the critical Re region are considered to be due to the K arm an vortex weakening
that takes place. Intriguingly, bluntly turning this rationale to prevention measures,
it could mean that vortex shedding suppression devices are galloping inducers. This
norm is obviously served by ordinary splitter plates.
2.2. Circular galloping: myth or true? 33
Yeo and Jones [116, 117] and Zuo and Jones [118] are still ardent advocates of
the vortex type response. According to them the phenomenon is exclusive to inclined
cylinders and will be controlled by reduced wind velocity. Actually suggesting control
from reduced wind velocity, is another way of stating that any motions occurring, are
primarily forced vibrations. Yeo and Jones employed a hybrid numerical turbulence
modelling scheme to simulate the ow past a skewed horizontal circular cylinder at
Re=1.410
5
(calculated on freestream velocity), and recover the aerodynamic force
functioning. Increasing skew angle up to a critical value (with the ow thereafter fun-
damentally changing), K arm an shedding is gradually suppressed and the axial velocity
component rises. The weakened shedding interacts with the axial ow and develops so-
called swirling structures, advancing along the length in organised patterns. Shedding
among others, would also have to serve as a signal carrier. The resultant travelling
forces, when considered sectionally, have PSDs with low frequency content, and look
both frequency and amplitude modulated. With added skewness, the lowest acquired
peak in the across-wind PSDs, draws away from the ordinary K arm an value. It is pos-
tulated that such forcing component starts up a long period vortex resonance, which
could further get amplied due to motion. Eventually it is suggested that an ecient
plan for counteracting the whole process is to fully eliminate K arm an vortex shed-
ding. This will deprive axial ow from the potential to nucleate swirling structures.
Awkwardly this is exactly the opposite from what Matsumoto et al. [122] advise.
Critical Reynolds number and galloping
In the previous section, critical Reynolds number was assigned a secondary role in cable
instabilities inuencing them only due to contributing to vortex shedding mitigation.
Yet historically there have been cable-like examples which ideally t reasoning exclu-
sively on Reynolds number. The high Re, along-current vibrations of an oil jetty at
Immingham [126], could convincingly be captured if the instantaneous drag force act-
ing on it, steeply decreases with Re, exactly as shown for the mean C
D
in Fig. 2.12.
Steam generator tubes in a number of nuclear power plants had identical issues [127].
Martin et al. [128] modelled in a quasi-steady way the ow-structure interaction and
obtained what looks like the benchmark mass-on-moving-belt problem [129]. Imagine a
circular cylinder oscillating while sweeping a smooth critical drag drop region. During
the part-cycle that it accelerates against the ow, Re becomes larger due to the relative
velocity increasing. Therefore, the drag force acting on the cylinder has progressively
smaller values, establishing a drag dierential that points in the direction of the cylin-
der motion. Likewise, when the cylinder velocity reverses, the new dierence in drag is
again in the direction of motion. This process will continuously pump energy into the
34 Chapter 2. The Aeroelasticity Framework
system until a stable limit-cycle is reached. Being in need of negative dC
D
/dRe, such
oscillations are evidently feasible only in the critical Reynolds number range.
The unusual feature with the similar vibrations of stranded power conductors over
the River Severn, was that motion could also be close to the vertical plane [130].
Richards [131] showed that Re related, shape-induced lift in skew winds and its subse-
quent changes (i.e. derivatives) would have a crucial function in the aeroelastic forcing.
Covering the strands with tape to form a smooth nish the instability disappeared.
Prophetically probably, Richards also warned about the actual non-zero lift measured
on the modied smooth cables. Macdonald et al. [132] applied a newly proposed gen-
eralised galloping theory [133135], which quasi-steadily accounts for Reynolds number
eects, and analytically proved the source of the stranded Severn conductor problem
to lie on excessively negative dC
D
/dRe and C
L
dC
L
/dRe terms. Note that charging
the instability on Re derivatives could not work for purely across-wind vibrations.
Re would then remain unchanged, thus any Reynolds eect in a rst approximation
negates. On a plain circular section steep gradients in both C
D
and C
L
exist at Re
nominally around 1-310
5
, where laminar separation bubbles start forming. Accord-
ingly excessive force derivatives with respect to Re can potentially cause the same type
of galloping instability found in stranded cables.
However, there is much hesitation in accepting Re eects as a major piece in the
dry galloping puzzle. One of the main arguments against this is that Reynolds numbers
in most recorded events fall short for being characterised critical. Typically they are
of the order of 10
4
, which with some (but not absolute) certainty classies them as
subcritical. A second more insightful reason has to do with a fundamental objection.
When a system approaches a critical transition point, as this can be envisaged the case
here, it will get increasingly slower in recovering from small perturbations [136]. At any
instant the dynamic state carries the additional inuence of adjacent (or even far apart)
previous states, establishing a memory eect. Quasi-steady theory, which idealises the
instantaneous force to be only a function of the instantaneous relative velocity, cannot
naturally cope with this requirement. Put into simpler words, Carassale et al. [137]
used the hysteresis evidence of Kamiya et al. [104] and Schewe [106] to question the
inclusion of Re derivatives in quasi-steady formulations. Such derivatives could be
non-unique or even indeterminate. However, they also utilised purely geometric quasi-
steady theory inside the critical Re region, where in principle this could be equally
inapplicable. Promoting a simple Reynolds number explanation for dry galloping, is
thought by many equivalent to proving the quasi-steady theory validity throughout
the critical range. Concluding this part, a paradoxical feature should be noted. The
strands-induced instability of the Severn conductor was cured by surface smoothening.
In modern cable-stayed bridges the instability-control tactics are to eectively roughen
2.3. State of the art in bridge wind design 35
the already smooth cables. More intriguingly the roughening measures could even seem
to resemble the old conductors shape, making up a clear inconsistency as commented
by Tanaka [138].
This extensive review of proposed mechanisms is a sine qua non for the scope of the
current thesis. The many contradicting elements presented, not adequately stressed
in many points, are the best indication that phenomena for which theory is only now
getting shaped are addressed. The diversity shown in explanations, dominates also
in eld-observations. Records spread over a wide range of wind conditions, geometric
congurations, and structural characteristics perplexing any analysis. At the Iroise
Bridge, a long monitoring campaign indicated large, unforeseen cable vibrations in
what categorically was said to be the critical Re range [139]. On the other hand Zuo
and Jones, reporting on the monitoring of the Fred Hartman and Veterans Memorial
Bridges [140], identify all sorts of large cable vibrations but none in the nominal critical
Re regime. Interestingly they record that a particular non-ordinary-K arm an event type
occurs for similar wind conditions, with similar modal features, for both dry and rainy
conditions. A connection is made with U
r
, which seems to consistently fall near 40,
and subsequently the scenario that rain-wind and dry-wind vibrations are the same
forced phenomenon is advanced. Finally, Matsumoto et al. [101, 122, 141] summarise a
number of large-scale cable incidents in a selection of Japanese bridges. Most of their
events have Re>10
5
, concentrate in a narrow dened cable-wind angle domain (i.e.
near 60

), refer to single low mode motion and occur at U


r
>100. Ordinary such high
reduced wind velocities would designate indubitable prevalence of quasi-steady theory,
but in this instance there also seem to be complications from complex unsteady eects.
2.3 State of the art in bridge wind design
After a long section focusing on aerodynamic issues only recently being explored, it is
time to illustrate how todays bridge engineering design deals with them as well as with
the earlier presented older and more rigorously studied wind problems. The biggest
concern in bridge aerodynamic performance is utter. Its interaction with gust loading
is inevitable in any real-case consideration. There are yet more interaction phenomena,
and fortunately many of them have reached an elaborate state of treatment. To this
end only the basic framework for combined utter-bueting evaluations, essential for
the following chapter, will be given.
In 2.1.5, utter was discussed in view of sectional aeroelastic forces, but no connec-
tion was made to the estimation of stability limits when a full-bridge, with subsequent
multi-modal behaviour is the application in hand. Following the succinct derivation of
36 Chapter 2. The Aeroelasticity Framework
Jain et al. [142], for the most general case of a geometrically complex bridge with mixed
modes, analysis proceeds as follows. The deection components, shown in Fig. 2.7(b),
are written by use of the dimensionless mode shapes h
i
(s),
i
(s), p
i
(s) as
h(s, t) =

i
h
i
(s)Bq
i
(t) , (s, t) =

i
(s)Bq
i
(t) , p(s, t) =

i
p
i
(s)Bq
i
(t) ,
(2.12)
where s is the distance along the deck span and q
i
is the ith mode generalised displace-
ment. The equation of motion becomes
I
i
_
q
i
+ 2
i

i
q
i
+
2
i
q
i

= Q
i
, (2.13)
where
i
is the modal cyclic frequency, I
i
is the generalised inertia and Q
i
is the
generalised aerodynamic force. The latter two are given by
I
i
=
l
_
0
_
m(s)h
2
i
B
2
+I(s)
2
i
+m(s)p
2
i
B
2
_
ds ,
Q
i
=
l
_
0
_
Lh
i
(s)B +M
i
(s) +Dp
i
(s)B
_
ds , (2.14)
m(s) and I(s) being the mass and mass moment of inertia (about the sections centre
of gravity) respectively and l is the deck span length. Let the lift, moment and drag per
unit span be linearly decomposed into the sum of self-excited and bueting components
L = L
se
+L
b
, M = M
se
+M
b
, D = D
se
+D
b
. (2.15)
In general there is also self-induced bueting action, which is herein discarded. For
the self-excited parts, Eqs.(2.11) are employed additionally assuming that all utter
derivatives are constant along l. For the modal integrals it is written
l
_
0
h
i
(s)a
j
(s)
ds
l
= G
h
i

i
, (2.16)
together with the rest ve obvious h
i
(s),
i
(s), p
i
(s) permutations. When Eqs.(2.13)
are Fourier-transformed into the reduced frequency (K) domain they can be expressed
in matrix form as
Eq =

Q
b
, (2.17)
2.3. State of the art in bridge wind design 37
where denotes the Fourier transformation. The components of E and

Q
b
are given
by
E
ij
= K
2

ij
+KA
ij
(K) +B
ij
(K) ,
A
ij
(K) = 2
i
K
i

ij

B
4
lK
2I
i
_
H

1
G
h
i
h
j
+H

2
G
h
i

j
+H

5
G
h
i
p
j
+A

1
G

i
h
j
+A

2
G

j
+A

5
G

i
p
j
+P

1
G
p
i
p
j
+P

2
G
p
i

j
+P

5
G
p
i
h
j

,
B
ij
(K) = K
2
i

ij

B
4
lK
2
2I
i
_
H

4
G
h
i
h
j
+H

3
G
h
i

j
+H

6
G
h
i
p
j
+A

4
G

i
h
j
+A

3
G

j
+A

6
G

i
p
j
+P

4
G
p
i
p
j
+P

3
G
p
i

j
+P

6
G
p
i
h
j

Q
b
i
=
B
3
U
2
I
i
l
_
0
_

L
b
h
i
(s) +

M
b

i
(s) +

D
b
p
i
(s)
_
ds , (2.18)
with
ij
=1 for i = j,
ij
=0 for i = j.
The multi-modal utter critical condition is determined by solving the homogeneous
equivalent of Eq.(2.17). The evaluation consists of varying K and deriving dierent
sets of . The values of K and for which both the real and imaginary parts of
the determinant of matrix E become simultaneously zero, are the eective negative-
damping thresholds. The minimum wind speed calculated by these pairs, dene the
utter velocity. Thereon expressing Eq.(2.17) in PSDs form and utilising standard
concepts of random vibrations, the characteristics of the combined bueting response
can be estimated. As seen by considering only the homogeneous solution, the utter
condition does not include explicitly the bueting inuence. A complimentary analysis
accounting also for bueting was presented by Scanlan [13]. He considers the classical
case of two interacting modes and calculates the average rate of change in their total
energy. This turning to positive may be taken as a sign of instability. Additionally a
slight variation regarding the generalised self-excited forces was proposed. According
to it, the utter derivative products of A
ij
(K) and B
ij
(K) in Eqs.(2.18) are no more
functions of a single K. Instead they are evaluated at the aerodynamically modied K
j
corresponding to the relevant participating mode j (see also [12]). This latter analysis
variant will be used in an inverse way for estimating the values of utter derivatives
for the CSB in the next chapter. The described utter framework found application in
the latest and most important bridge designs, in-short referring to the examples of the
Storeblt Bridge [143], the Akashi Kaikyo Bridge [15] and the Messina Strait planned
bridge [144].
A great deal of specialised analysis also exists for the rest individual elements of a
bridge, and particularly for the versatile cable-stays. Treatments are mostly based on
quasi-steady aerodynamic theory with any unsteady concerns being conned to ordi-
38 Chapter 2. The Aeroelasticity Framework
nary K arm an vortex shedding. As illustrated in 2.1.7 and 2.2.3, for typical circular
cables, many alarming phenomena occur beyond the reach of classical galloping and
classical vortex shedding. Subsequently they would be unaccounted in any typical de-
sign. Not having as yet hard evidence on the forcing attributes of such instabilities, the
only true solution for engineering practice is to recur to experiments. This obviously
serves both current and future design, since an empirical basis is constructed which
could then lead to the development of analytical tools. The core knowledge in prac-
tically dealing with dry cable galloping is expressed by Fig. 2.14, where curves in the
S
c
U
r
parameter space bound the regions of safety. The initial dynamic tests of Saito
et al. [145], on an inclined bridge-cable replica, designate that the necessary S
c
(mainly
seen as a structural damping requirement) for restraining vibrations is approximated
by 35

S
c
= U
r
. Such a connection implies that cables will always become unstable for
suciently high U
r
. The unrealistically high damping values imposed for most cases by
the Saito et al. relation, led to a new test campaign sponsored by the Federal Highway
Administration (FHWA) [146]. FHWA results, indicate that the threat should partly
be released since for Sc>3 no large events were recorded. The latter is reminiscent of
the actual design guideline against rain-wind phenomena. There Sc>10 [138], similarly
on purely empirical grounds.
Figure 2.14. Dry cable instability design criteria together with real-bridge unstable records.
Dotted lines are due to the uncertainty in dening structural damping values. Adapted from
Matsumoto et al. [122, 141] and Kumarasena et al. [146]
Thus the current state of understanding comprises two far apart instability criteria
without clear indication of which is the most appropriate. Large-scale records from
real bridges [122, 141], due to uncertainty over the exact structural damping values,
deceivingly seem to agree with both. Recently Matsumoto supported that the two cri-
2.4. Concluding Remarks 39
teria should be equally valid [122]; Saitos et al. criterion designates unstable galloping
(cf.2.2.3), while the FHWA criterion diverging galloping. FHWA data near S
c
=4.5 in
Fig. 2.14 do not seem to agree with this assertion. It should be mentioned that Fig. 2.14
crudely groups points corresponding to dierent congurations, and most importantly
that some of these (especially unstable ones) may be non-repeatable.
In 2.1.3 it was noted in passing that alternative numerical methods can be used for
dealing with the wind-structure interaction modelling. The two alternatives quoted (i.e.
discrete vortex models and solution of the Navier Stokes equations) have many variants
depending on their computational implementations. In any case, for bridge structures
where the combination of wind conditions and size results in very high Reynolds num-
bers, there are noticeable diculties with the operation of such schemes. Discrete
vortex methods suer from inconsistencies due to the poor knowledge of how to ad-
just the vortex arrangement to accurately describe the turbulent ow. On the other
hand Navier Stokes solvers need to resolve a wide range of turbulence scales, which
numerically becomes extremely tedious. Often simplications are put forward such as
time or spatial averaging of the equations, which necessitate explicit turbulence mod-
els. Tuning turbulence models for large separation problems remains ambiguous and
introduces non-robustness into calculations. A detailed review of functioning charac-
teristics and specic attributes of the methods is outside the scope of the current thesis.
Yet it is informative to briey refer to a number of bridge applications that illustrate
the use and value of these analytical tools. For the utter analysis earlier presented
the utter derivatives, which are yet to be determined, are possible to calculate with
any of the two methods. Larsen et al. [143, 147] and Starossek et al. [148] numerically
established sensible agreement with experimental utter derivatives for a large number
of both streamlined and blu cross sections. For the circular cylinder ow the near
critical Reynolds regime still remains rather challenging to simulate. Recently Yeo and
Jones [116, 117] have treated the high Re inclined cylinder ow and their results are
of specic interest to the current study. A general conclusive point to be drawn out
is that although the trust in such computational methods has seriously increased over
the last two decades, they are still in need of validation. The simplications and as-
sumptions made in order to reduce computational costs create uncertainties that can
only be waived through comparisons with experimental results.
2.4 Concluding Remarks
Apart from laying the ground for the analysis to follow, this chapter additionally tried to
establish questions. All of them should naturally fall into the broadest themes of: How
40 Chapter 2. The Aeroelasticity Framework
ecient and reliable is the current state of knowledge? and Are there new elements
to complement this knowledge?. This thesis will attempt to touch upon both.
As presented, the best available defence that bridge engineering has against wind
is multi-modal utter analysis. It aims primarily at keeping sound the deck, which
is probably the heart of a bridge. Although the method has reached a great level
of sophistication, concerns will always exist as to whether results are representative
and realistic. This is not surprising for a method that has developed inside wind
tunnels, based on scaled models. Jones et al. [149] and Katsuchi et al. [150], quote the
signicance that full-scale ambient recorded data should have in a verication scheme.
Actually modern bridges cope quite well with utter. Any size-eect or inconsistency
of modelling, if is there, does not seem to show up due to the structure operating really
far from the conditions where the phenomenon was estimated to unfold. Unfortunately
the best available real-scale example that could, when well instrumented, give us a
great lesson on utter and expose potential aws, lies 80m below water in the bottom
of Puget Sound (i.e. Tacoma Narrows). In any case, all monitoring attempts should
be seen as a chance of putting more reality into bridge modelling. This is exactly the
objective sought in the following chapter.
Further, it is also essential to understand the limitations and shortcomings of other
analytical tools currently in hand. Simple details (e.g. the exact geometric arrange-
ment), when correctly accounted for, can bring straightforward explanations on what
might have looked as an unforeseen aerodynamic event. And this is by no means an
easy task. In the context of this research, the quasi-steady galloping framework is re-
visited in order to produce an original contribution that recognises the true limits of
the galloping analysis and points out all past omissions and defects. Still there are also
truly new wind-structure interaction phenomena that cannot be cast into the existing
knowledge. For a latest F/A-18 ghter jet, unexpected unsteady events rose from the
newly shaped wings [151]. Similarly in long-span bridges new events of alarming am-
plitude came through their smooth cables, which were paradoxically shaped like that
to prevent old well-known dynamic instabilities. A number of proposed mechanisms
were devised to explain these large cable vibrations, hoping to become precursors of
a better analytical framework. Intriguingly, very few common grounds can be found
between dierent approaches. As illustrated, they seem to disagree even on basic prin-
ciples, often leading to exactly opposite results. The wind tunnel tests presented and
analysed in a later chapter, attempt to oer a dierent interpretation of this unique
phenomenon. The critical Reynolds number range by means of the associated com-
plex transitional behaviour is carefully examined to reveal all the attributes compatible
with the instability. Stepping towards a holistic approach of the aerodynamic bridge
2.4. Concluding Remarks 41
design it is essential that the unique parts that compose it are adequately claried and
understood. And this is exactly the notion that this thesis attempts to serve.
Chapter 3
Identication of utter derivatives from
full-scale data
The estimated response of large exible bridges to severe wind loads is prone to mod-
elling uncertainties that can only ultimately be assessed by full-scale testing. To this
end, ambient vibration data from full-scale monitoring of the historic Clifton Suspen-
sion Bridge (CSB) have been analysed in order to capture elements of true wind-bridge
interaction. The multi-modal utter framework earlier presented, is herein employed in
an inverse investigation. Flutter derivative identication, which has rarely previously
been attempted on full-scale data, was performed to seize any trends towards aerody-
namic instability. The chapter does not intend to be a meticulous dynamic description
of CSB. Instead, a number of useful notes for todays aeroelasticity will be drawn out,
while there is a clear potential for the old outdated CSB to become the test bed for
future advances.
3.1 Introduction
For large-scale structures the most rational way to proceed with predictions on the
reliability and operational safety, includes identication methods from response only
measurements. Especially for existing bridges, the risk of utter can substantially be
veried in this way. A blu bridge cross-section, unlike a at plate or an airfoil, has
no analytical expression for the uid forces exerted on it while in motion. Identifying
the critical wind speed for instability inevitably has to adopt some experimental, semi-
empirical or numerical foundation. Most commonly wind tunnel tests of scale models
are used for reproducing the utter phenomenon leaving the question of the eects of
scaling. It is well established that minor details such as deck railings or roadway grills
and vents can strongly alter the aerodynamic performance (see Scanlan and Tomko [73],
43
44 Chapter 3. Identication of utter derivatives from full-scale data
Jones et al. [152] and Matsumoto et al. [153]). Hence, analysis of the response of the real
bridge can clarify the validity of wind tunnel tests and even reveal aspects, which either
due to modelling assumptions or to loading irregularities, were previously concealed.
Aeroelastic parameters have rarely been obtained from full-scale bridge data. Okau-
chi et al. [154] were the rst to attempt something relevant. Building a bridge section
model (at a large scale, roughly 1/10) and setting it on-site against real wind conditions,
they compared results with smaller wind tunnel equivalents. They suggested that, al-
though for turbulent ow the relative dierences in the turbulence details can impose
inconsistencies, in general wind tunnel models produce representative results. Jakobsen
and Larose [155] addressed the problem on the H oga Kusten Bridge and presented a
comparative analysis with wind tunnel results using a subspace identication technique
for extraction of utter derivatives. Costa and Borri [156] essentially applied the same
identication routine, both on numerically simulated responses and on measured data
from the Iroise Bridge, quoting good performance of the method in each case. For all
of these bridge studies, the identication routine itself was found to be reliable, when
tested using simulated data produced with added variously coloured noise. Compar-
isons between full-scale and wind tunnel results were not unreasonable, but since the
full-scale bridges were far from utter, the trends in utter derivatives were not clear.
Another approach to the problem of identifying the aerodynamic eects on full-scale
bridge vibration characteristics was used by Macdonald [157] on the Second Severn
Crossing. Variations of eective damping ratios and natural frequencies with wind
speed were found, and some indications of aeroelastic modal coupling were identied on
the partially constructed bridge. In other full-scale studies, Littler [158] and Brownjohn
[159] on the Humber Bridge, Bietry et al. [16] on the Saint-Nazaire Bridge, Jensen et
al. [160] on the Great Belt Bridge, Ge and Tanaka [161] on the H oga Kusten Bridge
during construction and Nagayama et al. [162] on the Hakucho Bridge all found some
trends of eective aerodynamic damping with wind speed, but coupling between modes
and utter derivatives were not pursued.
The limited number of full-scale studies from which aeroelastic parameters have
been found, makes any new cases useful for extending knowledge on the viability of
system identication from site data and for interpreting actual bridge behaviour.
3.2 The case study
As part of this work, analysis is performed on full-scale vibration measurements from
the historic CSB, shown in Fig. 3.1. The CSB spans the Avon Gorge in Bristol, UK
and was designed by I.K. Brunel in 1830, although it was not completed until 1864
3.2. The case study 45
(Barlow [163]). It was one of the longest suspension bridges of that time, with a main
span of 214m. Wrought iron chains provide the suspension system, being the common
practice for such early long-span bridges. In the light of modern understanding of
bridge aerodynamics, the bridge cross-section (Fig. 3.2) and its light weight make it
potentially susceptible to wind-induced vibrations. Indeed, on a few occasions in its
lifespan large amplitude vibrations in strong winds have been reported.
FIGURES
Leigh
Woods
Clifton
Maintenance
craddle
Accelerometer reference cross-section
Accelerometer cross-section
Anemometer
214m
80.2m 107m
46m
Figure 3.1. Bridge elevation showing instrument locations. Based on gure after Barlow [163],
with permission from Thomas Telford Publishing.
On Christmas Day 1990 there was evidence of vertical motion at the bridge ends
of the order of 250mm, which translates to even larger amplitudes within the bridge
span. Both vertical and torsional deck motions were evident on a video recording of
the bridge towards the end of a storm. A similar large vibration event was reported
on 3 December 2006. Although no wind recordings exist from the bridge site itself on
these occasions, data from the nearest weather stations imply that the maximum 1h
mean wind speeds could have been around 20m/s. For recordings on site with wind
speeds up to 16m/s, coupling action between the rst vertical and torsional modes
seemed to occur and the maximum vertical displacement at the ends of the bridge was
35mm (maximum measured elsewhere 92mm). The coupling action between modes
and the rapid growth of vibration amplitudes for a modest increase in wind speed,
indicate strong aeroelastic eects and make the bridge behaviour rather interesting.
Such characteristics are reminiscent of features observed, in a more severe form, on the
Tacoma Narrows (Farquharson et al. [164]) and Deer Isle (Kumarasena et al. [165,166])
bridges.
It is worth noting that ten suspension bridges from the same era as the CSB failed
due to wind between 1818 and 1889, including the Menai Straits Bridge, the Wheeling
Bridge and a span of the Brighton Chain Pier (Farquharson et al. [164]). In contrast
the CSB has survived virtually unscathed for over 140 years. According to empirical
46 Chapter 3. Identication of utter derivatives from full-scale data
Figure 3.2. Sketch of the bridge cross-section.
estimates, similar in most bridge design rules, it is potentially susceptible to utter with
an estimated critical wind speed of not more than 20m/s. Therefore adverse aeroelastic
eects are expected to become signicant in moderately strong winds. For the current
study, the wind conditions and bridge response recordings over two winter periods, from
November 2003 to March 2004 [167] and from December 2007 to February 2008, are
used. The site monitoring was conducted by Macdonald initially under a contract from
the Clifton Suspension Bridge Trust. Later the monitoring was continued for research
purposes with the assistance of the author. The data include several occasions with
moderately strong winds, and reasonable ranges of wind speeds and directions, thus
enabling a meaningful assessment of the wind eects on the bridge dynamics.
Two ultrasonic anemometers were mounted 61m either side of midspan, more than
5m above road level. Nine accelerometers were positioned at a series of cross-sections
along the bridge during an earlier analysis of the CSB dynamic response, enabling mode
shapes to be identied (Macdonald [168]). The rst three vertical and torsional modes
are shown in Fig. 3.3. For the records considered here, two sets of three accelerome-
ters were positioned at midspan and at a cross-section slightly o centre (26.8m from
midspan) as illustrated in Fig. 3.1. This location was chosen as the reference cross-
section since all signicant vibration modes could be measured there. All motion related
measurements below, refer to this location. Signals from all instruments were passed
through anti-aliasing lters with a cut-o frequency of 4Hz and were recorded at a
sampling rate of 12.5Hz. The primary aim here is to uncover the underlying mech-
anism of large amplitude response that the bridge was found to produce for certain
wind conditions, and explore the utter potential. To do this the variation of modal
characteristics with wind velocity has to be determined.
Modal parameter estimates from a frequency based curve tting technique (devised
by Macdonald [157]) are used here, together with a subspace stochastic identication
formulation especially modied to extract utter derivatives (Jakobsen [169]). Due
3.3. Wind characteristics 47
f
v1
=0.293Hz,
f
v2
=0.424Hz,
f
v3
=0.657Hz,

v1
=3.31%

v2
=1.99%

v3
=2.12%
f
t1
=0.356Hz,
f
t2
=0.498Hz,
f
t3
=0.759Hz,

t1
=2.60%

t2
=3.44%

t3
=2.16%
Vertical Torsional
Figure 3.3. Experimentally obtained vertical and torsional modes for CSB, adapted from
Macdonald [168].
to lack of wind tunnel data from a scale model, utter derivatives of other typical
bridge cross-sections, as presented by Scanlan and Tomko [73], are used for comparative
assessments. Cross-sections employed for this purpose are chosen to represent both the
low structural depth and the high parapets (perforated on the CSB), of the section in
hand.
In the following section, rst a short background of the acquired wind measure-
ments is given. Typical wind speeds and wind turbulence conditions are described and
the local terrain eects are discussed. Subsequently attention is moved to the bridge
response details. All necessary modal background is provided, and the wind parameter
on them is distinguished. The last part containing the utter derivative identication
scheme, starts by shortly commenting on the employed Covariance Block Hankel Ma-
trix (CBHM) formulation. Eventually a utter velocity estimate is sought to compare
with theoretical predictions.
3.3 Wind characteristics
The topography around the CSB has a considerable eect on the local wind attributes.
As shown in the polar plots from both anemometers in Fig. 3.4, the wind speeds follow
certain trends with orientations. (True North is at 31

relative to the axes shown).


