You are on page 1of 15

J. Wind Eng. Ind. Aerodyn.

131 (2014) 31–45

Contents lists available at ScienceDirect

Journal of Wind Engineering


and Industrial Aerodynamics
journal homepage: www.elsevier.com/locate/jweia

Characteristics of distributed aerodynamic forces on a twin-box


bridge deck
Q. Zhu n, Y.L. Xu
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Kowloon, Hong Kong, China

art ic l e i nf o a b s t r a c t

Article history: To facilitate an effective assessment of stress-related bridge performance and safety, stress-level
Received 30 August 2013 buffeting analysis of a long-span cable-supported bridge is necessary. A stress-level buffeting analysis
Received in revised form requires knowledge of characteristics of distributed aerodynamic forces on the surfaces of a bridge deck.
7 May 2014
This paper first presents a formulation for distributed aerodynamic forces on the surfaces of a bridge
Accepted 7 May 2014
Available online 3 June 2014
deck based on the quasi-steady theory. The characteristics of distributed aerodynamic forces, such as
pressure coefficients, pressure admittances and span-wise pressure coherences, are introduced in the
Keywords: formulation. In consideration of different characteristics of incident and signature turbulences, the
Distributed aerodynamic forces empirical mode decomposition is then adopted to separate their effects on the distributed aerodynamic
Twin-box bridge deck
forces. Wind tunnel pressure tests of a sectional motionless bridge deck model were conducted to
Pressure admittance
identify the characteristics of distributed aerodynamic forces on the surfaces of the Stonecutters cable-
Span-wise pressure coherence
Signature turbulence stayed bridge with a twin-box bridge deck as a case study. The results indicate that the separation of
Wind tunnel pressure tests incident and signature turbulence effects is necessary and that the proposed framework is feasible
although more works need to be done towards a complete stress-level buffeting analysis.
& 2014 Elsevier Ltd. All rights reserved.

1. Introduction time domain (e.g. Chen et al., 2000). The non-uniform spatial
distributions of aerodynamic forces and self-excited forces around
The increasing span length of modern cable-supported bridges the cross section of bridge deck are seldom considered. The disregard
leads to a significant increase in buffeting response. Excessive of the cross-sectional distribution of aerodynamic forces and self-
buffeting responses can cause fatigue damage in components and excited forces may have a considerable impact on the accuracy of
connections of long-span bridges (Gu et al., 1999; Li et al., 2002; Xu computed buffeting-induced stress responses, and will in turn affect
et al., 2009). In this regard, an accurate prediction of buffeting- the comparison with the measured stresses from a SHM system.
induced stresses should be an essential objective of buffeting To obtain the distributions of aerodynamic forces and self-
analysis (Liu et al., 2009). In the meantime, structure health excited forces on a bridge deck, wind tunnel pressure tests of the
monitoring (SHM) systems have been installed in a number of bridge deck are needed. Many studies have been performed with
long-span cable-supported bridges to assess bridge performance the pressure tests of sectional motionless bridge deck models (e.g.
and safety (Xu and Xia, 2012). The number of sensors in a SHM Larose et al., 1998). Most of these studies focused on the aero-
system is always limited so that not all the structural components dynamic admittance functions and/or the span-wise correlations
can be directly monitored. Therefore, to facilitate an effective of integrated aerodynamic forces rather than distributed aerody-
assessment of stress-related bridge performance and safety, namic forces. More recently, the pressure tests of sectional bridge
stress-level buffeting analysis is required so that responses in all deck models with forced motion have been carried out to explore
important structural components can be directly computed and the nonlinear mechanism of wind-deck interaction and the
some of them can be compared with measured ones for verification. phenomenon of self-excited forces and wind-induced unsteady
The traditional buffeting analysis of a long-span bridge is based on forces (Diana et al., 2010; Argentini et al., 2012).
integrated forces (e.g. drag, lift and pitching moment) instead of As a first step, this study focuses on the characteristics of
distributed forces on the bridge deck in either the frequency domain distributed aerodynamic forces on the surfaces of a bridge deck
(e.g. Davenport, 1962; Scanlan and Gade, 1977; Xu et al., 2000) or the without considering distributed self-excited forces and vortex-
shedding associated lock-in. The formulation for distributed aero-
dynamic forces on the surfaces of a bridge deck is first derived
n
Corresponding author. Tel.: þ 852 27666008. based on the quasi-steady theory. The characteristics of distributed
E-mail addresses: stevenzq@gmail.com (Q. Zhu), ceylxu@polyu.edu.hk (Y.L. Xu). aerodynamic forces, such as distributed force coefficients, pressure

http://dx.doi.org/10.1016/j.jweia.2014.05.003
0167-6105/& 2014 Elsevier Ltd. All rights reserved.
32 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

