2001 A Unified Approach For 3D Stability and Time Domain Response Analysis With Application of Quasi-Steady Theory

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Wind Engineering

and Industrial Aerodynamics 89 (2001) 1591–1606

A unified approach for 3D stability and time


domain response analysis with application
of quasi-steady theory
Stoyan Stoyanoff *
Rowan Williams Davies & Irwin Inc., 650 Woodlawn Road West, Guelph, Ont.,
Canada N1K 1B8

Abstract

This paper focuses on a general quasi-steady approach for evaluation of three-


dimensional (3D) galloping stability and buffeting response of elastic structures such as
bridges and towers. The buffeting response is obtained as a time history, solving the
generalised state equation of dynamic equilibrium. Its homogeneous solution allows
evaluation of the galloping stability (linear formulation). The time histories of the wind
velocities required for the buffeting analysis are numerically generated. Buffeting responses of
the Millennium Footbridge over the Thames River in London, are simulated and the statistics
of the results are compared with a conventional frequency domain method. Stability and
accuracy of the integration are examined to establish the range of the solution parameters. The
galloping stability analysis is verified on the I235 Pedestrian Bridge, Des Moines, Iowa, where
the results are compared with the wind tunnel predictions. r 2001 Elsevier Science Ltd. All
rights reserved.

Keywords: Quasi-steady theory; Turbulence; 3D response; Galloping; Buffeting; Bridges; Towers; Time
domain analysis; Generalised state equation; Numerically generated wind speeds; Frequency domain
method; Integration accuracy; Millennium Footbridge; I235 Pedestrian Bridge; Comparison with wind
tunnel

*Tel.: +1-450-375-0899; fax: +1-519-823-1316.


E-mail address: sts@rwdi.com (S. Stoyanoff).

0167-6105/01/$ - see front matter r 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 0 1 ) 0 0 1 5 7 - X
1592 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

1. Development of quasi-steady galloping and buffeting approach

Den Hartog [1] explained galloping of a simple one degree-of-freedom (SDOF)


oscillator using static force coefficients. This approach was later extended by
Parkinson and Brooks [2], Novak and Davenport [3], and Iwan [4]. Subsequently,
Blevins and Iwan [5] studied coupled across-wind and torsional galloping, and later
Jones [6] studied both along- and across-wind two-degree-of-freedom (2DOF)
galloping. Among other effects, the turbulence effects on galloping were investigated
experimentally by Novak and Davenport [3], Laneville and Parkinson [7] and also
simulated numerically by Lindner [8]. Parallel to these studies, and again based on
quasi-static assumptions, the theory of wind-induced buffeting was developed by
Liepmann [9], Davenport [10], Holmes [11], and Irwin [12]. In general, the quasi-
steady methods apply to line-like structures where, on a representative fixed-length
of the cross-section (strips), the wind induces fluctuating moment, lift and drag
forces. In reality, the wind–structure interaction is quite a bit more complex than
this. The fluctuating 3D wind forces excite the structure and those forces are
modified by the motions and also by the inherent turbulence generating properties of
the structure. Scanlan [13] took into account the self-excitation effects in terms of
aerodynamic derivatives. In an attempt to simplify the buffeting problem, the
equilibrium equations are typically solved in frequency domain where the fluctuating
wind is represented in general statistical terms by its spectra. Additionally, modal
coordinates are applied, and after integration along the structural span (all strips),
the problem is reduced from a 3D multi-degree-of-freedom (MDOF) problem, to a
SDOF generalized case. When analysing structures with coupled modes, the single
mode representation requires certain simplifications to be made. One illustration is
the neglecting of the aerodynamic damping induced by the coupling, which typically
leads to conservative predictions. In spite of these simplifications and other
shortcomings however, the quasi-steady buffeting method has proved to be quite
useful and has been applied for derivation of wind loads on many structures
[12,14,15]. An alternative to frequency-domain method (FDM), the time-domain
method (TDM) appears to be the most general approach for theoretical simulation
of time-dependent structural responses. Beliveau et al. [16] directly solved the
generalised buffeting problem in time domain via Duhamel type integrals while other
authors, as for example Arzoumanidis and Bieniek [17], preferred to numerically
integrate the full system of finite element equations [18]. The latter approach has
advantages over the former if structural or flow/structure nonlinear interaction
effects are considered. Typically, TDM requires numerically simulated time histories
of wind speeds with statistical properties similar to those of the natural wind [19–22].
These wind series include longitudinal UðtÞ; lateral vðtÞ; and vertical wðtÞ wind
velocity components. Today, most of the available load functions are based on static
force coefficients. Thus it is practical to use the accumulated experimental data on
various cross-sections, apply simulated wind velocity histories, derive wind loads as
time series, and integrate for the buffeting response.
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1593