These trends dier markedly from the general wind pattern in the region away from the
local inuences. The strongest winds in the absence of topographic eects are typically
from the south-west direction (at about 250

on Fig. 3.4 axes). The on-bridge measured


stronger winds, are aligned along the gorge and can be probably explained by funnelling
of the ow in these orientations and sheltering due to the high ground near the bridge
ends. Evidently strong winds from the south-west are greatly attenuated at the bridge
site, and virtually no wind from the north-east quadrant is experienced. It was also
48 Chapter 3. Identication of utter derivatives from full-scale data
observed that the correlation of the wind directions and wind speeds measured by the
two anemometers, was strongly inuenced by the wind orientation. In particular, for
wind directions close to along the gorge even a small change in wind direction, results
a large but consistent variation in the ratio between the two individual anemometer
wind speed values.
A typical assumption is made that that wind loading is approximately stationary for
records up to one hour [2]. The maximum wind speed, averaged over one hour (unless
stated all wind information hereafter refers to 1h means), did not exceed 16m/s at the
bridge site, although higher speeds were measured at the nearest weather stations, for
winds from the south-west. A histogram of 1h average wind speeds at the bridge is
given in Fig. 3.4. Maximum 1s gusts were of the order of 26m/s for both anemometers.
In addition, the wind turbulence and angle of attack parameters were considered.
For wind turbulence there was a strong dependence on wind direction and a weaker
one on wind speed. High levels of turbulence (up to 60% longitudinal turbulence
intensity, I
u
) were experienced, particularly for wind not along the gorge and for lower
wind speeds. In winds over 8m/s, which only occurred along the gorge, approximately
normal to the bridge, the mean longitudinal turbulence intensity was 21% and the mean
vertical turbulence intensity I
w
, 10%. The vertical and across-wind (I
v
, in the deck-
wind plane) turbulence intensities followed very similar patterns to the longitudinal
turbulence. For longitudinal turbulence up to 40%, the across-wind turbulence was
approximately equal to it and the vertical turbulence intensity approximately half the
value. These are typical relationships between the three components of turbulence. For
higher turbulence intensities measured, generally in lower wind speeds, the vertical and
across-wind turbulence intensities were relatively larger. For the vertical angle of attack
there was also strong dependence on the wind direction, and there were noticeable
dierences in the measurements from the two anemometers. The presence of the bridge
itself is likely to have aected these measurements, as well as the topography of the
gorge, since the anemometers were relatively close to the deck. High vertical angles of
attack occurred, up to approximately 10

. It should be reminded that these values are


averaged over one hour periods. There was no signicant dierence in vertical angles
of attack for dierent wind speeds.
A wind aspect signicant for the subsequent analysis, refers to the frequency content
of wind bueting. Although the trac loading seems to be reasonably well captured by
a white noise approximation (predominantly from step loading as vehicles drive onto
or o the suspended span), the same does not hold for wind. By comparing spectral
estimates deduced for various combinations of wind and trac, it was found that above
3.3. Wind characteristics 49
Figure 3.4. (a) Histogram of wind speeds during the 2003-04 recording period. (b) Polar plots
of 1h mean wind velocities from both anemometers.
50 Chapter 3. Identication of utter derivatives from full-scale data
approximately 0.25Hz the wind loading spectra tailed o as f

with around -8/3,


producing a general loading spectrum of the form

S
load
(f) =

S
w
f
8/3
+

S
t
, (3.1)
where

S
w
is a constant for a given record, being a function of the wind parameters,
and

S
t
is the magnitude of white noise trac loading specic for each record. The
frequency power exponent of -8/3 when compared with the -5/3 value corresponding
to isotropic K arm an turbulence (referring to both I
u
and I
w
) customarily employed in
design, it implies that the product of the aerodynamic admittance and joint acceptance
functions in Eq.(2.2) should be inversely proportional to f.
3.4 Response and modal parameters
3.4.1 Response Characteristics
Fig. 3.5 shows the 1h average wind speeds over the whole of the rst monitoring pe-
riod, and the corresponding Root Mean Square (RMS) vertical accelerations at the
reference cross-section. The RMS amplitudes normally show a clear daily cycle with
the varying trac load, with a maximum vertical response of approximately 0.02m/s
2
.
By comparison it can be seen that only in wind speeds exceeding approximately 8m/s
does the response noticeably exceed the maximum trac-induced response. The maxi-
mum wind-induced acceleration measured was approximately four times the maximum
trac-induced acceleration. The torsional and lateral acceleration responses at the
reference cross-section followed very similar patterns to the vertical response over the
monitoring period, although the magnitudes of the responses were lower. The maxi-
mum instantaneous value of each component was found to be approximately six times
the 1h RMS value.
Dynamic displacements were calculated from the measured accelerations by double
integration and it was noticed that the response is dominated by low frequency modes.
The dominance of the low frequency modes is caused by the relatively higher wind
loading at low frequencies and the eect of the integration, which exaggerates low fre-
quency components. Whereas the maximum RMS acceleration due to wind loading was
approximately four times the maximum due to trac loading, in terms of displacement
the maximum RMS response to wind was approximately 10 times that for trac.
The inuence of wind loading on the measured vertical accelerations is shown in
Fig. 3.6. Similar gures were obtained for the lateral and torsional accelerations. The
3.4. Response and modal parameters 51
10 17 24 1 8 15 22 29 5 12 19 26 2 9 16 23 1 8 15 22
0
2
4
6
8
10
12
14
16
November December January February March
Date (2003-04)
1
h
r
a
v
e
r
a
g
e
w
i
n
d
s
p
e
e
d
(
m
/
s
)
(a)
10 17 24 1 8 15 22 29 5 12 19 26 2 9 16 23 1 8 15 22
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0.09
November December January February March
Date (2003-04)
R
M
S
v
e
r
t
i
c
a
l
a
c
c
e
l
e
r
a
t
i
o
n
(
m
/
s
2
)
(b)
Figure 3.5. (a) 1h average wind speed over the monitoring period. (b) 1h RMS vertical
accelerations at the reference location over the monitoring period.
scatter of results, particularly at low wind speeds, is largely due to the varying traf-
c contribution. The varying wind turbulence intensity also proved, as expected, to
have an eect. Excluding records dominated by trac, and normalising by the verti-
cal turbulence intensity, gives a much clearer relationship with wind speed as shown
in Fig. 3.6(b). Vertical turbulence intensity is considered more appropriate for such
scaling, when bridge sections with relatively low mean lift coecient C
L
(cf. Eq.(2.2))
are considered. Normalised RMS responses become a power law functions of the wind
speed as shown in Eq.(2.5). The power exponent is very close to the theoretical 2.83
bueting value, which could imply that response is solely due to bueting. Yet as shown
in Fig. 2.8 for turbulent conditions, a full bridge experiences the combined bueting-
utter action with a power law too. Also it is notable that no sharp peaks, that could
signify vortex-induced response, exist in Fig. 3.6(b).
3.4.2 Modal Analysis
Modal parameters were previously calculated from curve tting acceleration Power
Spectral Densities (PSDs), using the Iterative Windowed Curve tting Method (IWCM)
of Macdonald [157]. IWCM was specically developed for the analysis of ambient
vibration data with in general non-white loading spectrum. The method iteratively
curve ts in the frequency domain a series of idealised single-degree-of-freedom (SDOF)
transfer functions, taking into account the shape of the loading spectrum, the eect of
windowing on the spectra (both from the measured data and from the idealised transfer
functions) and the contributions of multiple modes (in a linear sense).
Measurements were only taken on the suspended bridge deck, but all modes in-
evitably involve vibrations of other parts of the structure, particularly the chains.
52 Chapter 3. Identication of utter derivatives from full-scale data
0 2 4 6 8 10 12 14 16
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0.09
Wind speed (m/s)
R
M
S
v
e
r
t
i
c
a
l
a
c
c
e
l
e
r
a
t
i
o
n
,

v
(
m
/
s
2
)
(a)
0 2 4 6 8 10 12 14 16
0
0.2
0.4
0.6
0.8
1
1.2
Wind speed (m/s)
N
o
r
m
a
l
i
s
e
d
R
M
S
v
e
r
t
i
c
a
l
a
c
c
e
l
e
r
a
t
i
o
n
,

v
/
I
w
(
m
/
s
2
)


12:30am - 6:30am
Wind speed >8m/s
No vehicles
No vehicles or pedestrians
(Wind speed)
2.8
(b)
Figure 3.6. (a) RMS vertical accelerations
v
, in relation to wind speed for all 1h records. (b)
Same as (a) for 1h records dominated by wind loading, with RMS vertical accelerations now
divided by the vertical turbulence intensity. The power-law approximating the obtained trend
is also plotted.
Analysis was performed for frequencies up to 3Hz with twelve vertical, eleven tor-
sional and four lateral modes being identied in this range, based on measurements
in low wind speeds (Macdonald [168]). Typical PSDs for three dierent loading sce-
narios for vertical, torsional and lateral accelerations are given in Fig. 3.7 to present
the eect of wind loading on the bridge response. An important detail is the proxim-
ity of the rst vertical and torsional modes, with natural frequencies of 0.293Hz and
0.356Hz respectively (ratio 0.82). These are the rst antisymmetric modes of each type
as depicted in Fig. (3.3). It appears that in the stronger winds they start to couple
in a possibly incipient utter motion as evidenced by the small hump at 0.35Hz in the
vertical spectrum seen in Fig. 3.7(a) and more clearly in Fig. 3.8.
For further analysis it was desirable to reduce the actual multi-mode response to
a simpler two-degree-of-freedom equivalent, concerning only the relevant rst vertical
and torsional modes. Low-pass ltering could not fully remove the contributions of the
second mode of each type, since they were very closely located. However, using the
measured responses at both accelerometer cross-sections (see Fig. 3.1) together with the
known mode shapes, the following operation was performed. With the second section
in the midspan, the rst pair of modes being anti-symmetric is showing as zero. The
second pair of modes on the other hand, containing symmetric modes that maximise
there, could be readily identied. Transferring these modal displacement values to the
reference cross section and subtracting them from the total signal, results in almost
pure rst mode motions.
Fig. 3.8(a) hence shows responses of the rst pair of modes (z vertical and tor-
sional) for the highest recorded 1h wind speed (15.3m/s). The peak in the vertical PSD
3.4. Response and modal parameters 53
0 0.5 1 1.5 2 2.5 3
10
7
10
6
10
5
10
4
10
3
10
2
10
1
Frequency (Hz)
V
e
r
t
i
c
a
l
a
c
c
e
l
e
r
a
t
i
o
n
P
S
D
(
(
m
/
s
2
)
2
/
H
z
)


Maximum wind, with trac
Minimal wind, Rush hour trac
Moderate wind, only
(a)
0 0.5 1 1.5 2 2.5 3
10
9
10
8
10
7
10
6
10
5
10
4
10
3
10
2
Frequency (Hz)
T
o
r
s
i
o
n
a
l
a
c
c
e
l
e
r
a
t
i
o
n
P
S
D
(
(
r
a
d
/
s
2
)
2
/
H
z
)


Maximum wind, with trac
Minimal wind, Rush hour trac
Moderate wind, only
(b)
Figure 3.7. PSDs for dierent loading condi-
tions for (a) vertical (b) torsional and (c) lat-
eral accelerations.
0 0.5 1 1.5 2 2.5 3
10
7
10
6
10
5
10
4
10
3
10
2
10
1
Frequency (Hz)
L
a
t
e
r
a
l
a
c
c
e
l
e
r
a
t
i
o
n
P
S
D
(
(
m
/
s
2
)
2
/
H
z
)


Maximum wind, with trac
Minimal wind, Rush hour trac
Moderate wind, only
(c)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0
2
4
6
x 10
3
f (Hz)
P
S
D

(
r
a
d
/
s
2
)
2
/
H
z
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0
0.03
0.06
0.09
f (Hz)
P
S
D
z
(
m
/
s
2
)
2
/
H
z
(a)
0 0.1 0.2 0.3 0.4 0.5 0.6
10
4
10
3
10
2
10
1
Frequency (Hz)
V
e
r
t
i
c
a
l
a
c
c
e
l
e
r
a
t
i
o
n
P
S
D
(
(
m
/
s
2
)
2
/
H
z
)


15.3m/s
13.8m/s
11.9m/s
9.8m/s
7.9m/s
6.1m/s
1
st
Torsional mode
(b)
Figure 3.8. (a) PSDs of ltered data for rst vertical and torsional modes for the maximum
wind speed record. (b) The evolution of the coupling action is evident in the vertical PSD for
wind speeds above 11m/s.
54 Chapter 3. Identication of utter derivatives from full-scale data
at the torsional frequency is strong evidence of coupling action. The coupling was not
evident for low winds and became stronger only for higher winds (Fig. 3.8(b)), setting
it most probably to be an aeroelastic eect. The coherence between the vertical and
torsional accelerations showed similar evidence, with the value around 0.35Hz rising to
approximately 0.6 in high winds compared with typical values below 0.4 in low winds
and at other frequencies. It is also worth noting that the second pair of modes showed
a similar tendency for coupling action in strong winds, due to their also similar shapes
(see Fig. 3.3) and close natural frequencies (ratio 0.424/0.4980.85). The next section
discusses in more detail the identication of the CSB utter derivatives, so as to be
able to quantify the observed coupling signs and the tendency towards utter.
3.5 Flutter derivatives
3.5.1 Flutter Analysis
According to the semi-empirical Selberg [170] equation for bridge sections resembling
at plates, an estimate for the utter speed is given by
U
f
f
0
B
= C

r
g
m
B
3
[1 (
f
z0
f
0
)
2
] , (3.2)
where U
f
is the utter speed, B the deck width, r
g
the radius of gyration, m the total
mass per unit length, C a constant depending on the sections behavioural resemblance
to a at plate, the air density and f
z0
and f
0
the still air vertical and torsional
natural frequencies. For the rst pair of natural frequencies, described above, substi-
tuting =1.2kg/m
3
, B=9.46m, C=2, m=5370kg/m, r
g
=4m, the utter onset speed is
estimated at approximately 18m/s, or 14m/s using the more conservative variant of
Eq.(3.2) in the British design rules [171]. These very low values are due to the low
torsional natural frequency, the close neighbourhood of vertical and torsional modes
and the low mass per unit length. However, the bridge cross-section is not a at plate
and the uncertainty in the value of C (containing an experimental correction factor)
means Eq.(3.2) is not very reliable. Still, it gives a rough rst approximation for the
utter speed and this is within the range of wind speeds encountered at the site.
For evaluating the utter behaviour, the classical multi-modal formulation presented
in 2.3 is employed for expressing the aeroelastic forces. Further it is assumed that the
considered modes have no considerable mixed components. Substituting z and to
the generalised displacements h
1
,
2
in Eqs.(2.18) (whereas
1
=p
1
=h
2
=p
2
=0), the
3.5. Flutter derivatives 55
motion-dependent lift and moment, L
se
and M
se
become
L
se
=
1
2
U
2
B
_
KH

1
G
zz
z
U
+KH

2
G
z
B

U
+K
2
H

3
G
z
+K
2
H

4
G
zz
z
B
_
M
se
=
1
2
U
2
B
2
_
KA

1
G
z
z
U
+KA

2
G

U
+K
2
A

3
G

+K
2
A

4
G
z
z
B
_
, (3.3)
where following the convention of Scanlan [12,13], the reduced frequencies K(=2fB/U)
and all utter derivatives H

i
, A

i
are calculated based on the frequency of the mode
they refer to. Any amplitude or structural damping dependence of utter derivatives
was not accounted for. In this basic 2D formulation, the motion-induced drag force,
D
se
, and the eects of the along-wind motion are ignored, meaning that P

i
deriva-
tives are disregarded. Actually the standard linear decomposition of D
se
was recently
questioned by Starossek et al. [148]. According to them for a bridge section symmetric
along the centreline, both positive and negative rotations will result the same exposed
area to the wind. Thus it is more natural that drag forcing with double the motion
frequency will result. Their assertion was corroborated by numerical simulations. In
any case along-wind displacements were quite low for all records in CSB.
3.5.2 Identication Method
A state space formulation of the dynamic problem was assembled using the Covariance
Block Hankel Matrix Method (CBHM method), which is founded on the widely used
Eigenvalue Realization Algorithm (ERA) described by Juang and Pappa [172]. The
formulation in CBHM is identical to ERA with the exception that instead of the Markov
parameters containing Impulse Response Functions (IRFs), covariance estimates of
output measured random data are employed. Jakobsen [169] rst applied CHBM in the
estimation of utter derivatives from wind tunnel tests and the method has since found
extensive use in aerodynamic applications and testing. Peeters and De Roeck [173],
Qin and Gu [174] and Siringoringo and Fujino [175] all describe the matrix derivation
in detail. Brownjohn et al. [176] found a relative advantage of the method over other
operational modal analysis approaches. The method is based on the Singular Value
Decomposition (SVD) and appropriate factorisation of a Hankel matrix built up by
covariance estimates of the output time series (i.e. displacements or accelerations in
this application). If y stands for the displacement matrix with z and in rows, then the
unbiased sample cross-covariance matrix to be used in the Hankel block construction
is given by
C
yy
(i +n, i) = C
yy
(n) =
1
N n

Nn
y(i +n)y
T
(i), n = 0, 1, .., . (3.4)
56 Chapter 3. Identication of utter derivatives from full-scale data
where n is the number of sampling intervals for the discrete time delay nt, N is the
number of samples in the time series, is the maximum number of lags considered and
i is a counting index. The biased estimate, which only diers in the use of the denomi-
nator N instead of N n, can be used instead with negligible dierences for long time
records. The method exclusively handles white noise loading but here it was attempted
to also account for the actual shape of the loading spectra given in Eq.(3.1). Imagine
the SDOF system of Fig. 3.9 with frequency response function H
i
(f). When loaded
by coloured noise, indicated by the non-at PSD

S
load
, then the produced response
PSD will not preserve the shape of H
i
(f) as the identication method necessitates. A
remedy would be to apply a lter function on the response PSD with magnitude equal
to the inverse square root of the loading PSD and zero phase lag. Thus the corrected
PSD, dashed line in Fig. 3.9, would look as if produced by a ctitious white noise load.
For the two-degree-of-freedom system similarly, ordinary ltering on the response data
in the frequency domain could account for the coloured lift and moment spectra. A new
attribute that has to be taken care of though, is preserving after ltering the exact cou-
pling between modes, that was articially altered. The lift lter function F
L
in Fig. 3.9
for instance, will introduce increased coupling. This is not realistic because bueting
action for each mode is assumed uncorrelated to the coupling action between modes
due to self-excited forcing. The correction for this deciency, was to modify the nal
extracted coupled derivatives with the ratio of the lter values in order to try reverting
to the initial coupling. Thus, the obtained H

2
and H

3
in the illustrated example, are
adjusted by dividing them with the ratio of F
L
(f

)/F
L
(f
z
). Additional information on
the performance of identications with coloured noise and specic attributes on the
relative sensitivity of each derivative were described by Jakobsen [169]. H

2
, H

4
and
A

4
were found to be aected the most from such a correction procedure. Interestingly
H

4
is not a coupling derivative. In practice, for the specic problem in hand the eect
of the actual wind spectra on the utter derivatives was found to be insignicant.
The decomposition of the Hankel matrix recovers simultaneously all parameters of
the discrete time realisation. Knowing the modal stiness and damping matrices for
the in wind and still air cases (pure structural stiness and damping contributions)
allow one to separate the utter derivatives. The whole method (having here the
dimensionality for the problem already decided as trivially two degrees-of-freedom)
relies on the choice of two parameters; the length of the individual time record N and
the number of time delays for which the covariance matrix is evaluated and stored
in the block Hankel matrix. The choice of both is investigated through a sensitivity
analysis together with inspection of the time evolution of the auto and cross-covariance
functions.
3.5. Flutter derivatives 57

S
load
(f)
f
f
f
|H
i
(f)|
2

S
resp
(f)
f
z
f

F
L
(f)
f

S
z
(f)
f
z
f
f
F
L
(f

)
F
L
(f
z
)
1DOF 2DOF
Figure 3.9. Decolouring process. In the 1DOF system, lter application will produce corrected
spectra, see dashed line, simulating white noise loading. For the 2DOF system, ltering with
F
L
will erroneously modify the true aeroelastic coupling, see dashed line versus greyed area.
Adapted from Jakobsen [169].
3.5.3 Application to the Clifton Suspension Bridge
The proximity of the fundamental vertical and torsional modes seems to encourage some
coupling action, which could potentially be the initial sign of classical utter. The PSDs
in Fig. 3.8 indicate some non-negligible values of the H

2
or H

3
utter derivatives, since
coupling occurs in the vertical PSD at the torsional motion frequency. The utter
derivative identication was performed in one case with recorded acceleration data and
in another with displacements evaluated by double integration of the accelerations in
hand. Both cases produced identical results.
For the N and identication parameters, time records from 10 minutes to 1 hour
and lag ranges between 10 and 40 seconds were used, preserving analogies with sim-
ilar previous treatments. Example covariance functions, for moderately strong wind,
are plotted against time lag in Fig. 3.10. As previously demonstrated by Jakobsen
et al. [177], the suitable number of maximum time lags is strongly inuenced by the
response character at dierent wind speeds. Higher wind speeds usually allow only a
shorter meaningful portion of the covariance function for accurately reproducing the
two-degree-of-freedom interaction, e.g. due to high aerodynamic damping of the pure
vertical response in the case of streamlined box-girders. For the example in hand an
optimum set of values, producing representative results, was found to be the combi-
nation of 15 minute records (N=112
10
) with time delays up to 20 seconds (=250).
To justify the choice, such a lag value is slightly higher than the beating period of the
58 Chapter 3. Identication of utter derivatives from full-scale data
two frequencies in hand, while as seen in Fig. 3.10, C
z
peaks at approximately 10s.
Sensitivity of the identication for the range quoted, was only weak
0 10 20 30 40
1
0.5
0
0.5
1
C
z
z
0 10 20 30 40
1
0.5
0
0.5
1
C
z