admittances and span-wise pressure coherences, are introduced in


the formulation. In consideration of different characteristics of
incident and signature turbulences, the empirical mode decom-
position is then adopted to separate their effects on the distributed
aerodynamic forces. Wind tunnel pressure tests of a sectional
motionless bridge deck model were conducted to identify the
characteristics of distributed aerodynamic forces on the surfaces of
the Stonecutters cable-stayed bridge with a twin-box bridge deck Fig. 1. Structural coordinates, wind axes and wind pressure on section outline.
as a case study.
functions of the fluctuating pressure at the ith surface point of the
bridge deck with respect to the fluctuating wind u and w,
2. Distributed aerodynamic forces on a bridge deck respectively. The pressure admittance functions are the functions
of reduced frequency and dependent on the geometrical config-
2.1. Formulation of distributed quasi-steady aerodynamic forces uration of the cross section of the bridge deck as well as their
locations on the deck surface.
Aerodynamic forces on a bridge deck result from wind pres-
sures acting on the surfaces of the bridge deck. The distribution of 2.2. Relationships between distributed and integrated aerodynamic
aerodynamic forces can thus be represented by pressure distribu- forces
tion. Based on the quasi-steady assumption, wind pressure on a
surface point of a motionless bridge deck section can be expressed In this section, the aerodynamic lift force is taken as an
as: example to show the relationships between distributed aerody-
1 h   namic pressures and traditionally-used integrated aerodynamic
P i ðtÞ ¼ ρ ðU þ uðtÞÞ2 þ wðtÞ2   C pi α0 þ Δα ð1Þ forces.
2
The aerodynamic lift force per unit length can be calculated as
where Pi(t) is the time-history of wind pressure on the ith surface
the integration of the distributed aerodynamic pressures on the
point of the bridge deck section; ρ is the density of air; U is the mean
same section as
speed of the incoming wind flow; u and w are the longitudinal and
N
vertical turbulence component, respectively; Cpi is the pressure
F L;b ðtÞ ¼ ∑ P i;b ðtÞ  cos ðβi þ α0 Þ  δi ð7Þ
coefficient defined in the structural coordinates; α0 is the mean angle i¼1
of incidence; and Δα is the additional angle of incidence caused by
where FL,b is the aerodynamic lift force; βi is the angle between
turbulence. The definitions of structural coordinates, wind axes and
pressure and vertical structural axis z (see Fig. 1); δi is the
angles of incidence can be found in Fig. 1.
characteristic length on the deck section outline for the aerody-
Because Δα is a comparatively small angle and u is much
namic pressure Pi,b; and N is the total number of wind pressure
smaller than U for a normal wind resistance design, it can be
points on the section where the pressures are measured.
assumed that
Substituting Eq. (6) into Eq. (7) yields
wðtÞ wðtÞ   
Δα  tan ðαÞ ¼  ð2Þ 1 2 N
F L;b ðtÞ ¼ ρU  ∑ 2C pi ðα0 Þ  χ Pui
uðtÞ
þ C 0pi ðα0 Þ  χ Pwi
wðtÞ
 cos ðβi þ α0 Þ  δi
U þ uðtÞ U 2 i¼1 U U
and therefore ð8Þ
wðtÞ The relationship between the integrated aerodynamic lift
C pi ðα0 þ ΔαÞ  C pi ðα0 Þ þ C 0pi ðα0 Þ  Δα  C pi ðα0 Þ þ C 0pi ðα0 Þ  ð3Þ
U coefficient CL and the pressure coefficients Cpi can be derived
based on the mean wind lift and the mean wind pressures as
where C 0pi ¼ dC pi =dα is the derivative of pressure coefficient with
respect to the angle of incidence. N
C L ðα0 Þ  B ¼ ∑ C Pi ðαÞ  cos ðβ i þ α0 Þ  δi ð9Þ
Substituting Eq. (3) into Eq. (1) and neglecting quadratic terms i¼1
of u(t) and w(t) yields
where B is the deck width as a reference length.
1 2 1 2
P i ðtÞ ¼ ρU  C pi ðα0 Þ þ ρU Let us define the distributed lift coefficient CLi and drag
2 2 coefficient CDi in wind coordinates for wind pressure at the ith
 
uðtÞ wðtÞ surface point as
 2C pi ðα0 Þ  þ C pi ðα0 Þ 
0
ð4Þ
U U
C Li ðα0 Þ ¼ C Pi ðα0 Þ  cos ðβi þ α0 Þ  δi ð10Þ
The first part of the right side of Eq. (4) is the static pressure,
which has been well studied. Thus, aerodynamic pressure can be C Di ðα0 Þ ¼ C Pi ðα0 Þ  sin ðβ i þ α0 Þ  δi ð11Þ
expressed as Eq. (8) can then be rewritten as
 
1 2 uðtÞ wðtÞ uðtÞ  0
P i:b ðtÞ ¼ ρU  2C pi ðα0 Þ  þ C 0pi ðα0 Þ  1 2 N wðtÞ
ð5Þ F L;b ðtÞ ¼ ρU  ∑ 2C Li ðα0 Þ  χ Pui þ C Li ðα0 Þ þ C Di ðα0 Þ  χ Pwi ð12Þ
2 U U 2 i¼1 U U

Since the quasi-steady assumption does not hold for most wind The aerodynamic lift force acting on a bridge deck section is
pressures acting on a bridge deck, the aerodynamic admittance traditionally expressed as
function of wind pressure, which is similar to the aerodynamic  
1 2 uðtÞ wðtÞ
admittance functions of integrated aerodynamic forces, should be F L;b ðtÞ ¼ ρU B 2C L χ Lu þ ðC 0L þ C D Þχ Lw ð13Þ
2 U U
introduced into Eq. (5). Therefore, the aerodynamic pressure can
be expressed as where χLu and χLw are the aerodynamic admittance functions of
  the integrated aerodynamic lift force with respect to the fluctuat-
1 2 uðtÞ wðtÞ
P i:b ðtÞ ¼ ρU  2C pi ðα0 Þ  χ pui  þ C 0pi ðα0 Þ  χ pwi  ð6Þ ing wind u and w, respectively.
2 U U
The comparison Eq. (12) with Eq. (13) gives us the relationships
where χpui and χpwi are the aerodynamic pressure admittance between the integrated force admittance functions and the
Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 33

pressure admittance functions. χpi ¼ χpui ¼ χpwi, and the square of the module of the equivalent
 aerodynamic pressure admittance can then be expressed as
∑N χ  C Li ðα0 Þ ∑N χ  α α
C 0Li ð 0 Þ þ C Di ð 0 Þ
χ Lu ¼ i ¼ 1 Pui
; χ Lw ¼ i ¼ 1 Pwi

