Effects of errors in flutter derivatives on the wind ...811/fulltext.pdfEFFECTS OF ERRORS IN FLUTTER...

187
EFFECTS OF ERRORS IN FLUTTER DERIVATIVES ON THE WIND- INDUCED RESPONSE OF CABLE-SUPPORTED BRIDGES A Dissertation Presented by Dong-Woo Seo to The Department of Civil and Environmental Engineering in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Civil Engineering in the field of Structural Engineering Northeastern University Boston, Massachusetts February 2013

Transcript of Effects of errors in flutter derivatives on the wind ...811/fulltext.pdfEFFECTS OF ERRORS IN FLUTTER...

EFFECTS OF ERRORS IN FLUTTER DERIVATIVES ON THE WIND-INDUCED RESPONSE OF CABLE-SUPPORTED BRIDGES

A Dissertation Presented

by

Dong-Woo Seo

to

The Department of Civil and Environmental Engineering

in partial fulfillment of the requirements

for the degree of

Doctor of Philosophy

in

Civil Engineering

in the field of

Structural Engineering

Northeastern University

Boston, Massachusetts

February 2013

ii

Abstract

This dissertation discusses the development and implementation of a methodology for the

buffeting response of cable-supported bridges, including uncertainty in the aeroelastic input

(i.e., flutter derivatives, FDs). Flutter derivatives are the most important part of the loading

and are estimated in a wind tunnel experiment.

A second order polynomial model (“model curve”) for the flutter derivatives is

proposed. The coefficients of this polynomial are random variables, whose probability

distribution is conditional on the reduced wind speed. For computational reasons in

subsequent analysis, however, this dependency is neglected and the probability of these

random variables is treated as independent of the reduced wind speed. For analysis purposes

the first- and second-order statistics are estimated from experiments, treating all the wind

speed data as part of the same population. Wind tunnel experiments are conducted at

Northeastern University and a section model of a truss-type bridge deck is used.

The simplified polynomial model for the FDs, including the second order description

of its variability, is employed in the derivation of the probability of the onset of flutter using

Monte-Carlo (MC) simulations.

The simplified stochastic “model curves” for FDs are used to estimate the buffeting

bridge response. In the standard approach the result of the buffeting analysis is the value of

the RMS dynamic response at a given wind speed. In the proposed probabilistic setting one

iii

estimates the probability that a given threshold for the variance of the response is exceeded.

This probability is later used, together with information on the probability of the wind

velocity at a given site, to predict the expected value of the loss function due to the buffeting

response of a 1200-meter suspension bridge (a function proportional to the cost associated

with interventions needed to ensure safety).

iv

Acknowledgments

I would like to most sincerely thank my advisor Dr. Luca Caracoglia for his constant

guidance, teaching, and support throughout this rather complicated and intense research. He

has motivated and encouraged me towards improvement and excellence in research. I have

been grateful to work with him and to be part of his research team.

I would also like to thank my PhD committee members, Professor George G. Adams,

Professor Dionisio P. Bernal and Professor Mehrdad Sasani for their overall encouragement

through the PhD studies and for valuable comments and recommendations.

The PhD studies, described by this research, were supported in part by the National

Science Foundation of the United States (NSF), Award No. 0600575 from 2008 to 2010.

Any opinions, findings, conclusions and recommendations are those of the writer and do not

necessarily reflect the views of the NSF. The Department of Civil and Environmental

Engineering is also acknowledged for providing support from 2010 until the completion of

the program in the form of teaching assistantship.

Finally, I am deeply grateful to my parents and younger brother, whose love and

support has always been a tremendous source of strength and encouragement for me. They

never lost faith in me and are always willing to provide a helping hand. Their love and

invaluable support gave me the motivation to accomplish many goals in my life.

v

Table of Contents

Abstract ii 

Acknowledgments ................................................................................................................. iv 

Table of Contents ................................................................................................................... v 

List of Tables ...................................................................................................................... viii 

List of Figures ....................................................................................................................... ix 

Nomenclature ..................................................................................................................... xiii 

Chapter 1 1 

Introduction ........................................................................................................................ 1 

1.1 Motivation .................................................................................................................... 4 

1.2 Outline .......................................................................................................................... 6 

Chapter 2 9 

Wind-Induced Response of Long-Span Bridges: Review .................................................. 9 

2.1 General Formulation .................................................................................................... 9 

2.2 Background on Flutter ............................................................................................... 18 

2.3 Background on Buffeting ........................................................................................... 20 

2.4 Effect of Wind Directionality: Skew Wind Theory .................................................... 27 

Chapter 3 31 

A Second-order Polynomial Model for Flutter Derivatives ............................................. 31 

3.1 Description of the Polynomial Model and Discussion on its Physical Interpretation 32 

3.1.1 Description of the Polynomial Model ................................................................. 32 

3.1.2 Discussion on the Selection of the Polynomial Model, based on Physical Behavior of Flutter Derivatives ................................................................................................... 33 

3.2 Description of the Wind Tunnel, used for Experimental Verification of the Polynomial Model ............................................................................................................................... 35 

3.3 Description of the Experimental setups, used for Verification .................................. 36 

3.4 Description of the Aeroelastic Section-Model, used for Verification ........................ 37 

vi

3.5 Description of the Tests and Experimental Identification .......................................... 38 

3.6 Reason for the Use of the Polynomial Model in the Context of Random Flutter Derivatives ....................................................................................................................... 41 

3.6.1 Estimation of Variance and Co-variance of Cj and Dj coefficients of the “Model Cures” from Experiments ............................................................................................. 43 

3.7 Summary of Experimental Results and Comparison with Literature Data (“Jain’s Data”) ............................................................................................................................... 46 

Chapter 4 58 

A Methodology for the Analysis of Long-Span Bridge Buffeting Response, accounting for Variability in Flutter Derivatives ...................................................................................... 58 

4.1 Introduction ................................................................................................................ 58 

4.2 Multi-Mode Buffeting Analysis (“Deterministic Case”) ........................................... 60 

4.2.1 Validation for Closed-Form Solution .................................................................. 61 

4.2.2 Monte-Carlo and Quasi-Monte-Carlo Methods .................................................. 62 

4.2.3 Examination of the Computational Efficiency of the MC and QMC Methods for Calculating the Double Integral in Eq. (2.24) .............................................................. 64 

4.3 Monte-Carlo-based Methodology for Buffeting Analysis Considering Uncertainty in the Flutter Derivative (“Statistical Case”)........................................................................ 68 

4.3.1 Description of the Bridge Example and RMS Threshold Levels (“Probabilistic Setting”) ....................................................................................................................... 72 

4.3.2 TEP Curves using Literature Data ...................................................................... 73 

4.3.3 TEP Curves using NEU’s Flutter Derivative Data .............................................. 77 

4.4 Effect of Wind Directionality on “Statistical Buffeting” Response: TEP Surfaces ... 78 

4.5 Exploratory Performance Analysis on a Full-Scale Structure.................................... 79 

4.6 Summary .................................................................................................................... 81 

Chapter 5 120 

Lifetime Cost Analysis due to Buffeting Response on a Long-Span Bridge, accounting for Variability in Flutter Derivatives .................................................................................... 120 

5.1 Introduction .............................................................................................................. 120 

5.2 Peak Estimation via RMS Response ........................................................................ 121 

5.3 Lifetime Expected Cost Analysis ............................................................................. 122 

5.4 Description of the Structural and Aeroelastic Model ............................................... 126 

vii

5.5 Estimation of Peak Dynamic Response during Buffeting ........................................ 127 

5.6 Monte-Carlo-based Methodology for “Statistical Buffeting” Analysis considering Uncertainty in the FD ..................................................................................................... 128 

5.6.1 Wind-Direction-Independent Analysis .............................................................. 128 

5.6.2 Wind-Direction-Dependent Analysis ................................................................ 129 

5.7 Flutter Analysis: Numerical Results ........................................................................ 130 

5.8 Lifetime Expected Intervention Cost Analysis - Numerical Results ....................... 132 

5.8.1 Estimation of the Limit-State Probabilities Pj from TEP Analysis ................... 132 

5.8.2 Expected Intervention Cost - Description of the Simulations ........................... 134 

5.8.3 Expected Intervention Cost - Numerical Results using NEU’s FD Data.......... 135 

5.9 Discussion and Remarks .......................................................................................... 135 

Chapter 6 158 

Summary and Conclusions ............................................................................................. 158 

6.1 Summary .................................................................................................................. 158 

6.2 Conclusions .............................................................................................................. 160 

6.3 Recommendations for Future Research ................................................................... 161 

6.4 Outcome of the PhD Studies: List of Deliverables .................................................. 162 

6.4.1 Journal Publications (Published/under review) ................................................. 162 

6.4.2  Other Journal Publications (not related to the main topic of this Dissertation) .. ...................................................................................................................... 163 

6.4.3 Full Papers in Conference Proceedings............................................................. 163 

6.4.4 Other Papers Published as Conference Proceedings (not related to the main topic of the Dissertation) ..................................................................................................... 163 

6.4.5 Poster Presentations .......................................................................................... 164 

References .......................................................................................................................... 165 

viii

List of Tables

Table 3.1 The static coefficients and their derivatives at α0 (Jain et al., 1998). ................... 48 

Table 4.1 Natural frequencies and mode types of Golden Gate Bridge (Jain 1996). ........... 83 

Table 4.2 Comparison of closed-form numerical solution with literature results at l/4. ...... 83 

Table 4.3 Bias and relative errors in the MC case: (a) for heave σhh; (b) for torsion σαα. .... 84 

Table 4.4 Bias and relative errors in the QMC case: (a) for heave σhh; (b) for torsion σαα. . 85 

Table 4.5 Threshold values for σhh and σαα, employed in the TEP analysis with flutter derivatives from the literature. ............................................................................ 86 

Table 4.6 Study cases used for serviceability on full-scale structure. .................................. 86 

Table 5.1 Structural performance thresholds for vertical deck response (Tj). .................... 137 

Table 5.2 Probabilities of each damage state (Pj) due to buffeting response based on the structural performance thresholds (T = Tj) using NEU’s FD data..................... 137 

ix

List of Figures

Figure 1.1 Tacoma Narrows Bridge collapsed in 1940 due to wind-induced “torsional flutter” (reproduced from Simiu and Scanlan, 1996). ....................................................... 8 

Figure 2.1 A suspension bridge and a section of the deck (Schematic view of a generic finite-element model of the structure). .......................................................................... 29 

Figure 3.1 NEU-MIE wind tunnel (Brito 2008). ................................................................. 49 

Figure 3.2 Experimental setup: (a) NEU’s small-scale wind tunnel; (b) NEU’s Aeroelastic Force Balance with the truss-type bridge deck model. ....................................... 50 

Figure 3.3 The Golden Gate Bridge (Photo courtesy of Google Image). ............................ 51 

Figure 3.4 Truss-type deck section model, replicated the features of the Golden Gate Bridge at a scale 1:360; model width is B = 76 mm and the aspect ratio is B/D = 3.5:1. .................................................................................................................... 52 

Figure 3.5 Flutter derivatives of a truss-type section model with aspect ratio B/D= 3.5:1 measured at NEU: (a) H1

*; (b) H2*; (c) H3

*; (d) H4*; (e) A1

*; (f) A2*; (g) A3

*; (h) A4

*. ....................................................................................................................... 56 

Figure 3.6 Flutter derivatives of a truss-type section model(the Golden Gate Birdge) derived from (Jain et al., 1998): (a) heave Hi

* (i=1,…,4); (b) torsion Ai* (i=1,…,4). ...... 57 

Figure 4.1 Flowchart describing the MC-based methodology for buffeting analysis. ......... 87 

Figure 4.2 Two-dimensional sample points = 1,000: (a) MC with uniform distribution, (b) QMC with Halton sequence. ............................................................................... 88 

Figure 4.3 Ten simplified (sinusoidal-like) mode shapes used in the multi-mode buffeting analysis: (a) LS, 0.049 Hz; (b) VAS, 0.087Hz; (c) LAS, 0.112 Hz; (d) VS, 0.129 Hz; (e) VAS, 0.134 Hz; (f) VS, 0.164 Hz; (g) TAS, 0.192 Hz; (h) TS, 0.197 Hz; (i) VAS, 0.199 Hz; (j) VS, 0.202 Hz. (Note: L is Lateral, V is Vertical, T is Torsional, S is Symmetric, AS is Anti-symmetric). ............................................ 92 

Figure 4.4 MC-based scatter plots of RMS response for deck section at x= l/4 and for U = 22.2 m/s as a function of NMC: (a) heave σhh; (b) torsion σαα. ............................. 93 

Figure 4.5 Coefficient of variation of the RMS response at x = l/4 for U = 22.2 m/s, computed by MC algorithm: (a) heave σhh; (b) torsion σαα. ................................................. 94 

x

Figure 4.6 Tolerance intervals for vertical RMS response (σhh) of 100 MC simulations: (a) NMC = 1,000; (b) NMC = 5,000; (c) NMC = 10,000; (d) NMC = 50,000; (e) NMC = 100,000. ............................................................................................................... 97 

Figure 4.7 Tolerance intervals for torsional RMS response (σαα) of 100 MC simulations: (a) NMC = 1,000; (b) NMC = 5,000; (c) NMC = 10,000; (d) NMC = 50,000; (e) NMC = 100,000. ............................................................................................................. 100 

Figure 4.8 QMC-based scatter plots of RMS response for deck section at x= l/4 and for U = 22.2 m/s as a function of NQMC: (a) heave σhh; (b) torsion σαα. .......................... 101 

Figure 4.9 Coefficient of variation of the RMS response at x = l/4 for U = 22.2 m/s, computed by QMC algorithm: (a) heave σhh; (b) torsion σαα. ............................................ 102 

Figure 4.10 Tolerance intervals for vertical RMS response of 100 MC simulations (σhh): (a) NQMC = 1,000; (b) NQMC = 5,000; (a) NQMC = 10,000. ....................................... 104 

Figure 4.11 Tolerance intervals for RMS response of 100 MC simulations (σαα): (a) NQMC = 1,000; (b) NQMC = 5,000; (a) NQMC = 10,000. .................................................... 106 

Figure 4.12 RMS values for vertical buffeting response as a function of wind speed U corresponding to a given confidence level. ....................................................... 107 

Figure 4.13 Flutter derivatives H1* (a) and A2

* (b) of the Golden-Gate Bridge girder with aspect ratio B/D = 3.5:1. Data sets were reproduced from (Jain 1996; Jain et al. 1996) with α0=0°. The (“reference”) coefficients of the “Polynomial Model” were derived by regression of the data sets, according to Eqs. (3.1) and (3.2). Tolerance limits (dashed lines) were based on approximate evaluation of one standard deviation. ........................................................................................................... 108 