0 10 20 30 40
1
0.5
0
0.5
1
C

z
lag (s)
0 10 20 30 40
1
0.5
0
0.5
1
C

lag (s)
Figure 3.10. Example covariance functions (scaled with variance) for the combined two
degrees-of-freedom plotted against time lag.
Results for the CSB utter derivatives are given in Fig. 3.11. No wind tunnel tests
have been undertaken on the CSB deck section, but where possible the site data are
compared with available wind tunnel results of other deck cross-sections. Sign conven-
tions for aerodynamic forces are as in Fig. 2.7, i.e. lift force and vertical displacement
pointing downwards and overturning moment with rotation positive for the windward
side of the bridge girder moving upwards. A sensitivity analysis on the measured wind
characteristics, such as the turbulence and the angle of attack, proved not to be able
to reproduce a clear picture of their eect. The identied trends of utter derivatives
remained unaltered, but data were insucient to quantify a systematic impact of the
investigated parameters.
Some of the derivatives in Fig. 3.11 appear to have an oset for still air wind
conditions. This has also been encountered in previous treatises both in wind tunnel
[169], and on site [155, 156]. Here it can mostly be attributed to eects such as the
distortion from trac, inuencing both the loading and the mass distribution, as well as
uncertainties in the modal masses and inaccuracies in the still air structural matrices.
The non inclusion of P

i
derivatives may also be of some inuence. For o-diagonal
still air values there is no direct control on this oset, which can be purely a noise
artefact even in the absence of gyroscopic terms [178]. For diagonal values there was
3.5. Flutter derivatives 59
Figure 3.11. Flutter derivatives of Clifton Suspension Bridge from full-scale data, compared
with wind tunnel extracted utter derivatives for various cross-sections (after Scanlan and
Tomko [73] and adapted to Eqs.(3.3)). A

1
and A

3
for section G5 are negligible. H

4
and A

4
were not measured in the wind tunnel tests. Identied values correspond to binned and averaged
identied values.
60 Chapter 3. Identication of utter derivatives from full-scale data
an attempt to minimise the oset since for damping especially, the absolute values are
of great importance.
Although the identied utter derivatives are noisy, unsurprisingly for full-scale am-
bient data, some trends are apparent. Consistent with the observed bridge behaviour,
the results indicate that, within the range of wind speeds recorded (maximum 15.3m/s),
the bridge is not susceptible to torsional utter (so called damping-driven utter as
presented by Matsumoto et al. [179]), which was the reason for the famous Tacoma
Narrows Bridge collapse. Neither galloping
1
seems an issue. This is due to having
close to negative A

2
and H

1
(direct damping derivatives), which will probably need a
further increase in reduced wind speed to initiate such alarming behaviour, if indeed it
does occur. However, H

1
apparently shows a steep positive gradient near the highest
wind speed recorded, suggesting it could become positive for higher wind speeds, pos-
sibly leading to instability. This trend persists regardless of the selected identication
parameters (N and ), indicating it is not due to numerical errors, although the last
few points in the gure are from averages over few records, so their accuracy may be
limited. Fig. 3.12 shows the H

1
utter derivative estimates from each 15-minute record,
before the averaging used for Fig. 3.11. Although only the last few points show the
apparent positive gradient, these points depart signicantly from the trend at lower
reduced velocities and the dierences are greater than the general scatter of points,
implying this is most probably a real eect. If this is conrmed, the eect of possible
positive H

1
(i.e. negative aerodynamic damping of vertical motion) could provide a
feasible explanation for the occasional observations of large vibrations of the bridge in
strong winds.
Actually such reversing H

1
behaviour, tending to self-excited oscillations, was rst
met for the Tacoma Narrows Bridge section by Scanlan [73]. It was postulated then,
that the phenomenon is vortex shedding related. Yet oddly in the same study, an H-
section, with depth to width ratio equal to the Tacoma Narrows Bridge did not produce
any pronounced H

1
turn. Results from later wind tunnel tests by Neuhaus et al. [180],
plotted in Fig. 3.12, show an almost ever decreasing H

1
. Additionally the expected
reduced wind velocity for vortex oscillations on Tacoma should be U
r
= U/fB0.55
1
according to [143]. This sets vortex shedding quite far apart (Fig. 3.12) to be capable
of causing lock-in and imposing a strong eect on H

1
. Analogously for the CSB, if the
captured phenomenon is assumed the same, the high turbulence intensities on-site make
quite improbable the explanation on a vortex-shedding basis. Thus, this monotonicity
inversion or in some cases simply multi-valuedness, seems an interesting unresolved
1
Purely translational instabilities, according to the denition introduced in Chapter 2, are termed
galloping. However, the instability described here may elsewhere be found with the name SDOF
translational utter.
3.5. Flutter derivatives 61
0 1 2 3 4 5 6 7
15
10
5
0
5
U
rz
(= 2/K =
U
f
z
B
)
H
1


identied
Tacoma [73]
Tacoma [180]
Tacoma nominal S
1
r
Figure 3.12. Identied H

1
utter derivative from each 15-minute record. H

1
for the Tacoma
Narrows Bridge from Scanlan and Tomko [73] and from Neuhaus et al. [180] are given for
comparison.
issue. It is worth noting that the numerical derivation of H

1
for the Tacoma Narrows
section, practised by Larsen [143], did not reproduce the experimental convexity.
Dimensionally assessing the possibility of unstable motion in the pure vertical re-
sponse, a value of H

1
6 needs to be reached. This estimate is based on the low
amplitude structural damping ratio of =3.3% for the vertical mode [168] and on the as-
sumption that no benecial amplitude-dependent increase in structural damping takes
place. Structural damping is believed to be so high, compared with modern suspen-
sion bridges, because of the many joints in the structure, particularly the wrought
iron suspension chains. The possibility that bias errors could have lead to erroneous
overestimation should be ruled out, since IWCM handles eciently most artefacts.
The H

2
and H

3
derivatives, which control the coupling from torsional to vertical
motion, have small values. However at the higher wind speeds there is a noticeable
growing negative trend in H

3
, in line with the curves for other bridge proles, which
potentially explains the previously illustrated coupled spectra in Fig. 3.8. The relative
inuence of H

3
for the most coupled response record, translates to a vertical force
approximately 1/10 of the peak restoring elastic force for the mode.
The evolution of H

4
(aeroelastic direct vertical stiness) reects a reduction of
vertical natural frequency with increasing wind speed, although this could alternatively
be due to an amplitude dependence rather than the wind. Similarly A

3
(aeroelastic
direct torsional stiness) illustrates a reduction in the torsional natural frequency with
62 Chapter 3. Identication of utter derivatives from full-scale data
increasing wind speed. The highest identied values of A

3
and H

4
translate to an actual
frequency drop of less than 7% in each case (cf. variation of up to 4% from trac [168]).
Such values are higher than the 1%-3% found in [73], which could imply that the
recorded frequency shifts in CSB are actually due to a combinatory cause. In any
case variations are quite low. This reinforces previous observations that unlike airfoil
utter, for bridges with blu sections aeroelasticity inuences more the damping than
the frequencies of the modes (see Scanlan and Tomko [73] and Billah and Scanlan [181]),
although in this study coupling could also become apparent.
In any case, from the sections in hand, a qualitative similarity was found with the
blu section of Tacoma Narrows. For a proper estimation of the critical utter wind
speed, through Complex Eigenvalue analysis (CEV) described in 2.3, data inclusive of
higher wind speeds are needed to extend the plots of Fig. 3.11. However, considering
that the Tacoma Narrows bridge failed under pure torsional utter, due to A

2
(negative
torsional damping), focus is also put on this utter derivative. Actually an estimate of
its value at higher wind speeds, can be obtained by utilising the relationships between
utter derivatives initially proposed by Matsumoto [179] and here written as in Scanlan
et al. [182]
H

1
= KH

3
, H

4
= KH

2
, A

1
= KA

3
, A

4
= KA

2
. (3.5)
These suggested relationships are based on the assertion that twisting and the
apparent angle of attack associated with the bridge girder vertical velocity generate
similar motion dependent forces. Scanlan et al. [182] present a simple elegant proof
based on the classical Theodorsen utter treatise [63]. The relationships should be
mostly applicable for streamlined sections but were found to yield also accurate match
for many blu cross sections too [179]. In [182] the streamlined section of the Tsurumi
Bridge was found to comply well with Eqs.(3.5), while the blu section from Golden
Gate showed a much worse t. Eqs.(3.5) are here employed to investigate the possible
extension of estimates of A

2
to higher reduced wind speeds through the measured values
of A

4
. This can be achieved since the scaling of the reduced wind speed for A

4
uses the
lower vertical frequency giving higher reduced wind speeds than A

2
, which is expressed
in relation to the torsional frequency. Consequently it is possible to better review the
possibility of single-degree-of-freedom torsional utter on the CSB. Fig. 3.13 shows the
A

2
initial data with the additional points. Fitting the polynomial A

2
=0.12U
r
(0.3U
r
1) provides an estimate for pure torsional utter at U
r
6.3 i.e. at U21m/s. This
estimate is based on the low amplitude structural damping estimate for the torsional
mode where =2.6%. No allowance for a potential benecial increase of structural
damping with amplitude was considered. This, combined with the uncertainty in the
estimation of small aerodynamic forces represented by A

4
and with the uncertainty in
3.6. Concluding Remarks 63
the relationships in Eqs.(3.5), expand the error margins of the estimation. Thus the
actual utter wind speed could be higher, although results of this magnitude are still
signicantly below todays standards. For comparison, the modern Ting Kau Bridge
has an expected utter velocity of more than 60m/s [78], while the latest Sutong and
Stonecutters bridges raise this value to an impressive 88m/s and 140m/s respectively
[77].
0 1 2 3 4 5 6 7
0.4
0.2
0
0.2
0.4
0.6
0.8
1
U
r
(= 2/K =
U
f

B
)
A
2


A

2
A

4
/(K) (Matsumoto)
0.12U
r
(0.3U
r
1)
Tacoma
Figure 3.13. Flutter derivative A

2
with additional points from A

4
as suggested by Matsumoto
et al. [179]. The tted polynomial, indicated by the broken line, was used for estimation of a
pure torsional utter wind speed.
3.6 Concluding Remarks
A similar utter derivative identication scheme has been attempted also for the Ting
Kau Bridge, which benets from a state of the art health monitoring system. Its utter
derivatives H
i
,A
i
for i=1,2,3,4 were previously measured during scaled sectional wind
tunnel tests, thus providing the analysis a reference to compare with. Additionally,
data records are inclusive of the period that the bridge was hit by Typhoon York.
Evidently it should have been expected that the Ting Kau Bridge case study is far
more interesting and didactic than CSB. As a matter of fact it is not. The maximum
wind speed acquired at deck level on the Ting Kau Bridge was only 24.1m/s, too low in
comparison with the modelled utter capacity of around 60m/s. Any utter signs are
too weak and probably get masked by other inuences, which contaminate the signal
parts assigned to utter derivatives. Further all structural modes are three-dimensional
with their mixed character seriously complicating any analysis.
64 Chapter 3. Identication of utter derivatives from full-scale data
On the other hand the CSB is the simplication anyone would try to have for the
full-scale utter problem. Modal complexities and unknowns are reduced, and on top
of that the operational wind velocities span an extensive regime closing to what could
be an instability threshold. It was proved that the historic bridge showed some similar
trends with the Tacoma Narrows Bridge, in terms of both the direct damping utter
derivatives H

1
and A

2
. Although it is now well accepted by the engineering society that
single-degree-of-freedom torsional utter led the Tacoma Narrows Bridge to destruction,
there have also been some dierent views. K arm ans initial diagnosis was that vortex
shedding was the reason for the catastrophe, while Plaut [183] employing nonlinear
logic argued that initial small damage to a cross-tie introduced an asymmetry without
which the bridge would never go into unstable torsional motion, whatever many wind
tunnel tests unanimously support. Such ideas attempt to refute the whole bridge utter
analysis in use today. CSB being a unique large scale example seemingly close to utter,
can be used in exploring how the phenomenon truly unfolds (progress or die down by
leaking energy to other modes) and answer many remaining questions.
U
r
H

dC
L
d
+C
D

Re
2
< 0

dC
L
d
+C
D

Re
1
= c
1
> 0

dC
L
d
+C
D

Re
0
= c
0
> c
1
Figure 3.14. H

1
for dierent H-sections [73] and a possible Reynolds number based expla-
nation for the H

1
inversion. Crossovers can initiate when Reynolds number changes alter the
multiplier of U
r
in Eq.(3.6).
With the opportunity of the CSB identication another aerodynamic feature was
highlighted. Positive A

2
is the signature sign of H-sections. Yet, although often ne-
glected, H

1
for many H-proles seems to change its gradient sign, also tending to a
positive unstable value, see the left of Fig. 3.14. Such behaviour was explained by
Scanlan as lock-in from vortex shedding. This would be quite unlikely for the inversion
in CSB that has not been seen to experience vortex shedding in any case. An alternate
explanation never suggested before could be founded on Reynolds number eects as
3.6. Concluding Remarks 65
follows. H

1
could be quasi-steadily estimated by [184]
H

1
(U
r
) =
_
dC
L
d
+C
D
_

0
U
r
2
, (3.6)
where C
D
is the mean drag coecient, the angle of attack and
0
the equilibrium
angle of attack to which all utter derivatives refer to. This linear estimate was found
to be very close to the true H

1
value for a number of bridge sections [169,184]. Suppose
that the critical Reynolds number Re
1
is reached. Then C
D
will reduce and Eq.(3.6)
will result a less steep approximation. This is shown in Fig. 3.14. Potentially at Re
1
a
crossover could occur that replicates the H-section behaviour. This could even lead to
a sign change for H

1
, when the Reynolds number change induces the classical galloping
condition, cf. Eq.(2.9). A C
D
drop with concomitant Strouhal number increase was
previously recorded in wind tunnel tests of the Storeblt bridge section by Schewe [185].
His attempt was to reason the 20% higher Strouhal number that was experienced for
the bridge on-site.
In general, this work attempted to contribute to the literature on the analysis of
aeroelastic eects from ambient vibration data on full-scale bridges, being one of very
few similar studies. On the specic issue of CSBs utter potential, the identied A

2
was extrapolated at higher reduced velocities by use of the A

4
results, and an apparent
trend was recovered. On this basis, a wind speed estimation for pure torsional utter
has been made. Although uncertain, it is believed this is the rst time this has been
achieved based solely on the actual full-scale dynamic performance of a bridge.
Concluding this chapter a future perspective can be given. It was early seen that
small shape modications could enforce dramatic changes over utter derivatives, at
least on scale models. Scanlan and Tomko [73] presented results from tests on two truss-
stiened blu sections for which A

2
on the addition of railroad tracks, or a short middle
trac barrier became positive in where they were negative. This was considered a wind
tunnel artefact. For the CSB application recorded data after restoration work has taken
place on the bridge, indicate that there have been changes in utter derivatives. Note
that none of the works induced any explicit shape alteration to the deck. The process of
understanding if such changes are truly of aerodynamic source is still ongoing. In any
case this could hold the key for increasing the safety limit of a true bridge monument.
Chapter 4
Quasi-steady galloping analysis revisited
The utter derivative based denition of aeroelastic loads, shown in the previous chap-
ter, entails a series of laborious dynamic tests. Fortunately such descriptions become
redundant in the case of galloping vibrations, where simpler static tests of the blu
body in hand together with quasi-steady theory can accurately capture the ensuing
dynamic behaviour. Specically the condition U
r
Sr
1
is considered sucient for
enabling this simplication. Yet classical galloping analysis deals mostly with cases of
across-wind vibrations, leaving aside the more general situation where the wind and
motion may not be normal. This can arise in many circumstances, such as the motion
of a power transmission cable about its equilibrium conguration, which is swayed from
the vertical plane due to the mean wind, or for a tall slender structure in a skewed wind.
Furthermore the generalisation to such situations, when this had been made, has not
always been performed correctly.
This chapter aims at elucidating such shortcomings, thus naturally it is separated
from the introductory Chapter 2. Initially the correct equations for the quasi-steady
aerodynamic damping coecients for the rotated system or wind are re-derived, and
dierences from other variants are highlighted. Motion in two orthogonal structural
planes is considered, potentially giving coupled translational galloping, for which pre-
vious analysis has often been limited or has even arrived at erroneous conclusions. For
the two-degree-of-freedom case, the behaviour is dependent on the structural as well
as aerodynamic parameters, in particular the relative natural frequencies in the two
planes. Dierences in the aerodynamic damping and zones of galloping instability are
quantied, between solutions from the correct perfectly tuned, well detuned and clas-
sical Den Hartog equations (and also an incorrect generalisation of it), for a variety of
typical cross-sectional shapes. The presentation intends to introduce concepts essential
for the subsequent parts of this thesis.
67
68 Chapter 4. Quasi-steady galloping analysis revisited
4.1 Introduction
Quasi-steady theory allows aerodynamic problems to be simplied vastly by replacing
the actual unsteady condition in hand to an equivalent static one, where only the
relative ow velocity is considered for capturing the relevant aerodynamic forcing. Its
most famous application is probably the galloping criterion put forward by Den Hartog
[186, 187] setting the condition for dynamically unstable behaviour of a single degree-
of-freedom (DOF) oscillator as:
F

= sin
_
L +D

_
+ cos
_
L

+D
_
< 0 (4.1)
which is often (e.g. see Holmes [188, p117], Hemon and Santi [189, p856]) expressed in
terms of static force coecients as:
S
sc
= sin
_
C
L
+C

D
_
+ cos
_
C

L
+C
D
_
< 0 (4.2)
where is the angle of attack and L, D, C
L
, C
D
are the static lift and drag forces
and static lift and drag coecients respectively, assumed for nominally 2D ow to be
functions only of , and the prime indicates the derivative with respect to . The
criterion is only a statement to avoid an undamped oscillator becoming negatively
damped due to aerodynamic action. Thus the whole problem reduces to determin-
ing the aerodynamic damping contribution and setting it equal and opposite to the
available structural damping. Eq.(4.2) presented as such is strictly not valid since its
trigonometric terms only apply for = 0, corresponding to 1D across-wind oscillations,
which is tacitly ignored. In the general case (i.e. = 0), the principal structural axes
may not be aligned along the ow direction and normal to it, for example for a vertical
section in skewed wind (or for a horizontal section in inclined wind), or considering the
static sway of a catenary due to the mean wind force. Then Eq.(4.2) fails to accurately
account for the eect of aerodynamic damping, as in the across-wind galloping scenario,
and even if employed in its correct Den Hartog stated form then it fails to describe the
true condition. The correct treatise, although partly presented in the literature already
(Richardson and Martuccelli [190], Blevins [1]) is generally not followed in practice
and the resulting errors in calculations rising from the mishandling have not previously
been quantied. To this end a number of benchmark cross-sections are considered to
illustrate the dierences emerging in dening instability bounds.
In addition, the full extension of a generalised translational galloping criterion of
sections with principal axes arbitrarily inclined to the ow, has to cover motion in
both axes, including their coupling, which is especially important when they have close
natural frequencies. Such an analysis has previously been performed by a number of
4.1. Introduction 69
authors with specic scopes and sometimes with erroneous conclusions. In particular,
Jones [191] addressed coupled motion in two planes in some detail, but only for the
special case of = 0 with identical natural frequencies in the vertical and horizon-
tal directions, and she concludes that no vertical galloping can occur when horizontal
motion is restrained. Although suggesting that this may be a reason for experimental
behaviour observed, it results only from mistreating boundary conditions and neglecting
the eect of the envisaged restraining force in the analysis. Liang et al. [192] and Li et
al. [193] used the formulation in terms of body co-ordinates, following Davenport [194],
and covered the seemingly more general case of = 0 2D perfectly tuned coupled
motion. The fact that the frequencies are restrained to be tuned renders arbitrary,
thus the results should only be equivalent to Jones [191] case and not a generalisa-
tion of it. By using force coecients dened in body co-ordinates rather than wind
co-ordinates, the connections of their work to the Den Hartog criterion are unclear. In
the analysis a special behavioural subcase is missed and the inaccurate quote is put
forward that 2D coupled galloping oscillations may occur only when the fundamental
natural frequencies of a structure in the two orthogonal principal axial directions equal
each other. Macdonald and Larose [134, 135], focusing on the dry inclined galloping
of circular cables, accurately provided the full 2D aerodynamic damping contribution,
including also terms due to Reynolds number and 3D geometric eects. Also, both
resonant and non-resonant conditions between vibrations in the two planes were taken
into account. A similar treatise, though waiving (and questioning) the use of Reynolds
number dependent terms, was presented by Carassale et al. [137], who utilised Kro-
necker products and matrix calculus to derive a full aerodynamic damping matrix. In
both these research works the interest focused on circular cylinders and the objective
was beyond deriving a simple Den Hartog analogue for motion in two orthogonal planes
and testing it against the 1D requirement as is pursued here. Especially in Carassale
et al. [137], due to concentrating on circular sections, derivatives with respect to
were not included in the analysis, rendering the formulation inapplicable to other clas-
sical galloping cases. It is worth referring also to Luongo and Piccardo [195], who
use bifurcation analysis to capture the limit cycle behaviour of detuned congurations,
again limited to Jones [191] schematic case. These references so far, alongside broader
3DOF treatises
1
, with dierent perspectives and not focusing on subtle translational
interaction details (e.g. [197, 198]), roughly cover the full range of available literature
on modelling two-degree-of-freedom (2DOF) galloping vibrations.
Other previous analysis of explicit 2DOF coupled quasi-steady instabilities has pri-
marily been concerned with the combination of translational and rotational motions
e.g. [1, 199201]. Still, there is an inherent incompatibility of the quasi-steady formula-
1
Flutter galloping according to the terminology of Chabart and Lilien [196].
70 Chapter 4. Quasi-steady galloping analysis revisited
tion with rotations. There would always be some arbitrariness on selecting the point to
which a unique eective wind vector refers to. On the other hand, in many cases there
are similar structural conditions for translational motions in the two orthogonal planes
normal to the cylinder axis, rotational motion may not occur simultaneously, and there
is no complication in extending the classical 1DOF galloping treatment. These render
natural the analysis of 2DOF galloping. The work should be very relevant to the later
studied bridge cable motions. In what follows the correctly modied version of the
Den Hartog criterion for an arbitrary angle of attack for 1D motion and a solution
for the more generic motion in two orthogonal planes are re-derived. For the analysis
the full aerodynamic damping matrix is formed and the scenario of coupled galloping
oscillations is considered, which is a function of the structural parameters as well as
the aerodynamic ones and can lead to elliptical trajectories.
4.2 Quasi-steady derivations
The novel contribution of this chapter is not to propose new galloping criteria but, em-
ploying the current state-of-the-art [134, 135, 190], to quantify the dierence between
the generalised 2DOF galloping scenario shown in Fig.4.1(a) and the normally con-
sidered special case in Fig.4.1(b) for pure across-wind motion. For completeness, the
quasi-steady aerodynamic damping derivations, yielding galloping criteria, are briey
repeated hereafter with an added intention of highlighting the errors in Eq.(4.2). For
succinctness the variation of the force coecients with Reynolds number is neglected,
but its incorporation is straightforward.
Figure 4.1. Geometry of a blu section indicating lift and drag forces (L, D), relative angle
of attack () and principal structural axes (x, y). (a) The general case with
0
= 0 and the
2DOF motion potential. (b) The special case for 1DOF across-wind oscillations.
The classical Den Hartog derivation starts typically by writing the mean aerody-
namic force, per unit length, along the y-axis in Fig.4.1, as a function of the mean drag
4.2. Quasi-steady derivations 71
and lift forces:
F
y
= Lcos +Dsin . (4.3)
where L =
1
2
U
2
rel
BC
L
, D =
1
2
U
2
rel
BC
D
, is the uid density, B is a reference di-
mension of the section and U
rel
is the relative velocity. For motion limited to the y
direction, expanding F
y
around y=0 the standard damped equation of motion becomes
m y +c y +m
2
y
y = F
y
= F
y
|
y=0
+ y
dF
y
d y

y=0
, (4.4)
where m is the cylinder mass per unit length,
y
is the circular natural frequency
(in the absence of wind), c is the structural damping coecient, and dots represent
dierentiation with respect to time. Noting that exclusively when the free-stream wind
velocity U and the motion velocity y are orthogonal and thus = arctan( y/U)
(Fig.4.1(b)),
dF
y
d y

y=0
=
1
U
F

=0
. (4.5)
From Eq.(4.3), the derivative of F
y
with respect to , which can be used for Taylor
expanding around any initial inclination
0
, is
F

y
= L

cos Lsin +D

sin +Dcos . (4.6)


In Eq.(4.4), the aeroelastic force (the last term on the right hand side), is equivalent
to a linear viscous damping force. The condition for dynamic instability is simply that
the total eective damping coecient is negative. Hence from Eqs.(4.5&4.6), noting
that for = 0, U

rel
= 0, it is easily shown that the galloping criterion is

dF
y
d y

y=0
=
UB
2
_
C

L
+C
D
_
< c , (4.7)
where the threshold U is proportional to c as noted in 2.1.4. For zero structural
damping, this reduces to the classical Den Hartog criterion presented earlier in Eq.2.9
S
DH
=
_
C