B  C L ðα0 Þ B  C 0L ðα0 Þ þ C D ðα0 Þ 2
2 SCpi ðωÞ  U
ð14Þ χ pi ðωÞ ¼ 2 2
ð16Þ
4C 2pi ðα0 Þ  Su ðωÞ=U þ C 0pi 2ðα0 Þ  Sw ðωÞ=U
Eq. (14) shows that the integrated force admittance is the average
of the corresponding pressure admittances weighed by the distributed 2
where SCpi ðωÞ ¼ 4Spi ðωÞ=ρ2 U B2 denotes the normalized auto-
aerodynamic coefficients in wind coordinates, indicating that the
spectrum of the ith pressure coefficient Cpi.
information on the non-uniform cross-sectional distribution of pres-
The aerodynamic pressure admittance functions directly identified
sure admittances is lost in this averaging process. The pressure
from the measured wind and pressure time histories based on Eq. (16)
admittance and its distribution, on the other hand, will provide more
often involves high frequency peaks due to signature turbulence,
information on wind effects on a bridge deck.
which is the turbulence produced by the structure itself in the flow
even if the incoming flow is perfectly smooth. Such pressure admit-
2.3. Identification of pressure admittance tance functions, which are difficult to be fitted with rational functions,
cause difficulty in buffeting analysis (Zhu et al., 2009). In this study, a
The auto-spectrum of aerodynamic pressure can be obtained decomposition method is proposed in Section 4.4 to decompose
based on Eq. (6) through Fourier transformation and by ignoring measured pressure time-histories into incident and signature turbu-
the cross-spectrum between the turbulence components u and w lence induced components. Because signature turbulence effects (high
(Larose, 1999). frequency) are largely separated from incident turbulence effects (low
  frequency), the admittance functions can be identified for them
1 2 2 2
SPi ðωÞ ¼ ρ2 U  4C 2pi  χ pui  Su ðωÞ þC 0pi 2  χ pwi  Sw ðωÞ separately assuming these two parts are uncorrelated.
4
Suppose a pressure time-history is decomposed into incident
ð15Þ
and signature turbulence induced components as
where SPi(ω) is the auto-spectrum of aerodynamic pressure; Su(ω)
and Sw(ω) are the auto-spectra of turbulence components u and w, P i:b ðtÞ ¼ P I;i ðtÞ þ P S;i ðtÞ ð17Þ
respectively; and | | is the operation of module. Due to the practical
difficulty in the identification of aerodynamic admittances, it is where PI,i and PS,i are the incident turbulence induced component
assumed that the equivalent aerodynamic pressure admittance and signature turbulence induced component, respectively.

Fig. 2. Typical cross section of bridge deck in main span.

Fig. 3. Positions of the pressure strips and taps. (a) Position of pressure-tapped strips (Unit: mm) and (b) Position of pressure taps.
34 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

The admittance function of each component can be identified Both Eqs. (23) and (24) can be fitted with exponential decay
from its corresponding time-history as functions.
2
2 SCpI;i ðωÞ  U
χ pI;i ðωÞ ¼ 2 2
ð18Þ
4C 2pi ðα0 Þ  Su ðωÞ=U þ C 0pi 2ðα0 Þ  Sw ðωÞ=U 3. Sectional bridge deck model and wind tunnel tests

2 SCpS;i ðωÞ  U
2 3.1. Stonecutters Bridge and its pressure-tapped sectional deck
χ PS;i ðωÞ ¼ 2 2
ð19Þ model
4C 2pi ðα0 Þ  Su ðωÞ=U þ C 0pi 2ðα0 Þ  Sw ðωÞ=U
The Stonecutters Bridge is a two cable-plane cable-stayed
where χpI,i and χpS,i are the admittance functions of incident and
2 bridge with a twin-box deck carrying dual 3-lane highway traffic.
signature components, respectively; SCpI;i ðωÞ ¼ 4SpI;i ðωÞ=ρ2 U B2
2 2 2 The bridge is currently the world's third longest cable-stayed
and SCpS;i ðωÞ ¼ 4SpS;i ðωÞ=ρ U B denote the non-dimensional
auto-spectrum of PI,i and PS,i, respectively.

2.4. Span-wise coherence function of pressure

Span-wise coherence function of aerodynamic pressures can


commonly be defined and calculated as

Scr
α1 α2 ðK Δ Þ
1=2
Cohα1 α2 ðK Δ Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð20Þ
Sα1 ðK Δ ÞSα2 ðK Δ Þ

where Coh1/2(KΔ) is the root coherence function; α1 and α2 can be


turbulence components, aerodynamic forces or pressures; KΔ ¼ fΔ/U is
the reduced frequency with respect to the span-wise distance Δ
between α1 and α2; Scr(KΔ) is the cross-spectrum between α1 and α2;
Sα1 and Sα2 are the auto-spectrum of α1 and α2, respectively.
The root coherence function is conventionally fitted with an
exponential decay function (Davenport, 1961). The root coherence
function can be fitted as