2. Quasi-steady response analysis

2.1. Basic assumptions

The present approach, concentrating on practical procedures able to predict


galloping and wind buffeting of structures, is based on the quasi-steady assumptions
([23] or elsewhere) that:

(a) the instantaneous forces exciting the moving structure are almost equal to the
static forces measured at the same effective angle of wind attack; and
(b) the phase sift between the fluid force and the body velocity is negligibly small.
These conditions could be met at wind speeds where no vortex shedding lock-in and/
or flutter occur. Other important assumptions are:

(1) The structure is situated in an open terrain, where the influence of the other
adjacent structures on the fluctuating part of the wind is small, i.e., only the
effect of wind buffeting is considered.
(2) The cross-sectional dimensions of the structural elements are small compared to
the free lengths exposed to the wind and the aerodynamic influence of one part
of the structure on the others is not considered. Therefore, the line-strip
approach is applicable and permits a division of the exposed elements into a
finite number of sections and nodes. All properties, structural and aerodynamic
are assumed to be constant and lumped into the nodes.
(3) The aerodynamic properties of each section are governed by the wind field where
its modification effects due to the section geometry are incorporated in the wind
model by means of aerodynamic admittance. Thus the wind speeds are
converted into instantaneous wind forces and moments where the time lag
effects are neglected as small compared to the overall response time lengths.
(4) The mean wind direction is assumed to be approximately normal to the exposed
areas.
(5) The structure is expected to vibrate elastically around its mean deflected position
with small amplitudes. Hence the resulting structural nonlinear effects are
presumed to be insignificant.
(6) An internal reduction from Cartesian to generalized coordinates is applied
which enables analysis of large MDOF systems at reasonable time.
(7) The process involves step-by-step integration in time of buffeting responses.
Thus the wind load should be given explicitly as a function of time. Here the
wind load series are generated as multi-correlated stochastic processes by auto-
regressive (AR) filtering technique [22]. Aerodynamic admittance is incorporated
in the simulation using Irwin’s equation [12].

2.2. Quasi-static wind loads

Fig. 1 shows a 2D cross-section of a horizontal elastic structure loaded by


turbulent wind. After neglecting the higher order terms, the relative wind velocity
1594 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

Fig. 1. Positive speeds, deflections, forces, and moment.

vector Urel and its effective angle of attack a can be presented [12] as

wðtÞ  zðtÞ
ðtÞ ¼ U% þ 2UuðtÞ
%  2U% xðtÞ;
2
2
Urel ’ a0 ðtÞ ¼ ;
%
U ð1Þ
aðtÞ ¼ a0 ðtÞ þ yyy ðtÞ; yyy ðtÞ ¼ y% yy þ y* yy ðtÞ;
where u and w are the along- and across-wind fluctuating speeds, and U% is the mean
speed in the mean wind direction. Since the cross-section is free to oscillate, for
convenience its translation x can be taken in the direction of the mean wind, then z is
the across-wind translation, and yyy is the rotation also taken relative to the mean
wind direction where y% yy is the mean, and y* yy the fluctuating rotation angles.
According to the quasi-steady fundamentals, the instantaneous forces and moments
acting on a unit length cross-section at different angles of attack can be estimated
from the corresponding static force and moment coefficients measured in the wind
tunnel as
FD ðtÞ ¼ 12rbUrel
2
ðtÞCD ðaðtÞÞ; FL ðtÞ ¼ 12rbUrel
2
ðtÞCL ðaðtÞÞ;
ð2Þ
MðtÞ ¼ 12rb2 Urel
2
ðtÞCM ðaðtÞÞ;
with r the air density, or in body coordinates as
FX ðtÞ ¼ FD ðtÞ cosðaðtÞÞ  FL ðtÞ sinðaðtÞÞ;

FZ ðtÞ ¼ FD ðtÞ sinðaðtÞÞ þ FL ðtÞ cosðaðtÞÞ; ð3Þ

MðtÞ ¼ MðtÞ:
It can be proved that for small angles of attack, up to about 710 degrees (with error
of about two percent)
FX ðtÞ ¼ 12rbUrel
2
ðtÞðCD ðaðtÞÞ  CL ðaðtÞÞa0 ðtÞÞ;