Figure 4.14 TEP curves of RMS response with respect to thresholds T1 to T3 at the deck section l/4: (a) σhh; (b) σαα (DFV: “Deterministic” Flutter Velocity). ................ 109 

Figure 4.15 Procedure for rescaling the TEP curves in Fig. 6.11(c) based on Eq. (6.3): (a) prior probability or TEP; (b) marginal likelihood function; (c) Posterior probability or TEP (DFV: “Deterministic” Flutter Velocity). ........................... 111 

Figure 4.16 TEP curves of RMS response at deck section l/4 (T2 threshold only) before (T2) and after rescaling (T2M): (a) σhh; (b) σαα (DFV: “Deterministic” Flutter Velocity). ........................................................................................................................... 112 

Figure 4.17 TEP curves of RMS responses with thresholds based on the RMS displacement, deck section at l/4 and NEU’s flutter derivatives: (a) σhh; (b) σαα. .................... 113 

Figure 4.18 TEP surfaces of RMS displacement for T2M threshold as a function of wind accounting for effects of skew winds at l/4 with literature flutter derivatives: (a) σhh at l/4; (b) σαα at l/4; (c) σhh at l/2; (d) σαα at l/2. ............................................ 115 

xi

Figure 4.19 TEP surfaces of RMS displacement for T2 threshold as a function of wind accounting for effects of skew winds at l/4 with NEU’s flutter derivatives: (a) σhh at l/4; (b) σαα at l/4; (c) σhh at l/2; (d) σαα at l/2. ................................................. 117 

Figure 4.20 National Data Buoy Center (NOAA Station 9414290, Latitude: 37.807 N, Longitude: 122.465 W) (Photo reproduced from NOAA, http://www.ndbc.noaa.gov/). ............................................................................. 118 

Figure 4.21 PDFs of “parent” (continuous time) mean wind velocity and annual maxima of mean wind velocity, data from NOAA (NOAA). .............................................. 119 

Figure 5.1 Reference peak vertical dynamic response, (“deterministic” without variability in FD) as a function of wind velocity at θ = 0° with both Jain’s flutter derivatives and NEU’s flutter derivatives at l/4: (a) displacement; (b) acceleration. .......... 138 

Figure 5.2 TEP curves of the peak dynamic response with respect to thresholds T1, T2, T3 deck section at l/4 using Jain’s FD data: (a) vertical response; (b) torsional response. ............................................................................................................ 139 

Figure 5.3 Recaled TEP curves (modified by Eq. 4.3) of the peak dynamic response with respect to thresholds T1, T2, T3 deck section at l/4 using Jain’s FD data: (a) vertical response; (b) torsional response. ....................................................................... 140 

Figure 5.4 Rescale TEP curves (modified by Eq. 4.3) of the peak dynamic response with respect to thresholds T1, T2, T3 deck section at l/4 using NEU’s FD data: (a) vertical response; (b) torsional esponse. ........................................................... 141 

Figure 5.5 A comparison between two sets of curves for vertical response based on NEU’s FD data (continuous lines) and Jain’s FD data (dotted lines): (a) T1=20 milli-g; (b) T2=1m; (c) T3=2m.............................................................................................. 143 

Figure 5.6 A comparison between two sets of curves for torsional response based on NEU’s FD data (continuous lines) and Jain’s FD data (dotted lines): (a) T1=20 milli-g; (b) T2=1m; (c) T3=2m.............................................................................................. 145 

Figure 5.7 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic vertical response with various intervention levels at l/4 using Jain’s FD data: (a) T1=20 milli-g; (b) T2=1m; (c) T3=2m.............................................................................................. 147 

Figure 5.8 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic (normalized to B/2) torsional response with various intervention levels at l/4 using Jain’s FD data: (a) T1=20 milli-g; (b) T2=1m; (c) T3=2m. .......................................................... 149 

Figure 5.9 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic vertical response with various intervention levels at l/4 using NEU’s FD data: (a) T1=20 milli-g; (b) T2=1m; (c) T3=2m.............................................................................................. 151 

xii

Figure 5.10 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic (normalized to B/2) torsional response with various intervention levels at l/4 using NEU’s FD data: (a) T1=20 milli-g; (b) T2=1m; (c) T3=2m. ................................................. 153 

Figure 5.11 Probability distributions of flutter velocity, Ucr, using NEU’s flutter derivatives: a) PDF (probability density function); b) CDF (cumulative density function). 154 

Figure 5.12 Resolution of the Monte-Carlo-based flutter procedure vs. standard error. ... 155 

Figure 5.13 Intervention costs normalized to the initial construction cost for user comfort level threshold T1=20 milli-g over time based on NEU’s FD data

: (a) 3D PMF (probability mass function stem plot); (b)

2D expected normalized cost, - discount rate/year λ=0.05................................ 156 

Figure 5.14 Expected intervention costs normalized to the initial construction cost due to

deformation in deck based on NEU’s FD data : (a) lower

tolerance case (T2=1m); (b) higher tolerance case (T3=2m), - discount rate/year λ=0.05. ............................................................................................................... 157 

0 0EC E C t C C

0 0EC E C t C C

xiii

Nomenclature

The following symbols were used in this dissertation:

* *1 4

1 5

damping matrix of the system

,.., flutter derivatives per unit length, torsional moment

,..., constant parameters in Eqs. (5.11a) to (11e)

stiffness matrix of the system

b

A A

b b

B

A

B

0

1

ridge deck width

decay factor

total cost of the structure at time (years)

initial cost of the structure

, random parameters in Eq. (4.1) ( = 1,3,5,7)

cost in present dolli i

j

c

C t t

C

C C j

C

,

ar value

, , drag, lift, moment static coefficients

expected value of maintenance and repair cost

infinitesimal inertia

search direction at step n (FORM)

/ bridge deck, as

D L M

M E

n

C C C

C

dm

B D

d

1

1, 4,

' '1 4

pect ratio

bridge deck height

, random parameters in Eq. (4.2) ( = 1,3,5,7)

,..., "reference" parameters in Eqs. (5.14) (average solution with no errors)

,..., mea

i i

ref ref

D

D D j

D D

D D

1 4

*

n-removed random ,...,

[] expectation operator

system matrix for two-mode aeroeastic instability analyses

complex conjugate transpose of matrix

loss of performance for the bridgv

C C

E

F

Ε

Ε E

,

e

marginal probability density function of

marginal probability density function of

joint probability density function of ,

gust effect factor

g( , ) limit state fu

U

U

cr site

f U

f

f U

g

U U

1, 1 1, 1

nction

,..., modal integrals, simulated bridgest t v vG G

xiv

1

( , ) vertical bridge oscillation, simulated bridges

dimensionless eigen-function of -th mode shape

( ), ( ) modal eigenfunctions, vertical oscillation ( -th mode and mode 1)

/

i

g v

h x t

h i

h x h x g v

h dh

* *1 4

*

0

ˆ peak vertical acceleration

,.., flutter derivatives per unit length, lift force

vertical flutter derivatives ( = 1,...,4)

ˆ imaginary unit

I identity matrix

mass mo

i

dt

h

H H

H i

i

I

g

1 1

ment of inertia per unit length of the deck, simulated bridges

I indicator function

, generalized modal inertias (modes , )

, generalized modal inertia for modes 1 and 1

, re

i j

t v

cr

I I i j

I I t v

K K

1 1

0

duced frequency, reduced critical frequency

, reduced modal frequencies for modes 1 and 1

effective reduced frequency

central span length of the simulated bridges

mass per unit

t v

n

K K t v

K

l

m

length of the bridge deck, simulated bridges

, , aeroelastic drag, lift, moment forces per unit length (Fig. 2.2)

, , bufetting drag, lift, moment forces per unit length (Fig. 2.2)

,

ae ae ae

b b b

D L M

D L M

n

1 1, still-air natural frequency (any, torsional mode 1, vertical mode 1), Hz

number of wind tunnel data points

number of Monte Carlo points

( , ) lateral vibration of the simulat

t v

MC

n n t v

N

N

p x t

ed bridges

dimensionless eigen-function of the -th mode

[] probability function

direction-dependent flutter probabilities

loss of performance of the bridge (threshold )

probab

i

F

T

j

p i

P

P

P T

P

41 1 1 1

ilities of each damage state

flutter probability

flutter probability

-mode shape of the bridge

, dimensionless modal groups (e.g., 0.5 / )

generalized force v

crU

f

i

t v t t

b

j

P

P

q i

q q q B L I

Q ector

generalized force of the -th modeiQ i

xv

standard Gaussian vector

standard error

, , cross PSD of modal loading between two generic sections and

, , auto and cross PSD function of the dynamic response

i j

E

F A B A B

hh h

Q Q

S

S x x K x x

S S S

S

S

modal force cross-spectra

Kaimal spectrum

Lumley-Panofsky spectrum

generic threshold

time

( , ) lateral component of turbulent wind

velocity variable used in conditional

uu

ww

S

S

T

t

u x t

u

*

probability functions (with )

friction velocity

* design point

wind speed, m/s

critical torsional-flutter speed, m/s

normal component of wind speed to the deck, m/s

cr

n

p

u U

u

u

U

U

U

U

,

parallel component of wind speed to the deck, m/s

reduced velocity ( /( ))

reduced critical velocity (at flutter); real root of Eq. (5.12)

wind speed at the bridge site, m/s

(

R

R cr

site

U U nB

U

U

v x

0

, ) horizontal component of turbulent wind

( , ) vertical component of turbulent wind

longitudinal coordinate along the bridge axis

terrain roughness length

( , ) torsional vibrati

t

w x t

x

z

x t

0

1

on of the simulated bridges

dimensionless eigen-function

mean-wind attack angle of the deck

( ), ( ) mode-shape functions, torsional oscillation of the simulated bridges

ˆ unit-gra

i

g t

n

x x

α dient row vector (FORM)

generalized safety index

Kronecker delta function

tolerance parameter (FORM)

dimensionless variance reduction coefficient

ij

xvi

Subscripts and superscripts:

generic mode index for the simulated bridges

generic step of iteration (FORM)

1, 1 fundamental bridge modes, torsional and vertical

† Moore-Penrose pseudo-inverse

g

n

t v

1 1

1 1

1 1

, structural modal damping with respect to critical, modes 1 and 1

generalized coordinate

( ), ( ) generalized modal coordinate, modes 1 and 1

( ), ( ) Fourier transfor

t v

i

t v

t v

t v

t t t v

K K

1 1

,

m of the generalized modal coordinate, modes 1 and 1

two-mode flutter eigenvector = , , with transpose operator

amplitude parameter in FORM

single-mode torsional instabilit

T T

v t

n

t I

t v

ξ

3 4 5 6

1

,

, , 3 4

y equation (see Eq. 5.9)

torsional-mode dimensionless ratio ( / )

value of at critical velocity

air density

, correlation coefficients of random parameters ,

t t

t cr t

D D D D

K K

D D

5 6

' ''

1 1

and ,

standard Gaussian comulative density function

frequency ratio, simple harmonic motion

, maximum positive and minimum negative differences

, , still-air circular fi i

t v

D D

requency (any, torsional and heaving mode), rad/s

critical-flutter circular frequency (torsional mode), rad/s

, , vertical, torsional and lateral RMS displacements

variance of thf

cr

hh pp

P

e flutter-probability MC estimator

still-air circular frequency, rad/s

-th mode natural circular frequency, rad/si i

1

Chapter 1

Introduction

The quantitative analysis of aerodynamic effects on long-span bridges has been considered

(Simiu and Scanlan 1996), since the collapse of the Tacoma Narrows Bridge in 1940 (shown

in Fig. 1.1). Wind engineering researchers have devoted great efforts to understand wind-

induced or “aeroelastic” phenomena (Davenport 1962; Scanlan and Tomko 1971), associated

with the vibration of long-span bridges. “Aeroelastic” is used to indicate fluid-structure

interaction between a flexible structural system and wind air flow. Investigations have also

been performed by many researchers (Kwon 2010; Namini et al. 1992; Scanlan 1987; 1993;

Scanlan and Jones 1990b) to prevent these loading mechanisms from adversely affecting the

satisfactory performance of long-span bridges.

The accurate assessment of fluctuating wind loads on long-span bridges is necessary

to avoid failures or undesired vibrations. Two types of “aeroelastic” phenomena, namely

flutter and buffeting response, are considered in this study. There are the two phenomena

that are relevant for the analysis of bridge deck. Both the potential collapse of the structure

2

due to flutter instability and the dynamic vibration due to wind turbulence (buffeting) at

moderate to high wind speeds are important for bridge design.

Flutter is defined as an oscillatory instability, induced in the bridge deck, when a

bridge is exposed to a wind speed above a certain critical threshold. Beyond this limit,

diverging vibration of the deck is possible, which may result in a catastrophic structural

failure. A classic example of such failures is illustrated in Fig 1.1. Instability can be predicted

through the analysis of flutter derivatives (FDs), and needless to say must be avoided by all

means.

Buffeting is defined as the dynamic vibration regime due to fluctuating loading,

promoted by wind turbulence, which is also influenced by the interaction with structural

deck motion. The bridge vibration is stochastic due to oncoming-flow turbulence and

“signature” turbulence, produced around the deck girder through flow separation and air

recirculation (e.g., (Jones and Scanlan 2001)). The dynamic amplification of vibration,

which causes buffeting, is often observed on long-span bridges (Miyata et al. 2002; Xu et al.

2007; Xu and Zhu 2005). Buffeting does not usually lead to catastrophic failure of the bridge.

However, vibrations cannot be avoided but need to be monitored since they can affect the

serviceability; in fact, damage, fatigue in selected structural elements and user discomfort

are possible.

Both phenomena may occur either separately or together, and can be predicted by

utilizing experimentation in wind tunnel (Bienkiewicz 1987; Ehsan et al. 1993; Huston et al.

1988). Such experimentation is essential for the derivation of wind-induced forces,

especially the loading terms on the deck triggered by fluid-structure interaction; these can

3

be expressed in terms of “flutter derivatives”, originally developed by (Scanlan and Tomko

1971), which are non-dimensional aerodynamic force coefficients per unit deck length as a

function of the “reduced velocity”. Flutter derivatives are the essential parameters in the

estimations of the critical wind velocity of flutter instability and the buffeting response of

long-span cable supported bridges.