Cohα1 α2 ðK Δ Þ ¼ ¼ A  e  CK Δ
1=2
ð21Þ

where A is the root coherence peak value at zero reduced


frequency, which is adopted to cater for the situation where the
root coherence is not equal to 1 when KΔ is 0 (Hui, 2006); and C is
the decay factor.
After the decomposition of a pressure time-history into the two
parts, the span-wise coherence of incident turbulence component can
be directly calculated by Eq. (20) and fitted by Eq. (21). Unlike the
Fig. 4. Wind tunnel simulation of turbulent wind fields (unit: mm). (a) Open ocean
incident turbulence component, the peak value of span-wise coher-
fetch (T.I.¼ 6%) and (b) Over-land fetch (T.I.¼ 14%).
ence of signature turbulence component appears at the predominant
signature frequency, which is not zero, of the deck section. As a result,
the shape of span-wise coherence function of signature component
varies with Δ (see the discussion in Section 4.4). To avoid this problem,
the span-wise root coherence of signature turbulence induced pres-
sure is presented with a new reduced frequency KΔS defined as
f f s Δ
K ΔS ¼ ð22Þ
U
where fs is the predominant frequency of the signature turbulence
induced pressure.
The new reduced frequency KΔS is then used instead of KΔ so
that the peak value of the root coherence function occurs at zero
reduced frequency and the root coherence value decays exponen-
tially with the increase of the new reduced frequency. Fig. 5. Sectional deck model in the wind tunnel.
In summary, the span-wise coherence functions of incident and
signature pressure components can be calculated as

Scr
P I;1 P I;2 ðK Δ Þ
1=2
CohP I;1 PI;2 ðK Δ Þ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð23Þ Table 1
SP I;1 P I;1 ðK Δ ÞSP I;2 PI;2 ðK Δ Þ Measured turbulence intensities and integral length scales.

Flow field Iu (%) Iw (%) Lux (m) Lwx (m)


Scr
P S;1 P S;2 ðK ΔS Þ
1=2
CohP S;1 PS;2 ðK ΔS Þ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð24Þ T.I¼ 6% 7 6 0.325 0.175
SP S;1 P S;1 ðK ΔS ÞSP S;2 P S;2 ðK ΔS Þ T.I¼ 14% 17 14 0.375 0.175
Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 35

Fig. 6. Auto-spectra of turbulent components u and w. (a) Spectrum of u (T.I.¼6%), (b) Spectrum of w (T.I.¼6%), (c) Spectrum of u (T.I.¼14%) and (d) Spectrum of w (T.
I.¼ 14%).

Table 2
Mean aerodynamic force coefficients and their derivatives.

α CD CL CM C0 D C0 L C0 M

 31 0.0458  0.2704  0.0044  0.1386 4.0027 0.6524


01 0.0426  0.1003 0.0295  0.0490 2.8741 0.8643
þ 31 0.0484 0.1038 0.0703 0.2553 3.4827 1.2044

model was installed with 7 pressure-tapped acrylic strips to


measure time-histories of surface pressures. The strips were
spaced at 1/8, 1/4, 1/2, 1, 2 and 4 times the chord length of a
single box, which is 0.244 m, to investigate the span-wise correla-
tion of aerodynamic pressures. Each strip was fitted with 64
pressure taps distributed around the twin-box deck. Locations of
the pressure taps are shown in Fig. 3b with the coordinate systems
of aerodynamic forces. The pressure taps on the windward box are
denoted with numbers 101 to 132 and the pressure taps on the
leeward box are denoted with 201 to 232.

3.2. Simulation of turbulent wind flow field

Turbulent wind flow fields were simulated with horizontal and


vertical fences at 8.5 m upstream of the model. Two turbulent flow
fields with different turbulence intensities (T.I.) were simulated as
shown in Fig. 4: one represents an open ocean fetch and the other
an over-land fetch. Details of the simulated turbulent fields are
summarized in Section 4.1.
Fig. 7. Span-wise root coherence of turbulence (T.I.¼6%).
3.3. Wind tunnel pressure tests
bridge with a main span of 1018 m. The typical cross section of the
bridge deck in the main span is shown in Fig. 2. Fig. 5 shows the pressure-tapped sectional model mounted in
All pressure measurements were conducted on a motionless the wind tunnel and the locations of pitot tube and cobra probe
sectional deck model which represents the typical deck geometry which were used to calibrate mean wind speed and record
of the Stonecutters Bridge. The model was 3 m in length with a transient 3D turbulences, respectively. The model was further
length scale of 1:80 (see Fig. 3a). The 1 m central portion of the stabilized with guy wires to avoid vibration. Pressure tests were
36 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

conducted in two types of turbulent wind flow fields at wind derivatives with respect to angle of incidence. The time-histories
speed of 15 m/s with 31, 01 and þ31 angles of incidence for the of aerodynamic pressures from totally 448 pressure-taps were
time-histories of aerodynamic pressures induced by turbulent acquired and collated. The time-histories of aerodynamic forces
wind. Additional tests were conducted in smooth flow field at could also be obtained by integration of the pressure. The time
wind speed of 15 m/s with 751, 741, 731, 721, 711 and 01 histories of 3-D wind speeds were recorded by cobra probes
angles of incidence for mean aerodynamic coefficients and their simultaneously with the pressure measurements.

4. Results and discussions

4.1. Characteristics of simulated turbulent field

Two turbulent flow fields with different turbulence intensities


(T.I.) were simulated to represent the open ocean fetch and the
over-land fetch, respectively. Measured turbulence intensity is
defined as
pffiffiffiffiffiffi pffiffiffiffiffiffiffi
su u2 sw w2
Iu ¼ ¼ ; Iw ¼ ¼ ð25Þ
U U U U
where su and sw are the standard deviation of u and w,
respectively.
The vertical fluctuation is usually more dominant in buffeting
analysis so the two turbulence fields are denoted by the vertical
turbulence intensity as T.I¼ 6% and T.I¼ 14%. The integral length