L
+C
D
_
< 0 . (4.8)
which agrees with Eq.(4.2) for = 0.
If the motion is not normal to the wind direction, Eq.(4.7) does not hold and there
are two problems with Eqs.(4.1&4.2). Firstly = arctan( y/U) so Eq.(4.5) is not
valid, which aects both Eqs.(4.1&4.2). Secondly in nding L

and D

, U

rel
= 0, so
extra terms are introduced. For the general case and for extending the analysis to
cover two orthogonal translational DOFs (see Fig.4.1(a)), which potentially can lead
72 Chapter 4. Quasi-steady galloping analysis revisited
to coupled response, the derivation follows.
Eq.(4.3) still holds and also
F
x
= Lsin +Dcos . (4.9)
Expanding F
y
and F
x
around zero motion, to nd equivalent viscous damping terms,
similar to before,
F
y
= F
y
|
x= y=0
+ x
dF
y
d x

x= y=0
+ y
dF
y
d y

x= y=0
,
F
x
= F
x
|
x= y=0
+ x
dF
x
d x

x= y=0
+ y
dF
x
d y

x= y=0
. (4.10)
For evaluating the derivatives the chain rule is employed
d()
d x
=
()
U
rel

dU
rel
d x
+
()


d
d x
, (4.11)
and similarly for y.
Keeping in mind the relations
U
rel
=
_
(U
y
y)
2
+ (U
x
x)
2
, tan =
U
y
y
U
x
x
, tan
0
=
U
y
U
x
, (4.12)
nally the unit length full 22 aerodynamic damping matrix of a blu section is ob-
tained:
C
aero
=
_
_
c
xxa
c
xya
c
yxa
c
yya
_
_
=
_

dF
x
d x

dF
x
d y

dF
y
d x

dF
y
d y
_

_
x= y=0
=
BU
2
_
_
S
xx
S
xy
S
yx
S
yy
_
_
,
(4.13)
with
S
xx
= C
D
(1 + cos
2

0
) (C
L
+C

D
) sin
0
cos
0
+C

L
sin
2

0
, (4.14a)
S
xy
= C
L
(1 + sin
2

0
) + (C
D
C

L
) sin
0
cos
0
+C

D
cos
2

0
, (4.14b)
S
yx
= C
L
(1 + cos
2

0
) + (C
D
C

L
) sin
0
cos
0
C

D
sin
2

0
, (4.14c)
S
yy
= C
D
(1 + sin
2

0
) + (C
L
+C

D
) sin
0
cos
0
+C

L
cos
2

0
. (4.14d)
The derivation is also valid for wind skewed to the cylinder axis, by employing the
wind component normal to the cylinder and the force coecients with respect to that
component, as long as the independence principle is a viable approximation.
4.2. Quasi-steady derivations 73
The 1DOF instability thresholds for galloping in the x or y planes are simply when
the diagonal terms of C
aero
become negative (or more generally equal to minus the
structural damping coecient). Evidently the non-dimensional aerodynamic damping
coecients, S
xx
and S
yy
in Eqs.(4.14a&d), dier from S
sc
in Eq.(4.2), conrming that
it is incorrect. For
0
= 0, S
yy
in Eq.(4.14d) reduces to S
DH
in Eq.(4.8) (as does S
xx
in Eq.(4.14a) for
0
= 90

), as expected.
For the instability condition of the coupled response, an eigenvalue analysis has to
be performed for the 2DOF system, which is a function of the structural, as well as
the aerodynamic, parameters. In general a numerical solution is required, but for the
special case of equal mass (m), structural damping coecient (c) and natural frequency
(
x
=
y
=
n
) for both DOFs, a closed form result is derived as below. The equations
of motion
m x + (c +c
xxa
) x +m
2
x
x = c
xya
y ,
m y + (c +c
yya
) y +m
2
y
y = c
yxa
x , (4.15)
are assumed to possess a solution of the form
x = X exp(t) , y = Y exp(t) , (4.16)
where the eigenvalues and the amplitudes X, Y are in general complex valued.
Eqs.(4.15&4.16) yield
Y
X
=

2
+
c
xx
m
+
2
x

c
xya
m
=

c
yxa
m

2
+
c
yy
m
+
2
y
, (4.17)

4
+
_
c
xx
+c
yy
m
_

3
+
_
c
xx
c
yy
c
xya
c
yxa
m
2
+
2
x
+
2
y
_

2
+
_
c
xx

2
y
+c
yy

2
x
m
_
+
2
x

2
y
= 0 , (4.18)
where c
xx
= c + c
xxa
and c
yy
= c + c
yya
. For
x
=
y
=
n
, solving the biquadratic
Eq.(4.18) for the stability boundary (i.e. purely imaginary eigenvalues, =i) results
in
=
n
, together with c
xx
c
yy
c
xya
c
yxa
= 0 , (4.19)
or c
xx
+c
yy
= 0 (with not restricted to equal
n
) . (4.20)
74 Chapter 4. Quasi-steady galloping analysis revisited
Eq.(4.19) translates, by analogy with Eqs.(4.2, 4.8&4.14a&d), to the criterion for cou-
pled galloping (for no structural damping):
S
2D
=
1
2
_
3C
D
+C

_
_
C
D
C

L
_
2
+ 8C
L
_
C

D
C
L
_
_
< 0 , (4.21)
where S
2D
denotes the non-dimensional eective aerodynamic damping coecient of
the coupled motion (equivalent to S
xx
and S
yy
for uncoupled motion) and the nega-
tive square root obviously gives the critical case. Here Y/X is real, indicating planar
trajectories. As expected Eq.(4.21) does not explicitly include
0
.
The solution in Eq.(4.20) corresponds to the so-called complex response [134, 191],
which arises when the term under the square root in Eq.(4.21) is negative. Then the
criterion for coupled galloping becomes
S
2D
=
1
2
_
3C
D
+C

L
_
< 0, (4.22)
which coincides with the real part in Eq.(4.21), but in addition, the frequencies of
the resulting two in-wind modes are released from being equal. This solution is often
missed (as in [192,193]) by constraining X and Y to be real. From Eq.(4.17), for = i
with =
n
, Y/X is complex, indicating elliptical trajectories. Since the two modal
responses occur simultaneously at dierent frequencies, a 2D beating-type behaviour
occurs, as described in [134, 191, 195].
More generally, in the presence of structural damping (the same in both planes),
the right hand side of Eqs.(4.21&4.22) becomes 2c/BU (equivalent to Eq.(4.7)).
Also of interest is the case where the initial natural frequencies in the two DOFs
(
x
and
y
) are not equal. Then for the stability boundary, similar to Eqs.(4.17&4.18),
is given by [135]:
(c
xx
c
yy
c
xya
c
yxa
)(c
xx
+c
yy
)(
2
c
xx
+c
yy
) +c
xx
c
yy
(1
2
)
2
m
2

2
x
= 0, (4.23)
where =
y
/
x
. This can generally only be solved numerically. For all detuned cases,
the trajectories become elliptical, similarly to the complex response for the perfectly
tuned system and to actual occurrences of galloping in the eld.
4.2.1 Relevance to uniform continuous systems
It is worth noting that all the derived aerodynamic damping estimates (and hence the
galloping criteria), although referring explicitly to a unit length section, are often also
4.3. Application: quantifying dierences 75
applicable for a uniform continuous system allowing motion in two orthogonal planes,
in a uniform ow. This can be easily proved by applying in Eqs.(4.15) the standard
separation of time and space variables,
x(s, t) =

xn
(s)q
xn
(t) , y(s, t) =

yn
(s)q
yn
(t), (4.24)
where s is the distance along the continuous system,
xn
(s) and
yn
(s) are the n
th
undamped mode shapes in the x, y planes and q
xn
(t), q
yn
(t) the corresponding gener-
alised displacements. Employing apart from the standard orthogonality conditions for
same plane modes
_

xn
1
(s)
xn
2
(s)ds = 0 for n
1
= n
2
,
_

yn
1
(s)
yn
2
(s)ds = 0 for n
1
= n
2
, (4.25)
the condition that the mode shapes in the two planes are the same (i.e.
xn
(s) =

yn
(s) =
n
(s)), for any pair of modes in the two planes, Eqs.(4.15) transforms to
_
m(s)(s) q
x
(t)ds +
_
c
xx
(s)(s) q
x
(t)ds +
2
x
_
m(s)(s)q
x
(t)ds
=
_
c
xya
(s)(s) q
y
(t)ds ,
_
m(s)(s) q
y
(t)ds +
_
c
yy
(s)(s) q
y
(t)ds +
2
y
_
m(s)(s)q
y
(t)ds
=
_
c
yxa
(s)(s) q
x
(t)ds . (4.26)
Since for a uniform section in a uniform wind, the generalised coordinates, mass per unit
length and damping coecients are not functions of s, the integral term
_
(s)(s)ds
cancels out, yielding back again Eqs.(4.15) and thus rendering the deduced instability
thresholds in Eqs.(4.14a&d, 4.21&4.22) still valid. It is noted that when the aero-
dynamic damping coecients cannot be deemed to be constants over the structural
length, as for instance for high rise bridge towers where the wind velocity prole is
signicant, or for a varying section, then the integrals in Eq.(4.26) should be calculated
explicitly. However, in many cases the simplifying approach of uniform wind velocity
and section is adequate, and in the present situation it allows comparison between the
relatively simple dierent criteria presented above.
4.3 Application: quantifying dierences
The dierences in the results arising from the dierent galloping criteria are quantied
by utilising data of aerodynamic force coecients for a variety of cross-sectional shapes.
76 Chapter 4. Quasi-steady galloping analysis revisited
Fortunately such data are available in the literature for many shapes (e.g. see [190,191,
202207]), although they have almost exclusively been used in the Den Hartog criterion
only, which, as previously stressed, is not always the case in hand. For the current study
a number of representative shapes, as illustrated in Fig.4.2, were chosen to span a whole
range of possible relative dierences between the dierent galloping criteria. The last
three iced cable shapes (Figs.4.2(j,k,&l)), although only specic examples of the innite
number of possible iced geometries, were chosen for direct comparison with the work
of Jones [191], since, although she attempted to dene the worst case for 1DOF or
perfectly tuned coupled galloping, she chose a presentation method that did not make
the actual dierences in the results clear.
[190] [204] [204] [203] [205] [206]
[203] [190] [207] [191] [191] [191]
Figure 4.2. Sections with aerodynamic coecients provided in the literature [190, 191, 203
207], used in the galloping analysis herein.
Fig.4.3, presents the non-dimensional aerodynamic damping coecients for each
galloping criterion, for each section, for the whole angle range of angles of attack that are
available. Negative values identify aerodynamically unstable regions, where galloping
would occur in the absence of structural damping. More generally, with structural
damping, galloping occurs when the non-dimensional aerodynamic damping coecient
is below 2c/BU. For each shape, three lines are plotted, corresponding to the
aerodynamic damping contributions from: i) the classical Den Hartog summation, S
DH
,
in Eq.(4.8); ii) the more adverse of the two rotated 1DOF cases, S
xx
or S
yy
, as given
in Eqs.(4.14a&d); and iii) the perfectly tuned 2DOF coupled motion case, S
2D
, in
Eqs.(4.21&4.22). It is pointed out that these correspond to three conceptually dierent
motion scenarios: i) applies to dierent aerodynamic angles of attack of the cross-
section, , but with the motion always constrained to be purely across-wind; ii) applies
to the instance where the principal structural axes and cross-sectional shape are xed to
each other and rotate together with respect to the wind (or the wind rotates relative to
the structural axes and shape) as in Fig.4.1(a), with
0
becoming the variable; and iii)
applies to combined 2D motion with perfect tuning of the structural natural frequencies
in the two planes, in which case the orientation of the structural axes is arbitrary,
4.3. Application: quantifying dierences 77
reected by Eqs.(4.21&4.22) being independent of
0
, and only the orientation of the
cross-sectional shape with respect to the wind direction is then relevant. The sub-case
of the 2DOF complex response is distinguished in Fig.4.3 by plotting open circles. It
is notable that there is no case of instability linked to this scenario (i.e. S
2D
from
Eq.(4.22), when it applies, is never negative). This is in keeping with the suggestion by
Macdonald et al. [132], for galloping of a skewed stranded cable in the critical Reynolds
number range, that the combination of parameters required for galloping of a complex
response is unlikely to occur in practice. In addition, comparing Eqs.(4.8&4.22), since
C
D
is always positive, the condition for 2DOF complex galloping is less onerous than the
condition for pure across-wind galloping. Hence, in contrast to Jones [191] suggestion
that observations of elliptical galloping trajectories in the eld can be attributed to
complex galloping, here it seems most probable that this is not the actual case.
The results presented here are for the cases of the wind and motion direction xed at
right angles to each other or for the principal structural axes and cross-sectional shape
xed to each other (or for perfect tuning in 2DOF). The full generalisation allows the
wind direction, principal structural axes and the orientation of the cross-sectional shape
to all be independent. This could occur, for example, for a transmission line, where
the wind is close to horizontal, the structural axes are given by the inclination of the
cable catenary in the mean wind, and the cross-sectional shape can rotate due to the
inuence of gravity on any accreted ice. Such a situation is still covered by Eqs.(4.13-
4.15), where in Eq.(4.14)
0
is the angle between the wind direction and the structural
x-axis (Fig.4.1(a)), but C
D
, C
L
and their derivatives should be evaluated at the angle
of attack between the wind direction and the reference direction of the cross-sectional
shape (not necessarily equal to
0
due to the cross-section rotating).
Commenting on the individual plots in Fig.4.3, the rst impression is that in most
cases all the values follow roughly similar trends and predict close instability zones with
respect to the angle of attack. Especially for sections being or resembling rectangles, in-
cluding the square in Fig.4.3(d), the rectangle with side ratio 3:1 in Fig.4.3(g), and the
rectangle with rounded ends in Fig.4.3(h), the instability zones from the three dierent
criteria are almost indistinguishable, showing some insensitivity of the susceptibility to
galloping for the dierent cases. The square and rectangle were actually chosen for this
study for exhibiting dierent characteristics, with the square galloping for zero angle
of attack and the 3:1 rectangular not (for a classical treatise on the instabilities of
rectangular sections with dierent side ratios see Nakamura and Hirata [57]), although
the most severe zone of instability is for an angle of attack near 10

in both cases. This


is the case for the section in Fig.4.3(h) also. A similar connection exists between the
rectangle in Fig.4.3(g) and the ellipsoid in Fig.4.3(c) with a strong instability near 70

for both, showing that sections very close to circular can still exhibit negative aerody-
78 Chapter 4. Quasi-steady galloping analysis revisited
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(a)
0 20 40 60 80 100
-3
-2
-1
0
1
2
3
4
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(b)
0 20 40 60 80 10 30 50 70 90
-3
-2
-1
0
1
2
3
4
5
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(c)
0 20 40 60 80 10 30 50 70 90
-20
-15
-10
-5
0
5
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(d)
0 20 40 60 80 90 10 30 50 70
-8
-6
-4
-2
0
2
4
6
8
10
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(e)
0 10 20 30 40 50 60
-4
-2
0
2
4
6
8
10
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(f)
0 20 40 60 80 100 120 140 160 180
-2
-1
0
1
2
3
4
5
6
Figure 4.3. Non-dimensional aerodynamic damping coecients (S
DH
, min(S
xx
, S
yy
), S
2D
)
as a function of angle of attack for the cross-sectional shapes given in Fig.4.2 (inset letters
link the two gures). Negative values indicate unstable behaviour (in the absence of structural
damping).
4.3. Application: quantifying dierences 79
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(g)
0 20 40 60 80 10 30 50 70 90
-5
0
5
10
15
20
25
30
35
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(h)
0 20 40 60 80 100 120 140 160 180
-25
-20
-15
-10
-5
0
5
10
15
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(i)
0 20 40 60 80 100 120 140 160 180
-4
-2
0
2
4
6
8
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(j)
100 120 140 160 180 200 220 240 260
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(k)
0 20 40 60 80 100 120 140 160 180
-8
-6
-4
-2
0
2
4
6
8
Angle of attack [

]
S
D
H
,
m
i
n
(
S
x
x
,
S
y
y
)
,
S
2
D
(l)
0 50 100 150 200 250 300 350
-0.5
0
0.5
1
1.5
2
2.5
Figure 4.3 (continued)
80 Chapter 4. Quasi-steady galloping analysis revisited
namic damping values and consequently unstable behaviour, in this case because, after
separation at the sharp corner, the ow then does not reattach, causing a rapid drop
in lift. In any case, for all the gures referenced so far, the dierences are suciently
small to consider that any of the above instability bounds works quite well in any ac-
tual case. Indeed the previous lack of a study to quantify the dierences arising from
the use of dierent criteria can probably be linked to the fact that benchmark studies
of galloping analysis have often been performed on rectangles or rectangle-like shapes,
where the dierences are unimportant.
On the other hand, another section, equally widely studied in wind tunnel tests and
historically connected with galloping, the D-section in Fig.4.3(a), shows more diverse
behaviour. For zero angle of attack, the Den Hartog summation (S
DH
) predicts zero
aerodynamic damping (the D-section is a hard oscillator [208] that, for motion exceeding
a certain amplitude, will gallop even for this angle of attack, but this is beyond the
scope of the current presentation). Evidently the lesser of S
xx
and S
yy
(in this case S
yy
)
from Eqs.(4.14a&d) (referred to as the 1D rotated case hereafter) becomes the same as
S
DH
when
0
= 0, and slightly more unexpectedly the 2DOF solution (S
2D
) also falls
on the same value giving a common start for all three. As the angle of attack increases,
S
DH
departs from the other two, which have a negative peak near 40

, representing
almost twice the negative aerodynamic damping as for S
DH
. This is a signicant
dierence and it clearly demonstrates that the appropriate galloping criterion should
always be applied carefully to the actual the problem in hand. Increasing the angle
of attack further, a smaller instability zone is expected near 100

where now the most


severe condition is for 2D motion and the Den Hartog summation estimates slightly
more negative aerodynamic damping than for the 1D rotated case. Near 80

it is seen
that S
DH
gives extremely positive aerodynamic damping. Looking more broadly it is
seen that actually for nearly all sections the Den Hartog summation gives the highest
positive aerodynamic damping value. Such extreme values are often very close to the
ones coming from the alternate 1D rotated case (the greater of S
xx
or S
yy
), which
is not shown in Fig.4.3 that presents only the worse case. This also explains why in
Figs.4.3(f,g&h) the Den Hartog summation does not match the 1D rotated case for
zero angle of attack - the aerodynamic damping of along-wind vibrations is lower than
for across-wind, although both are positive.
Other sections considered also show notable dierences between the outputs for
the dierent cases. The results for the triangle with a vertex angle 30

as shown in
Fig.4.3(f) show that near 20

the 1D rotated case is close to being stable while the


other two cases are clearly unstable, and around 30

the 2D case is unstable whereas


the other two are not. The same section near 180

(presenting a at face to wind)


on the other hand shows all the three lines in Fig.4.3(f) coinciding. Similarly the
4.3. Application: quantifying dierences 81
Angle of attack [

]
S
s
c
,
S
y
y
Ssc, Eq.(4.2)
Syy, 1D rotated, Eq.(4.14d)
(a)
0 20 40 60 80 90 10 30 50 70
-8
-6
-4
-2
0
2
4
6
8
10
Angle of attack [

]
S
s
c
,
S
y
y
Ssc, Eq.(4.2)
Syy, 1D rotated, Eq.(4.14d)
(b)
0 20 40 60 80 100 120 140 160 180
-4
-2
0
2
4
6
Figure 4.4. Comparison between the erroneous S
sc
(Eq.(4.2)) and the correct value for the
1D rotated y-axis case, S
yy
(Eq.(4.14d)), for (a) the square in Fig.4.2(d) and (b) the triangle
in Fig.4.2(f).
equilateral triangle (Fig.4.3(e)) exhibits signicant dierences, although the 1D and
2D rotated cases generally give close results. In addition it is of interest to note that
the two triangles behave quite dierently (Fig.4.3(e&f)) when presenting their at faces
to the wind although only a small vertex angle change has occured. Dierences are also
apparent in Figs.4.3(b,j&k), with the 1D rotated case giving the least unstable results,
thus rendering the simple Den Hartog calculation to be unnecessarily conservative if
the structural axes rotate with the section. Conversely for the L-section in Fig.4.3(i),
in the most critical region near 60

, the Den Hartog summation is unconservative.


Drawing some general conclusions, although in most cases the broad picture from
the three criteria is similar, the absolute aerodynamic damping values at certain angles
can be quite dierent. There are many instances where a section stable according to
one criterion can be unstable according to another, and there is no set sign for the
relative dierences, with changes being possible even for the same shape in a dierent
range of angle of attack. Still, in almost all the examples interestingly the worst case
occurs for the 2DOF criterion.
It should be noted that the accuracy of the results is of course limited by the
accuracy of the available data. But additionally there is the need to convert the discrete
point measurements of static force coecients into a continuous or piecewise continuous
function in order to determine their derivatives. Many options were pursued towards
this goal, including polynomial and harmonic curve ts of dierent orders. In any case
there is a need for a very high order for any function to accurately t the measured
points, which was recognised early by Blevins [199]. In Blevins analysis the nonlinear
82 Chapter 4. Quasi-steady galloping analysis revisited
terms enter the equation of motion and subsequently solutions are sought to yield the
steady state amplitudes, thus it is detrimental that the introduction of dierent non-
linearity, from dierent tted functions, strongly inuences the results. However for the
purposes of the present analysis, dierent choices make little change to the results of the
comparison (except making the plots smoother). For this reason the simplest possible
piecewise linear assumption was picked for estimation of the derivatives in Fig.4.3.
One of the main initiatives of this section was to correct Eq.(4.2), but it is equally
important to show the error arising from its use. At rst sight it is clear that for 180

rotation it is the opposite of the Den Hartog criterion, while for 90

any correlation with


the Den Hartog criterion should be deemed as fortuitous. Judging from the general
similarity that was earlier found between the Den Hartog case and the 1D rotated case,
it is expected that great dierences from Eq.(4.2) can emerge. Actually it should be
noted that the correct direct equivalent of Eq.(4.2) is not the worse of Eqs.(4.14a&d),
as considered earlier, but only Eq.(4.14d), for motion in the y direction. As can be seen
in Fig.4.4(a), for plotted results for the square section, although Eq.(4.2) inaccurately
estimates a weak instability zone near 80

it otherwise predicts instabilities in agreement


with the correct result. This is only because a squares critical zone occurs for small
angles of attack, where Eq.(4.2) should be close to the Den Hartog criterion. On the
other hand, for the triangular section in Fig.4.2(f) the errors are alarming (Fig.4.4(b)).
Apart from the initial coalescence of the two curves, Eq.(4.2) is consistently unable to
capture not only the value of the aerodynamic damping but even its correct sign. This
is true for most of the section shapes considered (Fig.4.2).
4.4 The detuning eect
The 2DOF solution presented earlier is restricted to perfectly tuned natural frequencies
in the two motion planes. It is evident that in a great number of situations dierent
natural frequencies exist for the dierent directions of motion. Such a case is found for
instance on cables, where, due to sag, the frequencies of odd in-plane modes are higher
than for the corresponding out-of-plane ones. For any detuning the coupled Eqs.(4.15)
still apply, but for increasing detuning the coupling terms on the right hand side move
further from the relevant natural frequency and hence have a reduced eect on the
behaviour. Eventually, for greatly detuned systems the coupling becomes irrelevant
and the system behaves like two uncoupled 1DOF systems in the orthogonal planes.
This poses the interesting questions of: i) what happens for close but not equal natural
frequencies and ii) what will the behaviour be for quite large detuning values.
Utilising Eq.(4.23) for all the dierent sections in hand it is found, as intuitively
expected, that the two 1DOF and the tuned 2DOF solutions dene an envelope within
4.4. The detuning eect 83
Frequency ratio, =
y
x
S
d
e
t
u
n
e
d
,
S
x
x
,
S
y
y
,
S
2
D
Sdetuned, 2D detuned solution, Eq.(4.23)
Sxx, Syy, 1D solutions, Eqs.(4.14a&d)
S2D, 2D tuned solutions, Eq.(4.21)
(a)
1D y
1D x
0.8 0.9 1 1.1
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
Frequency ratio, =
y
x
S
d
e
t
u
n
e
d
,
S
x
x
,
S
y
y
,
S
2
D
Sdetuned, 2D detuned solution, Eq.(4.23)
Sxx, Syy, 1D solutions, Eqs.(4.14a&d)
S2D, 2D tuned solutions, Eq.(4.21)
1D y
1D x
(b)
0.8 0.9 1 1.1
-1
-0.5
0
0.5
1
1.5
2
Figure 4.5. Evolution of the aerodynamic damping solution for dierent values of detuning, ,
for (a) the section in Fig.4.2(j) and = 123

and (b) the section in Fig.4.2(k) and = 30

. The
lower branch is the important one. In (a) for perfect tuning the solution is unstable (negative
aerodynamic damping) and for detuning above about 7% it approaches the 1D solution, which
in this case is stable. In (b), on the other hand, for perfect tuning the solution is stable and for
detuning above 1% it becomes unstable while moving towards the 1D solution.
which the detuned coupled solutions always fall. The actual evolution with detuning
consists of starting from the tuned 2DOF solutions (using in Eq.(4.21)) and progres-
sively converging towards the 1DOF ones, as presented, for example, in Fig.4.5(a) for
the iced cable section of Fig.4.2(j) for an angle of attack of 123

(other parameters (m,

x
, , B, U) were taken from Jones [191]). The alternate path is given in Fig.4.5(b) for
the iced cable section of Fig.4.2(k) for an angle of attack of 30

, where the 1D solution


is more onerous than the 2D tuned one. The actual rate of convergence is found to be
strongly dependent on the force coecients and hence the angle of attack. The smaller
the initial distance between the coupled and uncoupled solutions in Fig.4.3, the slower
the rate of convergence seems to become. As clearly shown in Fig.4.5(a), for detuning
as low as 7% the 2D solution, which is unstable when tuned, becomes stable and ap-
proaches the 1D solution (see the lower critical branch in the gure and also Fig.4.3(j)
at 123

).
Incidentally, this particular example gives the opportunity to correct an erroneous
conclusion reached by Jones [191]. Results from tests on this section by Nigol and
Buchan [209], where no galloping occurred, were taken as support for an assertion that
in general when along-wind motion is restrained then no instability can ever occur.
However, the reasoning was only a result of forgetting the associated external restraining
force in the balance of forces in the equation of motion. As seen in Fig.4.3(j) at 123