Recently, it has been demonstrated that experimentally-derived FDs are random in

nature with uncertainty affected by measurement errors (Sarkar et al. 2009); this uncertainty

in the measurement of the FDs are unavoidable during testing in the wind tunnel. To assess

such uncertainties and the effects on both flutter and buffeting response, it is necessary to

develop specific “analysis tools” which could enable accurate bridge performance

assessment.

This dissertation proposes to develop a methodology for deriving the solution to

buffeting problem on long-span bridges, which will involve the direct representation of the

above-described sources of uncertainty in the aeroelastic input through appropriate statistical

analysis of the FDs.

A second order polynomial model for the FD is proposed and labeled as “model curve”

in this study. The coefficients of this polynomial are treated as random variables, whose

probability distribution is conditional on the reduced wind speed. For computational reasons

in subsequent analysis, however, this dependency is neglected and the probability of these

random variables is treated as independent of the reduced wind speed. For analysis purposes

the first- and second-order statistics are estimated from experiments, treating all the wind

speed data as part of the same population. Wind tunnel experiments are used to validate the

4

proposed methodology and to confirm the relevance of measurement errors in these

aeroelastic force terms. Wind tunnel tests have been conducted at Northeastern University

(NEU) for this purpose.

1.1 Motivation

The prediction and simulation of long-span cable-supported bridge dynamic response due to

wind hazards are particularly difficult in consideration of the inherent complexity of the wind

field, turbulence fluctuations and pressure distributions around the deck (which is the most

vulnerable part of the structure). Experimentation is essential for the derivation of this

dynamic loading. Despite the efforts of the research community towards the development of

refined techniques to simulate full-scale bridge response (Ozkan 2003; Ozkan and Jones

2003), observations can differ from the simulations of the response, based on wind-tunnel

experiments and measurements of equivalent loading. These discrepancies are associated

with the experimental procedures (and their errors) in wind engineering and must be

carefully accounted for in the existing simulation methods.

Reliability analysis against flutter, considering uncertainty in the FDs, has been

investigated by a few researchers (Dragomirescu et al. 2003; Ge et al. 2000; Kwon 2010;

Mannini and Bartoli 2007; Ostenfeld-Rosenthal et al. 1992; Pourzeynali and Datta 2002;

Scanlan 1999). Recently, it has also been shown how to experimentally estimate the

statistical moments of the FDs (variance, co-variance, etc.) from data extracted in wind

tunnel tests (Kwon 2010; Mannini et al. 2012).

5

However, very limited emphasis has been given to structural serviceability due to

uncertainty in buffeting loading (Caracoglia 2008a). With the aging bridge inventory in the

United States, it is therefore important to advance the current analysis approaches to include

the “statistical buffeting” response. The term “statistical buffeting” analysis is coined for the

first time in this dissertation to differentiate from the standard buffeting analysis in the

absence of uncertainty in the FD.

This dissertation focuses on the development of a methodology for “statistical

buffeting” analysis, including the uncertainty in the FD. To accomplish these tasks and, most

importantly, a second order polynomial model (“model curve”) for the FD is proposed. The

model curve is a second order polynomial description of the FDs where uncertainty is

associated with coefficient of the polynomial. This curve is used to describe the behavior of

the flutter derivatives as a function of reduced velocity. The physical justification for this

selection stems from the observation that the FDs tend to follow a general trend, especially

for moderately bluff deck sections (Simiu and Scanlan 1996). Therefore, postulating such a

“model curve” for FDs was selected as an appropriate assumption for projecting the

variability of the coefficients of the model curve into the analysis of the buffeting response

of the bridge.

In the standard approach the result of the buffeting analysis is the value of the RMS

dynamic response at a given wind speed. In the proposed probabilistic setting one estimates

the probability that a given threshold for the variance of the response is exceeded. This

probability is later used, together with information on the probability of the wind velocity at

a given site, to predict “lifetime expected cost” (Wen and Kang 2001) due to the buffeting

6

response of a 1200-meter suspension bridge (a function proportional to the cost associated

with interventions needed to ensure safety or for maintenance) is affected by the variability

of the FDs.

Even though the ultimate goal of the research is the development of a generalized

methodology for the solution to buffeting problems on long-span bridges for the analysis of

the effects induced by various sources of uncertainty; this should involve the extension of

the procedures and should include both wind-loading input and selected structural properties.

This dissertation represents a first step towards this objective.

1.2 Outline

This dissertation is divided into the following chapters, after providing a general introduction

and motivation in Chapter 1.

Chapter 2 summarizes the background theory of wind-induced response of long-span

bridges and reviews the fundamental aspects of aerodynamics and aeroelasticity of long-

span suspended bridge decks.

Chapter 3 describes the development of a “model curve” for representation of the

behavior of flutter derivatives as a function of reduced wind velocity. This chapter also

describes the experimental setup, measurements and experimental results, used in this

research. Flutter derivatives were measured in the wind tunnel at NEU.

Chapter 4 describes the standard buffeting analysis of long-span bridges

(“deterministic case”) as well as “statistical buffeting” analysis which includes the variability

in the FD. The “statistical buffeting” response was evaluated by adopting the concept of

7

“fragility”; this was employed in the calculation of the exceedance probability of pre-

selected vibration thresholds, conditional on mean wind speed and direction at the deck level.

Chapter 5 discusses the lifetime estimation of monetary losses for a long-span bridge,

designated as “cost analysis”, due to buffeting response. The expected value of the loss

function (lifetime cost estimation) of a 1200-meter suspension bridge is evaluated by

applying the “statistical buffeting” analysis. Summary of the work, conclusions of the

dissertation and directions for future research are discussed in Chapter 6.

8

Figure 1.1 Tacoma Narrows Bridge collapsed in 1940 due to wind-induced “torsional

flutter” (reproduced from Simiu and Scanlan, 1996).

9

Chapter 2

Wind-Induced Response of Long-Span Bridges: Review

This chapter reviews the fundamental aspects of aerodynamics and aeroelasticity of long-

span suspended bridge decks. Figures 2.1 and 2.2 show a section view of the bridge deck to

be analyzed. The wind-induced dynamic response of a long-span bridge close to aeroelastic

instability is most conveniently analyzed in the frequency domain, as described by (Jain

1996; Jones and Scanlan 2001; Katsuchi et al. 1999); this method, referred to as “multi-

mode” approach in wind engineering, is reviewed in this chapter.

2.1 General Formulation

The deflection components of the bridge deck (i.e., h(x,t), p(x,t) and α(x,t) in Fig. 2.2) can

be expressed in terms of the generalized coordinate of the mode ξi(t), the deck width B and

the dimensionless representations of the i-th mode form along the deck hi(x), pi(x) and αi(x)

as

10

vertical : ( , ) ,i ii

h x t h x B t (2.1a)

(2.1b)

torsional : (x,t)

ix i

t i . (2.1c)

In Eq. (2.1), hi(x), pi(x) and αi(x) are dimensionless eigen-functions associated with

the i-th mode shapes of the deck, ξi(t) are generalized coordinates. x is the coordinate along

the deck span and t is time (Fig. 2.2). The linear dynamic response of the bridge deck is

derived by standard modal expansion of the vibration in terms deck modes, as indicated in

Eqs. (2.1); the original formulation, described for example by Jain (1996), is based on the

representation of the deck girder as a continuous (i.e., beam-type) element oriented along

the x axis, transversely rigid section by section, and with lateral and vertical vibration with

respect to the centroid C of the deck section (Fig. 2.2) in the directions orthogonal to x (h

and p) and torsional rotation about x. Only vertical, lateral, torsional components of the deck

section are used herein because these three are considered as primary deflection components

due to wind loadings for long-span suspension bridges.

A review of modal expansion techniques for structural dynamic analysis of

continuum systems is not included in this chapter but may be found, for example, in

(Meirovitch 1970). It must be noted that normalization of the eigen-functions in Eq. (2.1) is

performed with respect to a reference dimension B, coincident with the deck width. This

normalization was first introduced by (Scanlan and Jones 1990a) to enable the subsequent

derivation of the dynamic modal response equation in a general form, regardless of the

specific features of the generic mode i (lateral, vertical, torsional, etc.).

lateral : ( , ) ,i ii

p x t p x B t

11

The governing generalized equation of motion of mode i (ξi) therefore becomes (e.g.,

Scanlan and Jones, 1990)

I

i

i 2

i

i

i

i2

i Q

it , (2.2)

where Ii and Qi(t) are the generalized inertia and modal force of the i-th mode, ωi and ζi are

the i-th mode natural circular frequency and the modal damping ratio.

The generalized inertia Ii is defined as

2 , , , , ,i istructureI q x y z dm x y z (2.3a)

where qi (x, y, z) are the i-th mode shape of the bridge and dm is an infinitesimal inertia; this

equation is written in a general form to emphasize that the integration may be carried out

over the entire structure to also account for modal mass contributions from portions of the

bridge other than the deck itself, which are also involved in the vibration; these terms depend,

for example, on the moving cables in a suspension bridge or the tower motion. If the mass

and moment of inertia of the moving deck and cables are assumed as constant along x, and

respectively expressed by m0 and I0 per unit deck length, the modal inertia, using the

expansion in Eq. (2.3a), simply becomes

2 2 2 2 20 0 00

,l

i i i iI m h x B m p x B I x dx (2.3b)

12

in which l is the total deck length.

The generalized force Qi(t) due to wind loading is given in a similar form by Eq.

(2.4).

0

, , , ,l

i i i iQ t L x t h x B D x t p x B M x t x dx (2.4)

where L(x,t), D(x,t) and M(x,t) represent the lift, drag and pitching moment per unit span

length (in Fig. 2.2). They are assumed to be separable into motion-dependent loads and

turbulence-induced loads (motion-independent) and are defined as

lift : ae bL L L (2.5a)

drag: ae bD D D (2.5b)

moment : ae bM M M (2.5c)

where the subscripts ae and b refer to aeroelastic and buffeting loads, respectively. The

aeroelastic (or self-excited forces) are assumed to be linearizable. For purely sinusoidal

motions of frequency ω, the aeroelastic forces can be expressed as (Scanlan and Tomko

1971).

Lae

1

2U 2B KH

1*h

U KH

2* B

U K 2H

3* K 2H

4* h

B KH

5* p

U K 2 H

6* p

B

, (2.6a)

13

Dae

1

2U 2B KP

1* p

U KP

2* B

U K 2P

3* K 2P

4* p

B KP

5*h

U K 2P

6* h

B

, (2.6b)

Mae

1

2U 2B2 KA

1*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

, (2.6c)

where is the air density, U the mean velocity of the oncoming wind (which is turbulent

in general) at the deck level, K (= ωB/U) is the reduced frequency with ω circular frequency;

/h dh dt , d / dt and /p dp dt pertain to the deck section at x. In Eq. (2.6), Hi* Pi

*

and Ai* (with i = 1,…,6) are flutter derivatives of the deck cross-section. As noted by Scanlan

and Tomko (1971), the previous expressions are written in a mixed time-frequency form and

are valid for simple harmonic motion of the deck at a given ω; this assumption is strictly

valid at flutter but is acceptable in the case of vibration induced by turbulence disturbances

on low-damping systems, such a suspension bridge. Equations were derived by extension of

the standard airfoil theory (Theodorsen 1935) to non-aerodynamic (bluff) bridge deck

sections. It must be noted that the frequency-time duality is only apparent since it disappears

once Fourier-domain analysis is employed to derive the bridge response (Scanlan and Tomko

1971), as later described.

Buffeting forces per unit length are fluctuating loads which can be described by

“quasi-steady theory” and turbulence disturbances for vibration about a static equilibrium

configuration of the deck due to mean wind loads. For mean incident wind orthogonal to the

bridge axis, the loading depends on the vertical (w), and lateral (u) turbulence, which are

stochastic quantities as a function of time t and position x along the deck (Fig. 2.2). For small

vibration amplitudes these can be obtained by first-order expansion about on equilibrium

14

position under mean wind, described by an angle of attack α0 in the vertical plane (Fig. 2.2).

As an example, the lift force Lb, drag force Db and moment Mb per unit deck length are

defined as (Jones and Scanlan 2001)

2 , ,1, 2 ,

2L

b L D

u x t w x tdCL x t U B C C

U d U

(2.7a)

2 , ,1, 2 ,

2D

b D

u x t w x tdCD x t U B C

U d U

(2.7b)

2 2 , ,1, 2 ,

2M

b M

u x t w x tdCM x t U B C

U d U

(2.7c)

with CL, CD and CM being lift, drag and moment static coefficients (referred to deck width

B) of a typical deck section, evaluated at mean-wind attack angle of the deck α0; u(x,t), w(x,t)

are the along-wind and vertical components of turbulent wind. Span-wise correlation loss of

the turbulence-induced forces along the x direction (from Eq. 2.7) was incorporated into the

formulation to calculate the generalized loading Qi. The standard exponentially decreasing

coherence model was employed (e.g., Jones and Scanlan, 2001). This model is later

described in Section 2.3.

For example, for single-mode analysis, Eq. (2.1) can be further simplified as

( , ) ( ) ( ),h x t h x B t (2.8a)

( , ) ( ) ( ),p x t p x B t (2.8b)

( , ) ( ) ( ).x t h x t (2.8c)

15

Similarly, the dynamic loading Qi in Eq. (2.4) becomes a scalar term. The multi-mode

system of equations can be formed by separating the generalized loading Qi(t) into

aeroelastic and buffeting components Qi(t)=Qae,i(t)+Qb,i(t) and by recognizing that the

loading induced by Qae,i(t) is linearly dependent quantities related to dynamic motion and

velocity of the deck sections; therefore, Qae,i(t) can be expressed as a linear function of the

generalized displacements and velocities through Eq. (2.1) and the effect of Qae,i(t) on the

bridge can be interpreted as a modification to the generalized stiffness and damping of the

structural modes which depend on wind speed U. A more detailed description may be found

in Scanlan and Jones (1990). The generic scalar Eq. (2.2) can be simplified as

i 2

i

i

i

i2

i Q

ae,it I

i Q

b,it I

i and the left-hand side rewritten by regrouping

terms as a function of Qae,i(t)/Ii as wind-induced stiffness and damping equivalent quantities.

In matrix notation the dynamic system becomes, after modal truncation to a significant set

of modes,

'' ' ,b s I A B Q (2.9)

where ξ is the generalized coordinate vector, ( )’ represents a derivative with respect to a

dimensionless time (Jain et al. 1996) s = Ut/B, I is the identity matrix, A and B are the

damping and stiffness matrices of the system, which are no longer diagonal since mode

coupling is induced by fluid-structure interaction via aeroelastic loads; finally, Qb is the

generalized buffeting force vector.