Fig. 8. Aerodynamic admittance and PSD of integrated forces in different flow Fig. 9. Aerodynamic admittance of integrated forces with different angles of
fields with 01angle of incidence. incidence (T.I¼ 6%).
Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 37

scales listed in Table 1 are estimated by using Eq. (26) with the taken into consideration in the subsequent analysis to simplify the
assumption that the vortex patterns do not change as wind problem.
sweeps them leeward. As shown in Fig. 8, the signature turbulence has a significant
Rt RT influence on the aerodynamic forces at high reduced frequency,
Lux ¼ U 0 cux ðτÞdτ; cux ðτÞ ¼ T1 0 =½uðtÞ  uðt þ τÞdt s2u and such influence is larger with lower turbulence intensity. It can
Z t Z be seen that the signature turbulence effect can be separated from
1 T
Lwx ¼ U cwx ðτÞdτ; cwx ðτÞ ¼ =½wðtÞ  wðt þ τÞdt s2w ð26Þ the incident turbulence effect in the frequency domain as the
0 T 0
latter mainly dominates the low reduced frequency range. This
where Lux and Lwx are the integral length scales of u and w in the provides a possibility of analyzing the two types of turbulence
longitudinal direction; cux and cwx are the auto-covariance of u and w. effects separately. Besides, the fact that signature turbulence
Fig. 6 shows the auto-spectra of turbulent components u and w. mainly affects high reduced frequency range suggests that its
For the convenience of comparison between the force and pres- influence is more critical at low wind speed.
sure spectra, the turbulence spectra are presented with respect to Fig. 10 shows the span-wise root coherence of integrated
the reduced frequency K ¼ f B=U. The auto-spectra indicate that aerodynamic forces. As mentioned above, the signature turbulence
the energy of turbulence is mainly concentrates in the reduced effect on the deck section has a fixed predominant reduced
frequency range from 0.0 to 0.6. frequency K ¼ f B=U around 4, which should also be the reduced
Fig. 7 shows the measured and fitted span-wise root coherence frequency of the signature turbulence coherence peak regardless
of turbulence components u and w with 6% turbulence intensity. of the span-wise distance Δ. As a result, when multi-groups of
The root coherence presented in Fig. 7 combines 5 groups of measured coherence are presented with respect to K Δ ¼ f Δ=U in
coherence data which are obtained for the span-wise distance Δ of the same figure, multiple signature turbulence peaks appear. The
1/8, 1/4, 1/2, 1, and 2 times the chord length of a single box, multi-peak coherence is difficult to be fitted with either the
respectively (so as other root coherence figures presented in this
paper). The root coherence of wind turbulence can be fitted quite
well with exponential decay curves.

4.2. Characteristics of integrated aerodynamic forces

Mean aerodynamic force coefficients and their derivatives with


respect to angle of incidence are calculated for integrated aero-
dynamic forces on the sectional model, and the results are listed in
Table 2. The admittances of the integrated aerodynamic forces
(drag, lift and pitching moment) are depicted in Fig. 8a–c and
Fig. 9. Fig. 8a–c shows the admittances of the integrated aero-
dynamic forces with angle of incidence of 01 for the two turbu-
lence fields. Fig. 8d compares the power spectrum density (PSD)
functions of the lift force in the turbulent flow with those in the
smooth flow at different wind speeds. Fig. 9 shows the aerody-
namic admittance of the integrated forces with  31, 01 and þ31
angle of incidence for the low turbulent field of 6% turbulence
intensity.
As shown in Figs. 8 and 9, all the force spectra and aerodynamic
admittances have two notable peaks. One peak appears at the reduced
frequency around 0.3, which is consistent with the spectral peak of
vertical turbulence component w. The other peak appears at a reduced
frequency about 3.95 for both 15 m/s and 12 m/s winds in the smooth
flow. The proportionality of frequency and wind speed indicates that
this peak results from the predominant signature turbulence fre-
quency of the deck section. The Strouhal number St as defined in
Eq. (27) of the deck can be estimated from the force spectra as
around 0.26, which is generally consistent with the study carried out
by Kwok et al. (2012).
fH
St ¼ ð27Þ
U
where H is the height of the section.
As shown in Fig. 8d, the signature turbulence peak of turbulent
flow is slightly smaller than that of smooth flow. This may
attribute to the interference by the incoming turbulence.
It should be noted that excitation due to signature turbulence
includes all wake-induced excitations and not just those asso-
ciated with critical velocities (vortex shedding) (Singh, 1997).
It should also be noted that the signature turbulence of such a
complicated twin-box section yield more than one predominant
signature frequencies. As shown in Figs. 8 and 9, smaller peaks
appear within the reduced frequency range from 1.6 to 2.4. In this
study, only the predominant signature turbulence frequency is Fig. 10. Spanwise root coherence of integrated aerodynamic forces (T.I. ¼ 6%).
38 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

conventionally used exponential decay function or other simple The derivatives were calculated by central difference method. It
rational functions. This phenomenon further justifies the need to can be seen from the comparison of Fig. 11 with Fig. 12 that the
separately the analysis of incident and signature turbulence value of C0 p is much larger than the value of 2Cp, which is the
induced pressures. The decomposition of these two types of effects quasi-steady multiplier on the turbulence component u. This fact
on aerodynamic pressures is introduced in Section 4.4. indicates that the vertical turbulence w has a much larger impact
on the aerodynamic forces than the longitudinal turbulence u. As a
result, Fig. 12 largely represents the quasi-steady cross-sectional
4.3. Distribution of mean and r.m.s value of pressure distribution of the aerodynamic forces. Fig. 12 also indicates that
from a quasi-steady point of view, the aerodynamic forces on the
Fig. 11 shows the cross-sectional distribution of mean pressure windward box are larger than those on the leeward box.
coefficients. Most of the pressure taps yield negative pressure Fig. 13 depicts the root mean square (r.m.s) value of instantaneous
coefficients. The largest negative pressure coefficients occur on pressure coefficients Cp and shows the general cross-sectional dis-
windward corners of both windward and leeward boxes at  31, 01 tribution of fluctuating pressure. Although it can be concluded that the
and þ 31 angle of incidence. This indicates the flow separates at windward box bears a larger fluctuating forces, the difference between
these locations. Positive pressures occur at pressure taps 101, 102 the fluctuating pressure acting on windward and leeward boxes is not
and 217, which contribute the largest part to mean drag force. The as large as suggested by Fig. 12. This is mainly because that the quasi-
cross-sectional distribution of Cp of this deck section has also been steady theory does not hold in this case where signature turbulence
studied by Kwok et al. (2012). The results are generally consistent effect is strong. The results indicate that the cross-sectional distribu-
with this study. tion of aerodynamic pressure is not uniform, and the exact distribution
Fig. 12 shows the derivatives of pressure coefficients with can hardly be estimated by quasi-steady aerodynamic coefficients. It is
respect to angle of incidence, which are the quasi-steady multi- necessary to investigate the fluid-motionless structure interaction in
plier on the turbulence component w, as indicated in Eq. (5). terms of the cross-sectional distribution of aerodynamic pressure.