,
picked as being the most critical orientation, the tuned 2D solutions and Den Hartog
84 Chapter 4. Quasi-steady galloping analysis revisited
criteria produce indistinguishable values. Thus in this case, with allowance for across-
wind motion, presence or absence of along-wind motion, whatever the frequency ratio
between the two, hardly changes the galloping threshold. The true reason for lack of
observed galloping in this case, which is theoretically only slightly unstable, is likely
related to the level of structural damping and/or slight inaccuracies in quasi-steady
theory, as discussed by Bearman et al. [202] or Luo et al. [210]. Bearman et al. ,
while studying the square sections aerodynamic behaviour, noted interaction eects
between galloping and vortex shedding that could delay galloping. Quite unexpectedly
the structural damping values during limit cycle oscillations were at least 13% higher
than the wind-o measured values. This was not a mere amplitude eect, and showed
strong interplay with turbulence conditions. Further the transition to high U
r
, where
quasi-steady theory assumes wind forcing in quadrature with motion was not recorded.
Even for the highest attained U
r
the phase advance of the aerodynamic force was
approximatelly 80

ahead of the motion. Luo et al. had a similar nding with unstable
galloping sections only slowly reaching the theoretical phase advance of 90

.
Closing this short parenthesis there should be a note on the response trajectories,
which as already mentioned can range from planar to elliptical. As discussed above
and shown in Fig.4.5, the detuned solutions quickly approach the aerodynamic damping
values that correspond to the 1D solutions. However, they still remain qualitatively
dierent from them in terms of the trajectories, which are described by Eq.(4.17).
Such dierences are presented for the case of Fig.4.5(a) in Fig.4.6. When perfectly
tuned (Fig.4.6(a)) two planar modes occur (not necessarily in orthogonal planes). For
small detuning values, the planar motion of the modes turns into ellipses with growing
magnitudes of their minor axes as the detuning increases. This is shown for = 1.005
in Fig.4.6(b). For larger detuning the axes of the elliptical modes also rotate as in
Fig.4.6(c) where = 1.05. This rotation continues until the principal structural axes
are reached, and when that is virtually accomplished (for detuning of the order of 10%
as in Fig.4.6(d)) the width of the trajectories reduces as they converge on the uncoupled
planar solutions. The detuning values required to produce essentially planar responses
are in fact much larger (of the order of 200% in this case). The relevant aerodynamic
damping estimates are also shown in Fig.4.6 to indicate instabilities and to establish
the link to Fig.4.5(a). It is a noteworthy conclusion that the elliptical galloping paths
observed in the eld (see discussion in [191]), are almost certainly due to a detuning
eect between the structural axes.
4.4. The detuning eect 85
x
y
S2D = 1.24
S2D = 1.76
(a)
-1 -0.5 0 0.5 1
-1
-0.5
0
0.5
1
x
y
Sdetuned = 1.22
Sdetuned = 1.74
(b)
-1 -0.5 0 0.5 1
-1
-0.5
0
0.5
1
x
y
(c)
Sdetuned = 0.13
Sdetuned = 0.65
-1 -0.5 0 0.5 1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
x
y
(d)
Sdetuned = 0.48
Sdetuned = 0.04
-1 -0.5 0 0.5 1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
Figure 4.6. Modal trajectories corresponding to Fig.4.5(a) for (a) = 1, (b) = 1.005, (c)
= 1.05 and (d) = 1.1. The applicable S
detuned
value is also indicated. Unstable modes
are plotted with solid lines while stable ones are dotted. Note for comparison that S
xx
= 0.45,
S
yy
= 0.06. For all plots the structural damping value was c = 0.
86 Chapter 4. Quasi-steady galloping analysis revisited
4.5 Concluding Remarks
Quasi-steady aerodynamic theory, which lies at the core of the current analysis, has
known limitations and it has long been recognised that in certain operational regimes
(only approximately identiable) it breaks down. Still, for a broad range of conditions
for low frequency modes, quasi-steady analysis has a proven ability to predict aerody-
namic damping and galloping behaviour. The current chapter identies cases where
limitations are introduced only because of shortcomings in the actual application of the
method.
In particular, the case sketched in Fig.4.1(a), with the structural axes inclined to
the wind direction, has hardly been properly considered before and the dierences in
the behaviour from classical across-wind galloping had not been quantied. The Den
Hartog criterion is correct for pure across-wind vibrations, but not otherwise. Its ap-
plication, or even worse a faulted extension of it presented in Eq.(4.2), for the rotated
system or wind, can yield solutions that can range from close to even the opposite of the
correct ones. Although the Den Hartog summation often gives reasonable estimates of
the aerodynamic damping, it can in some circumstances give negative estimates of only
around half the magnitude of the real values, which is potentially unsafe. Conversely
in some other cases it can be unnecessarily conservative. Furthermore, the dynamic
stability of a section can be determined not only by its shape, the aerodynamic ori-
entation and the orientation of the principal structural axes (which may or may not
follow the aerodynamic orientation), but also by the proximity of the structural natural
frequencies in the two planes.
The correct equations for the non-dimensional aerodynamic damping coecients,
and hence the instability criteria, have been devised for arbitrary relative orientations
of the system with respect to wind, and for perfectly tuned and well detuned natu-
ral frequencies. Such a presentation is essential in realising the way galloping theory
embodies geometric arrangement details and detuning. In particular detuning is a
parameter almost entirely neglected in the existing literature. Here, detuned results,
numerically obtained, always fall between the solutions for the 2D perfectly tuned case
and the more adverse of the two uncoupled 1D cases, so the more critical of the 1D
or 2D cases can be used conservatively. The equations provided are almost as easy to
apply as the classical Den Hartog equation, yet they avoid potential errors and give
accurate estimations of the aerodynamic damping and the propensity of a cylinder
to gallop. But it is important to use the particular equation relevant to the specic
problem being addressed.
A main application of the above design tool in bridges are the bridge cables. Being
of relatively small sectional dimension, having low motion frequencies due to their long
4.5. Concluding Remarks 87
lengths, and moving both in and out of the bridge-cable plane they seem an excellent
practice eld for the above analysis. An interesting illustration of a case where the
above description becomes relevant is the lightning protection system of a cable-stayed
bridge. This typically consists of ordinary stranded cables as in transmission lines. The
Rion Antirion Bridge, which suered a cable stay failure due to a lightning strike, had
several issues with its lightning protection cables [211]. A simple galloping analysis as
above would be able to eciently mitigate the problem. On the other hand, in view
of Eqs(4.14a,d&4.21) a perfectly circular cable with C
L
=C

L
=C

D
=0, C
D
>0 should not
gallop. The reasons why bridge stays do seem to gallop in reality are subsequently
pursued.
Chapter 5
Experiments on galloping vibration of a
circular cylinder
This chapter addresses large galloping-like vibrations of bridge cables, generically in-
clined and yawed to the ow. To this goal, wind tunnel experiments were performed for
various geometric arrangements of a rigid circular cylinder covered with a smooth high
density polyethylene (HDPE) duct as in real bridge stays. Both static and dynamic
congurations of the cylinder model were tested, while Reynolds numbers spanned the
estimated critical range 10
5
4.510
5
. In what follows, initially the experimental appa-
ratus is presented and subsequently the ow features and excessive motions observed
are discussed. For motion frequencies far from K arm an vortex resonance, unsteady
pressure measurements are utilised in order to uncover the aeroelastic forcing func-
tion. It is shown that a fundamental dierence between the inclined and non-inclined
cylinder aerodynamics exists producing dierent pressure distributions and resulting in
alternate dynamic behaviour. Reynolds number-induced transitional behaviour seems
to be crucial in the recorded instability phenomenon in a way that was never suggested
before in the literature. Intermittent jumps in overall sectional forces, wake discon-
tinuities between cell structures along the cylinder span-wise direction and axial ow
are conjectured to be key elements towards realising the complex mechanism of dry
galloping.
5.1 Introduction
As it was shown in 2.2, there are many conicting theories on the explanation of dry
galloping. Most of them are qualitative and schematically demonstrate complex ow
structure interaction mechanisms that could instigate response. However, they are un-
able to seize the magnitude of the involved aerodynamic actions that is essential in
89
90 Chapter 5. Experiments on galloping vibration of a circular cylinder
designing mitigation measures. One exception to this rule, that allows for quantifying
damping thresholds, analogously to Chapter 4, is the Reynolds number based gener-
alised quasi-steady model initially proposed by Macdonald and Larose [133]. Circular
cylinders on dierent inclination-yaw congurations, could have a varying Reynolds
number dependence of their mean static force coecients. This renders the existence
of dedicated sectional asymmetry terms, a sucient but not necessary condition for
galloping.
The rst wind tunnel tests, specically designed to cover the range where Reynolds
number eects on a dry circular cylinders mean forcing become dominant (i.e. critical
Reynolds numbers), took place in 2001. Previous tests on moving cylinders in high
Reynolds numbers are extremely limited, mostly refer to marine applications and focus
on ordinary self-limited K arm an vortex shedding motions e.g. [110, 212]. Most impor-
tantly they do not assess the inuence of a cylinders inclination. The series of the 2001
tests, hereafter referred as Phase 1, was conducted on a 6.687m long inclined rigid ca-
ble model, spring mounted (in two perpendicular directions) in the National Research
Council Canada (NRC) open circuit propulsion wind tunnel. The tests produced data of
signicant vibrations, including a record that was quoted being of diverging character,
see Cheng et al. [213, 214]. This case, which appeared to be dry galloping, proved to be
non-reproducible in a single repeated run. The main reason for the non-repeatability,
was suspected to lie on ambient weather conditions change. The only explicit infor-
mation reporting on this is that during the unstable test it was raining outside while
the unsuccessful run arose on a dry sunny day. Yet it could have also been a system-
atic behaviour of a complex coupled system that even under identical conditions can
produce alternate results. The Phase 1 tests applied a range of parametric scenarios,
varying the cable inclination angle, critical structural damping ratio, support spring
orientation and surface roughness, with wind speeds covering the full nominal critical
Reynolds number range. Unfortunately there is no direct evidence of the actual ow
regimes encountered. Various attempts have been made to explain the signicant vi-
brations observed. To complement these unique Phase 1 dynamic tests, a new test
series of static tests, referred to as Phase 2, was undertaken in the NRC 2m3m closed
circuit wind tunnel. The cylinder this time was much smaller and testing parameters
such as end conditions, ow smoothness, aspect ratio and Mach numbers also diered.
The aim was to gain further insight into the cable forcing by testing a similarly inclined
static cylinder equipped with pressure measurement taps [215, 216].
The mean static force coecients subsequently deduced were used by Macdonald
and Larose [133135] in their earlier referenced galloping framework to identify insta-
bility regions. The largest response incident from Phase 1 was adequately predicted but
obviously not also its non-repeatability. Other Phase 1 large response cases, of smaller
5.2. Wind tunnel tests 91
amplitude though, could not be traced by the galloping analysis and were thus waived
the galloping characterisation. Another very similar variant of quasi-steady analysis
was independently put forward by Carassale et al. [137]. In contrary to Macdonald and
Larose they consider the Reynolds number to have a slow eect on the aerodynamic
forcing. This translates in disregarding all derivatives with respect to Re when form-
ing aerodynamic damping expressions (a process as in Eqs.(4.10&4.11)). The method
was still found capable of reproducing the observed dry galloping incident. Cheng et
al. [217] attempted to formulate a simplied Den Hartog-like criterion, dierent from
the above two, for studying the instability attributes. As a matter of fact the plausible
expression acquired is only product of mishandling derivatives estimates [217, p2272
Eqs.(A.11&A.12)]. When the correct derivation is performed results revert to the orig-
inal Macdonald and Larose, Carassale relations. In a distinct approach, Jakobsen et
al. [218] utilised the unsteady force components contained in the static measurements
to show that they could also cause large responses. And this is without employing any
force lock-in or amplitude dependent action. The ultimate decisive test for all these
treatises would be a comparison with the real dynamic forcing, which has never been
measured before.
In this chapter a newer series of wind tunnel tests, Phase 3, carried out under a
collaborative project between NRC, Canada, the University of Stavanger, Norway and
the University of Bristol, UK is described. It used the same large-scale aeroelastic model
as in Phase 1, but additionally instrumented with pressure taps and covering a dierent
range of parameters, strategically selected for large response to emerge. The main
objectives pursued herein are to identify the local characteristics of the aerodynamic
forcing causing the vibrations, and to highlight the true eects of Reynolds number.
5.2 Wind tunnel tests
5.2.1 Preliminary
Imagine the case of a real inclined bridge cable lying on a vertical plane subject to
horizontal arbitrary yawed wind as in Fig. 5.1. The cable orientation against the wind
can be uniquely described by the set of cable inclination angle and yaw angle . When
modelling cable wind induced vibrations three directions are of major importance: the
mean wind direction, the cable orientation and the motion direction. Preserving the
way the wind sees the cable should leave unaltered the relative angles between them. In
wind tunnels, space restrictions often put limitations in the range of angle sets , that
can be reproduced. This is also the current case. Since for the aerodynamic problem in
hand gravity is not relevant, as in rain-wind induced vibrations for instance, a rotational
92 Chapter 5. Experiments on galloping vibration of a circular cylinder
Figure 5.1. Transformation from real cable to wind tunnel model.
5.2. Wind tunnel tests 93
transformation that will best t the testing facility is performed, while keeping relative
angles constant. Thus and convert to , which is the angle between the wind and
cable axis vectors, and , which is the angle between the out-of plane motion direction
x
1
and the wind-cable plane normal x
2
. It is straightforward to prove that the mapping
operation yields the relations
cos = sin cos ,
tan = cot / sin . (5.1)
In the setup of both Phase 1&3 series, was implemented as the vertical angle inclina-
tion i.e. its horizontal projection is parallel to the wind. This corresponds to rotating
the real cable in Fig. 5.1 around the AB axis until the cable-wind plane coincides with
the vertical plane. Evidently will be realised as the angle between the spring support
axes and the across-wind direction. Every geometric conguration ever tested with its
analogy to prototype set of angles and its naming convention as introduced in [146] is
given in Table 5.1.
Table 5.1. Orientation angles for studied cases.
P
h
a
s
e
1
Setup

P
h
a
s
e
3
Setup

(

) (

) (

) (

) (

) (

) (

) (

)
1B 45 90 45 0 2A 60 90 60 0
1C 30 54.7 45 54.7 2C 45 45 60 54.7
2A 60 90 60 0 4A 90 90 90 0
2C 45 45 60 54.7 4B 90 60 90 30
3A 35 90 35 0 5B 75.1 61 77 30
3B 20 60.6 35 58.7 5A 77 90 77 0
The (NRC) open circuit propulsion wind tunnel is an open loop wind tunnel blowing
air from the outside environment. Air is sucked via a conical intake from an electrically
powered fan with diameter 7.9m. The screens available for the inlet entry were not in
place during the time of Phase 3 tests. Before entering into the settling chamber the
ow passes additional screens, various stators and straightening vanes to counteract the
fan-induced swirl. Following, the air goes through a 1:6 contraction prior to entering
the test section. The test section is 12.2m long, 6.1m high and 3.1m wide. The model
was positioned within the second third of the tunnel after the end of the contraction.
A view of the facility and its inner section with the model in place is given in Fig. 5.2.
Right after completion of Phase 3 a calibration operation was performed by NRC sta,
acquiring local ow characteristics at 105 stations across the tested section area. Results
from data acquired, are illustrated in Fig.5.3. Turbulence intensities, referring to the
94 Chapter 5. Experiments on galloping vibration of a circular cylinder
Figure 5.2. View of the NRC wind tunnel facility and its test section with the model in place.
total wind vector, were of the order of 0.5%. As it can be seen a vertical asymmetry
exists in the tunnel. This was quoted to be mainly due to the removal of the lower
llets.
5.2.2 Setup details
In an attempt to reduce any potential scaling artefacts, the cable model had typical
properties of a full-scale bridge cable. The actual cylinder was a rigid steel pipe with
diameter d=140mm and thickness 16mm. It was covered with a smooth close t HDPE
pipe of the same type that is used for covering cable strands on site. The duct had a
nominal outside diameter of 159mm. The nal model was measured to have d=161.7mm
(though quoted as 160mm in all later discussion) and a total mass of 60.3kg/m. For
simply supported ends, these numbers give a natural bending frequency estimate of 7Hz.
The aspect ratio (l/d) was approaching 40, which according to West and Apelt [219]
is sucient to minimise contamination of pressure data due to end eects. Still the
tested models of West and Apelt were horizontal and tted with end plates, which is
not the case here. The wind tunnel blockage ratio is calculated at 5.3%. In line with
Blackburn [220], this value is suciently low not to cause added articial span-wise
coherence. The wind speed (U) ranged for most tests between 10-40m/s, corresponding
roughly to Reynolds numbers of 10
5
to 4.510
5
, fully covering the expected critical
range. The maximum Mach number tested was less than 0.12.
5.2. Wind tunnel tests 95
Figure 5.3. Turbulence intensity (U
rms
/U) and mean wind velocity variation with height
across the wind tunnel section.
The cable assembly was mounted on spring supports at both top and bottom,
allowing motion in the sway and heave directions normal to the cable axis as shown
in Fig. 5.4. A cable support was additionally provided at the top to carry the axial
component of the cable weight. The natural frequency for translational motions (f) was
set at approximately 1.4Hz. An end to end rotation also existed with a frequency around
2.21Hz. For practical reasons, the lower (down-wind) end of the cable support was
positioned inside the wind tunnel, whereas the upper one was outside the section (see
Fig. 5.4). The cable passed through a rectangular opening in the roof (approximately
0.43m along-wind 0.42m across-wind) to allow motion. The orthogonal sets of springs
at both ends were connected to the cable body with spacer plates that indented to
inhibit any rotational motion. Structural damping ratios were attempted to be very low
to allow for aeroelastic eects to be distinguished. Values ranged between =0.06-0.33%
for low amplitude response. It should be noted that the quoted numbers signicantly
increased during high levels of response. For identical natural frequencies in the two
vibration planes, the orientation of the principal axes is arbitrary. However, real cables
have slight detuning of their natural frequencies in the vertical and horizontal directions
due to the cable sag. This was modelled with slightly dierent spring stiness in the
two perpendicular directions. With detuning, the orientation of the principal axes
becomes signicant. Former experience from Phase 1, indicated the importance of the
96 Chapter 5. Experiments on galloping vibration of a circular cylinder
detuning parameter. The range of detuning between the natural frequencies in the two
planes was varied up to approximately 3%. To cover a variety of possible geometric
arrangements, the cable was tested at three dierent inclination angles attempting to
expand the previous Phase 1 range. As presented in Table 5.1, values of were 60

, 77

and 90

, with 60

being tested again to cover the previously quoted divergent response


record [214]. The support spring rotation angles were chosen as 0

and 54.7

for =
60

, and 0

and 30

for the other inclinations. The rationale behind angle selections


is mostly founded on analysis from previous Phase 2 results [135, 137]. Further the
vertical cylinder scenario is a benchmark case that can allow for comparisons with
existing literature. It was identied that the principal axes in cases with =54.7

did
not coincide with the rotated spring axes, indicating the existence of structural coupling.
For each of the inclination angles, a set of tests was also taken with the model xed
in position, to measure the pressures on the stationary cylinder and compare with
the corresponding dynamic cases. The static tests for the inclinations of 60

and 90

were conducted with the hole around the top end of the cable both open and sealed
in order to assess the inuence of the top end condition. Judging from the attained
mean pressure proles, very small dierences between the two alternatives established
that the top end eects were not signicant, at least locally in the vicinity where the
pressures were measured.
The pressure measurement instrumentation was arranged in four rings of 32 pressure
taps each, more densely spaced on the leeward side, as illustrated in Fig. 5.4. Additional
lines of pressure taps were positioned close to the expected separation and back pressure
points at 100

and 150

from stagnation (measurements for when =0) at one diameter


span-wise intervals. Taps were connected to four electronic pressure scanners embedded
in the model via 1mm diameter urethane tubing of varying length. All data had to be
corrected for the tubing frequency response. In short, the frequency response function
of each tap was measured prior to the experiments. Typical examples of them shown in
Fig. 5.5, look very similar to previous results obtained by Irwin et al. [221]. Using these
transfer functions the recorded measurements were corrected for magnitude and phase
distortions in the frequency domain. Details about the pressure measuring technique
on a similar application are given by Gatto et al. [222].
Both the displacements and accelerations of each end of the model were measured
in the two spring directions (termed heave and sway throughout, dened in Fig. 5.4).
Data sampling was performed at two dierent frequencies, 500Hz and 1250Hz, to assess
the signicance of very high frequency components. It was found that the higher
sampling frequency was redundant. Being interested in ne details of the ow, the
turbulence intensity was measured with two dynamic three-component cobra probe
instruments, upstream and downstream of the cable model. Most of the upstream
5.2. Wind tunnel tests 97
cobra measurements were later found to be corrupted by noise and were discarded. An
additional eect that could inuence the studied ow transitions was dirt accumulation
on the cable surface particularly on the windward face. This was due to the open
return design of the wind tunnel, driving in air from outside. The size of these random
roughness anomalies (up to 1mm high) could be large enough (according to Achenbach
and Heinecke [223] and Shih et al. [224]) to alter the ow transition behaviour. To
minimise this eect the cable was regularly cleaned. The pattern established was that
an unclean cable would not present large motion. All the tests described below are
believed to not have been exposed to the inuence of this uncontrolled parameter. On
a clean cable, surface roughness corresponded to a roughness-to-diameter ratio (/d)
of 610
6
.
Figure 5.4. Elevation of cable model showing instrumentation arrangement. Some distances
are not in scale. For accurate positioning details consult Table 5.2.
98 Chapter 5. Experiments on galloping vibration of a circular cylinder
Table 5.2. Position details for the model. For rings and lowest cable end distance from oor
refers to stagnation points, while for cobra probes distance from model refers to along-wind
distance.
(

) Distance from oor (m) Distance from model (m)


target/actual Ring1 Ring2 Ring3 Ring4 Low end Cobra1 Cobra2 Cobra1 Cobra2
60 / 59.4 3.87 3.594 3.043 2.63 0.859 4.34 3.19 3.018 0.133
77 / 76.7 4.231 3.919 3.296 2.829 0.827 4.34 3.24 2.381 0.305
90 / 90.9 4.152 3.832 3.192 2.712 0.655 4.34 3.19 1.702 0.304
Figure 5.5. Typical frequency response curves for three pressure taps.
5.3. Results 99
5.3 Results
5.3.1 Overview and large responses
Large responses of primarily across-wind character were observed, only for the incli-
nation of =60

, for both cases of spring rotations examined, = 0

and 54.7

, and
for various combinations of spring tunings in the two vibration planes. All events refer
to the lowest translational modes with end to end motion always yielding much less
signicant amplitudes. Results are consistent with previous ndings by Matsumoto et
al. [120], who, for a horizontal cylinder restricted to planar across-wind motion, identi-
ed the range of yaw angles for which large cable vibrations occurred to be [22.5

,
45

]. Note that cable orientations relatively close to parallel to the wind direction were
not considered and that the test wind speeds were intended to be in the subcritical
Reynolds number range. In the current study, the large responses for a cable-wind
angle of =60

fell in the unstable region, as previously identied in Phase 1, while


inclinations of =77

, 90

did not produce any similarly large responses.