16

The general terms of A, B and Qb can be expressed as (Jones and Scanlan 2001)

4

* * * *1 2 5 1

* * * * *2 5 1 2 5

A 2 [2

],

i j i j i j i j

i j i j i j i j i j

ij i i ij h h h h p p pi

p p h h p

B lKK K H G H G H G P G

I

P G P G A G A G A G

(2.10)

4

2 * * * *3 4 6 1

* * * * *4 6 3 4 6

B [2

],

i j i j i j i j

i j i j i j i j i j

ij i ij h h h h p pi

p p p h h p

B lKK K H G H G H G P G

I

P G P G A G A G A G

(2.11)

4

0Q {L , D , M , } ,

2i

l

b b i b i b ii

B l dxs x s h x s p x s

I l

(2.12)

where Ki=Bωi/U and the dependency on K=Bω/U is due to the flutter derivatives; δij is the

Kronecker delta function defined as

1.

0ij

i j

i j

(2.13)

Note that the diagonal terms (i = j) in Eqs. (2.10) and (2.11) represent the single

degree of freedom (and uncoupled) equations. The off-diagonal terms permit the aeroelastic

coupling through the flutter derivatives and through the cross-modal integrals among

different modes.

The modal integrals between the ri normalized displacement or rotation of mode i

and the corresponding sj component of mode j are denoted by i jr sG and are obtained by

17

0( ) ( ) ,

i j

l

r s i j

dxG r x s x

l (2.14)

where ri = hi, pi or αi, sj = hj, pj or αj.

The new system of equations can be Fourier transformed into the reduced frequency

(K) domain (Scanlan and Jones 1990a); for example a generic time-dependent function f(s)

with s = Ut/B becomes in the frequency domain with ˆ 1i :

ˆ

0.iKsf K f s e ds

(2.15)

Consequently, a system of equations, exclusively dependent on K and, is derived

from Eq. (2.9) as

,b E Q (2.16)

where and bQ are the Fourier-transformed ˆ

0

iKsK s e ds and bQ vectors,

respectively. A general term of the “impedance matrix” E is

2 ˆ .ij ij ij ijE K iKA K B K (2.17)

18

2.2 Background on Flutter

A review of flutter theory is presented in this section. Information was derived from (Jain

1996; Katsuchi 1997). Flutter is an oscillatory instability induced when a bridge is exposed

to a wind speed above a certain critical threshold. Beyond this limit, diverging vibration of

the deck is possible, which may result in a catastrophic structural failure. One particular

example was the Tacoma Narrows incident in 1940 when such phenomenon was clearly

recognized. Flutter instability must be avoided by all means in bridge engineering.

Aeroelastic instability can be predicted by analyzing the aeroelastic coefficients of bridge

decks (flutter derivatives in Eq. 2.6) developed by (Scanlan and Tomko 1971), which are

employed for simulating the dynamic response of the bridge. Flutter derivatives in Eq. (2.6)

are force coefficients per unit length, routinely measured in wind tunnel tests. As briefly

outlined in a previous section, these expressions of the shape-dependent force coefficients

of the moving deck section have been strongly influenced by airfoil theory for streamlined

bodies (Theodorsen 1935), from which they have been derived for use in civil engineering

applications.

Frequency-domain analysis was used in this study; this is the preferred method of

bridge dynamic researchers because this analysis can be related to direct physical

interpretation through flutter derivatives, obtained experimentally, as opposed to time-

domain analysis, which requires modeling of the Qae,i(t) loading in terms of convolution

integrals (Scanlan et al. 1974) and a more complex formulation (on occasion suffering from

the lack of physical interpretation; (Caracoglia and Jones 2003).

19

The flutter condition is identified by solving the aeroelastically influenced effective

damping problem derived from Eq. (2.16) by setting the turbulence-induced loading 0bQ

(Jones and Scanlan 2001; Katsuchi et al. 1998).

. E 0 (2.18)

Equation (2.18) can be reduced to the nontrivial solution of the complex algebraic

system. However, a direct method for the solution of det[E] is not available because the

matrix E includes two unknown variables, K and ω. The matrix E also consists of complex

numbers so that the condition of det[E]=0 must be satisfied with both the real and imaginary

parts of the determinant simultaneously equal to zero. An iterative procedure is needed to

solve for det[E]=0 (Jones and Scanlan 2001). This can be accomplished by fixing a value of

K and seeking a value of ω, in the frequency range of interest, for which the determinant is

zero, and changing the value of K until both the real and imaginary determinants are zeros

at the same ω. Once the values K (= ωB/U) and ω are obtained with satisfying Eq. (2.18),

the flutter speed can be calculated.

For a multi-mode problem, the same procedure is required and the largest value of K

of all solutions gives the flutter-critical condition. The mode corresponding to the solution

of ω is the leading mode in the flutter condition. Moreover, the eigenvector ξ at the flutter

condition gives the flutter mode shape which indicates the relative participation of each

structural mode in flutter vibration.

20

2.3 Background on Buffeting

A review of buffeting theory is presented in this section; the material has been derived from

(Jain 1996; Katsuchi 1997). Buffeting is a dynamic phenomenon, in which the wind-induced

loading is dynamic and due to wind turbulence. The bridge vibration is stochastic due to

oncoming-flow turbulence or signature turbulence, enhanced by flow separation and

recirculation around the deck girder (e.g., (Jones and Scanlan 2001)).

The vector of buffeting forces on the right hand side of Eq. (2.16) is

1

2

01

40

2

0

1F

1F

,2

1F

n

l

b

l

b

b

l

bn

dx

I l

dxB l

I l

dx

I l

Q

(2.19)

where the integrands in the vector above are the Fourier transforms of the Eq. (2.12).

F , L , D , M , .ib b i b i b ix K x K h x x K p x x K x (2.20)

By substituting the terms L b, Db

and Mbfrom Eqs. (2.7a – 2.7c) at the span location

xA, Eq. (2.20) leads to

21

' ' '

1F , [{2 ( ) 2 ( ) 2 ( )} ( )

{( ) ( ) 2 ( ) ( )} ( )],

ib A L i A D i A M i A

L D i A D i A M i A

x K C h x C p x C x u KU

C C h x C p x C x w K

(2.21)

where notations CL’=dCL/dα, CD’=dCD/dα and CM’=dCM/dα are used with the derivatives of

the static coefficients with respect to the angle of attack evaluated at α0.

Taking the complex conjugate transpose of the j-th equation at the location xB results

in

*

' ' '

1F , [{2 ( ) 2 ( ) 2 ( )} ( )

{( ) ( ) 2 ( ) ( )} ( )],

ib B L j B D j B M j B

L D j B D j B M j B

x K C h x C p x C x u KU

C C h x C p x C x w K

(2.22)

where the ( )* represents the complex-conjugate-transpose operation. From Eq. (2.19), using

Eqs. (2.21) and (2.22) the following matrix equation can be obtained (Jones and Scanlan

2001)

1 1 1

1

* *

0 0 0 02 1 1 14

*

* *

0 0 0 01

1 1F F F F

,2

1 1F F F F

n

n n n

l l l lA B A B

b b b bn

b b

l l l lA B A B

b b b bn n n

dx dx dx dx

I I l l I I l lB l

Udx dx dx dx

I I l l I I l l

Q Q

(2.23)

22

from which the turbulence-induced dynamic response of the bridge can be derived by

standard random vibration techniques (e.g., (Newland 1993). The power spectral density

(PSD) matrix of the generalized loading can be calculated, a general term of which is

24

0 0

1, ,

2

, , ,

b bi j

l l

Q Q i A j B uu A Bi j

A Bi A j B ww A B

B lS K q x q x S x x K

U I I

dx dxr x r x S x x K

l l

(2.24)

where

2 ,i L i D i M iq x C h x C p x C x (2.25)

' ' '2 .i L D i D j M jr x C C h x C p x C x (2.26)

The equations above depend on the cross-power spectral densities of the lateral

turbulence Suu(xA,xB,K) and vertical turbulence Sww(xA,xB,K). The uw-cross spectrum of the

wind Suw has not been included in the equation above as it is usually of secondary importance

on the dynamic response. A more complete formulation may be found in (Jain 1996).

If the auto-power spectral density of the wind components is independent of the

location x along the deck axis, the span-wise cross-spectral densities of the wind components

in conventional form as (Scanlan and Tomko 1971)

23

, , exp ,2A B A B

cKF x x K F K x x

B

(2.27)

where c is a decay factor, the range of which is generally taken as

8 16nl nlc

U U . (2.28a)

Therefore, the cross-power spectral densities Suu(xA,xB,K) and Sww(xA,xB,K) can be

simplified as

( , , ) exp ,2

( , , ) exp .2

uu A B A Buu

ww A B A Bww

cKS x x K S K x x

B

cKS x x K S K x x

B

(2.28b)

The limits in Eq. (2.28b) can be used for force calculations to reflect the higher span-

wise correlation in the pressure loading than the one seen in the velocity components (Larose

1992). Using the following expressions

0 0

exp ,2i j

l lA B

rs i A j B A B

dx dxcKH K r x s x x x

B l l (2.29)

where ri and si = hi, pi or αi, the ij-th term of the buffeting force matrix can be expressed as

24

24 1

,2b bi j

Suu SwwQ Q ij uu ij ww

i j

B lS K Y K S K Y K S K

U I I

(2.30)

where

2 2 22 2 2

4 4

4 ,

i j i j i j

i j i j i j i j

i j i j

Suuij L h h D p p M

L D h p p h L M h h

D M p

Suuij

Suuij p

Y K C H C H C H

C C H H C C H H

C C H

K

K H

Y

Y

(2.31)

2 2 2' ' '

' ' ' '

' ' .

i j i j i j

i j i j i j i j

i j i j

Swwij L D h h D p p M

L D D h p p h L D M h h

D M p p

Y K C C H C H C H

C C C H H C C C H H

C C H H

(2.32)

The power spectral of the wind components u and w in the atmospheric boundary

layer, expressed as functions of K, are assumed as (Simiu and Scanlan 1996)

2*

5/3

200,

1 502

uu

zuS K

KzU

B

(2.33)

2*

5/3

3.36.

1 102

ww

zuS K

KzU

B

(2.34)

25

The first equation above is the “Kaimal spectrum”; the second expansion is the

“Lumley-Panofsky spectrum”. In the previous equations z is the elevation above ground and

u* is the friction velocity, a function of the surface roughness. The friction velocity u* can be

determined by using

*

1ln ,

o

zU z u

k z (2.35)

where U(z) is the mean wind velocity at elevation z (usually taken as the velocity at the deck

level, U), k is the von Kármán constant which is generally assumed to be 0.4k (Simiu and

Scanlan 1996) and z0 is the terrain roughness length.

The power spectral density matrix for the generalized displacements ξ is developed

in dimensionless form using Eq. (2.16) as

11 *( ) ( ) ( ) ,b bQ QK K K K

S E S E (2.36)

where E* is the complex conjugate transpose of matrix E.

The PSD of the physical displacements (Eqs. 2.1a – 2.1c) can be obtained from the

PSD of the respective generalized displacement components through

2, , ,i jhh A B i A j B

i j

S x x K B h x h x S K (2.37)

26

2, , ,i jpp A B i A j B

i j

S x x K B p x p x S K (2.38)

, , ,i jA B i A j B

i j

S x x K x x S K (2.39)

where i and j are summed as the summation over the number of modes being used. Cross-

spectral densities can be developed in a similar manner.

Evaluation of the spectral densities of the displacements at combinations of discrete

xA and xB will result in a matrix. The mean-square values of these displacements can be

evaluated in terms of their respective PSD functions.

2

0

2,hh hh

BS K dK

U

(2.40a)

2

0

2,pp pp

BS K dK

U

(2.40b)

2

0

2,

BS K dK

U

(2.40c)

where K is the reduced frequency.

A covariance matrix for h, p and α is thus obtained, from which statistics of the

displacement components h, p and α can be calculated. A use of the reduced frequency, K,

as the variable of integration results in an additional factor of 2πB/U in the estimation of

mean-square value.

27

2.4 Effect of Wind Directionality: Skew Wind Theory

In the previous sections, considering the wind as it approaches the bridge orthogonal to the

deck axis derives basic flutter and buffeting studies. However, in nature, the highest winds

of record at a given site is very likely to be skew to the bridge (Scanlan 1999). Therefore,

flutter and buffeting analyses are modified to account for directionality θ (mean-wind yaw

angle in Fig. 2.3), as originally proposed in (Scanlan 1993).

A review of skew wind buffeting theory is presented in this section. The material is

derived from (Scanlan 1993). In Fig. 2.3, the plan view of the deck is shown with steady

mean wind velocity U and accompanying turbulence components (i.e., u(t), v(t) and w(t))

and approaching the bridge at an angle θ with respect to the direction orthogonal to the deck

axis indicated by the normal y (or y(n) in Fig. 2.3). As explained in (Scanlan 1993), the effect

of a skew wind on the bridge vibration can be estimated by considering a component Un

normal (across-deck) to the deck of mean wind velocity and two turbulence components u

and v as

co s sin ,nU U u v (2.41)

s in co s .pU U u v (2.42)

The effect of the parallel (along-deck) component to the deck Up, usually of minimal

relevance, was neglected in this research. As an example, vertical and torsional aeroelastic

loadings due to two components of skew wind U shown in Fig. 2.3 can be constructed by

replacing the original expressions by means of a reduced skew frequency Kn:

28

Lae

1

2U 2 B K

nH

1*h

U K

nH

2* B

U K

n2H

3* K

n2H

4* h

B

, (2.43)

Mae 1

2U 2B2 Kn A1

*h

U Kn A2

* B U

Kn2 A3

* Kn2 A4

* h

B

, (2.44)

where given the reduced K = Bω/U, the normal component becomes Kn = K/cosθ. Flutter

derivatives in previous equations are also evaluated at Kn (sometimes referred to as the

“cosine rule”).

Similarly, buffeting loads can be constructed due to skew wind effect as follows:

21cos 2 2 sin ,

2L

b L D L

v tdCu wL U B C C C

U d U U

(2.45)

2 21cos 2 2 sin ,

2M

b M M

v tdCu wM U B C C

U d U U

(2.46)

with static coefficients evaluated at α0, CD, CL and CM.

This formulation can be recast in Eqs. (2.37 – 2.40) as a function of u, v, w, θ and Kn

and can be used directly into the framework of the multimode algorithm, introduced in the

previous sections.