FZ ðtÞ ¼ 12rbUrel
2
ðtÞðCD ðaðtÞÞa0 ðtÞ þ CL ðaðtÞÞÞ; ð4Þ

MðtÞ ¼ 12rb2 Urel


2
ðtÞCM ðaðtÞÞ:
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1595

The force and moment coefficients can be then expanded in Taylor series around the
assumed mean angle a0 ¼ y% yy of attack which for many cases is negligibly small.
After retaining only the linear part of that expansion
 
dCD  dC 

aðtÞ ;
L
CD ðaðtÞÞ ¼ CD þ CL ðaðtÞÞ ¼ CL þ aðtÞ ;
da %yyy da  %
yyy
 ð5Þ
dCM  
CM ðaðtÞÞ ¼ CM þ aðtÞ ;
da y% yy

which are the slopes of the corresponding force and moment coefficients [1]. Using
Eqs. (4) and (5) the load on each finite-strip j along the structure is
f f gj ¼ ½ca j fzg ’ j þ ½ka j fzgj þ ½o j fwgj ; ð6Þ
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflffl{zfflfflfflffl}
Self-excited forces Buffeting forces

where j ¼ 1; 2; y; n strips or element section, and ½ca ; ½ka ; and ½o are the
aerodynamic damping, stiffness and buffeting load matrices
2 0
3 8 x’ 9
2CD ðCL  CD Þ 0 < U% >
> =
6 7 z’
½ca j ¼ plj 4 2CL ðCD þ CL0 Þ 0 5 ; ’ j¼
fzg U%
;
0 :’ >
> ;
2bCM bCM 0 j yyy j

2 0
3 8 9
0 0 CD < x >
> =
6 7
½ka j ¼ plj 4 0 0 CL0 5 ; fzgj ¼ z ;
>
: >
;
0 0 bCM 0
yyy j ð7Þ
j

2 0
3 8 9
CD 2CD ðCL  CD Þ <1>
> =
6 7 u
½o j ¼ plj 4 CL 2CL ðCD þ CL0 Þ 5 ; fwgj ¼ U% ;
0 :w>
> ;
bCM 2bCM bCM j U% j
dCDðL;MÞ
p ¼ 12rbU% j ;
2 0
CDðL;MÞ ¼ :
da
where lj is the exposed wind strip length. In terms of body coordinates these matrices
are
2 0
3 2 0
3
2CX CX 0 0 0 CX
6 7 6 7
½ca j ¼ plj 4 2CZ CZ0 05 ; ½ka j ¼ plj 4 0 0 CZ0 5 ;
0 0
2bCM bCM 0 j 0 0 bCM j
ð8Þ
2 0
3
CX 2CX CX
6 7 dCXðZ;MÞ
½o j ¼ plj 4 CZ 2CZ CZ0 5; 0
CXðZ;MÞ ¼ :
0
da
bCM 2bCM bCM j
1596 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

Considering that for vertical strips vðtÞ replaces wðtÞ; y  z; yzz  yyy ; and expanding
½ca j ; ½ka j ; ½oa j to 6DOF per node, the overall matrices assemble
2 3
½a 1 0 : 0
6 0 ½a 0 7
6 2 : 7
½A ¼ 6 7 ; ð9Þ
4 : : : : 5
0 0 : ½a n 6n 6n
where ½A ¼ ½Ca ; ½Ka ; or ½O ; and ½a ¼ ½ca ; ½ka ; or ½o :
The unloaded DOF are filled with zeroes and for those sheltered from wind strips,
f f gj ¼ 0: The overall vectors are
fFgT ¼ ff f g1 ; f f g2 ; y; f f gn gT ; fWgT ¼ ffwg1 ; fwg2 ; y; fwgn gT ;
ð10Þ
’ T ¼ ffzg
fZg ’ 1 ; fzg ’ n gT ;
’ 2 ; y; fzg fZgT ¼ ffzg1 ; fzg2 ; y; fzgn gT :