All large vibrations in the Phase 1&3 tests only occurred within a limited range of
wind speeds. This feature is reminiscent of typical K arm an vortex shedding, although
the frequency content was very far from K arm an vortex resonance (cf. K arm an vortex
shedding corresponds to a reduced velocity U
r
=U/fd 5 or lower, while in the tests
exhibiting large vibrations it was over 100). The feature also distinguishes the observed
response from classical Den Hartog galloping, which occurs for all wind speeds exceeding
a certain threshold (see Fig. 2.5). However such behaviour does not necessarily mean
that the term galloping is inappropriate. Galloping response of such a transient type,
though of dierent attributes, has previously been presented in Fig.2.6. Conjecturing
that instability may be triggered by some sustained boundary layer asymmetry in the
critical Reynolds number range, the excitation mechanism may be similar to normal
galloping but the restoration of symmetry in the supercritical Reynolds number range
for increased wind speeds may bring it to an end. This is exactly the underlying
mechanism implication of the quasi-steady analysis proposed by Macdonald and Larose
[133135]. Their predicted unstable vibrations are limited to the critical Reynolds
number range, as observed, in contrast to classical galloping and to other proposed
mechanisms that are not limited to a specic wind speed range. However it should be
pointed out that the vibrations expected from the quasi-steady analysis for the setup
2A (see Table 5.1) were more along-wind, rather than across-wind as observed in Phase
3. The quasi-steady implementation of Carassale et al. [137] cannot predict any large
amplitude response for this case.
The maximum amplitude of vibrations observed was around 0.75d, which reached
the maximum available clearance of the hole through the wind tunnel roof and corre-
100 Chapter 5. Experiments on galloping vibration of a circular cylinder
sponded to signicantly increased structural damping values. The aerodynamic limit
state, if one exists, would occur at higher amplitude. In every case the trajectories of
the cable for large vibration events were elliptical, generally with the major axis at an
angle to the spring directions, suggesting coupling action induced by aeroelastic forces.
The motion trajectories evolution with detuning shown in Fig. 5.6 is qualitatively in
excellent agreement with theoretical quasi-steady analysis of coupled translational gal-
loping oscillations as presented earlier in Chapter 4. Relatively large detuning values
(in this case of only 2%) produce an ellipse aligned with one of the uncoupled degrees
of freedom (the more excited one) as in Fig. 5.6a. Closer tuning of the system (in this
case roughly 1% detuning) rotates the ellipse as in Fig. 5.6b&c, and perfect tuning
would lead to planar motion in the direction of the divergent coupled mode. The width
of the ellipse is controlled by the coupling action induced by the wind. Since in all
large amplitude cases the motion was predominantly across-wind, and in some cases
almost exclusively so (e.g. Fig. 5.6a), it seems that the dominant aerodynamic forcing
is across-wind and it is likely that the along-wind component arises as a secondary ef-
fect from coupling, giving a signicant response in that direction for very close tuning.
5.3.2 Pressure data
When correlating large responses to Reynolds number it is observed that major events
were grouped in two distinct regions at approximately Re=2.510
5
and 3.510
5
, falling
inside the boundaries of the drag crisis. Proper Orthogonal Decomposition (POD) was
performed on the pressure tap data and interestingly it was found that some consis-
tency of loading exists in the aforementioned regions. As seen in Fig. 5.7, which presents
the cumulative variance explained by a relatively small number of Proper Orthogonal
Modes (POMs), two peaks emerge, approximately coinciding with the Reynolds num-
bers for large responses. The number of POMs selected was such that the cumulative
variance could reach a value of around 90%, which is routinely selected when POD is
employed for ltering purposes. The trends presented remained even with alternate
choices of POM numbers. Very noticeably, the behaviour is similar for both the static
and dynamic tests indicating organisation of the loading even for the static cylinder,
although the actual POMs themselves were found to dier in the two cases. An actual
set of relevant modeshapes in both the static and dynamic tests is illustrated in Fig. 5.8.
The POMs selected are the most energetic ones and correspond to the points indicated
by arrows in Fig. 5.7. Interestingly modeshape coordinates for Ring 3 are quite similar
in both dynamic and static cases. The greatest variance contributions are located at
what seems to be the separation region on the left side. Actually for the dynamic tests
5.3. Results 101
0.25 0 0.25
1
0
1
A
c
r
o
s
s
w
i
n
d
d
i
s
p
l
a
c
e
m
e
n
t
(
/
d
)
0.25 0 0.25
1
0
1
0.25 0 0.25
1
0
1
heave
sway
a
heave
sway
b
heave
sway
c
Along wind displacement (/d)
U=30m/s
Re=3.110
5
U=32m/s
Re=3.410
5
U=35m/s
Re=3.710
5
Figure 5.6. Motion traces for cases a) =60

, =0

, with frequency ratio in heave/sway


f
h
/f
s
=0.979 and structural damping ratio in heave/sway
h
/
s
=1.91; b) =60

, =0

,
f
h
/f
s
=1.007,
h
/
s
=1.18 and c) =60

, =54.7

, f
h
/f
s
=1.014,
h
/
s
= 0.94. The heave and
sway spring axes are also presented. Along-wind really means normal to the cable in the
cable-wind plane rather than along the wind itself.
102 Chapter 5. Experiments on galloping vibration of a circular cylinder
Figure 5.7. Proportion of total variance from 20 POMs (from all pressure tap data) against
Reynolds number. Model setup: =60

and, for dynamics tests, =0

.
all the Rings peak at around this location. On the right side there is a much broader
contribution, which seems to attain its maximum near the across-wind pressure tap.
The POD analysis indicated the existence of both localised and widespread energetic
modes to explain data variances. There were cases where time coecients of certain
POMs from the static tests were primarily harmonic with a frequency close (but not
equal) to the structural frequency, raising the question as to whether this component
has the ability to lock in with motion to cause large response in the later dynamic tests.
The subcritical Reynolds number behaviour is dominated by periodic vortex shedding
and accordingly the possibility for fewer periodic modes to relatively accurately describe
the pressure tap measurements. Increasing Reynolds number into the critical range it
seems to destroy coherent structures and cause an increase in the dimensionality of the
underlying dynamics. Still, there are two breaks in the expected monotonic decrease in
the variance, which provide the system with energetic mechanisms that most probably
accommodate the large responses. The dierences presented in Fig. 5.7 may seem small
but they are consistent and moreover seem to be amplied under the inuence of large
scale motion, thus giving an indication of a possible lock-in action.
Power Spectral Densities (PSDs) of the uctuating lift coecient (C
L
), averaged
over the four pressure rings, for dynamic tests in dierent wind speeds are presented
in Fig. 5.9. The PSDs are normalised by multiplying by U/d, the inverse of the nor-
malisation of the frequency axis, to preserve the variance magnitude. The plot shows
the evolution of the wind forcing in four characteristic behavioural cases for dierent
5.3. Results 103
Figure 5.8. First Proper Orthogonal modeshapes for a set of dynamic and static tests. Tests
correspond to the cases indicated by arrows in Fig. 5.7. Interestingly the greatest variance
contributions originate from the near separation regions.
Figure 5.9. Spectra of the lift coecient, averaged over all four pressure rings. Star-marked
points, 32m/s case, show the resonant peaks that correspond to the structural frequency of
1.4Hz and twice this value. Model setup 2A.
104 Chapter 5. Experiments on galloping vibration of a circular cylinder
Reynolds numbers. For the subcritical range, represented by Re=1.210
5
(U=11m/s),
clear vortex shedding can be identied at a Strouhal (Sr) number of 0.17 corresponding
to 12Hz (cf. the natural frequency of 1.4Hz). When applying the independence prin-
cipal using the component of wind normal to the cable (Usin ) the estimate matches
the expected value for a static cylinder normal to the wind, Sr=0.2. Also the cable
inclination broadens the frequency range of the forcing, relative to the normal wind
case, which looks very similar to the eect of added turbulence [225]. Increasing the
Reynolds number, e.g. Re=2.510
5
(U=23m/s), leads to the vortices becoming inco-
herent, thus reducing the spectrum in the Strouhal reduced frequency region. At the
same time some very low frequency components emerge fd/U <0.05) and the spec-
trum becomes at for a range of reduced frequencies from 0.1 to 0.2. This regime was
reached for much higher Reynolds numbers for the cable normal to the ow. For the
record for which there was a large cable response, at Re=3.410
5
(U=32m/s), there
are large sharp peaks in the spectrum at reduced frequencies corresponding to the mo-
tion frequency and twice this value, indicated by stars in Fig. 5.9. It seems that the
energy in the broad low frequency band locks in to the structural frequency and large
motion builds up, while some broader band low frequency forcing, probably related to
weak vortex shapes, still survives for fd/U <0.1. In the supercritical Reynolds number
regime, e.g. Re=4.310
5
(U= 40m/s), the low frequency components have vanished
and only low-level broadband excitation remains for fd/U <0.3 not having consistent
uniform periodic components. On the contrary, as expected according to Roshkos
(normal cylinder) tests [108], the static equivalents did show up a narrow band process
near fd/U=0.2. It is important to note that the lift coecients above were calculated
as the average from all four rings, whist there were marked dierences between the
rings, particularly for the =60

inclination case. It is believed that this should not be


a consequence of end-eects. As a matter of fact there was some consistency between
rings for the two dierent sets of end conditions, which possibly implies that this be-
haviour is mainly sourced by the way the 3D ow pattern establishes itself regardless
of boundaries.
To investigate the relationship between the aerodynamic forces at the dierent rings,
lift and drag cross-correlation functions were estimated. Fig. 5.10 gives an illustration
of such functions showing the systematic force delays along the cable, indicated by the
asymmetry between the top right and bottom left of the gure. Such delays denote
some propagation in the axial direction only along with the ow. When the time lags
of the peak absolute values of the cross-correlation functions are transformed into a
propagation velocity, they compare quite well with the axial component of the free-
stream wind, although they are not constant along the length between dierent rings
but show a variation. This result is in excellent agreement with the numerical ndings
5.3. Results 105
Figure 5.10. Correlation functions (R) of lift coecients (C
Li
C
Lj
) between rings (i, j) for the
subcritical dynamic test case of U=13m/s, Re=1.410
5
. Propagation is evidently one-sided.
Model setup, =60

and =0

.
of Yeo and Jones [117]. Indications of axial propagation were clear in both the static
and dynamic tests for the cable inclined at =60

and, as possibly expected, ceased (or


almost ceased) for higher inclination values. Fig. 5.10 intentionally displays a subcritical
case with apparent vortex shedding, and was picked in order to match the Reynolds
number of the numerical simulations by Yeo and Jones [116, 117]. To continue on the
inhomogeneity point raised earlier, it is indicative from the auto-correlation functions,
along the diagonal of Fig. 5.10, that the state of shedding for each ring is very dierent.
For Rings 1-3 vortex shedding is highly damped in contrast to the strong periodic
phenomenon that is well known to cause large responses of cylinders normal to the
ow. The periodicity translates typically to Sr=0.17 which is around 20% lower than
the value recovered by Yeo and Jones. At Ring 4, most interestingly, vortex shedding
is almost entirely masked (but not vanished) by a low-frequency broadband process.
The case is reminiscent of so-called swirling structures and their associated pressure
distributions [117]. In any case, signature indications of some axial correlation persisted
throughout the wind speed range along the whole cable. Another possible eect of this
apparent secondary ow can be conjectured in view of static pressure (C
p
) proles such
as the ones given in Fig. 5.11. Ring 3, presents near 120

what looks as a sustained


attached ow or an axial-vortex type disturbance. Its pressure contribution is reaching
only up to the base pressure point (180

). There the behaviour changes abruptly, which


106 Chapter 5. Experiments on galloping vibration of a circular cylinder
appears to be consistent with the suggestion by Matsumoto et al. [120] that axial ow
on the lee side of the cable acts as a splitter plate. Still it should be highlighted that
only Ring 3 receives locally a splitting action if truly this is the case.
5.4 Discussion
5.4.1 Symmetry considerations
The force measuring technique of pressure taps employed (integrating pressures over
tributary areas) has the feature of allowing the assessment of the contribution of par-
ticular segments of the cross section to the total drag
1
and lift forces. Dealing with
an instability that is fundamentally fed by asymmetries, it would be useful to identify
where and how these arise. As indicated in 2.2, which meticulously described the
evolution of the ow characteristics for smooth ow past a normal circular cylinder
over the whole Reynolds number range, asymmetries can occur during the initiation of
critical transition. There the drag suddenly drops and considerable mean lift appears,
since half the section contains a laminar separation bubble while the other half does
not. The modied treatise, also allowing for the eects of turbulence and roughness
on ow ranges, is much more complex. For the current case, to assess the transitional
ow symmetry, the section illustrated in Fig.5.4 (for =0

) is split into a right part


containing taps 4-3-2. . . -20 (with only half the 4 and 20 contributions) and a left part
containing taps 4-5-6. . . -20 (again with half the 4 and 20 contributions).
There is some conceptual dierence in the behaviour of individual rings while vary-
ing the inclination angle, . As presented in Fig.5.12, for the inclination of 60

, the
drag crisis zone during static tests has very distinct features for dierent rings. For
Ring 2 the half-perimeter drag contributions (C
D
1/2
) from the right and left parts
almost coincide while for Ring 3 there is consistent spacing during the drag drops.
These are signs of two bubbles forming simultaneously or sequentially for increasing
Reynolds number. Thus for a given Reynolds number inside the drag crisis region, two
states exist together along the cylinder length (Rings 2 and 3 are at a spacing of 4
diameters). But according to previous studies on smooth cylinders in smooth ow, the
simultaneous formation of two bubbles, as for Ring 2, does not occur in the critical
Reynolds number range but only in the supercritical or transcritical range.
A similar situation to the above observation was described by Zdravkovich [109],
who found that roughness of the order of only /d

=0.003, or alternatively turbulence,


not only shifts the drag crisis to lower Reynolds numbers but actually obliterates the
1
Drag is here taken to be normal to the cable in the cable-wind plane
5.4. Discussion 107
Figure 5.11. Mean pressure coecient distribution around cylinder for large response case of
U=32m/s, Re=3.410
5
(Fig.5.6b). Model setup =60

, =0

.
critical state and causes a transition directly from the subcritical to the supercritical
state. More strikingly, in the presented results the two conditions are found to co-
exist stably (i.e. pressure proles locally did not change state) along the cylinder
length. Neighbouring sections, with one and two bubbles respectively, will shed wake
vortices at dierent frequencies, thus becoming the source of vortex dislocations similar
to the ones identied by Bearman and Owen [40] in the wake of rectangular sections
with sinusoidal variation of the along width dimension (i.e. by introduction of wavy
front additions). Furthermore, the observation that a beyond-critical state emerges
very early is consistent with the acquired forcing spectra. As shown by Schewe [106],
the unstable supercritical state (supercritical to transcritical transition specically) is
characterised by lift uctuations at reduced frequencies in a broad band around 0.2,
along with stronger low frequency peaks, similar to the spectra found here (Fig. 5.9).
Note that dierences between the two bubbles (or even inconsistency in the timing of
their existence) during the two bubble state can still give the opportunity for mean lift
to arise. Finally it should be stressed that what would appear as a second bubble can
yet be a non-conventional ow structure as earlier conjectured.
Considering the 77

and 90

inclination angles, dierent behaviour is found. All


rings seem to exhibit very similar behaviour between them for the mean drag force
coecients. This is consistent with a previous nding of closer agreement for the
higher inclinations, between the mean coecients from the four pressure tap rings
and those back-calculated from mean static displacements during the dynamic tests,
108 Chapter 5. Experiments on galloping vibration of a circular cylinder
1 1.5 2 2.5 3 3.5 4 4.5
x 10
5
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
Re

C
D
1
/
2



Ring 2 right
Ring 2 left
1 1.5 2 2.5 3 3.5 4 4.5
x 10
5
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
Re

C
D
1
/
2



Ring 3 right
Ring 3 left
Figure 5.12. Drag evolution acquired for right (taps 4-3-. . . 20) and left (taps 4-5-. . . 20)
parts of Rings 2 and 3, during static tests. Model setup =60

. Double points correspond to


increasing and decreasing wind speed runs.
5.4. Discussion 109
0.5 1 1.5 2 2.5 3 3.5 4 4.5
x 10
5
0
0.1
0.2
0.3
0.4
0.5
0.6
Re

C
D
1
/
2



Ring 3 right
Ring 3 left
Figure 5.13. Drag evolution acquired for right (taps 4-3-..20) and left (taps 4-5-..20) parts
of Ring 3, during static tests. Model setup, =90

.
which give a measure of the mean force over the whole cylinder length [226]. This
could well be due to the conjecture of increased inclination angles having a consistent
characteristic behaviour for the transitions of individual cylinder sections. Thus a
simple non-weighted averaging can be performed that closely matches the global forcing
coecients. On the other hand, for =60

, since dierent states exist along the cylinder


length, averaging over the four pressure tap rings with predened weightings does not
so accurately represent the global forcing.
The drag crisis for Ring 3 for =90

is illustrated in Fig. 5.13. Similar one-bubble


formations exist on all rings in a (closely spaced) narrow dened range of Reynolds
number. It is signicant that the Reynolds number corresponding to minimum drag is
increased relative to the previous inclined cable case (Fig. 5.12) as e.g. also in [111].
The dierence is amplied further when the modication of calculating Re based on the
wind component normal to the cable, as successfully performed earlier for Sr, is used.
Logically some additional mechanism should be operating for the early drag reduction
for the inclined cable. Apart from the evident curvature change of streamlines on the
cable surface, which translates to a higher eective Reynolds number, this could also
be due to the action of the quoted vortex dislocations, which when steadily located
along the length can act eectively towards drag reduction [40]. An interesting case of
enhanced drag reduction connected with stable ow patterns in the critical Reynolds
number range was previously given by Humphreys [103] when he performed static tests
tying ne silk threads at a cylinders stagnation points. One last feature not previously
explicitly reported was that of sudden avalanche-like (intermittency for quiescent ow
110 Chapter 5. Experiments on galloping vibration of a circular cylinder
Figure 5.14. Ring 1 C
L
and C
D
transitional avalanche-like behaviour. Setup =90

,
Re=3.210
5
, U=32m/s. Pressure distributions for time instants a, b, c, d are provided in
Fig.5.15.
conditions lends them this denition) jumps in the lift (and drag) coecient, denoting
alternating transitions between symmetric and asymmetric states. Such a case is shown
in Fig.5.14, where Re=3.210
5
, U=32m/s. These jumps, reminiscent of driven chaotic
vibrations (cc. the magnetoelastic problem [227]), could occur independently on all
rings, during both static and dynamic tests. Interestingly they were not recorded
as such for =60

. Clearly due to this nding, when high-angle arrangements are


concerned, averaging of force coecients should be performed cautiously. Analysis of
the pressure distributions around the cylinder conrmed that they were due to real
changes in the ow state. To illustrate this, four points (a,b,c,d) were selected in the
record of Fig.5.14 and their instantaneous pressures were drawn in Fig.5.15. As it can
be evidenced, all possible lift signs were realised. Time steps a and d do not only
produce opposite C
L
, but also quite distinct C
D
. This could well be due to a structural
asymmetry of the cylinder, however the fact that the lift ips sides during steady ow
seems as an indication that an unbiased circular behaviour prevails. It could thus be
a case where an ensemble of dierently sized laminar separation bubbles (see 2.2)
are possible. The sequence bcd was picked to surround a sudden drop in C
L
.
Noticeably during the prior to the jump instant b, a small one-sided bump develops
in the back pressure region. When the intermittent process is midway through, at c,
two bumps can be seen in the same region in both sides. Strikingly this yields an
instantaneous C
p
prole exactly as in Ring 3 in Fig.5.11.
5.4. Discussion 111
Figure 5.15. Pressure distributions during the state vibration indicated in Fig.5.14. Steps
b, c, d describe the actual dynamic changes during a state jump. In the plots the radius
corresponds to C
p
=1.
5.4.2 Mechanism implications
With all these observations in hand it is possible to speculate on possible ow mecha-
nisms underlying the dynamic excitation and on the inuence of the cable inclination
angle. Some secondary ow that was identied in the case of =60

is evidently propa-
gating disturbances axially along the cable. Taking into account that no extreme state
jump, as in Fig. 5.14, was observed for the numerous recorded cases it seems reasonable
to suggest that, for this inclination, some reasonably stable dynamic balance builds
up between sections along the cylinder. Still, the dynamic balance is built from cell
components that conceptually dier from each other thus becoming sources for vortex
dislocations. Each section is prone to the eects of turbulence and surface roughness,
as demonstrated by the frequent omission of the critical state in the transition from
subcritical to transcritical. Interaction between zones along the cylinder could possi-
bly allow for stable local formations of high lift to develop in the system (being an
indication of a structured asymmetry).
On the other hand, for =90

and 77

the sections appear to be less inuenced


by the similar turbulence intensity and surface roughness conditions and all transition
states can clearly be distinguished. The existence of delicate single bubble states
is probably proof of the good ow conditions during the whole series of tests. It is
suggested that the lack of signicant secondary ow, providing the means for coupling,
inhibits the formation of a stable dynamic balance along the cylinder, allowing each
section to behave independently from its not too distant neighbours. The fact that
112 Chapter 5. Experiments on galloping vibration of a circular cylinder
there seems to be some characteristic behaviour for the drag evolution, although data
are only available from a few sections, is not inconsistent with this conjecture. Although
considerable lift could occur locally, if the sign of lift is random, the expected total lift
on the cable could be zero, and in any case have unstable features. Vortex dislocations
if occurring (due to intermittent response) would also be unstable. Of course non-zero
total mean lift can stably occur due to the nite length of the cylinder or some exogenous
driving (e.g. blockage) or evidently if the lift sign statistics are not Gaussian. Yet
the above scenario could possibly explain the lack of large vibration incidents for high
inclination values in the Phase 3 tests. For lower cable-wind angles, the suggested
coherent dynamic behaviour over the length of the cable is more likely to result in
stable structured asymmetry (e.g. non-zero lift), which could be a signicant factor in
a galloping mechanism.
5.5 Concluding remarks
Large amplitude vibrations of dry inclined cables are still an unravelled mystery with
ongoing research trying to establish connections and reasoning about the dierent pa-
rameters that may trigger instabilities. This chapter has attempted to add to the
previous observations of the behaviour, using the latest ndings from a series of static
and dynamic tests on a realistic cable model equipped with pressure measurement taps
and identifying features not existing in the current literature.
An interesting nding was the identication of behaviour indicating two separation
bubbles existing very close to the subcritical Reynolds number range. This was only
found for the lowest cable-wind angle of =60

and ceased for =77

and 90

. It
was also established that the cable at =60

could retain dierent ow states along


its length in a very stable way, while at the greater cable-wind angles the ow often
intermittently jumped between states, altering the lift sign and/or value. Moreover,
the Reynolds number value corresponding to minimum drag, nominally designating the
end of the critical state, was found to be signicantly lower for the lowest cable-wind
angle. Accepting the limitation that only a few measurement sections were used, it
is suggested that some stable dynamic balance of dierent states builds up along the
cable for some inclination angles, probably connected with a specic range of along-
cylinder wind component values. Introduction of structures such as vortex dislocations,
resulting from a sustained asymmetry (e.g. as a distribution of pressure proles similar
to Ring 3 in Fig. 5.11) along the cable, could be an important factor in the excitation
mechanism. On the other hand, it is thought that weak coupling, as appears to occur for
ow normal to the cylinder and more importantly avalache-like disrupting behaviour,
may tend to inhibit galloping.
5.5. Concluding remarks 113
Simultaneous existence of dierent ow states was suggested to be a case of dierent
sensitivity of inclination angles to (even low) turbulence intensity and roughness, but it
could alternatively be an eect of the inclination angle itself in smooth ow. The mean
features identied are indicative of the ow characteristics. In any case the critical
Reynolds number regime has been shown able to provide ow structures that could
well be responsible for dynamic instabilities of inclined cables. Ongoing work is aiming
to shed light on the links between these structures and the actual instability mechanism.
Chapter 6
Conclusion and outlook
The present work is a collection of aerodynamic studies concerning dierent parts of
exible long-span bridges. In all cases an attempt was made to elucidate and understand
a number of distinct features that arise due to the complex structural interplay with
the wind. Specically threatening self-excitation phenomena that aect bridges were
addressed. This concluding chapter aims to summarise the main ndings and suggest
areas for further research.
Initially the utter potential of the Clifton Suspension Bridge (CSB) was considered.
It is well accepted that plate girder sections as on the CSB are extremely vulnerable to
aerodynamic eects, yet in a modernised variant they are still in use. Utilising ambient
vibration measurements from a long-term monitoring campaign, it was possible to per-
form the conventional utter analysis of Scanlan [12, 13] in an inverse way and deduce
the utter derivative description of the experienced aeroelastic loading. Records in-
cluding a quite wide variety of wind conditions gave a good range of reduced velocities
making possible further aeroelastic assessments. The results, obtained under uncon-
trolled conditions, seem to reliably follow values from wind tunnel sectional models of
similar sections. This is despite the fact that scaled tests are performed under homo-
geneous ow without the variations of natural wind. Interestingly it was found that
there is some similarity to the Tacoma Narrows Bridge, with single-degree-of-freedom
torsional instability being a possibility for the bridge for wind speeds not far beyond
the range experienced. For the highest obtained reduced velocities, utter derivative
A

2
becomes positive (contributing negative torsional aerodynamic damping), reducing
the total available torsional damping. The actual A

2
function may be increasing even
more rapidly than estimated, since the possible amplitude-dependent increase of struc-
tural damping was not accounted for. Notably the analysis was successful in nding
an estimate of the critical utter speed, based solely on full-scale measurements, for
the rst time. The estimated value of approximately 21m/s is only slightly above the
115
116 Chapter 6. Conclusion and outlook
maximum recorded wind speed on the bridge. The ndings also raised an interest-
ing question regarding the H

1
utter derivative (responsible for vertical aerodynamic
damping). On the Tacoma Narrows Bridge the sign reversal of H

1
, similar to A

2
,
was assigned to vortex shedding. However, it occurred far from the expected resonant
vortex shedding condition, which raises doubts about this explanation. On the CSB
there was no recorded vortex-induced response throughout the testing period, but it
shows a similar trend for H