Fig. 11. Distribution of mean pressure coefficients Cp.


Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 39

Fig. 12. Distribution of derivatives of mean pressure coefficients C0 p.

4.4. Decomposition of aerodynamic pressure time-series represent other minor signature turbulence components. There-
fore, the first three IMFs can be added up to represent the
Like the spectra of aerodynamic forces, the spectra of aerodynamic signature turbulence induced component of the pressure in this
pressures also have signature turbulence peaks. The span-wise coher- case. The fourth to fifteenth IMFs generally fall into the frequency
ences of aerodynamic pressures also have multiple signature turbu- range of incident turbulence (the seventh to fifteens IMFs have
lence peaks. To further investigate the cross-sectional distribution of comparatively smaller values and therefore are not depicted in
signature turbulence effects, the empirical mode decomposition Fig. 14). Therefore, they were added up to represent the incident
(EMD) method (Huang et al., 1998; Xu and Chen, 2004) is employed turbulence induced component of the pressure. Fig. 15 depicts the
to decompose each fluctuating pressure time-history into incident time-histories of the original and decomposed pressure time
turbulence and signature turbulence induced parts. First, a pressure histories. Fig. 16 shows the PSDs of the original and decomposed
time-history is decomposed by EMD into several intrinsic mode pressures. It should be noted that the original incident component
functions (IMFs). Then, the IMFs representing the low frequency as shown in Fig. 15 contains the mean pressure while in the
incident turbulence effect and high frequency signature turbulence following analyses the mean values were removed from all the
effect are added up respectively to form two time-histories: one time-histories.
mainly caused by incident wind turbulence and the other by signature
turbulence.
Figs. 14–16 give an example of the decomposition of a pressure 4.5. Distribution of aerodynamic pressure admittances
time-history. Fig. 14 shows the power spectral density (PSD)
functions of a measured pressure time-history and its first 6 IMFs After the decomposition of pressure time-histories, aerody-
after EMD. As shown in Fig. 14, the first decomposed IMF can namic pressure admittances are calculated for incident and sig-
largely represent the predominant signature turbulence compo- nature turbulence induced pressures respectively by using the
nent in the pressure. The second and third IMFs probably method presented in Section 2.3. Then, each pressure admittance
40 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

Fig. 13. Distribution of the r.m.s values of the measured instantaneous pressure coefficients (T.I. E6%).

are the fitting parameters for incident and signature admittance,


respectively. The first part of the equivalent admittance function
represents the incident admittance while the second part repre-
sents the signature admittance.
Eq. (28) is a modified version of the admittance functions
proposed by Zhu et al. (2009) to fit the aerodynamic admittance
with signature turbulence effect. Fig. 17 shows an example of
measured and fitted pressure admittances. In the identified
pressure admittance, the predominant signature turbulence com-
ponent is more apparent. The curve fitting result is satisfactory.
After fitting the identified pressure admittance of each pressure
tap, the cross-sectional distribution of ci1, which reflects the
distribution of incident admittance peak value, can be obtained.
Fig. 14. EMD results of a pressure time-history. The results are illustrated in Fig. 18. The distribution patterns of ci1
are very similar in the two turbulence field at 01 angle of
function can be fitted with a rational equation as incidence. The incident admittance peak value varies markedly
ci1 cs1 along the section outline. ci1 values larger than 1 indicate that the
χ P ðKÞ 2 ¼ þ ð28Þ incident turbulence effect is larger than that estimated with the
1 þ ci2  ðK K I Þ2 1 þ cs2  ðK  K s Þ2
quasi-steady assumption. At 01 angle of incidence, ci1 value is less
where KI and KS are the predominant reduced frequencies for than 1 on the windward box for most locations except the loca-
incident and signature admittances, respectively; ci(1–2) and cs(1–2) tions of pressure taps 115 and 132. On the leeward box, the value is
Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 41

Fig. 15. Decomposition of a pressure time-history in time domain. (a) Total


pressure, (b) Incident component and (c) Signature component.