2.3. The equilibrium equations

The equation of equilibrium governing the dynamic response of the system is


. þ ½C fZg
½M fZg ’ þ ½K fZg ¼ fFg; ð11Þ
where fZg is the nodal displacement vector and matrices ½M ; ½C ; and ½K ; have their
usual meaning as in structural mechanics. Generally, the total number of DOF is
N ¼ 6n; where n is the number of nodes (strips). As per Section 2, the wind forces
and moments are three per node, two being normal to the strip’s along-axis, drag
and lift, and a moment around it. A conventional approach to reduce the number of
unknowns which we have to solve for, is the modal synthesis technique that implies a
linear superposition of modes [26]. This reduction is applicable provided that a
particular system vibrates within small amplitudes about its current equilibrium
state. The modal superposition technique requires a preliminary step of solving the
free undamped eigensystem problem. The reduction transformation is then
fZðtÞg ¼ ½X fFðtÞg; where ½X N m is the modal matrix composed of the first m
eigenvectors fXgk ; k ¼ 1; 2; y; m of the undamped system, and fFðtÞg is the vector
of the generalized coordinates. Using this reduction transformation, Eqs. (9) and
(10) and multiplying from the left equation (11) by ½X T ; results in a reduced system
of scalar equations of motion where typically m5N
.
½M n fFðtÞg þ ½C* n fFðtÞg
’ þ ½K* n fFðtÞg ¼ fFgn ð12Þ
and ½M n ; ½C* n ; ½K* n ; and fFgn are the generalised mass, damping, and stiffness
matrices, and buffeting load vector. The reduction requires also solving for the
orthogonal eigenmatrix ½X of the undamped homogeneous system which serves then
as an orthonormal basis for normalization of Eq. (12). The generalised quantities are
½M n ¼ ½X T ½M ½X ¼ I; ½C* n ¼ diag½2zj oj  ½X T ½Ca ½X ;
ð13Þ
fFgn ¼ ½X T ½O fWg; ½K* n ¼ diag½o2j  ½X T ½Ka ½X ;
where zk ; and ok are the viscous damping ratio and the circular frequency of mode k:
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1597

Eq. (12) can be rewritten in a canonical form as


      
½C* n ½K* n ’
fFðtÞg .
fFðtÞg fFðtÞgn
¼ ’  ; ð14Þ
I 0 fFðtÞg fFðtÞg 0
|fflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflffl{zfflfflfflfflfflffl} |fflfflfflfflfflffl{zfflfflfflfflfflffl} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl}
½S fyðtÞg ’
fyðtÞg fYðtÞg

called the generalised state equation of the system [25]. Adopting the notations in the
braces, this equation takes the standard form

fyðtÞg ¼ ½S fyðtÞg þ fYðtÞg: ð15Þ

3. Stability and response solutions

3.1. Stability problem

The homogeneous solution of Eq. (15) or when fYðtÞgE0 is in fact the well-
known linear stability problem. It is obtained from the eigenvalue solution as a set of
conjugated complex eigenvalues lk and vectors fUgk ; for the k ¼ 1; 2; y; 2m
retained modes
( ) ( )

fFðtÞg lffg
lt
fyðtÞg ¼ ¼ fUge ; fUg ¼ ; ð16Þ
fFðtÞg ffg
where ffg are the generalised eigenvectors of the transformed generalised system.
The variable
lk ¼ ðzk þ iÞok ; ð17Þ
where zk is the total damping ratio (comprising aerodynamic plus structural, d ¼ 2pz
is the logarithmic decrement) and ok is the circular frequency of response. The
criterion for stability of any structure is the condition if
(k: lk p0 divergent or non-decaying oscillations (i.e., galloping) or if
8k: lk > 0 decaying oscillations of a dynamically stable system.
That is, a system is regarded as dynamically stable if and only if the lk > 0 stability
conditions are fulfilled for all modes.

3.2. Buffeting response

Obtaining the response of a stable system to turbulent wind requires solution of a


system of order 2m (yet 2m5N) that is Eq. (15), in the form [25]
Z t
t½S
fyðtÞg ¼ e fyð0Þg þ eðttÞ½S fYðtÞg dt; ð18Þ
0
ðtÞ½S
where e is referred to as the transition matrix constant in time and ½S 2m 2m is a
real non-symmetric matrix. For calculation of this transition matrix, it was also
suggested [25] using the computationally attractive biorthonormal properties of the
1598 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

eigensystem of ½S solving the adjoint problem


½S fUgk ¼ lk fUgk ; ½S T fVgk ¼ lk fVgk : ð19Þ

One can then find the left ½V and ½U right complex generalised eigenmatrices
½U ¼ ½fUg1 ; fUg2 ; y; fUg2m ; ½V ¼ ½fVg1 ; fVg2 ; y ;
ð20Þ
½l ¼ diag½lk ;

where the biorthonormal properties imply that


½V T ½U ¼ I; ½U ½V T
¼ I; ½V T ½S ½U ¼ ½L ;
T
ð21Þ
½U ½L ½V ¼ ½S :