1
(also at wind speeds far from the expected critical speed
for vortex shedding), suggesting that indeed it is due to some other cause. Aeroelastic
coupling action was also found but it was not a serious concern, at least for the range of
reduced velocities considered. This study essentially adds to the experience of aeroelas-
tic identication of bridges, which is very limited for full-scale structures, and provides
a practical example for any similar future study. In any case the main achievement
is that the best aeroelastic data of a real-scale bridge near critical behaviour to date
were clearly identied. The obvious means of complementing the analysis would be to
perform scaled wind tunnel tests or numerical simulations for comparison. There has
been no previous full-scale validation of bridge utter analysis. This could substantiate
the empirical utter framework currently in use and reassure engineering practice that
the safety margins that modern bridges are designed for are realistic. An added benet
that could surface from the current analysis in the near future is putting forward a
convincing response to the long-standing question of how small localised changes can
improve the overall aerodynamic performance. This could be benecial to assess the
wind risk of many existing bridges that were not designed using the recently devised
utter-resisting framework.
Next a topic that has often been mishandled in the existing literature was consid-
ered;the generalisation of galloping in two dimensions. When the structural axes are
inclined to the wind direction, the original Den Hartog derivation for galloping motion
is not valid. The root of the error in some previous treatments (e.g. [189]) was iden-
tied and succinct expressions for generalised galloping modelling were devised. The
study was extended to cover two-degree-of-freedom motion with allowance for arbitrary
detuning between vibrations along the two principal axes. Although the foundations
for generalised 2DOF galloping were laid by Richardson and Martuccelli [190] in 1965
it has not previously been correctly and clearly presented and the implications quan-
tied. The presentation claries the way quasi-steady theory incorporates geometric
and structural details. Turning the derived force descriptions into instability criteria,
three benchmark galloping scenarios were considered: single-degree-of-freedom motion
normal to the wind and inclined to the wind and perfectly tuned two-degree-of-freedom
motion. These cover the range of possible instability boundaries. Employing published
data of static force coecients it was possible to quantify the dierences in the eective
117
aerodynamic damping (positive or negative) in the dierent cases. It was apparent that
large dierences occur. Furthermore the inuence of detuning on the evolution of the in-
stability has rarely been considered before. However it was shown that it is an essential
parameter in dening the true stability boundary. The investigation was able to refute
the suggestion that the introduction of an along-wind degree-of-freedom [137, 191] will
necessarily stabilise a purely across-wind unstable motion. As presented, the potential
for the opposite behaviour also exists. A number of similar shortcomings in previous
analyses were also addressed, making the suggested updated galloping framework a
potentially valuable tool for wind studies of slender elements such as cables and bridge
towers.
For both the preceding sections, although Reynolds number (Re) was accepted as
a potentially inuential variable, it was excluded from the analysis per se. The nal
piece of this thesis attempted to tackle the controversial aerodynamic problem of dry
galloping of circular sections, in particular of inclined cables on cable-stayed bridges,
where the development of Reynolds-induced lift has been suspected of having a key role.
Therefore the study of Re eects was the focus of the nal part of the investigation. An
experimental approach was adopted, testing both dynamic and static models of cables
with sectional dimensions as on real cable-stayed bridges. Observed large amplitude
responses were primarily of across-wind character and occurred close to the critical Re
and nowhere else. Quasi-steady theory, even in its most elaborate form [133], is unable
to fully interpret the details of most of the responses recorded. The results indicate
that only cylinders inclined at a limited range of angles exhibited large amplitude vi-
brations, while cylinders close to normal to the ow only experienced limited responses.
Strikingly it was recorded that for the near normal cable, aerodynamic forces vary in a
discontinuous non-stationary way, jumping between dierent laminar separation bubble
ow states and resulting in intermittent abrupt steps in the lift and drag time series.
This so-called avalanche-like behaviour was apparent during both static and dynamic
tests. The observed intermittent state jumps could not be identied on the cable in-
clined at angles that produced galloping-like responses, so they were conjectured to
have the function of a quenching disorder that eectively mitigates vibrations. In his
seminal work addressing the transitional Re behaviour of a smooth circular cylinder
normal to the ow, Schewe [228] was in search of period doubling phenomena to ac-
company the turbulent transitions on the cylinders surface. In the context of the new
ndings it seems no surprise that he did not recover any. The current results conrm
that the route to the chaotic turbulent state in the boundary layers follow the alterna-
tive path of intermittency. Interestingly Schewe also found individual state jumps, but
these were global and were not in an oscillating mode (i.e. once a jump occurred it
did not reverse) as in the present study. It would be most intriguing to discover how
118 Chapter 6. Conclusion and outlook
and why locally erratic behaviour with numerous asynchronous jumps can combine
over the cable length to result in the global spatio-temporal stable structure recorded
by Schewe. Due to a lack of global force measuring equipment in the experiments this
is not feasible from the current study. Intriguingly in 2.1.3 something similar was il-
lustrated for classical vortex shedding past the lock-in region(Fig. 2.3). Local pressure
uctuations and global motion did not agree in terms of frequency content.
During large response events non-conventional ow structures near the back pres-
sure region that induce asymmetries were identied. Similarities were established be-
tween the current tests and the numerical simulations of Yeo and Jones [116, 117],
though no signicant dynamic response occurred in the experiments near their simu-
lation region as they anticipated. Finally it was summarised how correlation changes
along the length of the cable, for cables inclined at various angles, inuence the global
dynamics. It was surmised that the dry galloping region overlaps with the critical Re
regime not only because of an expected appearance of lift but due to the multitude of
complex ow features that emerge. If such features are possible to appear in other ow
regimes, which seems highly unlikely, then these would also be potential regions for dry
galloping.
In a study with similar aims to this one, Symes [229] conducted static wind tunnel
tests for a smooth circular cylinder normal to the ow, in the critical Re range. He
concluded that although nominally circular cylinders are normally treated as perfectly
symmetric, they could have a consistent rotational asymmetry that may lead to a clas-
sical galloping response similar to the one addressed in Chapter 4. It is uncertain how
such a behaviour would combine with the more complex picture captured here to form a
mixed origin instability. Additionally the range of Reynolds numbers where the critical
regime was found was unusually low, raising questions as to whether nominally precrit-
ical wind speeds on bridges could actually be critical. Future wind tunnel tests should
also verify the circularity of the model cylinder, which would enable consideration of
the role of the dierent possible galloping-like mechanisms involved.
The intermittent and noisy aerodynamic force uctuations with the distinct abrupt
ow state jumps had a disturbance role since they did not allow motion-induced load-
ing to set in. Yet is is questionable whether this is always true. Potentially the self-
excitation mechanism, if existing, can carry on despite the appearance of noise or even
use the energy content in the noise spectrum to develop. Actually non-harmonic forces
with sharp noisy peaks have been seen during large rain-wind vibrations [97] (Chap-
ter 2, Fig. 2.10). Preliminary results from circular cylinders forced to vibrate with
large amplitudes indicate that there may well be such a dierent function even under
dry conditions. It should be pointed out that as Parkinson comments [23] (based on
the unique studies of Staubli [230] and Bearman [202]), in regions such as the critical
119
Re one, the equivalence between forced and free vibration tests could be invalid. Thus
it would be useful to complement the series of free vibration tests studied here with
equivalent forced ones and compare the relevant aerodynamic force characteristics. In
the spirit of Bishop and Hassan [32], who rst suggested modelling the uid-structure
interaction during vortex-shedding with a nonlinear Van der Poll oscillator, this study
wishes to conclude by proposing an alternative modelling for dry galloping vibrations.
Combining the actual noisy intermittent state jumps with the suspected contribution
when driving the cylinder during forced vibration tests, it is conceivable that a formu-
lation based on stochastic resonance could reproduce the dry galloping phenomenon.
The existence of a second harmonic in the motion-induced lift presented in Fig. 5.9, is
an additional detail that can also justify this choice.
Publications
Authors publications related to the present thesis.
Journal papers
Nikitas N, Macdonald JHG, and Jakobsen JB. Wind induced Vibrations of the Clifton
Suspension Bridge. Wind Struct., 14(3):221-238, 2011.
Nikitas N, Macdonald JHG, and Jakobsen JB, Andersen TL. Critical Reynolds number
and galloping instabilities Experiments on circular cylinders. Accepted in Exp. Fluids.
Nikitas N, and Macdonald JHG. Misconceptions and generalisations of the Den Hartog
galloping criterion. Submitted for publication in J. Eng. Mech.-ASCE.
Conference Proceedings
Nikitas N, Macdonald JHG, and Jakobsen JB. Full Scale Identication of Modal and
Aeroelastic Parameters of the Clifton Suspension Bridge. In 6th International Col-
loquium on Blu Bodies Aerodynamics & Applications (BBAA VI), pages 135138,
Milano, Italy, 2008.
Macdonald JHG, Nikitas N, Symes JA, Jakobsen JB, Andersen TL, Savage MG, and
McAulie BR. Large-scale wind tunnel tests of inclined cable vibrations- Preliminary
ndings. In 8th UK Conference on Wind Engineering, Guildford, UK, 2008.
Nikitas N, Macdonald JHG, and Jakobsen JB., Andersen TL, Savage MG, McAulie
BR. Wind Tunnel testing of an inclined cable model-Pressure and motion characteris-
tics, Part I. In 5th European & African Conference on Wind Engineering (EACWE5),
Florence, Italy, 2009.
Jakobsen JB., Andersen TL, Macdonald JHG, Nikitas N, Savage MG, and McAulie
BR. Wind Tunnel testing of an inclined cable model-Pressure and motion characteris-
tics, Part II. In 5th European & African Conference on Wind Engineering (EACWE5),
Florence, Italy, 2009.
121
122 PUBLICATIONS
Andersen TL, Jakobsen JB, Macdonald JHG, Nikitas N, Larose GL, Savage MG,and
McAulie BR. Drag-crisis response of elastic cable-model. In 8th International Sym-
posium on Cable Dynamics, Paris, France, 2009.
Nikitas N, and Macdonald JHG. The Den Hartog galloping criterion revisited: a non-
classical case. In 9th UK Conference on Wind Engineering (WES-2010), Bristol, UK,
2010.
Zhang J, Au FTK, Li J, Nikitas N, Macdonald JHG, and Jacobsen JB. Identifying
bridge aeroelastic parameters from full-scale ambient vibration data. In 9th UK Con-
ference on Wind Engineering (WES-2010), Bristol, UK, 2010.
References
[1] Blevins RD. Flow-Induced Vibrations. Van Nostrand, New York, 1st edition,
1977.
[2] Simiu E and Scanlan RH. Wind Eects on Structures. John Wiley & Sons, New
York, 1st edition, 1978.
[3] Zdravkovich MM. Flow around circular cylinders Vol 1: Fundamentals. Oxford
University Press, Oxford, 1997.
[4] Dowell EH, Clark R, Cox D, Curtiss HC Jr, Edwards JW, Hall KC, Peters DA,
Scanlan RH, Simiu E, Sisto F, and Strganac TW. A Modern Course in Aeroe-
lasticity. Kluwer Academic Publishers, Boston, 4th edition, 2005.
[5] Davenport AG. The spectrum of horizontal gustiness near the ground in high
winds. Q. J. Roy. Meteo. Soc., 87(372):194211, 1961.
[6] Irwin PA. Wind tunnel and analytical investigations of the response of Lions
Gate Bridge to a turbulent wind. National aeronautical establishment report
LTR-LA-206, National Research Council Canada, 1977.
[7] Mann J, Kristensen L, and Jensen NO. Uncertainties of extreme winds, spectra,
and coherences. In Larsen A and Esdahl S, editors, Bridge Aerodynamics, pages
4956. Balkema, Rotterdam, 1998.
[8] von K arm an T. Progress in the statistical theory of turbulence. P. Natl. Acad.
Sci. USA, 34(11):530539, 1948.
[9] Liepmann HW. On the application of statistical concepts to the bueting prob-
lem. J. Aero. Sci., 19(12):793800, 1952.
[10] Davenport AG. The application of statistical concepts to the wind loading of
structures. P. I. Civil Eng., 19(4):449472, 1961.
[11] Davenport AG. The response of slender, line-like structures to a gusty wind. P.
I. Civil Eng., 23(3):389408, 1962.
123
124 REFERENCES
[12] Scanlan RH. The action of exible bridges under wind, I: Flutter theory. J.
Sound Vib., 60(2):201211, 1978.
[13] Scanlan RH. The action of exible bridges under wind, II: Bueting theory. J.
Sound Vib., 60(2):187199, 1978.
[14] Sears WR. Some aspects of non-stationary airfoil theory and its practical appli-
cation. J. Aero. Sci., 8(3):104108, 1941.
[15] Katsuchi H, Jones NP, and Scanlan RH. Multimode coupled utter and bueting
analysis of the Akashi-Kaikyo Bridge. J. Struct. Eng.-ASCE, 125(1):6070, 1999.
[16] Bietry J, Delaunay D, and Conti E. Comparison of full-scale measurements and
computation of wind eects on a cable-stayed bridge. In Proceeding of the In-
ternational Conference on Cable-stayed and Suspension Bridges, Vol. 2, pages
91100, Deauville, France, 1994. AFPC.
[17] Hay JS. Analyses of wind and response data from the Wye and Erskine Bridges
and comparison with theory. J. Wind Eng. Ind. Aerodyn., 17:3149, 1984.
[18] Strouhal V. Ueber eine besondere art der tonerregung. Ann. Phys., 241(10):216
251, 1878. (in German).
[19] Fey U, K onig M, and Eckelmann H. A new Strouhal-Reynolds-number rela-
tionship for the circular cylinder in the range 47<Re<210
5
. Phys. Fluids,
10(7):15471549, 1998.
[20] Feng CC. The measurement of vortex-induced eects in ow past stationary and
oscillating circular and D-section cylinders. Masters thesis, University of British
Columbia, 1968.
[21] Bearman PW. Vortex shedding from oscillating blu bodies. Ann. Rev. Fluid
Mech., 16:195222, 1984.
[22] Williamson CHK and Roshko A. Vortex formation in the wake of an oscillating
cylinder. J. Fluid Struct., 2:355381, 1998.
[23] Parkinson GV. Flow-induced vibrations of blu bodies. Prog. Aerosp. Sci.,
26:169224, 1989.
[24] Sarpkaya T. A critical review of the intrinsic nature of vortex-induced vibrations.
J. Fluid Struct., 19:389447, 2004.
[25] Jauvtis N and Williamson CHK. The eect of two degrees of freedom on vortex-
induced vibration at low mass and damping. J. Fluid Mech., 509:2362, 2004.
REFERENCES 125
[26] Morse TL and Williamson CHK. Prediction of vortex-induced vibration response
by employing controlled motion. J. Fluid Mech., 634:539, 2009.
[27] Durgin WW, March PA, and Lefebvre PJ. Lower mode response of circular
cylinders in cross-ow. J. Fluid Eng.-T. ASME, 102:183190, 1980.
[28] Pikovsky A, Rosenblum M, and Kurths K. Synchronization: A Universal Concept
in Nonlinear Sciences. Cambridge University Press, Cambridge, 2001.
[29] Minorsky N. Theoretical aspects of nonlinear oscillations. IRE T. Circ. Theor.,
7(4):368381, 1960.
[30] Ferguson N and Parkinson GV. Surface and wake ow phenomena of the vortex-
excited oscillation of a circular cylinder. J. Eng. Ind., 89:831838, 1967.
[31] Bishop RED and Hassan AY. The lift and and drag forces on a circular cylinder
in a owing uid. P. Roy. Soc. Lond. A Mat., 277:3250, 1964.
[32] Bishop RED and Hassan AY. The lift and drag forces on a circular cylinder
oscillating in a owing uid. P. Roy. Soc. Lond. A Mat., 277:5175, 1964.
[33] Minorsky N. Introduction to Non-Linear Mechanics. Edwards Brothers Inc, Ann
Arbor, 1947.
[34] Schmidt G and Tondl A. Non-Linear Vibrations. Cambridge University Press,
New York, 1986.
[35] Hartlen RT and Currie IG. Lift oscillator model of vortex-induced vibration. J.
Eng. Mech. Div.-ASCE, 96:557591, 1970.
[36] Skop RA and Grin OM. A model for the vortex-excited resonant response of
blu cylinders. J. Sound Vib., 27:225233, 1973.
[37] Berger E. On a mechanism of vortex-excited oscillations of a cylinder. J. Wind
Eng. Ind. Aerodyn., 28:301310, 1988.
[38] Berger E and Plaschko P. Hopf bifurcations and hysteresis in ow-induced vibra-
tion of cylinders. J. Fluid Struct., 7:849866, 1993.
[39] Facchinetti ML, de Langre E, and Biolley F. Coupling of structure and wake
oscillators in vortex-induced vibrations. J. Fluid Struct., 19:123140, 2004.
[40] Bearman PW and Owen JC. Reduction of blu-body drag and suppression of
vortex shedding by the introduction of wavy separation lines. J. Fluid Struct.,
12:123130, 1998.
126 REFERENCES
[41] Owen JC, Bearman PW, and Szewczyk AA. Passive control of VIV with drag
reduction. J. Fluid Struct., 15:597605, 2001.
[42] Virlogeux M. Recent evolution of cable-stayed bridges. Eng. Struct., 21:737755,
1999.
[43] Matsumoto M, Shiraishi N, Shirato H, Stoyano S, and Yagi T. Mechanism of
and turbulence eect on vortex-induced oscillations for bridge box girders. J.
Wind Eng. Ind. Aerodyn., 49:467476, 1993.
[44] Owen JS, Vann AM, Davies JP, and Blakeborough A. The prototype testing
of Kessock Bridge: response to vortex shedding. J. Wind Eng. Ind. Aerodyn.,
60:91108, 1996.
[45] Kumarasena T, Scanlan RH, and Ehsan F. Wind-induced motions of Deer Isle
Bridge. J. Struct. Eng.-ASCE, 117(11):33563374, 1991.
[46] Arzoumanidis SG and Birdsal B. Discussion: Wind-induced motions of Deer Isle
Bridge. J. Struct. Eng.-ASCE, 119(1):353355, 1993.
[47] Larsen A, Esdahl S, Andersen JE, and Vejrum T. Storeblt suspension Bridge
- vortex shedding excitation and mitigation by guide vanes. J. Wind Eng. Ind.
Aerodyn., 88:283296, 2000.
[48] Ge YJ and Xiang HF. Blu body aerodynamics application in challenging bridge
span length. In 6th International Colloquium on Blu Bodies Aerodynamics &
Applications (BBAA VI), pages 105108, Milano, Italy, 2008.
[49] Virlogeux M. State-of-the-art in cable vibrations of cable-stayed bridges. Bridge
Struct., 1(3):133168, 2005.
[50] Parkinson GV and Brooks NPH. On the aeroelastic instability of blu cylinders.
J. Appl. Mech.-T. ASME, 28:252258, 1961.
[51] Parkinson GV and Smith JD. The square prism as an aeroelastic non-linear
oscillator. Q. J. M. Appl. Math., 17:225239, 1964.
[52] Novak M. Aeroelastic galloping of prismatic bodies. J. Eng. Mech. Div.-ASCE,
95(1):115142, 1969.
[53] Novak M. Galloping and vortex induced oscillations structures. In Proceedings
of the Conference on Wind Eects on Building and Structures, pages 799809,
Tokyo, Japan, 1971.
REFERENCES 127
[54] Novak M. Galloping oscillations of prismatic structures. J. Eng. Mech. Div.-
ASCE, 98(1):2746, 1972.
[55] Scruton C. The use of wind tunnels in industrial aerodynamic research. Report
309, Advisory Group for Aeronautical Research and Development, Paris, France,
1960.
[56] Bearman PW. Investigations of the ow behind two-dimensional model with a
blunt trailing edge and tted with splitter plates. J. Fluid Mech., 21(2):241255,
1965.
[57] Nakamura Y and Hirata K. The aerodynamic mechanism of galloping. T. Jpn
Soc. Aeronaut. S., 36(114):257269, 1964.
[58] Virlogeux M. Wind design and analysis of the Normandy Bridge. In Larsen A
and Esdahl S, editors, Aerodynamics of Large Bridges, pages 183216. Balkema,
Rotterdam, 1992.
[59] Wardlaw RL. Cable supported bridges under wind action. In Ito M, Fujino Y,
Miyata T, and Narita N, editors, Cable-Stayed Bridges Recent Developments and
their Future, pages 213234. Elsevier, Amsterdam, 1991.
[60] Cai CS, Albrecht P, and Bosch HR. Flutter and and bueting analysis II: Luling
and Deer Isle Bridges. J. Bridge Eng., 4(3):181188, 1999.
[61] Wardlaw RL. Approaches to the suppression of wind-induced vibrations of struc-
tures. In Naudascher E and Rockwell D, editors, Practical Experiences with
Flow-Induced Vibrations, pages 650670. Springer-Verlag, Berlin, 1980.
[62] Gjelstrup H, Georgakis C, and Larsen A. A preliminary investigation of the
hanger vibrations on the Great Belt East Bridge. In 7th International Symposium
on Cable Dynamics, Vienna, Austria, 2007.
[63] Theodorsen T. General theory of aerodynamic instability and the mechanism of
utter. Report 496, NACA, Langley Field, United States, 1935.
[64] Scanlan RH. Problematics in formulation of wind-force models for bridge decks.
J. Eng. Mech.-ASCE, 119(7):13531375, 1993.
[65] Lazzari M. Time domain modelling of aeroelastic bridge decks: a comparative
study and an application. Int. J. Numer. Meth. Eng., 62:10641104, 2005.
[66] Pugsley A. Some experimental work on model suspension bridges. Struct. Eng.,
27(8):327347, 1949.
128 REFERENCES
[67] Bleich F. Flutter theory. In The Mathematical Theory of Vibration in Suspension
Bridges, pages 241280. U.S. Government Printing Oce, Washington, 1950.
[68] Rocard Y. Dynamic Instability. Crosby Lockwood & Son, London, 1957. Trans-
lated from French by Meyer ML.
[69] Hirai A and Okauchi I. Experimental study on aerodynamic stability of suspen-
sion bridges with special reference to the Wakato Bridge, pages 150. Bridge
Engineering Laboratory, University of Tokyo, Tokyo, 1960.
[70] Ukeguchi M, Sakata H, and Nishitani H. An investigation of aeroelastic instability
of suspension bridges. In Proceedings of the Symposium on Suspension Bridges,
pages 273284, Lisbon, Portugal, 1966.
[71] Tanaka H. Vibrations of blu-sectional structures under wind action. In Pro-
ceedings of the Conference on Wind Eects on Building and Structures, pages
899910, Tokyo, Japan, 1971.
[72] Scanlan RH and Sabzevari A. Experimental aerodynamic coecients in the an-
alytical study of suspension bridge utter. J. Mech. Eng. Sci., 11(3):234242,
1969.
[73] Scanlan RH and Tomko JJ. Airfoil and and bridge utter derivatives. J. Eng.
Mech. Div.-ASCE, 97(6):17171737, 1971.
[74] Sarkar PP, Jones NP, and Scanlan RH. Identication of aeroelastic parameters
of exible bridges. J. Eng. Mech.-ASCE, 120(8):17181742, 1994.
[75] Miyata T. Historical view of long-span bridge aerodynamics. J. Wind Eng. Ind.
Aerodyn., 91:13931410, 2003.
[76] Scanlan RH. On the state of stability considerations for suspended-span bridges
under wind. In Naudascher E and Rockwell D, editors, Practical Experiences with
Flow-Induced Vibrations, pages 595618. Springer-Verlag, Berlin, 1980.
[77] Ge Y and Xiang H. Aerodynamic challenges in major Chinese bridges. Recent
Major Bridges, IABSE Reports, Shanghai, 2009.
[78] King JPC and Davenport AG. A preliminary study of wind eects for the pro-
posed Ting Kau Bridge, Hong Kong. Research report, BLWT-SS16-1994, Uni-
versity of Western Ontario, Faculty of Engineering Science, Ontario, Canada,
1994.
REFERENCES 129
[79] Huston DR, Bosch HR, and Scanlan RH. The eects of fairings and of turbulence
on the utter derivatives of a notably unstable bridge deck. J. Wind Eng. Ind.
Aerodyn., 29:339349, 1988.
[80] Scanlan RH. Amplitude and turbulence eects on bridge utter derivatives. J.
Struct. Eng.-ASCE, 123(2):232236, 1997.
[81] Assi GRS, Bearman PW, and Meneghini JR. On the wake-induced vibration of
tandem circular cylinders: the vortex interaction excitation mechanism. J. Fluid
Mech., 661:365401, 2010.
[82] Zdravkovich MM. Flow-induced vibrations of two cylinders in tandem and their
suppression. In Naudascher E, editor, Flow-Induced Structural Vibrations, pages
631639. Springer-Verlag, Berlin, 1974.
[83] Price SJ and Padoussis MP. An improved mathematical model for the stability
of cylinder rows subject to cross-ow. J. Sound Vib., 97(4):615640, 1984.