29

Figure 2.1 A suspension bridge and a section of the deck (Schematic view of a generic

finite-element model of the structure).

Figure 2.2 Degrees of freedom and aeroelastic forces on a bridge deck (the p component neglected in this study).

30

Figure 2.3 Schematic plan view of bridge deck with skew wind approaching the girder at wind speed U with turbulence components u, v, w and skew wind angle θ (Scanlan 1993).

θ U+u(t)

v(t)

w(t)

Up

Un

BBridge Deck

y (n)

x

31

Chapter 3

A Second-order Polynomial Model for Flutter

Derivatives

This chapter introduces the concept of “model curve”, a polynomial-based function for

flutter derivatives in terms of reduced velocity and used to describe in a generalized form;

the coefficients of this polynomial are random variables considering the uncertainty in the

flutter derivative (FD). Probability distribution or the random variables is conditional on the

reduced wind speed. For computational reasons in subsequent analysis, however, this

dependency is neglected and the probability of these random variables is treated as

independent of the reduced wind speed. For analysis purposes the first- and second-order

statistics are estimated from experiments, treating all the wind speed data as part of the same

population. Experiments were conducted in the wind tunnel, maintained by the Department

of Mechanical and Industrial Engineering at Northeastern University (NEU).

This chapter describes the experimental methods and wind tunnel tests, employed for

the extraction of aeroelastic coefficient or flutter derivatives (FDs) and also for the

32

estimation of the first- and second-order statistics of the polynomial model. Section models,

representing a section of the deck in a long-span bridge, were used in the study. These models

are intended to represent only a portion of the deck of the bridge, i.e., the section

schematically shown in Fig. 2.2. The description of the wind tunnel and bridge models used

in this study are given. The FDs were found simultaneously from two-degree-of-freedom

(two-DOF) coupled motion section model tests.

The model curve will be later used in Chapters 4 and 5, which enables to directly

project the uncertainty in the FD into the analysis of the buffeting response of the bridge

3.1 Description of the Polynomial Model and Discussion on its Physical

Interpretation

3.1.1 Description of the Polynomial Model

It has been shown (e.g., Scanlan and Tomko, 1971) that most flutter-derivative experimental

curves tend to follow a similar trend, especially for relatively bluff deck sections (Simiu and

Scanlan 1996). Postulating a second order polynomial model for flutter derivatives (FDs) to

describe the evolution of flutter derivatives as a function of reduced velocity was proposed

as a physically acceptable assumption in the context of simulation. The polynomial, labeled

as “model curves”, for FDs as a function of reduced velocity UR=2π/K with i = 1,...,4 are

shown below:

* 21 ,i R j R j RH U C U C U

1,..., 4 1,3,5,7i j (3.1)

* 21 .i R j R j RA U D U D U

1,..., 4 1,3,5,7i j (3.2)

33

These parameters become Cj and Cj+1 for Hi*, Dj and Dj+1 for Ai

*. The general form

for all Hi* and Ai

* derivatives can be expressed as in Eqs. (3.1–3.2) with i=1,…,4 and

j=1,3,5,7. In Eq. (3.1), Cj and Cj+1 are constant parameters of the model, which are assumed

as random coefficients and can be related in a simple way to experimental errors. The mean

values of Cj and Dj can be determined from the mean of experimental points, extracted at

various wind speeds in wind tunnel; similarly, second-moment properties of Cj and Dj can

be related to the variances of measured FDs; the coefficients of the polynomial are

determined from statistical regression of the experimental data, described in a separate sub-

section.

3.1.2 Discussion on the Selection of the Polynomial Model, based on Physical

Behavior of Flutter Derivatives

It must be noted that the selection of the model curves, based on a second order polynomial,

is not arbitrary but has a direct interpretation with the physical phenomenon related to the

concept of FD. FDs are employed to describe the unsteady fluid-structure interaction.

Nevertheless, as a first approximation, the Hi* and Ai

* coefficients can be estimated by using

a suitable combination of the static lift and moment coefficients of the deck section model,

measured in a “static test” of the model, rigidly mounted on a fixed force balance (e.g., Simiu

and Scanlan, 1996). Using the general theory of quasi-stationary wind forces (e.g., Simiu

and Scanlan, 1996) and recalling that “air inertial” contributions are negligible in this

formulation (whence H*4, A*

4 derivatives cannot be evaluated), the approximate expressions

34

of the flutter derivatives as a function of reduced frequency K=2π/UR for initial angle of

attack α0 close to zero are (Singh 1997; Strømmen 2006)

0 0 00 0 0* * *1 2 3 2

, , ,

L L LdC dC dCd d d

H H HK K K

(3.3a)

0 0 00 0 0* * *1 2 3 2

, , .

M D MdC dC dC

d d dA A A

K K K

(3.3b)

In the expressions above the derivation with respect to the static angle of attack α is

applied to the static lift coefficient (CL) and moment coefficient (CM), which are constant

and independent of flow speed. Expressions above are inversely proportional to K (with

exponent at most equal to 2) or, in other words, proportional to the reduced velocity UR with

the same exponent. The equations above show that flutter derivatives can be theoretically

interpreted as a “monomial” in terms of reduced velocity (inverse of K), which is at most of

order two. Since the expressions above are theoretically valid for low K only (K of the order

0.2 to 0.4) only (Strømmen 2006) it is reasonable to assume a polynomial expression as the

most plausible model for the derivatives, based on physical evidence. Therefore, the use of

Eqs. (3.1) and (3.2) in the model curves is justified by physical evidence. This interpretation

is, however, valid for linear superposition of aeroelastic effects, because of the use of the

static coefficients in Eq. (3.3) and for small vibrations h or α compared to the reference

dimension of the deck (width B or depth D).

35

3.2 Description of the Wind Tunnel, used for Experimental Verification of the

Polynomial Model

The wind tunnel tests were performed in the small-scale low-speed wind tunnel of

Northeastern University (NEU). A two-DOF elastic force balance had been designed by a

former graduate student (Brito 2008) and built for free-vibration tests of scaled models of

bridge deck sections. The design of the balance was partially based on an existing setup,

developed by (Chowdhury and Sarkar 2003). The design of the NEU setup was tailored to

the specific characteristics of the NEU’s wind tunnel due to the limitation of the physical

dimensions of the facility and of the test chamber.

The experimentation was based on a “section-model”, which is a replica of a “section”

of the actual bridge. Section model consists of representative span-wise sections of the deck

constructed to scale, spring supported at the ends to allow for both vertical and torsional

motion (Simiu and Scanlan 1996). These are constructed at a reduced geometric scale in

comparison with the full-scale structure, usually of the order 1:50. Section models are widely

used since they have the important advantage of enabling the measurement of the

fundamental aerodynamic and aeroelastic characteristics of the bridge deck (flutter

derivatives).

The wind tunnel tests were carried out in a closed circuit wind tunnel. The tunnel has

a 305 mm × 305 mm (12 in × 12 in) test section and produces wind speeds up to 45 m/s (150

ft/sec) in smooth flow. The air flow is driven by a 15 hp DC motor which is connected to a

compressor blade that generates the air flow. The generated flow is controlled to pass from

36

the settling chamber toward the “drive section”, where the air is recalculated (see Fig. 3.1,

(Brito 2008)). This motor provides sufficient power to move the air though the tunnel.

The wind tunnel was originally designed for mechanical engineering applications

(i.e., for testing in smooth flow) so that it does not provide a long test section. For civil

engineering application the long test section is desirable to simulate atmospheric turbulence

(i.e., for simulating the boundary layer flow). However, it was not necessary, since smooth

flow was used in the tests. In fact, the purpose of this study was to interpret uncertainties on

measurement errors. Turbulent flows and boundary layer flows for section model analysis

of bridges can be experimentally obtained by simply adding passive devices (e.g., girds and

honeycomb mesh) at the exit of contraction cone, which can generate a uniform turbulence

field in front of the deck section. Uniform turbulence is acceptable in the contest of section

model testing of bridge decks (Jones and Scanlan 2001).

3.3 Description of the Experimental setups, used for Verification

The experimental setup allows for two-DOF free vibration test simulating vertical (h) and

torsional (α) dynamic response simultaneously at different wind velocities. Lateral

displacements (p) were not considered as these usually affect marginally the dynamic

response. This section briefly describes the setup, designated as NEU’s Aeroelastic Force

Balance (i.e., shown in Fig. 3.2), developed by former graduate student (Brito 2008).

The setup includes a T-shaped rig, mounted externally to the wind tunnel test

chamber (“the box”), for the suspension of the model, which can be vertically supported on

a set of extension springs. The springs were selected such that the spring combination,

37

configuration and spacing, for a particular DOF (either h or α) could produce the desired

stiffness (i.e., frequency) in the two-dimensional dynamic system calibrated to obtain a

reasonable duration of time history response in the free-decay tests. The target mechanical

frequencies of the dynamic system, selected for the design, were chosen as about 6 Hz and

10 Hz for vertical and torsional DOF, respectively. The extension springs can be mounted at

pre-selected distances from the center of the section model, allowing changes in the

frequency ratio between torsional and vertical oscillation of the section model. Initial pre-

tensioning in the springs was imposed to ensure that large displacements, even two to three

times the depth of the model, were possible during behavior in the mechanical model (e.g.,

slackening of the springs).

3.4 Description of the Aeroelastic Section-Model, used for Verification

The girder of the Golden Gate Bridge (truss-type deck girder shown in Fig. 3.3) was selected

as the benchmark in this work. A bridge section model, based on information derived from

(Jain 1996; Jain et al. 1996; 1998) was built at a geometric scale 1:360 and used in the wind

tunnel investigations; data from the experiments are later employed in the simulations in

Chapters 4 and 5.

The Golden Gate Bridge (full-scale structure) is a suspension bridge over the

“Golden Gate”, a strait between San Francisco Bay, California (USA) and the Pacific Ocean.

It connects the city with Marin County. The bridge consists of a center span of 1,280 m

(4,200 ft) and deck width of 27.5 m (90 ft). The cables are each supported on two steel towers,

38

each rising 227 m (746 ft) above water level, and are anchored in massive concrete

anchorage blocks at their ends.

Figure 3.2(b) depicts the “section model” of the deck used in this study, mounted on

the NEU’s balance. This model simulates the aeroelastic behavior of a truss-type deck girder

and approximately replicates the features of the Golden Gate Bridge at the scale 1:360. The

width-to-depth aspect ratio for the deck girder is B/D = 3.5:1; the geometric scale selection

was dictated by the maximum dimensions of the test chamber at NEU; the model scale was

kept as a minimum to avoid “blockage effects”, which would have required corrections of

the aeroelastic forces and flutter derivatives (Barlow et al. 1999). The mechanical

frequencies of the experimental apparatus (which include the mass of the model and of the

rig) are: 7.8 Hz (vertical) and 9.8 Hz (torsional); model width is B = 76 mm.

3.5 Description of the Tests and Experimental Identification

Eight flutter derivatives of this truss-type bridge section model (Fig. 3.5), H1*,…,H4

* (for lift

force) and A1*,…,A4

* (for moment), were extracted from 1-DOF and 2-DOF free-decay

dynamic tests; the setup is shown in Fig. 3.2(b). The “Iterative Least Squares Method (ILS)”

(Chowdhury and Sarkar 2003) was used for identifying the flutter derivatives for lift and

moment. The ILS method was developed for the extraction of flutter derivatives from free

vibration displacement time histories obtained from a section model testing in the wind

tunnel. The main features of the method are briefly summarized below.

The equations of motion for the section model, subjected to aeroelastic forces, can

be written as a function of the state vector ,T

h y as

39

My Cy Ky Fae

. (3.4)

The mass matrix M, damping matrix C and stiffness matrix K of the mechanically

suspended system (Fig. 3.2b) are of dimension 2 by 2 and can be assumed as diagonal if the

mechanical coupling between the degrees-of-freedom h and α is eliminated by appropriate

design of the suspension system and model in the experimental setup. The scalar terms of

the mass matrix include the contribution of the moving masses and suspension system. These

terms can be determined from a series of preliminary experiments in the absence of wind

flow as described in Brito (2008).

The vector of aeroelastic forces Fae in Eq. (3.4) includes lift force and moment, which

are related to flutter derivatives in Eqs. (2.6a) and (2.6c), respectively. Since Fae is a linear

function of h and α and their time derivatives (i.e., a linear function of y and y ), Eq. (3.4)

can be rewritten in terms of “aeroelastically modified” or “effective” damping and stiffness

matrices, as in Eq. (3.5b), by eliminating dependence on the force vector and after mass re-

scaling i.e., pre-multiplication by the inverse of the mass matrix (Eq. 3.5a). If this

interpretation of the fluid-structure interaction phenomena is employed, the effective

damping and stiffness matrices include both the contribution of mechanical part and

aeroelastic one. The aeroelastic part is isolated from the mechanical part by repeating the

measurements in the absence and in the presence of the flow at a pre-selected speed U.

The equations of motion for the section model subjected to aeroelastic forces can be

written as

40

y M1Cy M1Ky M1Fae

,

y CMech y K Mechy M1Fae . (3.5a)

Or, after rearranging the terms

y Ceff y K eff y 0. (3.5b)

Equation (3.5b) can be expressed in the state space as

y

y

0 I

K eff Ceff

A

y

y

. (3.6)

The A matrix contains the aeroelastically modified effective damping and stiffness

matrices, Ceff and Keff, respectively. I is the identity matrix of size 2 × 2. The A matrix can

be identified if displacement data are recorded and accelerations and velocities estimated by

numerical differentiation; the records should include all n=2 degrees of freedom for at least

2n different instants of time (Ibrahim and Mikulcik 1977).

In accordance with the ILS Method, tests are conducted at a given flow speed U to

extract the aeroelastically modified effective damping and stiffness matrices, Ceff and Keff.

Tests are also repeated at U=0 for replicating free vibration without air flow to extract the

mechanical matrices Kmech, Cmech. The flutter derivatives can be determined from the

41

difference (Keff – Kmech) and (Ceff – Cmech), as indicated in (Chowdury and Sarkar 2003, Brito

2008). For example, H1* and A3

* are given as (Chowdury and Sarkar, 2003)

*1 11 112

2,eff mechhm

H K C CB

(3.7)

*3 22 224 2

2.eff mechI

A K K KB

(3.8)

3.6 Reason for the Use of the Polynomial Model in the Context of Random

Flutter Derivatives

In the presence of measurement errors, flutter derivatives Hi*(K) and Ai

*(K) become random

functions in terms of K = (2π)/UR (or, equivalently, reduced velocity). The results of the

experiments (later discussed) suggested that the variance of the derivatives, experimentally

estimated by repeating measurements at the same reduced velocity in the wind tunnel (i.e.,

through “sample averaging”), can be a function of reduced wind speed. Therefore, it is

plausible to also conclude that the probability distribution of each Hi*(K) and Ai

*(K) may be

different at the various K (or reduced velocities).