The solution of Eq. (15) then simplifies to


Z t
fyðtÞg ¼ ½U et½L ½V T fyð0Þg þ ½U eðttÞ½L ½V T fYðtÞg dt; ð22Þ
0

where fyð0Þg is the initial state vector and et½L is a diagonal matrix. Provided the load
vector is given in discrete-time form at equal Dt time intervals, t ¼ pDt; p ¼ 0; 1; 2; y
the response can be calculated step by step, based only on the previous step, not as
required by the integral of Duhamel-type given in Eq. (18). The discrete-time
formula response calculation is
Z Dt
T T
½V fyðp þ 1Þg ¼ e Dt½L
½V fyðpÞg þ et½L ½V T fYðp þ 1Þg dt: ð23Þ
|fflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} 0 |fflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflffl}
fWðpþ1Þg fWðpÞg fQðpþ1Þg

Assuming that the load changes linearly from time p to time p þ 1; the exact solution
of Eq. (23) is
fWðp þ 1Þ ¼ eDt½L fWðpÞg þ ½L 1 ðeDt½L  IÞfQðp þ 1Þg; ð24Þ

where ½L 1 ¼ diag½1=lk is constant in time. At any time, the final response in


Cartesian coordinates can be obtained using the back transformation
( )

fZðpÞg
fZðpÞg ¼ ½w ½U ½WðpÞ ; fZðpÞg ¼ ;
fZðpÞg
" # ð25Þ
½X 0
½w ¼ ;
0 ½X

of size 2n and 2n 2m: From the given initial conditions, the starting state is
½Wð0Þ ¼ ½V ½*w 1 ½w T fZð0Þg;
T
½*w ¼ ½w T ½w ¼ diag½xk ;
ð26Þ
½*w 1 ¼ diag½1=xk ; xk ¼ xkþm ¼ fXgTk fXgk ;

which completes all necessary steps required for obtaining the response to wind
buffeting.
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1599

4. Examples

4.1. Response of London Millennium Footbridge

Fig. 2 shows a typical deck of the Millennium Footbridge over Thames River in
London which is used to illustrate the method. The bridge main span is 144 m with
total length of 325 m linking St. Paul’s Cathedral and the new Tate Gallery of
Modern Art. This unique elegant structure features a light-weight curved deck
supported on high-strength cables with variable width end elevation relative to the
deck curve. Force and moment coefficients of representative cross-sections were
obtained in the wind tunnel laboratory of RWDI, Guelph, Ontario, Canada. Wind
loads required for design of this bridge were obtained by applying the well
established FDM buffeting procedure of Irwin [12]. In parallel, the present method
was applied on a 41 nodes discrete model with typical strip length lj ¼ 8 m (Fig. 3).
Buffeting responses during storms with various durations of up to 54:61 min were
simulated. Fig. 4 shows a sample of longitudinal and vertical wind speeds generated
in the bridge middle (node 19). The turbulence model proposed by ESDU [24] was
used for the TDM analyses. Note that the FDM procedure was based on Irwin’s [12]
turbulence model. The aerodynamic admittance 2D function proposed by Irwin was

Fig. 2. Typical cross-section of Millennium Footbridge (courtesy of the Ove Arup & Partners, London,
UK).

l 17 = 8 m

Thames River

Fig. 3. The main span middle of the 41 node model.


1600
Velocity V+u(t) (m/sec) S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

40

30

20

10

0
0 10 20 30 40 50
Time (min)
10
Velocity w(t) (m/sec)

10
0 10 20 30 40 50
Time (min)

Fig. 4. Wind speeds at the bridge middle, elevation 15 m; time step Dt ¼ 0:1 s; turbulence properties:
Iu ¼ 18:9%; Iw ¼ 10:4%; x Lu ¼ 146:3 m; x Lw ¼ 12:2 m:

10 10
Normalized Power Spectra
Normalized Power Spectra

1 1

0.1 0.1

0.01 0.01
0.01 0.1 1 10 0.01 0.1 1 10
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 5. Power spectrum of longitudinal wind turbulence. (a) Admittance included, (b) Unity admittance.

also incorporated when simulating the wind series as a modification to the power
spectrum functions. Fig. 5 demonstrates the effect of including the aerodynamic
admittance on the target and simulated power spectra of the wind speeds. The
spectral folding above 2:5 Hz is due to the time step 0:1 Hz used for wind series
generation. However its effect was judged small for this study where the frequency of
Mode 20 is at 1:434 Hz: Table 1 shows a detailed comparison between FD and TD
methods.
Comparing the responses in single, uncoupled vibration modes, it can be judged
that the two methods (based on two different turbulence models) gave reasonably
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1601