Fig. 17. An example of measured and fitted pressure admittances. (a) Incident
admittance, (b) Signature admittance and (c) Total admittance.

aerodynamic pressure. The cross-sectional distributions of cs1/ci1 are


depicted in Fig. 19 for different turbulence fields and different angles
of incidence. The comparison of the results from different turbulence
Fig. 16. PSD of a pressure time-history and its components after decomposition. fields shows that the signature turbulence effect is significantly
larger when turbulent intensity is lower, but the cross-sectional
larger than 1 for most locations. The incident admittance signifi- distribution pattern of signature turbulence effect is generally alike
cantly deviates from the quasi-steady assumption on the leeward for different turbulent intensities at 01 angle of incidence. Signature
edge of the leeward box. The peak admittance values reach up to turbulence mainly affects wind pressures on the leeward box. At 01
more than 10 at the locations of pressure taps 203 and 204. The angle of incidence, for certain locations, the frequency-domain peak
cross-sectional distribution patterns of ci1 at  31 and þ31 angles value of signature turbulence induced pressure may reach up to
of incidence are different from that at 01 angle of incidence. This is about 12 times the value of incident turbulence induced pressure at a
particular true for the leeward box. At  31 angle of incidence, turbulent intensity of 6% or 3 times the value of incident turbulence
large values locate at pressure taps 203 and 204. At þ31 angle of induced pressure at a turbulent intensity of 14%. The ratio can reach
incidence, the largest ci1 value is only around 2.6 locating at the up to around 60 at  31 and þ31 angles of incidence. These results
windward edge of the deck plate. indicate that signature turbulence effect is somehow negligible for
The ratio cs1/ci1 between the frequency-domain peak values of most parts of the windward box but important for the leeward box.
incident and signature turbulence induced pressures can give Besides, the cross-sectional distribution patterns of signature turbu-
information about the proportion of signature turbulence effect on lence effect vary significantly with different angles of incidence.
42 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

Fig. 18. Cross-sectional distribution of the incident admittance peak value. (a) T.I. ¼ 14%, α0 ¼01, (b) T.I.¼6%, α0 ¼ 01, (c) T.I.¼ 6%, α0 ¼-31 and (d) T.I.¼ 6%, α0 ¼ þ31.

4.6. Span-wise pressure coherence To avoid multiple peaks in coherence function, the span-wise
root coherence of signature turbulence induced pressure is pre-
After the decomposition of pressure time-histories, the root sented with a reduced frequency KΔS as defined in Eq. (22).
coherence of incident and signature turbulence induced pressures An example of measured and fitted root coherence of signature
can be calculated using Eqs. (23) and (24) respectively and fitted by turbulence induced pressure is given in Fig. 20b. Fig. 22 illustrates
Eq. (21). Fig. 20 shows an example of measured and fitted root the cross-sectional distribution of fitted coherence coefficients
coherence of 7 aerodynamic pressures that lie on the same line of signature turbulence induced pressure. Because the signature
parallel to the deck axis. A total of 64 root coherence functions turbulence effects on the windward box and some parts of the
similar to Fig. 20 are obtained for either signature turbulence or leeward box are marginal, the coherence coefficients are hardly
incident turbulence. The cross-sectional distribution of fitted coher- identifiable on these locations. The coherence coefficients for the
ence coefficients of incident turbulence induced pressure at 01 rest locations are presented in Fig. 22. It can be seen that
angle of incidence with 6% turbulence intensity is shown in Fig. 21. signature turbulence induced pressure has a similar correlation
The root coherence peak value at zero reduced frequency is distribution pattern in both turbulent and smooth wind flow.
approximately 1 for most of the locations. The peak value decreases The correlation is only slightly stronger in the smooth flow.
to about 0.6 at the location of pressure tap 121 and the location Signature turbulence induced pressure has much weaker correla-
around pressure tap 209. The decay factor increases stream-wisely tion than incident turbulence induced pressure. Its correlation is
from about 6 to about 11 on the windward box which shows a negligible on most of the locations except the windward
stream-wise decline trend of pressure correlation. The pressure edge of the leeward box and the leeward edge of the windward
correlation is weaker in the leeward box and the distribution box. The result suggests that the flow separation on the
pattern is not as obvious as the windward one. The decay factors windward box is the main cause of the predominant signature
on the leeward box fluctuate from around 9 to around 12. turbulence.
Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 43

Fig. 19. Ratio between frequency-domain peak values of incident and signature turbulence induced pressures. (a) T.I. ¼14%, α0 ¼ 01, (b) T.I.¼ 6%, α0 ¼ 01, (c) T.I.¼ 6%, α0 ¼-31
and (d) T.I.¼ 6%, α0 ¼ þ 31.

5. Concluding remarks tests were conducted on the motionless pressure-tapped sectional


deck model of the Stonecutters Bridge with a twin-box deck.
The formulation for distributed aerodynamic forces (aerody- By separating the signature turbulence induced pressure from the
namic pressure) on the surfaces of a bridge deck is presented in measured pressure time-histories, the cross-sectional distribution
this paper based on the quasi-steady theory. Wind tunnel pressure of signature turbulence effects was investigated. The results show
44 Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45

Fig. 20. An example of measured and fitted root coherences of aerodynamic pressure. (a) Incident component and (b) Signature component.

Fig. 21. Span-wise coherence coefficients of incident turbulence induced pressures (T.I.¼ 6%). (a) Peak value A and (b) Decay factor C.

Fig. 22. Span-wise coherence coefficients of signature turbulence induced pressures. (a) Peak value A (T.I.¼6%) and (b) Decay factor C (T.I.¼6%), (c) Peak value A (smooth
flow) and (d) Decay factor C (smooth flow).
Q. Zhu, Y.L. Xu / J. Wind Eng. Ind. Aerodyn. 131 (2014) 31–45 45