Since the evaluation of PSD matrix of the generalized buffeting response is based on

Eq. (2.36), i.e., 11 *( ) ( ) ( )b bQ QK K K K

S E S E , a random set of flutter derivatives

would correspond to a stochastic matrix E (matrix of the aerodynamically-modified

frequency response “functions”), the statistical properties of which vary with K as a result

of the random Hi*(K) and Ai

*(K) being dependent on K. The coefficients Hi*(K) and Ai

*(K)

can in fact be seen as “non-uniform” random variables in terms of K. The random properties

42

of E lead to a stochastic PSD matrix of the generalized response Sξξ, “non-uniform” in terms

of K, which would be computationally very challenging to replicate in a Monte-Carlo setting

since the random properties of each Hi*(K) and Ai

*(K) (mean, variance and approximate

probability distribution) would be needed K by K. Needless to say, even a non-numerical

approach, for example by seeking an analytic solution of the stochastic problem by

expansion method about a “mean solution case”, would be very difficult to pursue since the

statistical properties would be needed to be specified (and estimated experimentally) at all

K values with acceptable fidelity.

For these reasons, the use of the polynomial model was preferred since it reduces the

complexity of the stochastic problem by “condensing” the evaluation of the uncertainty in

the flutter derivatives (dependent on K) into the randomness of the coefficients of the

polynomial model, which are treated independently of K in the proposed model (Cj and Cj+1,

Dj and Dj+1 in Eqs. 3.1 and 3.2). The coefficients of this polynomial therefore random

variables, whose probability distribution is not conditional on the reduced wind speed by

neglecting the dependency on K (or reduced velocity). The advantage is that, for analysis

purposes, the first- and second-order statistics of the coefficients can be estimated from the

combination of all experiments, treating all the wind speed data in the wind tunnel (or

equivalently, the K values) as part of the same population. This fact corresponds to an

increment in the size of the population, used to estimate the first- and second-order statistics

from experiments, which also leads to a better confidence on the estimates of the statistics

from the experiments.

43

3.6.1 Estimation of Variance and Co-variance of Cj and Dj coefficients of the “Model

Cures” from Experiments

Second order statistical moments of flutter derivatives (FDs) can be estimated from a sample

population, obtained by repeating the measurements at various wind speed U and the

identification process in Section 3.5. The variances and co-variances of each random

variable pair Cj and Cj+1, Dj and Dj+1 of the model curves in Eqs. (3.1) and (3.2) were

indirectly calculated from the second order statistical moments of the FDs, using the

experimental data as in Fig. 3.5. As an example, the following equation is valid for the k-th

experimental point of flutter derivative H1,k* at a reduced velocity UR,k:

* 2 2 4 2 2 31, 1 , 2 , , 1 22 .k R k R k R kE H E C U E C U U E C C (3.9)

In Eq. (3.9) the index k = 1,…,N is related to a measurement at a given UR,k. Since

measurements are repeated at the same (or very close) reduced velocity twenty times, it is

possible that UR,K ≅ UR,K’. The total second moments of C1 and C2 were estimated by

regression of Eq. (3.9), as explained below:

E H1,1*2

E H1,N*2

b

UR,14 U

R,12 2U

R,13

UR,N4 U

R,N2 2U

R,N3

R

E C12

E C22

E C1C

2

z

. (3.10)

44

The size of the b vector is N, the total number of measurements. The unknown vector

of the moments z (3×1) was calculated from Eq. (3.10) as z = R†·b, with R† being the Moore-

Penrose pseudo-inverse of R (i.e., by least squares).

Once the quantities in z are determined from Eq. (3.10), estimation of the variances

and co-variances of C1 and C2 (mean-removed) is subsequently enabled from the flutter

derivative data of H1*. Similarly, other statistical moments for Hi

* and Ai* were determined;

specific equations are similar to Eqs. (3.7) and (3.8) but are omitted for brevity.

The interval of reduced velocities 5 20RU , covered by experiments, was

considered as acceptable. A probabilistic model for Hi* and Ai

* (i=1,…,4) (Fig. 3.5) was

therefore obtained, as described in Eq. (3.1) for Hi*. The procedure for extracting the total

statistical moments is

At a given UR (or UR,k), the eight flutter derivatives (H1,k*,…,H4,k

* and A1,k*,…,A4,k

*)

were measured in the wind tunnel by repeating experiments and identification

procedure multiple times (Brito and Caracoglia 2009); after collection of the data

and results, flutter derivatives were treated as independent random variables;

The propagation of uncertainty was simulated by treating each of the coefficients in

the polynomial expansion in Eq. (3.1) and Eq. (3.2) as random parameters;

A total of the sixteen random coefficients, Cj and Cj+1, Dj and Dj+1, were considered;

Mutual dependency between the model curve parameters of each flutter derivative

(e.g., C1 and C2 of H1*) was considered, as described in Eq. (3.9);

45

Expectations and second-order moments of the sixteen random coefficients were

derived from NEU experiments.

It was discovered that a jointly gamma distribution of each pair of dependent random

variables Cj, Cj+1 and Dj, Dj+1 was a suitable model to describe the uncertainty found in the

experiments of the truss-type deck section model. This assumption was used in the reminder

of this study. It must be noted that this selection may be data-driven (e.g., two-type model)

and influenced by the type of measurement errors. Other investigations (Bartoli et al. 2009)

found that a Gaussian model was acceptable to describe uncertainty in the FD. Nevertheless,

the gamma type model were preferable since the probability distributions are sign

independent (one-sided) and are therefore more suitable to describe the variability in the Cj,

Cj+1 and Dj, Dj+1 due to uncertainty; a change of sign would result in a drastic change of

curvature or discrepancy of sign in the polynomial model curve, which would be physically

inconsistent with the actual definition of the FDs. This remark can also be clarified by

recalling the interpretation of a flutter derivative as an added “stiffness” or “damping” effect,

induced by fluid-structure interaction. Experimental errors induce a modification to this

“effect” by maintaining a general trend in the behavior of the curves (e.g., the negative

curvature in H1* related to a positive damping effect as UR increases, etc.), whereas a sign

reversal or curvature change would be incompatible with the phenomenon.

46

3.7 Summary of Experimental Results and Comparison with Literature Data

(“Jain’s Data”)

The first and second order statistics of the coefficients of the “model curve” were estimated

by repeating the tests twenty times (Brito and Caracoglia 2009). Experimental data for

H1*,…,H4

*, A1*,…,A4

* vs. reduced velocity UR = U/(nB) are shown in Fig. 3.5 (solid lines

with markers); H1* and H3

* are related to the aeroelastic lift associated with changes in

velocity of the vertical DOF ( h ) and the angular displacement (), respectively; the rest of

flutter derivatives correspond to the aeroelastic torque and depend on changes of h and

angular DOF velocity ( ) (Eq. 2.6).

Flutter derivatives provided from literature were also used in this study, which will

be utilized in the estimation of buffeting analysis in comparison with the results with using

FDs obtained from NEU. Flutter derivatives, Hi* and Ai

* along with the static coefficients of

lift and moment at α0 = 0° were derived from (Jain et al. 1998); these are assumed as more

reliable data and shown in Fig. 3.6 as a function of reduced wind velocity UR = U/(nB) with

n = ω/2π.

Static force coefficients CD, CL and CM and their derivatives are also required for

buffeting analysis (Chapter 2); the decision was to use the static coefficients measured by

Jain (1998) at α0 with both sets of flutter derivative data. The static coefficients and their

derivatives at α0 are shown in Table 3.1.

The results of the NEU’s experiment are very promising since the use of the “model

curves” and the repetition of the tests enabled the characterization of experimental variability

and the analysis of the second-order statistics of flutter derivatives (Fig. 3.5 for the examined

47

section-model of a truss-type deck). The figure also suggests a nonlinear dependence of the

error variance on reduced velocity (wind speed), which was not noted by other investigators.

Also, the experimental procedure offers an example of systematic examination and

quantification of the variances for most flutter derivatives in a simple way, based on Eqs.

(3.9) and (3.10). In spite of the results in Fig. 3.5, the estimation and quantification of such

errors is still an open question. In fact very few examples of error estimation are available

in the literature; these examples acknowledge this relevance even though they are very

limited and have often considered unrealistically simple deck shapes (Sarkar et al., 2009).

48

Table 3.1 The static coefficients and their derivatives at α0 (Jain et al., 1998).

C D 0.3042

C L 0.2113

C M 0.0044

dC D /d α 0 0.0000

dC L /d α 0 3.2487

dC M /d α 0 -0.0177

Static Coefficients

49

Figure 3.1 NEU-MIE wind tunnel (Brito 2008).

50

(a)

(b)

Figure 3.2 Experimental setup: (a) NEU’s small-scale wind tunnel; (b) NEU’s Aeroelastic

Force Balance with the truss-type bridge deck model.

51

Figure 3.3 The Golden Gate Bridge (Photo courtesy of Google Image).

52

Figure 3.4 Truss-type deck section model, replicated the features of the Golden Gate

Bridge at a scale 1:360; model width is B = 76 mm and the aspect ratio is B/D = 3.5:1.

D

53

(a)

(b)

0 5 10 15 20 25-14

-12

-10

-8

-6

-4

-2

0

U/(nB)

H1*

Experimental dataPolynomial model

0 5 10 15 20 25-14

-12

-10

-8

-6

-4

-2

0

U/(nB)

H2*

Experimental dataPolynomial model

54

(c)

(d)

0 5 10 15 20 25-8

-6

-4

-2

0

2

4

6

U/(nB)

H3*

Experimental dataPolynomial model

0 5 10 15 20 25-12

-10

-8

-6

-4

-2

0

2

U/(nB)

H4*

Experimental dataPolynomial model

55

(e)

(f)

0 5 10 15 20-4

-2

0

2

4

6

U/(nB)

A1*

Experimental dataPolynomial model

0 5 10 15 20-10

-8

-6

-4

-2

0

U/(nB)

A2*

Experimental dataPolynomial model

56

(g)

(h)

Figure 3.5 Flutter derivatives of a truss-type section model with aspect ratio B/D= 3.5:1

measured at NEU: (a) H1*; (b) H2

*; (c) H3*; (d) H4

*; (e) A1*; (f) A2

*; (g) A3*; (h) A4

*.

0 5 10 15 20-6

-4

-2

0

2

4

U/(nB)

A3*

Experimental dataPolynomial model

0 5 10 15 20-4

-2

0

2

4

6

U/(nB)

A4*

Experimental dataPolynomial model

57

(a)

(b)

Figure 3.6 Flutter derivatives of a truss-type section model(the Golden Gate Birdge)

derived from (Jain et al., 1998): (a) heave Hi* (i=1,…,4); (b) torsion Ai

* (i=1,…,4).

-10

-5

0

5

10

15

20

25

0 4 8 12 16

Flu

tter

Der

ivat

ive,

Hea

ve

U/(nB)

G.Gate H1*

G.Gate H2*

G.Gate H3*

G.Gate H4*

-0.4

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 2 4 6 8 10

Flu

tter

Der

ivat

ive,

Tor

sion

U/(nB)

G.Gate A1*

G.Gate A2*

G.Gate A3*

G.Gate A4*

58

Chapter 4

A Methodology for the Analysis of Long-Span Bridge

Buffeting Response, accounting for Variability in Flutter

Derivatives

4.1 Introduction

This chapter describes the development of a methodology for predicting the buffeting

response of a long-span bridge by Monte Carlo (MC) methods (Robert and Casella 2004;

Tempo et al. 2005). In the standard buffeting analysis (labeled as the “deterministic case” in

this work) the result is the value of the RMS dynamic response at a given wind speed. In the

proposed probabilistic setting (labeled as “statistical case” in this work) one estimates the

probability that a given threshold for the variance of the response is exceeded.

A Monte-Carlo-based methodology is proposed and implemented for predicting the

buffeting bridge response and for evaluating the variability due to uncertainty in the FDs

(“statistical buffeting” analysis). To accomplish these tasks, a second order polynomial

model (“model curve”) for the FD is utilized. The model curve is a second order polynomial

59

description of the FDs where uncertainty is associated with coefficient of the polynomial.

The coefficients of this polynomial are treated as random variables, whose probability

distribution is conditional on reduced wind speed. For computational reasons in subsequent

analysis, however, this dependency is neglected and the probability of these random

variables is treated as independent of the reduced wind speed. For analysis purposes the first

and second order statistics are estimated from experiments, treating all the wind speed data

as part of the same population.

A numerical procedure for multi-mode buffeting response (“deterministic case”) was

initially developed and its accuracy was validated by comparing with more reliable data from

the literature. In the standard multi-mode buffeting analysis, the power spectral density (PSD)

of the buffeting loads needs to be computed. This step, carried out by numerical integration,

is usually the “bottleneck” of the multi-mode buffeting analysis method in the modal space.

MC and Quasi-Monte-Carlo (QMC) methods were used to numerically compute the PSD of

the buffeting loads and to derive the root-mean-square (RMS) dynamic response of a long-

span bridge. A benchmark structure was utilized for this purpose (1,200 m suspension

bridge). A validation study was carried out by examining the performance of MC integration

methods using a series of standard buffeting analyses on the 1,200m bridge (Golden Gate

Bridge) and comparing the CPU time on a standard computer; also, the validation was

employed to determine the “optimal” number of sample points required by the MC and QMC

procedures in comparison with standard techniques for numerical integration. This part was

also based on a series of preliminary investigations (Caracoglia and Velazquez 2007).

60

Finally, the complete procedure (labeled as “MC-based methodology” in this

dissertation), which includes the probabilistic setting, is presented and implemented to

numerically evaluate for the probability of exceeding a set of pre-selected serviceability

thresholds, selected according to the RMS response of the deck as a function of mean wind

velocity and mean incident angle (skew wind). Such curves or surfaces were again derived

by MC sampling for two simulation examples, based on the same suspension bridge structure

(introduced above) and both literature data and measurements at NEU.

The MATLAB software environment was employed for coding purposes. The

detailed flowchart of the MC-based methodology for “statistical buffeting” analysis,

including the uncertainty in the FD, is shown in Fig. 4.1 and is described in the following

sub-sections.