Table 1
FDM vs. TDM peak dynamic responses, solid barriers, mean speed U% ¼ 23:3 m=s; time 13:65 min; time
step Dt ¼ 0:05 s; structural damping ratio z ¼ 0:5%

Span Mode Frequency Predicted responses in span middle (m/rad)


(Hz)
FDM TDM TDM
20 modes
Res. Total Res. Total

North Vertical 0.7014 0.0806 0.0960 0.0806 0.1007 0.1187


Lateral 0.7518 0.0204 0.0232 0.0211 0.0330 0.0375
Torsional 0.8871 0.0021 0.0024 0.0025 0.0030 0.0073

Main-node Vertical 0.4117 0.2055 0.2335 0.1921 0.2354 0.2446


(19) Lateral 0.4401 0.0523 0.0784 0.0586 0.0814 0.0788
Torsional 0.6379 0.0039 0.0043 0.0040 0.0043 0.0072

South Vertical 0.5317 0.1356 0.1594 0.1480 0.1805 0.1893


Lateral 0.7297 0.0202 0.0241 0.0271 0.0395 0.0376
Torsional 0.8871 0.0021 0.0024 0.0020 0.0024 0.0040
Lateral displacements (m)

0.05

0.05
0 10 20 30 40 50
Time (min)
Fig. 6. Lateral response at node 19, low solidity barrier, 20 modes, U% ¼ 23:3 m=s; total time ¼
54:61 min; Dt ¼ 0:05 s; z ¼ 0:5%:

close results. The TDM tended predicting higher responses in the lateral direction.
This was attributed to increased correlations in the wind loads. Simulations using the
first 20 modes and all cross-components showed no significant increases in lateral
and vertical predictions due to the fact that most of the added modes with
contributions in those directions had higher frequencies. Although the torsional
response predictions were small, the 20-mode simulation gave higher predictions
because most of the lower frequency modes had torsional components contributing
to the overall response.
Further analysis on the performance of the FD method was carried out using the
wind tunnel data of the actually built deck with low solidity barriers. Fig. 6 presents
54:61 min sample of lateral response at node 19. The power spectrum given in Fig. 7
1602 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

Fig. 7. Lateral responses F power spectrum and histogram.

0.001
Lateral displacements (m)

8 .10
4

6 .10
4

4 .10
4 - X - , ∆ t = 0.025 sec
-+- , ∆ t = 0.050 sec

2 .10
4
1100.2 1100.3 1100.4 1100.5 1100.6
Time (sec)

Fig. 8. Lateral responses for time steps: Dt ¼ 0:025 and 0:050 s:

shows that the dominating mode in the lateral response is the corresponding lowest
frequency mode (see Table 1). The histograms display Gaussian distributions with
peak factors of about 3.3. The numerical stability and accuracy of the FD method
was verified exploring in detail the motions predicted at node 19. Although the
response solution as provided by Eq. (24) is exact, its integration time step Dt is
preferable to be at least two times smaller than the wind simulation time step, 0:1 s in
this case, to preserve the spectral properties of the loads. A numerical experiment
was carried out where the time series step was reduced by splining the loads from
0.05 to 0:025 s:
Fig. 8 reveals that a double reduction of the time step was sufficient to achieve an
adequate response representation.
Fig. 9 compares the power spectra of the two responses where, as it could be
expected, the two power spectra collapse up to about 2:5 Hz (not shown here) since
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1603

Fig. 9. Response power spectra of lateral deflections: Dt ¼ 0:025 and 0:050 s:

there is no loading frequency components higher than 5 Hz (see Fig. 5). The
frequency of the highest mode used for analysis, mode 20 at f ¼ 1:434 Hz was
accurately estimated using either time step and the difference in the root-mean-
square (rms) response was about 0.5%.
The TD method relies on a discrete technique, where all aerodynamic and
structural properties are assumed piece-wise constant and lumped into a chain of
nodes. Consequently, a 79 node model was also created to estimate the possible
deviations in the simulation. Compared to the 41 node model, the predicted
responses showed differences of about 3–5%. Considering that the bridge geometry,
mass, and force and moment coefficients vary along the bridge, these response
differences were attributed to input data extrapolation rather than to the accuracy of
the discrete model. The typical strip length lj was 8 m for the 41 node model and
about 4 m for the 79 node model. Thus it can be concluded that a model with lengths
lj in the range of b to 2b (where b ¼ 5:3 m is the deck width in this case) enables
realistic predictions.
The peak deflections at node 19 were found creating response histograms (Fig. 7)
for various sample lengths. Fig. 10 shows that more than 10 min of response was
required for stable predictions. Therefore, considering the lowest structural
frequency, about 29 response cycles were needed.