that signature turbulence mainly affects the leeward box. For References
certain locations, the signature turbulence induced pressure may
be significantly larger than that of incident turbulence induced Argentini, T., Rocchi, D., Muggiasca, S., Zasso, A., 2012. Cross-sectional distributions
pressure. In view of the results from this study, signature turbu- versus integrated coefficients of flutter derivatives and aerodynamic admit-
tances identified with surface pressure measurement. J. Wind Eng. Ind.
lence induced forces are important for certain parts of the deck
Aerodyn. 104–106, 152–158.
section, where a considerable signature turbulence effect on local Chen, X., Matsumoto, M., Kareem, A., 2000. Time domain flutter and buffeting
aerodynamic responses exists. response analysis of bridges. J. Eng. Mech. 126 (1), 7–16.
The span-wise correlation of aerodynamic pressure on the Davenport, A.G., 1961. The spectrum of horizontal gustiness near the ground in
high winds. Q. J. R. Meteorolog. Soc. 87 (372), 194–211.
sectional deck model was also studied. For the incident turbulence Davenport, A.G., 1962. Buffeting of a suspension bridge by storm winds. J. Struct.
induced pressure, the span-wise correlation weakens stream- Div., ASCE 88 (3), 233–268.
wisely on the windward box, and the span-wise correlation on Diana, G., Rocchi, D., Argentini, T., Muggiasca, S., 2010. Aerodynamic instability of a
the leeward box is generally weaker than that on the windward bridge deck section model: linear and nonlinear approach to force modeling. J.
Wind Eng. Ind. Aerodyn. 98 (6), 363–374.
box. For the signature turbulence induced pressure, the span-wise Gu, M., Xu, Y.L., Chen, L.Z., Xiang, H.F., 1999. Fatigue life estimation of steel girder of
correlation is negligible for most parts of the deck except the Yangpu cable-stayed bridge due to buffeting. J. Wind Eng. Ind. Aerodyn. 80 (3),
windward edge of the leeward box and the leeward edge of the 383–400.
windward box. Huang, N.E., Shen, Z., Long, S.R., Wu, M.C., Shih, H.H., Zheng, Q., Liu, H.H., 1998. The
empirical mode decomposition and the Hilbert spectrum for nonlinear and
The cross-sectional and span-wise distributions of aerody- non-stationary time series analysis. Proc. R. Soc. London, Ser. A 454 (1971),
namic pressures provide more detailed information and deeper 903–995.
insight into the fluid-motionless structure interaction on the twin- Hui, M.C.H. (2006). Turbulent Wind Action on Long Span Bridges with Separated
box bridge deck. With the method proposed by this study, the Twin-girder Decks. Doctoral Dissertation. Tongji University, China.
Kwok, K.C.S., Qin, X.R., Fok, C.H., Hitchcock, P.A., 2012. Wind-induced pressures
incident and signature turbulence induced pressures can be around a sectional twin-deck bridge model: effects of gap-width on the
separated from the measured pressure time-histories. Their admit- aerodynamic forces and vortex shedding mechanisms. J. Wind Eng. Ind.
tances and span-wise coherences can be fitted with rational Aerodyn. 110, 50–61.
Larose, G.L., Tanaka, H., Gimsing, N.J., Dyrbye, C., 1998. Direct measurements of
equations separately. As a result, the distributed aerodynamic
buffeting wind forces on bridge decks. J. Wind Eng. Ind. Aerodyn. 74, 809–818.
forces can be represented by rational equations in the frequency Larose, G.L., 1999. Experimental determination of the aerodynamic admittance of a
domain for buffeting analysis. At this point, the authors would like bridge deck segment. J. Fluids Struct. 13, 1029–1040.
to address the comment from one of the anonymous reviewers Li, Z.X., Chan, T.H., Ko, J.M., 2002. Evaluation of typhoon induced fatigue damage for
Tsing Ma Bridge. Eng. Struct. 24 (8), 1035–1047.
that the signature turbulence here is clearly vortex shedding
Liu, T.T., Xu, Y.L., Zhang, W.S., Wong, K.Y., Zhou, H.J., Chan, K.W.Y., 2009. Buffeting-
which cannot be considered in buffeting analysis. Whether the induced stresses in a long suspension bridge: structural health monitoring
signature turbulence defined and described here should be taken oriented stress analysis. Wind Struct. 12 (6), 479–504.
into account in the buffeting analysis or not can be subject to Scanlan, R.H., Gade, R.H., 1977. Motion of suspension bridge spans under gusty
wind. J. Struct. Div., ASCE 103, 1867–1883.
further discussion. The method proposed in this study to separate Singh, L. (1997). Experimental Determination of Aeroelastic and Aerodynamic Para-
incident and signature turbulence effects is necessary and novel in meters of Long-span Bridges. Doctoral Dissertation. Johns Hopkins University,
the stress-level buffeting analysis. Baltimore.
Xu, Y.L., Chen, J., 2004. Characterizing non-stationary wind speed using empirical
mode decomposition. J. Struct. Eng., ASCE 130 (6), 912–920.
Xu, Y.L., Liu, T.T., Zhang, W.S., 2009. Buffeting-induced fatigue damage assessment
Acknowledgements
of a long suspension bridge. Int. J. Fatigue 31 (3), 575–586.
Xu, Y.L., Sun, D.K., Ko, J.M., Lin, J.H., 2000. Fully coupled buffeting analysis of Tsing
The authors wish to acknowledge the financial supports from The Ma suspension bridge. J. Wind Eng. Ind. Aerodyn. 85 (1), 97–117.
Xu, Y.L., Xia, Y., 2012. Structural Health Monitoring of Long-span Suspension
Hong Kong Polytechnic University through a Ph.D. studentship for the
Bridges, first ed. Spon Press, London.
first author and the financial support from the Research Grants Zhu L.D., Zhao C.L., Wen S.B., 2009. Signature turbulence effect on buffeting
Council of Hong Kong for the second author (PloyU5304/11E). The responses of a long-span bridge with a centrally-slotted box deck. In:Proceed-
support from the CLP Power Wind/Wave Tunnel Facility at the Hong ings of the International Symposium on Computational Structural Engineering.
Dordrecht, Heidelberg, London, New York: Springer, pp. 399–409.
Kong University of Science and Technology in the wind tunnel tests is
particularly appreciated. Any opinions and concluding remarks pre-
sented in this paper are entirely those of the authors.

You might also like