4.2 Multi-Mode Buffeting Analysis (“Deterministic Case”)

The closed-form solution for multi-mode buffeting analysis by using the second order

polynomial model for FDs was developed in this work. One of the major problems to apply

the closed-form solution is related to the numerical integration of Eq. (2.24), which may

become computationally demanding and impractical in the context of statistical buffeting.

In order to overcome such limitations, Monte-Carlo and Quasi-Monte-Carlo methods

(Robert and Casella 2004; Tempo et al. 2005) were introduced and employed for numerical

integration.

61

4.2.1 Validation for Closed-Form Solution

The benchmark structure was derived from the dynamic behavior of a suspension bridge

with main span l = 1200 m, modeled after the Golden Gate Bridge in San Francisco,

California (USA). This bridge has a deck width B = 28 m, deck torsional inertia I0 = 4.4×106

kg×m2/m (Jain et al. 1996). Modal structural damping ratio was selected as equal to 0.3%

for all modes as a deterministic constant by following the recommendations by Jain et al.

(1998) for this bridge. Lateral modes and responses were not analyzed, since the dynamic

bridge response for this bridge is mainly controlled by heaving and torsional motions (e.g.,

Jain et al, 1996). Flutter derivatives, Hi* and Ai

* along with the static coefficients of lift and

moment at α0 = 0° were reproduced from (Jain et al. 1998); the derivatives are shown in Fig.

3.6.

“Closed-form” (CF) estimation, which employs a standard integration algorithm (i.e.,

trapezoidal rule) for calculating the double integral in Eq. (2.24) to calculate the RMS

response based on standard multi-mode analysis, was employed for examination of the

computational efficiency of MC and QMC algorithms. The simulations included up to 10

fundamental structural deck modes of the benchmark structure. The natural frequencies and

mode types are shown in Table 4.1. Simplified (sinusoidal-like) mode shapes were assumed

and used to describe the main bridge motion (deck and cable vibration) shown in Fig. 4.3.

The CF solution was initially validated by comparison with the literature results,

provided by (Jain 1996) with 7 modes only. The RMS vibration was calculated for both

vertical σhh (Eq. 2.40a) and torsional σαα (Eq. 2.40c) responses at a quarter span of the bridge

deck x = l/4 and for wind speed at U = 22.2 m/s shown in Table 4.2. This bridge section was

62

selected since it corresponds to the anti-mode of the mode shapes in both the fundamental

vertical and torsional modes.

The vertical RMS response of the CF solution was in a good agreement with the

literature one, since the difference is equal to 4.1% in Table 4.2. Although the difference for

an overall torsional RMS displacement was quite large, equal to 182.1% in relative terms,

the result was accepted due to the fact that this value corresponds to an actual small rotation

in terms of degrees (i.e., angles 0.12° and 0.34° in Table 4.2) at full scale.

4.2.2 Monte-Carlo and Quasi-Monte-Carlo Methods

Monte-Carlo (MC) methods are a class of computational algorithms that rely on repeated

random sampling to compute their results. The MC numerical algorithm was employed to

compute the PSD of the buffeting loads Eq. (2.24) by integration and, later, to derive the

RMS response of the bridge Eq. (2.40) (Smith and Caracoglia 2011).

One of the major problems related to the numerical integration of Eq. (2.24), is the

fact that the integrand function needed to estimate Shh, Spp, Sαα in Eqs. (2.37–2.39) must be

evaluated numerically over a large portion of (xA, xB) for low frequency K. For low K the

whole (xA, xB) space contributes to the ( )i j

Q QS K function, whereas for moderate or large

frequencies the “non-zero part” of the integrand function concerns a tight zone, located along

the main diagonal in the (xA, xB) plane. This is a consequence of the decrease of coherence

with distance and frequency in the standard multi-mode formulation. Therefore, the

assessment of the integral for low-frequency structures becomes numerically demanding

since the resonant part of the loading is usually concentrated at low K.

63

In the MC-based approach for assessing the double integral in Eq. (2.24), two

samples of uniformly-distributed independent random variables are generated within an

interval of x, based on actual integration limits in Eq. (2.24) (0 ≤ x ≤ l) shown in Fig. 4.2(a).

These two sets of independent randomly-generated “points” (Robert and Casella 2004) are

used to find an approximation to the generic surface integral in Eq. (2.24). The use of Quasi-

Monte-Carlo (QMC) method was also investigated, as a “more deterministic” version of MC,

which is often preferred for reducing the variance of the estimated quantities and improving

the numerical efficiency (Tempo et al. 2005). In the QMC method the two-dimensional

coordinates of the “points” were chosen according to a deterministic criterion instead of a

“random selection” to distribute the integration points more evenly. The Halton sequence

(Tempo et al. 2005) was used in the QMC algorithm shown in Fig. 4.2(b).

An alternative way to reduce computational integration time would be possible if a

decomposition of the integral in Eq. (2.24) into “resonant” and “background” responses

(Davenport 1967) was used. This technique has been originally proposed in (Scanlan and

Jones 1990a) for long-span bridge aeroelastic analysis, providing approximate but still

acceptable results. Recently, this approach was re-proposed and used in a more generalized

form for bridge simulations (Denoël 2009; Gu and Zhou 2009), in an attempt to avoid

numerical integration issues. Nevertheless, the full integration is always more precise; it is

also preferable for the statistical buffeting analysis because accurate estimation of the

probability distribution of the error-contaminated RMS is crucial. Therefore, numerical

integration using MC and QMC methods has been chosen in this study.

64

4.2.3 Examination of the Computational Efficiency of the MC and QMC Methods for

Calculating the Double Integral in Eq. (2.24)

The accuracy and computational efficiency of the MC and QMC algorithms for double-

integration was assessed by repeating the estimation of the integral in Eq. (2.24) in the

absence of uncertainty in FDs (“deterministic case”). Simulations were repeated for various

MC integration points (NMC) to select an “optimal” value of NMC without compromising the

algorithmic accuracy. Two-mode analysis for the bridge example, simulating the structural

characteristics of existing bridge, was considered in this second investigation, restricted to

the first vertical and torsional skew-symmetric modes with frequencies nv1 = 0.087 Hz and

nt1 = 0.192 Hz (Mode n.2 in Fig. 4.3b and Mode n.8 in Fig. 4.3h).

Scatter plots, corresponding to 100 MC-based repeated estimates of the vertical and

torsional RMS responses as a function of NMC, were numerically evaluated. The performance

of the MC algorithm was also compared against the average relative run time (RT),

normalized to the value obtained for NMC = 50,000. The Coefficient of Variation (CoV) was

calculated and later employed as an indicator in the selection of the “optimal” NMC. The

maximum relative and bias errors (against the CF target value) were also calculated using

both MC and QMC methods.

Figure 4.3 shows the scatter plots corresponding to 100 repeated MC simulations

with the average RT also being indicated in the plots. The vertical (σhh in Fig. 4.4a) and

torsional (σαα in Fig. 4.4b) RMS responses of the deck of the simulated bridge for U = 22.2

m/s were compared to the target value at x = l/4.

In Fig. 4.5 the CoV, estimated for each sample population in Fig. 4.4 and

corresponding to each value of NMC, is shown in log-log scale. The CoV is a normalized

65

measure of dispersion in a distributed sample and was assessed as the ratio between the

standard deviation, σy,target, and mean of the sample, Yσ,mean.

As shown in Fig. 4.5(a), the CoV was very small for σhh, less than 1.4% in all cases.

The CoV for σαα was generally higher but very small for NMC > 10,000, less than 2.0%. In

Fig. 4.5(a) and Fig. 4.5(b), the decrement in the CoV was proportional to NMC on the log-log

chart with a constant negative slope in the curves; this behavior corresponds to a power-law

decrement for increasing NMC on a linear scale.

In Table 4.3 the maximum relative and bias errors for MC, estimated for each sample

population in Fig. 4.4 and corresponding to each value of NMC, are shown. The bias error

was estimated as the difference between the CF value of the RMS response and the mean of

the sample, σy,target - Yσ,mean. The relative error is the bias error divided by the mean of the

sample, σy,target / Yσ,mean. The relative errors were less than 0.7% for σhh at l/4 and -1.9% for

σαα, shown in Table 4.4, while they were almost negligible for larger sample sizes.

Tolerance intervals were employed to analyze the “fidelity” of the numerical

algorithms, as shown in Figs. 4.6 and 4.7. These tolerance limits were approximately

assessed assuming that the sample of RMS values follows a Gaussian distribution. In the

cause of a variable with Gaussian distribution and known population mean, μ, and standard

deviation, σ, the following tolerance limits can be used; μ ± z*σ. If z* = 1.96, these bounds

(covering the middle 95% of the population of observations) are essentially the confidence

intervals for a fixed proportion of the measurements (Walpole et al. 2002). When μ and σ are

unknown, which is typically the case, Eq. (4.1) can be applied.

(4.1) ,mean , target .yY k

66

The quantity k is the tolerance factor for a normal distribution. In this study, k is

defined such that there is a 99% confidence that the calculated tolerance limits will contain

at least 95% of the measurements, k = 2.36. The limiting confidence interval (e.g., 99%)

must be added to the statement since the bounds given by Eq. (4.1) cannot be expected to

contain any specified proportion (e.g., 95%) all of the time (Walpole et al. 2002). The

assumption of a normally-distributed population of σh and σα was used. Since the objective

was the section of an optimal sample size this hypothesis was accepted.

The tolerance interval results of are shown in Figs. 4.6 and 4.7. In both figures, the

horizontal axis represents the number of simulations conducted (i.e., 100 simulations) for a

given NMC, and the vertical axis is the RMS response. In each plot the tolerance intervals,

the target value (from CF simulations) and the results for each of the 100 simulations are

presented. All figures show a correlation between the number of integration points, NMC, and

the tolerance intervals, since the tolerance interval decreases as NMC increases. The tolerance

interval is approximately 9 times larger for NMC = 1,000 than it is for NMC = 100,000. The

benefit of using more integration points is clearly shown by the results of this investigation.

However, a smaller tolerance interval should not be the only criterion employed for the

selection of the optimal NMC. Figures 4.6 and 4.7 also suggest that, as the tolerance interval

decreases, the bias error can increase as the difference between the CF target value and the

estimated mean grows. Additionally, in Figs. 4.6(d) and 4.7(e) the tolerance limits no longer

contain the CF target value. This aspect must also be taken into consideration since the

overall goal of the MC simulations is to obtain results close to the CF “exact” value.

67

Figure 4.8 depicts the vertical (σhh in Fig. 4.8a) and torsional (σαα in Fig. 4.8b) RMS

responses of the same bridge example, using by QMC integration. Scatter plots and CoV of

σhh at l/4 correspond to 100 repeated simulations. The average RT was normalized to the

duration of the MC simulation with NMC = 50,000 for comparison with Fig. 4.4.

The CoV values in Fig. 4.9 (log-log scale) are much smaller than the ones observed

with the MC algorithm for both vertical and torsional response; however, the direct

proportionality on the log-log chart with a constant negative slope for all NQMC was not

observed. For example, the CoV for NQMC = 5,000 is less than 0.2%, which is clearly smaller

than approximately 0.5%, noticeable in Fig. 4.9(a) with NMC = 100,000. However, RT is

equal to 5.05, which is three times larger than the MC case with NMC = 100,000 (RT = 2.26).

The maximum relative and bias errors for QMC (Table 4.5) are shown. The relative

errors in the QMC case were less than 0.9% for σhh at l/4 and -3% for σαα, shown in Table

4.4. The relative and bias errors in the QMC case were larger than the ones in the MC case,

even though the QMC case had smaller CoVs. Tolerance intervals were also investigated

and are presented in Figs. 4.10 and 4.11.

From the interpretation of Fig. 4.4 through Fig. 4.11 it was concluded that QMC

integration, even with a medium sample size (NQMC > 10,000), was impractical because of

large RT. The QMC case with NQMC = 1,000 did not improve the numerical performance of

the standard algorithm in terms of CoV, with the RT being approximately eleven times larger

than the reference case with NMC = 1,000.

In the MC case the relative error of σhh consistently increases with an increment in

NMC; for the torsional response, the relative error also varies with NMC. The relationship

68

between the bias and the CoV was also utilized as an additional criterion for the selection of

the algorithm (MC or QMC). The QMC estimation showed smaller maximum relative errors

but higher bias errors (shown in Table 4.4).

As a consequence of the above observations, the MC integration algorithm was

selected in the subsequent stages of this study. By combining the results of Fig. 4.4 through

Fig. 4.7, NMC = 5,000 was recommended as the preferable choice for MC due to both

relatively small variance and good numerical efficiency. The computational time is

approximately one order of magnitude smaller than that of using a standard integration

algorithm (repeated trapezoidal rule).

4.3 Monte-Carlo-based Methodology for Buffeting Analysis Considering

Uncertainty in the Flutter Derivative (“Statistical Case”)

A Monte-Carlo-based methodology is proposed and implemented for predicting the

buffeting bridge response and for evaluating the variability due to uncertainty in the FDs

simulated by MC sampling. To accomplish these tasks, the second order polynomial model

(“model curve” in Eq. 3.1 and 3.2) for the FD is utilized, in which the coefficients of the

polynomial are random variables; more details can be found in Section 3.6.1.

In the MC-based methodology for buffeting analysis, the numerical procedure re-

calculates, at various wind speeds and skew wind angles, the buffeting loads by MC

sampling for each of the 5,000 realizations (e.g., the flowchart presented in Fig. 4.1). The

generalized power spectral density (PSD) of the buffeting loads is also calculated by MC

sampling, where a double integration (Eq. 2.24) is needed. Even though it is not necessary

to repeat the double integration for calculating the PSD of the buffeting loads, it is still

69

necessary to identify an efficient numerical procedure (Section 4.2.3) in order to be able to

generalize the method for future applications. In fact, extension of the method has been

investigated, in which also the effects of errors in the buffeting part of the loading must be

accounted for. These errors include, but are not limited to, simplifications in the modeling

of the wind turbulence spectrum, errors in the estimation of the span-wise loading parameters

Suu and Sww in Eq. (2.24). For more information on the effects of this category of errors, the

reader may refer to the recent publications in this area (Caracoglia 2008a; 2008b; 2011). In

the context of a future generalization of the method, in which the power spectral density of

the buffeting loading can become a random function in terms of K, repetition of the buffeting

loading estimation is required. Therefore, an efficient numerical procedure was needed.