4.2. Galloping stability of the I235 footbridge

Fig. 11 shows peak vertical response of the I235 Pedestrian Arch Bridge (span
L ¼ 80 m). The site at Des Moines, Iowa, USA, was characterized with extreme ice
accumulations of 40 mm once every 50 years. It was, therefore, found that the
pedestrian fences could be solid frozen, which raised concerns for galloping
instability. The RWDIs wind tunnel study, however, showed that the onset of
galloping speed is higher than the required speed criterion. Using a simplified
1604 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

0.5

Peak deflections (m/deg)


0.4 st
1 Torsional f=0.6379 Hz

0.3
st
1 Vertical f=0.4117 Hz
0.2

st
1 Lateral f=0.4401 Hz
0.1

0
0 10 20 30 40 50 60
Time (min)

Fig. 10. Predicted responses for various time intervals.

Iced Fences
d = 2.715 m

galloping speed 39.4 m/sec


0.015
Theoretically predicted
Peak vertical deflection (z/d)

0.010 b=6.0 m
f = 1.157 Hz
m=6765 kg/m
ζ =0.55% C =0.64 D
0.005
dC L /dα=-4.53

0.000
0 1 2 3 4 5 6 7U/fb

0 10 20 30 40 (m/sec)
Wind velocity
Fig. 11. Peak vertical response at mean angle 0 degree (courtesy HNTB Corporation, Kansas City,
Missouri, USA).

Eq. (27) of Den Hartog type


RL
8pSc mEq z 0 mðsÞFz ðsÞ ds
UR ¼ ; Sc ¼ ; mEq ¼ RL 2 ; ð27Þ
ðdCL =da þ CD Þ rb2
0 Fz ðsÞ ds

with Sc the Scruton number, mEq ; the equivalent mass of the first vertical mode,
Fz ðsÞ its mode shape, mðsÞ the deck mass, and s the coordinate along the deck; the
onset of galloping speed was found as UR ¼ 5:68 ðUR ¼ UGal =ðfbÞ; UGal ¼
39:4 m=sÞ: Exactly the same onset speed was also predicted by the proposed method
on a 10 node numerical model. This example was used as a simple means of checking
the galloping procedure.
S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606 1605

5. Conclusions

An extension to the existing quasi-steady methods for galloping and buffeting is


presented enabling simultaneous verification of stability and simulation of buffeting
responses. The accuracy of the proposed procedures is demonstrated on two bridge
structures. The important statistics of the response as peak resonant and total
deflections compared satisfactorily with those predicted by the well recognised
buffeting theory in the frequency domain, Irwin [12]. Stability and accuracy of the
computation procedures are examined in detail. A maximum time step t ¼ 0:05 s;
and minimum sample duration of 29 =(lowest structural frequency) seconds are
recommended.
Despite its theoretical and numerical complexity, the proposed procedures require
less manipulation of the input data compared to the FDM. Although for now only 3
load components per node are taken into account and the wind is assumed as being
normal to the structure, modal coupling and superposition are readily held by the
TDM permitting more realistic 3D response predictions. Simulated details in the
response as time histories of displacements, velocities, and acceleration, and precise
peak factors allow to refine the design of particular structural elements, for example,
supports, springs, and damper devices.
It should also be noted that the proposed simulation scheme is computationally
very effective. For example the whole simulation process for the longest time series of
55 min (20 modes, 41 nodes) takes about one half of that time on a PCF400 MHz:
Further studies on other line-like structures as bridges, pylons, and towers using
both FD and TD methods and variations of the turbulence properties are necessary
for additional verification of the proposed approach. Comparison with aeroelastic
model tests would permit fine tuning of the method and give new insights for
development.

Acknowledgements

The author is grateful to Drs. D. Novak, H. Herda, and K. Beronow, all visiting
scholars at Kyoto University back in 1992, for their help in creating the basics of this
work. Thanks to Dr. P. Irwin for his support and guidance of the RWDI’s Research
Project 94-052 where the method was largely developed. Part of that project was
supported by the National Science and Engineering Research Council of Canada
(NSERC). For the I235 Pedestrian Bridge example, thanks are due to the HNTB
Corporation, Kansas, Missouri, and to the Iowa Department of Transportation,
USA. The designers of the Millennium Footbridge, Ove Arup & Partners Consulting
Eng., London, UK are also acknowledged.