[84] Granger S and Padoussis MP. An improvement to the quasi-steady model
with application to cross-ow-induced vibration of tube arrays. J. Fluid Mech.,
320:163184, 1996.
[85] Zdravkovich MM and Pridden DL. Interference between two circular cylinder;
series of unexpected discontinuities. J. Ind. Aerodyn., 2:255270, 1977.
[86] Matsumoto M, Shiraishi N, and Shirato H. Aerodynamic instabilities of twin
circular cylinders. J. Wind Eng. Ind. Aerodyn., 33:91100, 1990.
[87] Tokoro S, Komatsu H, Nakasu M, Mizuguchi K, and Kasuga A. A study on
wake-galloping employing full aeroelastic twin cable model. J. Wind Eng. Ind.
Aerodyn., 88:247261, 2000.
[88] Fujino Y. Vibration control and monitoring of long-span bridges-recent research,
developments and practice in Japan. J. Constr. Steel Res., 58:7197, 2002.
[89] Caetano E. Structural Engineering Documents 9; Cable Vibrations in Cable-
Stayed Bridges. IABSE-AIPC-IVBH, Zurich, 2007.
[90] Hikami Y and Shiraishi N. Rain-wind induced vibrations of cables in cable stayed
bridges. J. Wind Eng. Ind. Aerodyn., 29:409418, 1988.
[91] Hardy C and Bourdon P. The inuence of spacer dynamic properties in the
control of bundle conductor motion. IEEE T. Power Ap. Syst., 99(2):790799,
1980.
130 REFERENCES
[92] Flamand O. Rain-wind induced vibration of cables. J. Wind Eng. Ind. Aerodyn.,
57:353362, 1995.
[93] Bosdogianni A and Olivari D. Wind-and rain-induced oscillations of cables of
stayed bridges. J. Wind Eng. Ind. Aerodyn., 64:171185, 1996.
[94] Matsumoto M, Shiraishi N, and Shirato H. Rain-wind induced vibration of cables
of cable-stayed bridges. J. Wind Eng. Ind. Aerodyn., 43:20112022, 1992.
[95] Cosentino N, Flamand O, and Ceccoli C. Rain-wind induced vibration of inclined
stay cables. Part I: Experimental investigation and physical explanation. Wind
Struct., 6(6):471484, 2003.
[96] Verwiebe C. Exciting mechanisms of rain-wind-induced vibrations. Struct. Eng.
Int., 8(2):112117, 1998.
[97] Flamand O, Peube JL, and Papanikolas P. An explanation of the rain-wind
induced vibration of inclined stays. In 4th International Symposium on Cable
Dynamics, pages 6976, Montreal, Canada, 2001.
[98] Yamaguchi H. Analytical study on growth mechanism of rain vibration of cables.
J. Wind Eng. Ind. Aerodyn., 33:7380, 1990.
[99] Geurts P, Vrouwenvelder T, van Staalduinen P, and Reusink J. Numerical mod-
elling of rain-wind-induced vibration: Erasmus Bridge, Rotterdam. Struct. Eng.
Int., 8(2):129135, 1998.
[100] Kitazawa M, Noguchi J, and Yamagami T. Design of the Higashi-Kobe Bridge,
Japan. Struct. Eng. Int., 3(4):226228, 1993.
[101] Matsumoto M, Daito Y, Kanamura T, Shigemura Y, Sakuma S, and Ishizaki
H. Wind-induced vibration of cables of cable-stayed bridges. J. Wind Eng. Ind.
Aerodyn., 7476:10151027, 1998.
[102] Bearman PW. On vortex shedding from a circular cylinder in the critical Reynolds
number regime. J. Fluid Mech., 37(3):577585, 1969.
[103] Humphreys JS. On a circular cylinder in a steady wind at transition Reynolds
numbers. J. Fluid Mech., 9(4):603612, 1960.
[104] Kamiya N, Suzuki S, Nakamura M, and Yoshinaga T. Some practical aspects
of the burst of laminar separation bubbles. In Singer J and Staufenbiel R, edi-
tors, Proceeding of the 12th Congress of ICAS, pages 418428, Munich, Federal
Republic of Germany, 1980.
REFERENCES 131
[105] Almosnino D and McAlister RW. Water-tunnel study of transition ow around
circular cylinders. Technical Memorandum 85879, NASA, 1984.
[106] Schewe G. On the force uctuations acting on a circular cylinder in crossow from
subcritical up to transcritical Reynolds numbers. J. Fluid Mech., 133:265285,
1983.
[107] Strogatz SH. Nonlinear Dynamics and Chaos. Perseus Books, New York, 1994.
[108] Roshko A. Experiments on the ow past a circular cylinder at very high Reynolds
number. J. Fluid Mech., 10:345356, 1961.
[109] Zdravkovich MM. Conceptual overview of laminar and turbulent ows past
smooth and rough circular cylinders. J. Wind Eng. Ind. Aerodyn., 33:5362,
1990.
[110] Humphries JA and Walker DH. Vortex-excited response of large-scale cylinders
in sheared ow. J. Oshore Mech. Art., 33:5362, 1990.
[111] Bursnall WJ and Loftin LK Jr. Experimental investigation of the pressure dis-
tribution about a yawed circular cylinder in the critical Reynolds number range.
Technical Note 2463, NACA, Langley Field, United States, 1951.
[112] Ramberg SE. The inuence of yaw angle upon the vortex wakes of stationary and
vibrating cylinders. NRL Memorandum Report 3822, Naval Research Laboratory,
1978.
[113] Shirakashi M, Hasegawa A, and Wakiya S. Eect of the secondary ow on K arm an
vortex shedding. B. JSME, 29(250):5362, 1986.
[114] Ruscheweyh H. Vortex-excited vibrations of yawed cantilevered circular cylinders
with dierent scruton numbers. J. Wind Eng. Ind. Aerodyn., 23:419426, 1986.
[115] Hayashi T and Kawamura T. Non-uniformity in a ow around a yawed circular
cylinder. Flow Mes. Instrum., 6(1):3339, 1995.
[116] Yeo DH and Jones NP. Investigation on 3-D characteristics of ow around a
yawed and inclined circular cylinder. J. Wind Eng. Ind. Aerodyn., 96:19471960,
2008.
[117] Yeo DH and Jones NP. A mechanism for large amplitude, wind-induced vibrations
of stay cables. In 11th Americas Conference on Wind Engineering, San Juan,
Puerto Rico, 2009. AAWE.
132 REFERENCES
[118] Zuo D and Jones NP. Wind tunnel testing of yawed and inclined circular cylinders
in the context of eld observations of stay-cable vibrations. J. Wind Eng. Ind.
Aerodyn., 97:219227, 2009.
[119] Matsumoto M, Shiraishi N, Kitazawa M, Knisely C, Shirato H, Kim Y, and Tsujii
M. Aerodynamic behavior of inclined circular cylinders Cable aerodynamics.
J. Wind Eng. Ind. Aerodyn., 33:6372, 1990.
[120] Matsumoto M, Saito T, Kitazawa M, Shirato H, and Nishizaki T. Response
characteristics of rain-wind induced vibration of cable-stayed bridges. J. Wind
Eng. Ind. Aerodyn., 57:323333, 1995.
[121] Zdravkovich MM. Review and classication of various aerodynamic and hydro-
dynamic means for suppressing vortex shedding. J. Wind Eng. Ind. Aerodyn.,
7:145189, 1981.
[122] Matsumoto M, Yagi T, Hasuda H, Shima T, Tanaka M, and Naito H. Dry
galloping characteristics and its mechanism of inclined/yawed cables. J. Wind
Eng. Ind. Aerodyn., 98:317327, 2010.
[123] Matsumoto M, Yagi T, , Shigemura Y, and Tsushima D. Vortex-induced cable
vibration of cable-stayed bridges at higher reduced wind velocity. J. Wind Eng.
Ind. Aerodyn., 89:633647, 2001.
[124] Bearman PW and Tombazis N. The eects of three-dimensional imposed distur-
bances on blu body near wake ows. J. Wind Eng. Ind. Aerodyn., 49:339350,
1993.
[125] Shirakashi M, Ishida Y, and Wakiya S. Higher velocity resonance of circular
cylinder in crossow. J. Fluid Eng.-T. ASME, 107:392396, 1986.
[126] Sainsbury RN and King D. The ow induced oscillation of marine structures. P.
I. Civil Eng., 49(3):269302, 1971.
[127] Ribeiro SVG and Meyer JE. Mechanisms of ow-induced vibrations of circular
cylinders in crossow. In Transactions of the 7th International Conference on
Structural Mechanics in Reactor Technology: Division B, pages 279281, Chicago,
USA, 1983. IASMiRT.
[128] Martin WW, Currie IG, and Naudascher E. Streamwise oscillations of cylinders.
J. Eng. Mech. Div.-ASCE, 107:589607, 1981.
[129] Harris CM and Piersol AG. Harris Shock and Vibration Handbook. McGraw-Hill,
New York, 5th edition, 2002.
REFERENCES 133
[130] Davis A, Richards DJW, and Scriven RA. Investigation of conductor oscillation on
the 275kV crossing over the Rivers Severn and Wye. P. I. Civil Eng., 110(1):205
219, 1963.
[131] Richards DJW. Aerodynamic properties of the Severn Crossing conductor, vol-
ume II, pages 687771. HMSO, London, UK, 1965.
[132] Macdonald JHG, Griths PJ, and Curry BP. Galloping analysis of stranded
electricity conductors in skew winds. Wind Struct., 11(4):303321, 2008.
[133] Macdonald JHG and Larose GL. A unied approach to aerodynamic damping
and drag/lift instabilities, and its application to dry inclined cable galloping. J.
Fluid Struct., 22:229252, 2006.
[134] Macdonald JHG and Larose GL. Two-degree-of-freedom inclined cable galloping
Part 1: General formulation and solution for perfectly tuned system. J. Wind
Eng. Ind. Aerodyn., 96:291307, 2008.
[135] Macdonald JHG and Larose GL. Two-degree-of-freedom inclined cable galloping
Part 2: Analysis and prevention for arbitrary frequency ratio. J. Wind Eng.
Ind. Aerodyn., 96:308326, 2008.
[136] Scheer M, Bascompte J, Brock WA, Brovkin V, Carpenter SR, Dakos V, Held
H, van Nes EH, Rietkerk M, and Sugihara G. Early-warning signals for critical
transitions. Nature, 461(7360):5359, 2009.
[137] Carassale L, Freda A, and Piccardo G. Aeroelastic forces on yawed circular
cylinders: Quasi-steady modeling and aerodynamic instability. Wind Struct.,
8(5):373388, 2005.
[138] Tanaka H. Aerodynamics of cables. In 5th International Symposium on Cable
Dynamics, pages 1125, Santa Margherita Ligure, Italy, 2003.
[139] Flamand O and Boujard O. A comparison between dry cylinder galloping and
rain wind induced excitation. In 5th European & African Conference on Wind
Engineering, Florence, Italy, 2009.
[140] Zuo D and Jones NP. Stay-cable vibration monitoring of the Fred Hartman
Bridge (Houston, Texas) and the Veterans Memorial Bridge (Port Arthur, Texas).
FHWA/TX-06/0-1401-2, Center for Transportation Research, The University of
Texas at Austin, Austin, USA, 2005.
[141] Matsumoto M, Yagi T, Hasuda H, Shima T, and Tanaka M. Sensitivity of dry-
state galloping of cable stayed bridges to scruton number. In 7th International
Symposium on Cable Dynamics, Vienna, Austria, 2007.
134 REFERENCES
[142] Jain A, Jones NP, and Scanlan RH. Coupled utter and bueting analysis of
long-span bridges. J. Struct. Eng.-ASCE, 122(7):716725, 1996.
[143] Larsen A. Advances in aeroelastic analyses of suspension and cable-stayed bridges.
J. Wind Eng. Ind. Aerodyn., 7476:7390, 1998.
[144] DAsdia P and Sepe V. Aeroelastic instability of long-span suspended bridges: a
multi-mode approach. J. Wind Eng. Ind. Aerodyn., 7476:849857, 1998.
[145] Saito T, Matsumoto M, and Kitazawa M. Rain-wind excitation of cables of
cable-stayed Higashi-Kobe Bridge and cable vibration control. In Proceeding of
the International Conference on Cable-stayed and Suspension Bridges, Vol. 2,
pages 507514, Deauville, France, 1994. AFPC.
[146] Kumarasena S, Jones NP, Irwin P, and Taylor P. Wind-induced vibration of stay
cables. FHWA-RD-05-083, Federal Highway Administration, U.S. Department of
Transportation, McLean, USA, 2007.
[147] Larsen A and Walther JH. Discrete vortex simulation of ow around ve generic
bridge deck sections. J. Wind Eng. Ind. Aerodyn., 77-78:591602, 1999.
[148] Starossek U, Aslan H, and Thiesemann L. Experimental and numerical identica-
tion of utter derivatives for nine bridge deck sections. Wind Struct., 12(6):519
640, 2009.
[149] Jones NP, Scanlan RH, Jain A, and Katsuchi H. Advances (and challenges) in
the prediction of long-span bridge response to wind. In Larsen A and Esdahl S,
editors, Bridge Aerodynamics, pages 5986. Balkema, Rotterdam, 1998.
[150] Katsuchi H, Saeki S, Miyata T, and Sato H. Analytical assessment in wind-
resistant design of long-span bridges in Japan. In Larsen A and Esdahl S, editors,
Bridge Aerodynamics, pages 8798. Balkema, Rotterdam, 1998.
[151] Nelson RC and Pelletier A. The unsteady aerodynamics of slender wings and
aircraft undergoing large amplitude maneuvers. Prog. Aerosp. Sci., 39:185248,
2003.
[152] Jones NP, Scanlan RH, Sarkar PP, and Singh L. The eect of section model details
on aeroelastic parameters. J. Wind Eng. Ind. Aerodyn., 54-55:4553, 1995.
[153] Matsumoto M, Nakajima N, Taniwaki Y, and Shijo R. Grating eect on utter
instability. J. Wind Eng. Ind. Aerodyn., 89:14871498, 2001.
REFERENCES 135
[154] Okauchi I, Tajima J, and Akiyama H. Response of the large scale bridge model to
natural wind. In 5th International Conference Wind engineering, pages 841852,
Fort Collins, USA, 1979.
[155] Jakobsen JB and Larose GL. Estimation of aerodynamic derivatives from ambient
vibration data. In 10th International Conference Wind Engineering, pages 837
844, Copenhagen, Denmark, 1999.
[156] Costa C and Borri C. Full-scale identication of aeroelastic parameters of bridges.
In 12th International Conference Wind Engineering, pages 799806, Cairns, Aus-
tralia, 2007.
[157] Macdonald JHG. Identication of the dynamic behaviour of a cable-stayed bridge
from full-scale testing during and after construction. PhD thesis, Department of
Civil Engineering, University of Bristol, Bristol, UK, 2000.
[158] Littler JD. Ambient vibration tests on long span suspension bridges. J. Wind
Eng. Ind. Aerodyn., 42:13591370, 1992.
[159] Brownjohn JMW. Estimation of damping in suspension bridges. P. I. Civil
Eng.-Str. B., 104:401415, 1994.
[160] Jensen JL, Larsen A, Andersen JE, and Vejrum T. Estimation of structural
damping of Great Belt suspension Bridge. In Proceedings of the 4th European
Conference Structural Dynamics (Eurodyn 99), pages 801806, Prague, Czech
Republic, 1999.
[161] Ge YJ and Tanaka H. Aerodynamics of long-span bridges under erection. J.
Struct. Eng.-ASCE, 126:14041412, 2002.
[162] Nagayama T, Abe M, Fujino Y, and Ikeda K. Structural identication of a
nonproportionally damped system and its application to a full-scale suspension
bridge. J. Struct. Eng.-ASCE, 135:15361545, 2005.
[163] Barlow WH. Description of the Clifton Suspension Bridge. Minutes P. I. Civil
Eng., 26:243257, 1867. reprinted P. I. Civil Eng.-Bridge Eng. 2003; 156(BE1):5
10.
[164] Farquharson FB, Smith FC, and Vincent GS. Aerodynamic stability of suspension
bridges with special reference to the Tacoma Narrows Bridge, volume I-V. Engi-
neering Experiment Station, University of Washington, Seatle, USA, 1949-1954.
[165] Kumarasena T, Scanlan RH, and Morris GR. Deer Isle: Ecacy of stiening
system. J. Struct. Eng.-ASCE, 115(9):22972312, 1989.
136 REFERENCES
[166] Kumarasena T, Scanlan RH, and Morris GR. Deer Isle: Field and computed
vibrations. J. Struct. Eng.-ASCE, 115(9):23132328, 1989.
[167] Macdonald JHG. Dynamic behaviour of Clifton Suspension Bridge: Response
to wind loading. Report CSB703/REP/2, Bristol Earthquake & Engineering
Laboratory LTD, 2004.
[168] Macdonald JHG. Pedestrian-induced vibrations of the Clifton Suspension Bridge,
UK. P. I. Civil Eng-Bridge Eng., 161(BE2):6977, 2008.
[169] Jakobsen JB. Fluctuating wind load and response of a line-like engineering struc-
ture with emphasis on motion-induced wind forces. PhD thesis, Norwegian Insti-
tute of Technology, Trondheim, Norway, 1995.
[170] Selberg A. Aerodynamic eects on suspension bridges, volume II, pages 462479.
HMSO, London, UK, 1965.
[171] British Standards Institution. Published Document: Background information to
the National Annex to BS EN 1991-1-4 and additional guidance, BS PD 6688-1-
4:2009, 2009.
[172] Juang JN and Pappa RS. An eigensystem realization algorithm for modal param-
eter identication and model reduction. J. Guid. Control Dynam., 8(5):620627,
1985.
[173] Peeters B and De Roeck G. Reference-based stochastic subspace identication
for output-only modal analysis. Mech. Sys. Signal Pr., 13(6):855878, 1999.
[174] Qin XR and Gu M. Determination of utter derivatives by stochastic subspace
identication technique. Wind Struct., 7(3):173186, 2004.
[175] Siringoringo DM and Fujino Y. System identication of suspension bridge from
ambient vibration response. Eng. Struct., 30:462477, 2008.
[176] Brownjohn JMW, Magalhaes F, Caetano E, and Cunha A. Ambient vibration
re-testing and operational modal analysis of the Humber Bridge. Eng. Struct.,
32:20032018, 2010.
[177] Jakobsen JB, Savage MG, and Larose GL. Aerodynamic derivatives from the
bueting response of a at plate model with stabilizing winglets. In 11th Inter-
national Conference Wind Engineering, Vol. 2, pages 673680, Lubbock, USA,
2003.
[178] Hoen C. Subspace identication of modal coordinate time series. In 24th Inter-
national Modal Analysis Conference, IMAC XXIV, St.Louis, USA, 2006.
REFERENCES 137
[179] Matsumoto M, Koboyashi Y, and Shirato H. The inuence of aerodynamic deriva-
tives on utter. J. Wind Eng. Ind. Aerodyn., 60:227239, 1996.
[180] Neuhaus C, Roesler S, H oer, Hortmanns M, and Zahlten W. Identication of
18 utter derivatives by forced vibration tests - a new experimental rig. In 5th
European & African Conference on Wind Engineering, Florence, Italy, 2009.
[181] Billah KY and Scanlan RH. Resonance Tacoma Narrows Bridge failure, and
undergraduate physics textbooks. Am. J. Phys., 59:118124, 1990.
[182] Scanlan RH, Jones NP, and Singh L. Inter-relations among utter derivatives. J.
Wind Eng. Ind. Aerodyn., 69-71:829837, 1997.
[183] Plaut RH. Snap loads and torsional oscillations of the original Tacoma Narrows
Bridge. J. Sound Vib., 309:613636, 2008.
[184] Larose GL and Livesey FM. Performance of streamlined bridge decks in relation
to the aerodynamics of a at plate. J. Wind Eng. Ind. Aerodyn., 6971:851860,
1997.
[185] Schewe G. Reynolds-number eects in ow around more-or-less blu bodies. J.
Wind Eng. Ind. Aerodyn., 89:12671289, 2001.
[186] Den Hartog JP. Transmission line vibration due to sleet. Trans. AIEE, 51:1074
1086, 1932.
[187] Den Hartog JP. Mechanical Vibrations. McGraw-Hill, New York, 3rd edition,
1947.
[188] Holmes JD. Wind Loading of Structures. Spon Press, New York, 1st edition,
2001.
[189] Hemon P and Santi F. On the aeroelastic behaviour of rectangular cylinders in
cross-ow. J. Fluid Struct., 16:855889, 2002.
[190] Richardson AS and Martuccelli JR. Research study on galloping of electric power
transmission lines, volume II, pages 611686. HMSO, London, UK, 1965.
[191] Jones KF. Coupled vertical and horizontal galloping. J. Eng. Mech.-ASCE,
118:92107, 1992.
[192] Liang SG, Li QS, Li G, and Qu W. An evaluation of the onset wind velocity for 2d
coupled oscillations of tower buildings. J. Wind Eng. Ind. Aerodyn., 50:329340,
1993.
138 REFERENCES
[193] Li QS, Fang JG, and Jeary AP. Evaluation of 2d coupled galloping oscillations
of slender structures. Comput. Struct., 66:513523, 1998.
[194] Davenport AG. The treatment of wind loading on tall buildings. In Symposium on
Tall Buildings, pages 345, University of Southampton, England, 1966. Pergamon
Press.
[195] Luongo A and Piccardo G. Linear instability for coupled translational galloping.
J. Sound Vib., 288:10271047, 2005.
[196] Chabart O and Lilien JL. Galloping of electrical lines in wind tunnel facilities.
J. Wind Eng. Ind. Aerodyn., 7476:967976, 1998.
[197] Yu P, Desai YM, Shah AM, and Popplewell N. Three-degree-of-freedom model
for galloping. Part I: Formulation. J. Eng. Mech.-ASCE, 119:24042425, 1993.
[198] Wang JW and Lilien JL. Overhead electrical transmission line galloping: A full
multi-span 3-DOF model, some applications and design recommendations. IEEE
T. Power Deliver., 13(3):909916, 1998.
[199] Blevins RD and Iwan WD. The galloping response of a two-degree-of-freedom
system. J. App. Mech., 41:11131118, 1974.
[200] Desai YM, Shah AH, and Popplewell N. Galloping analysis for two-degree-of-
freedom oscillator. J. Eng. Mech.-ASCE, 116:25832602, 1990.
[201] McComber P and Paradis A. A cable galloping model for thin ice accretions.
Atmos. Res., 46:1325, 1998.
[202] Bearman PW, Gartshore IS, Maull DJ, and Parkinson GV. Experiments on ow-
induced vibration of a square-section cylinder. J. Fluid Struct., 1:1934, 1987.
[203] Norberg C. Flow around rectangular cylinders: Pressure forces and wake fre-
quencies. J. Wind Eng. Ind. Aerodyn., 49:187196, 1993.
[204] Alonso G, Valero E, and Meseguer J. An analysis on the dependence on cross
section geometry of galloping stability of two-dimensional bodies having either
biconvex or rhomboidal cross sections. Eur. J. Mech. B/Fluids, 28:328334, 2009.
[205] Tatsuno M, Takayama T, Amamoto A, and Koji I. On the stable posture of a
triangular or a square cylinder about its central axis in a uniform ow. Fluid
Dyn. Res., 6:201207, 1990.
[206] Alonso G, Meseguer J, and Prez-Grande I. Galloping instabilities of two-
dimensional triangular cross-section bodies. Exp. Fluids, 38:789795, 2005.
REFERENCES 139
[207] ESDU 82007. Structural members: mean uid forces on members of various cross
sections. Engineering Sciences Data Unit, London, UK, 2004.
[208] Weaver DS and Veljkovic I. Vortex shedding and galloping of open semi-circular
and parabolic cylinders in cross-ow. J. Fluid Struct., 21:6574, 2005.
[209] Nigol O and Buchan PG. Conductor galloping part I - Den Hartog mechanism.
IEEE T. Power Ap. Syst., 100(2):699707, 1981.
[210] Luo SC, Chew YT, Lee TS, and Yazdani MG. Stability to translational galloping
vibration of cylinders at dierent mean angles of attack. J. Sound Vib., 215:1183
1194, 1998.
[211] Papanikolas P and Flamand O. Vibration of lightning protection cables on Rion-
Antirion Bridge. In 8th International Symposium on Cable Dynamics, pages
261268, Paris, France, 2009.
[212] Allen DW and Henning DL. Vortex-induced vibration tests of a exible smooth
cylinder at supercritical Reynolds numbers. In 7th International Oshore and
Polar Engineering Conference, Vol. 3, pages 680685, Honolulu, USA, 1997.
[213] Cheng S, Larose GL, Savage MG, Tanaka H, and Irwin PA. Experimental study
on the wind-induced vibration of a dry inclined cable Part I: Phenomena. J.
Wind Eng. Ind. Aerodyn., 96:22312253, 2008.
[214] Cheng S, Larose GL, Savage MG, and Tanaka H. Aerodynamic behaviour of an
inclined circular cylinder. Wind Struct., 6(3):197208, 2003.
[215] Larose GL, Jakobsen JB, and Savage MG. Wind-tunnel experiments on an in-
clined and yawed stay cable model in the critical Reynolds number range. In 5th
International Symposium on Cable Dynamics, pages 279286, Santa Margherita
Ligure, Italy, 2003.
[216] Jakobsen JB, Larose GL, and Savage MG. Instantaneous wind forces on inclined
circular cylinders in critical Reynolds number range. In 11th International Con-
ference Wind Engineering, Vol. 2, pages 21652173, Lubbock, USA, 2003.
[217] Cheng S, Irwin PA, and Tanaka H. Experimental study on the wind-induced
vibration of a dry inclined cable Part II: Proposed mechanisms. J. Wind Eng.
Ind. Aerodyn., 96:22542272, 2008.
[218] Jakobsen JB, Larose GL, and Andersen TL. Interpretation of wind forces moni-
tored on inclined stationary cylinder in critical Reynolds number range in relation
140 REFERENCES
to observed aeroelastic model response. In 6th International Symposium on Cable
Dynamics, pages 287294, Charleston, USA, 2005.
[219] West GS and Apelt CJ. The eects of tunnel blockage and aspect ratio on the
mean ow past a circular cylinder with Reynolds numbers between 10
4
and 10
5
.
J. Fluid Mech., 114:366377, 1982.
[220] Blackburn HM. Eect of blockage on spanwise correlation in a circular cylinder
wake. Exp. Fluids, 18:134136, 1994.
[221] Irwin PA, Cooper KR, and Girard R. Correction of distortion eects caused
by tubing systems in measurements of uctuating pressures. J. Wind Eng. Ind.
Aerodyn., 5:93107, 1979.
[222] Gatto A, Byrne KP, Ahmed NA, and Archer RD. Mean and uctuating pressure
measurements over a circular cylinder in cross ow using plastic tubing. Exp.
Fluids, 30:4346, 2001.
[223] Achenbach E and Heinecke E. On vortex shedding from smooth and rough cylin-
ders in the range of Reynolds numbers 610
3
to 510
6
. J. Fluid Mech., 109:239
251, 1981.
[224] Shih WCL, Wang C, Coles D, and Roshko A. Experiments on ow past rough cir-
cular cylinders at large Reynolds numbers. J. Wind Eng. Ind. Aerodyn., 49:351
368, 1993.
[225] Vickery BJ and Basu RI. Across-wind vibrations of structures of circular cross-
section. part I. Development of a mathematical model for two-dimensional con-
ditions. J. Wind Eng. Ind. Aerodyn., 12:4979, 1983.
[226] Andersen TL, Jakobsen JB, Macdonald JHG, Nikitas N, Larose GL, Savage MG,
and McAulie BR. Drag-crisis response of elastic cable-model. In 8th Interna-
tional Symposium on Cable Dynamics, Paris, France, 2009.
[227] Moon FC. Chaotic Vibrations: An Introduction for Applied Scientists and Engi-
neers. John Wiley & Sons, New York, 1st edition, 1987.
[228] Schewe G. Sensitivity of transition phenomena to small perturbations in ow
round a circular cylinder. J. Fluid Mech., 172:3346, 1986.
[229] Symes J. Dry inclined galloping of smooth circular cables in the critical Reynolds
number range. PhD thesis, Department of Civil Engineering, University of Bristol,
Bristol, UK, 2011.
REFERENCES 141
[230] Staubli T. Calculation of the vibration of an elastically mounted cylinder using
experimental data from forced oscillation. J. Fluids Eng., 105:225229, 1983.

S-ar putea să vă placă și