References

[1] J.P. Den Hartog, Transmission lines vibration due to sleet, AIEE 51 (1932) 1074–1076.
[2] G.V. Parkinson, N.P.H. Brooks, On the aeroelastic instability of bluff cylinders, ASME Trans.
J. Appl. Mech. 83 (1961) 252–258.
1606 S. Stoyanoff / J. Wind Eng. Ind. Aerodyn. 89 (2001) 1591–1606

[3] M. Novak, A.G. Davenport, Aeroelastic instability of prisms in turbulent flow, ASCE, EM1 96
(1970) 17–39.
[4] W.D. Iwan, Galloping oscillations of hysteretic structures, ASCE, EM99 6 (1973) 1129–1146.
[5] R.D. Blevins, W.D. Iwan, The galloping response of a two-degree-of freedom system, ASME,
J. Appl. Mech. 75 (1974) 1113–1118.
[6] K.F. Jones, Coupled vertical and horizontal galloping, ASCE, J. Eng. Mech. 118 (1992) 92–107.
[7] A. Laneville, G.V. Parkinson, Effects of turbulence on galloping of bluff cylinders, in: Conf. Proc.
Wind Effects Building and Structures, Tokyo, Japan, 1971, pp. 787–797.
[8] H. Lindner, Simulation of the turbulence influence of galloping vibrations, BBAA3, UWO, London,
Canada, 1990.
[9] H.W. Liepmann, On the application of statistical concepts to the buffeting problem, J. Aeronaut. Sci.
19 (1952) 793–822.
[10] A.G. Davenport, The response of slender, line-like structures to a gusty wind, Inst. Civ. Eng. 23 6610
(1962) 389–408.
[11] J.D. Holmes, Prediction of the response of a cable stayed bridge to turbulence, Fourth Wind Effects
Building Structures, Heathrow, UK, 1975.
[12] P.A. Irwin, Wind tunnel and analytical investigation of the response of Lions’ Gate Bridge to a
turbulent wind, NRCC Report LTR-LA-210, 1977.
[13] R.H. Scanlan, The action of flexible bridges under wind II: buffeting theory, Sounds Vibrations 60
(1978) 201–211.
[14] B.J. Vickery, K.H. Kao, Drag or along-wind response of slender structures, ASCE, ST1 98 (1972)
21–36.
[15] RWDI, Wind engineering studies for Cape Girardeau Bridge (steel alternate), RWDI Report
No. 94–197, 1994.
[16] J.-G. Beliveau, R. Vaicaitis, M. Shinozuka, Motion of suspension bridges subjected to wind loads, St.
Div., ASCE, ST6 103 (1977) 1189–1205.
[17] S. Arzoumanidis, G. Bieniek, M.P., Finite element analysis of suspension bridges, Comput. Struct. 21
(1985) 1237–1253.
[18] N.A. Newmark, A method for computation for structural dynamics, Trans. ASCE 127 (1962)
1406–1435.
[19] P. Spinelli, G. Solari, Time domain analysis of tall building response to wind action, Third
International Conference on Tall Buildings, Hong Kong, 1984, pp. 278–284.
[20] H.A. Buchholdt, S. Moossavinejad, A. Iannuzzi, Non-linear dynamic analysis of guyed masts
subjected to wind and guy ruptures, Inst. Civ. Eng. 2 (81) (1986) 353–395.
[21] D. Novak, S. Stoyanoff, H. Herda, Error assessment for wind histories generated by autoregressive
method, Struct. Safety 17 (1995) 79–90.
[22] S. Stoyanoff, P.A. Irwin, Simulation of wind loads in time domain, RWDI Research Project 94-052,
1995.
[23] E. Naudascher, D. Rockwell, Flow-Induced Vibrations: An Engineering Guide, A.A. Balkema,
Rotterdam, 1994.
[24] ESDU, Engineering Data Science Unit, Item Numbers ESDU 74030, 85020, 86010, London, UK,
1976–93.
[25] L. Meirovitch, Elements of Vibration Analysis, 2nd Edition, McGraw-Hill, New York, 1986.
[26] N.F. Morris, The use of modal superposition in nonlinear dynamics, Comput. Struct. 7 (1977) 65–72.

You might also like