In the proposed probabilistic setting (“statistical buffeting case”) one estimates the

probability that a given threshold for the variance of the response is exceeded. There are two

obvious formats to display the information that are useful in different ways. One way is to

plot the RMS value of buffeting response at a given confidence level of not being exceeded.

For example, Fig. 4.12 shows the RMS value for vertical buffeting response as a function of

wind speed at a given confidence level. More directly useful way for our purpose is to plot

the probability of exceedance at a given fixed RMS value as a function of wind speed. This

probability is designated as “threshold exceedance probability” (TEP) in this work, derived

by using a MC-based methodology with 5,000 sampling points (e.g., the flowchart presented

in Fig. 4.1) which projects the variability in the FD into the estimation of buffeting response.

The concept of using TEP was adopted from seismic engineering field (i.e., “fragility”).

70

Fragility analysis is a standardized methodology, utilized for performance-based

structural design. As a general statement, fragility curves measure (or quantify) the overall

structural vulnerability (Norton et al. 2008). The likelihood of structural damage due to

different “demand levels” – mean wind velocity levels in the case of wind engineering – is

usually expressed by a fragility curve (Saxena et al. 2000). A collection of these curves

describes the (conditional) probability of exceedance of representative structural response

indicators (“structural capacity”), corresponding to a specific feature of the dynamic

response at a given wind velocity (Bashor and Kareem 2007; Ellingwood 2000; Filliben et

al. 2002). A set of thresholds is usually selected to represent different levels of structural

performance derived from such indicators. As an example, in the case of a building these

indicators are either required or are prescribed by the designer, and can include inter-story

drift ratios, maximum lateral drift, and acceleration levels for occupant comfort (Bashor and

Kareem 2007; Filliben et al. 2002; Smith and Caracoglia 2011).

The TEP curves were developed in this dissertation by numerically deriving the

histogram of occurrences and the subsequent probability density function (PDF) of the RMS

dynamic response by a recursive procedure. The RMS response is a random variable in the

“probabilistic setting”. The probability of exceedance of pre-selected thresholds was later

calculated.

Equation 4.2 below relates the “loss of performance” of the structure, a bridge in this

study, which is the probability of exceeding of a threshold T, associated with the dynamic

response feature Y (performance indicator) at a given deck section for wind incidence angle

(θ) orthogonal to the longitudinal axis (initially assuming this response independent of wind

71

direction) as an exclusive function of mean wind velocity at deck level U (e.g., (Ellingwood

et al. 2004; Filiben et al. 2002)).

0

| .T T UP F Y T U u f u du

(4.2)

In Eq. (4.2), the conditional probability function, denoted as FT[], is the “TEP” for

threshold T; fU is the probability density function (PDF) of the mean wind velocity, which

can be derived from site wind data under the conservative assumption of constant wind

direction, always orthogonal to the longitudinal deck axis.

In this section the mean wind velocity at deck level was assumed as being always

perpendicular to the bridge longitudinal axis, as this direction usually corresponds to the

most unfavorable condition. The combined influence of mean wind speed (U) and incidence

angle (θ) is discussed in Chapter 4.5. These were derived after numerically assessing the

probability distribution of the RMS dynamic response at representative wind velocities (U).

In summary, the numerical procedure, which utilizes this methodology, combines the

estimation of RMS response via MC integration with “brute-force” uncertainty simulation

due to flutter derivative errors to estimate the RMS response and TEP curves. The complete

flow chart of the procedure with more details is shown in Fig. 4.1.

The RMS vertical and torsional response were utilized as an example, noting that the

RMS response can be directly related to the peak displacement through gust effect factor

(e.g., (Scanlan and Jones 1990a)) for serviceability analysis due to stationary winds (also

refer to Chapter 5 for discussion on peak response).

72

4.3.1 Description of the Bridge Example and RMS Threshold Levels (“Probabilistic

Setting”)

One bridge example was selected for “threshold exceedance probability” (TEP) analysis, the

same bridge model discussed in Section 4.2. Four-mode buffeting analysis was carried out

by considering the first two vertical (v1 and v2) and torsional modes (t1 and t2); it has

frequencies nv1 = 0.087 Hz and nv2 = 0.129 Hz, nt1 = 0.192 Hz and nt2 = 0.197 Hz. Simplified

(sinusoidal-like) mode shapes, shown in Fig 4.2(b, d, g and h) were used to describe the main

bridge motion (deck and cable vibration) in both models.

For example, as shown in Fig 4.2, the mode shapes of the vertical modes v1 and v2

were assumed as purely flexural, with shapes hv1(x) = sin(2πx/l) and hv2(x) = sin(πx/l), in

which l = 1,263 m is the central-span length for this bridge. Similarly, the shapes of torsional

modes t1 and t2 for the same model were simulated as purely torsional with αt1(x) = sin(2πx/l)

and αt2(x) = sin(πx/l).

The second order polynomial model (“model curve”) for flutter derivatives, proposed

in Section 3.4 to approximately account for effects of measurement errors in the FD, was

employed in the “statistical buffeting” analysis. Since eight flutter derivatives (H1*,…,H4

*

and A1*,…,A4

*) are measured in wind tunnel as a function of the reduced speed, UR = U/(nB)

= 2π/K, experimental data are usually available at discrete points on the UR (or K) axis. In

the TEP analysis, the coefficients of the model curves for each flutter derivative were

randomly perturbed to simulate the uncertainty in the FD; Hi*

= CiUR2 + Ci+1UR in Eq. (3.1)

with i=1,3,5,7 and Ai* = DjUR

2 + Dj+1UR in Eq. (3.2) with j=1,3,5,7.

The parameters of Hi* and Ai

* were assumed as a set of uncorrelated gamma-type

random variables. Description on the selection of this specific probability distribution may

73

be found in Section 3.4. The hypothesis of uncorrelated random variables is also described

in Section 3.4 The dispersion and shape parameters of the marginal probability of each

variable were associated with mean and standard deviation estimates of Hi* and Ai

*.

Three threshold levels were selected to derive the TEP curves. Table 4.5 shows the

threshold levels employed in the TEP analysis for σhh and σαα. Thresholds were based on a

median value of RMS displacements of the 5,000 buffeting analysis at wind velocity equal

to 20 m/s for the deck section at x = l/4, for example with T1 being equal to 50%, T2 equal to

100% and T3 equal to 150% of the corresponding the median value. For example, threshold

T2 corresponds to a dynamic displacement equal to 0.015D (i.e., D is being depth of the deck

equal to 7.83 m).

TEP analysis was carried out for the bridge example using flutter derivatives from

literature (Jain’s FD data) and flutter derivatives measured at Northeastern University (NEU).

4.3.2 TEP Curves using Literature Data

TEP curves, associated with the benchmark bridge model (Golden Gate Bridge) and based

on flutter derivatives reproduced from (Jain et al. 1998), were derived. The statistical

properties of the coefficients of the model curve were synthetically reproduced since no FD

error analysis was available. The coefficients of the Hi* and Ai

* model curves (Eqs. 3.1 and

3.2) were assumed as a set of uncorrelated gamma-type random variables. The dispersion

and shape parameters of the marginal probability of each variable were associated with the

mean and standard deviation estimates of Hi* and Ai.

For example, in Fig. 4.13 the Jain’s flutter derivatives H1* and A2

* are depicted; these

data were reproduced from the experiments described in (Jain 1996; Jain et al. 1996) and

74

from Fig. 3.6. The graphs also show the reference “model curves” based on the discrete data

points, assumed to be “average values” in the analysis. The dotted lines describe the upper

and lower limits of the reference curve that first statistics of the coefficients of the reference

curve was synthetically derived from such limits. Since no error analysis was carried out by

the investigators in their experiments, the variability (uncertainty) was indirectly estimated

in an approximate way, described above.

The TEP curves were calculated for both vertical and torsional vibrations at the

quarter span of the simulated bridge. The thresholds in Fig. 4.14(a) and Fig. 4.14(b) were

selected from Table 4.5. The TEP curves were developed by numerically deriving the

histogram of occurrences for each indicator at each wind velocity U, shown in Fig. 4.14(a)

and Fig. 4.14(b). The deterministic flutter velocity (DFV) without flutter derivative errors,

estimated as equal to 19.7 m/s, is also indicated a vertical dashed line.

In Fig. 4.14, the probability of exceeding T1, T2 or T3 increases as a result of a larger

perturbation (higher RMS response) for higher U. A small decrement in the TEP curves

related to T1 for U > 19.7 m/s in Fig. 4.14(a) may also be explained by potential inaccuracy

in the MC buffeting procedure in the proximity of flutter speed, as described below.

Beyond the deterministic flutter speed, the TEP curves are expressed as dotted lines

due to the onset of dynamic instability beyond a given critical velocity Ucr (deterministic

flutter velocity - DFV). Since the buffeting procedure numerically calculates the RMS

displacements of both vertical and torsional responses by spectral methods, it may fail by

predicting a finite RMS response at U beyond the velocity of the onset of flutter; finite

75

vibration amplitudes and stationary responses are compatible with a post-critical flutter

regime but linear random vibration cannot be used anymore to estimate the response.

In the context of the study of multiple realizations by using a sample random

population within a MC algorithm, this issue may be circumvented by conducting a

preliminary flutter analysis for each simulation before calculating the buffeting response.

Another possibility would be to increase the resolution of the discrete points used to

construct the TEP curves in the regions close to the deterministic flutter threshold for U <

Ucr. In either case, the numerical efficiency of the procedure is very limited also because, in

the context of a MC algorithm, a fully automated computer procedure for flutter analysis and

non-trivial solutions of Eξ=0 is not available.

In this study the Bayes’ Theorem was used as a simple yet efficient way to accurately

capture the buffeting response more realistically, by accounting for the probability of flutter

onset for U values close to Ucr. This approach utilizes the original (or “prior”) estimates of

exceedance probability, calculated by employing the standard MC procedure (i.e., Fig. 4.14)

and performs a simple “rescaling” of the probability. This operation leads to a modified TEP

curve or “posterior” estimate of exceedance probability.

The rescaling of the TEP curves was obtained by applying Bayes’ Theorem, as

described below.

P Y T |U |~ Flutter posterior

P Y T |U prior

P ~ Flutter Y T |U

likelihood

P ~ Flutter marginal likelihood

(4.3)

76

(4.4)

The symbol “~” is used in the previous equations to designate the non-occurrence

event, i.e., for the generic event A, P[NOT(A)] = P[~A] (flutter has not occurred). In Eq.

(4.3), is the modified TEP curve (“posterior probability”);

is the “prior probability”, coincident with the TEP curves before rescaling

which are shown in Figs. 4.14(a) and 4.14(b); is the non-flutter probability

(“marginal likelihood” in terms of Bayes’ Theory) and Ucr is the deterministic flutter velocity.

It must be noted that the calculation of the TEP | |~ FlutterP Y T U should only

include, among all random events in the probability, those events for which flutter has not

occurred; therefore, this probability may still be less than one for U beyond the DFV since

it is based on a sub-set of the cases, for which flutter has still not occurred.

Equation (4.4) describes the non-flutter probability conditional on the exceedance of

the given threshold and can be interpreted in terms of Bayes’ Theory as the likelihood

function. Since the use of the Eq. (4.3) is physically acceptable in the interval of U close to

deterministic flutter speed (15< U < 25 m/s), the validity of Eq. (4.3) is restricted to this U

interval and unacceptable at low U and very large U. This observation is compatible with

the statement that the likelihood function (Eq. 4.4) must be very close to unit value, which

means that flutter is likely to occur is a necessary condition for applicability of the re-scaling.

Therefore, the “true” and “fake” logical statements are used in Eq. (4.4) to designate this

T ru e 1 .0 , if ~ F lu tte r

F a lse 1 .0 , if c r

cr

U UP Y T U

U U

FlutterP Y T U

P Y T U

P Flutter

77

region of validity of Eq. (4.3). The rescaling, based on Eq. (4.3), is acceptable if the statement

in Eq. (4.4) is true and the conditional probability of flutter is 1.0.

In regions with crU U (DFV value) the rescaling is dubious since the assumption

on the hypothesis of a likelihood function close to the unit value is not satisfied; the practical

implication is that TEP greater than one may be observed.

Figure 4.15 depicts the procedure for rescaling based on Eq. (4.3). For example, the

discrete points on the initial curve (“prior” TEP) in Fig. 4.15(a) | ~ FlutterP Y T U ,

after multiplication by the likelihood function in Eq. (4.4), is divided by the marginal

likelihood function (i.e., the discrete points along the non-flutter probability

curve) in Fig. 4.15(b) to finally become Fig. 4.15(c) as a modified curve (“posterior” TEP).

Each curve in Fig. 4.14 was numerically evaluate by using the MC-based procedures and are

specific for the investigated case. The region of approximate validity of Eqs. (4.3-4.4) is

highlighted by a hatched area; it must be noted that the denominator in Eq. (4.3)

is strictly greater than zero in the region of validity (above 40% in Fig. 4.15(b)),

thereby confirming that | ~ FlutterP Y T U is not possible and that the proposed

rescaling is compatible with the assumptions.

The “posterior” TEP curves for T2, derived from Fig. 4.14 after rescaling, are shown

in Fig. 4.16.

4.3.3 TEP Curves using NEU’s Flutter Derivative Data

In order to further examine the performance of the MC-based methodology for TEP analysis,

flutter derivatives measured in the NEU’s wind tunnel were used. Eight flutter derivatives

P Flutter

P Flutter

78

(H1*,…,H4

* and A1*,…,A4

* in Fig. 3.5) were measured; the description of the wind tunnel

and experiments were presented in Chapter 3. Second-order statistical moments of flutter

derivatives were estimated from the wind tunnel measurements. Details may be found in

Section 3.6.

Structural properties of the model, the Golden Gate Bridge, were combined with

NEU’s experimental data in Fig. 3.5 to obtain a new set of TEP curves, based on threshold

levels in Table 4.5 shown in Fig. 4.17. For comparison purposes the T2 TEP curve, derived

from Jain’s flutter derivative data, was also reproduced from Figs. 4.16(a) and 4.16(b)

(dotted line, after the rescaling based on Eq. (4.3). These curves from Fig. 4.16 are shifted

to the right of the graph for vertical response and shifted to the left for torsion response. This

suggests that the variability in the FD play a significant role in the TEP analysis of this bridge

example and can reduce the dynamic performance, especially for the vertical DOF (Fig.

4.17a), if the FDs from literature were used.

4.4 Effect of Wind Directionality on “Statistical Buffeting” Response: TEP

Surfaces

In nature, the highest winds of record at a given site is very likely to be skew to the bridge

(Scanlan 1999). In this section, the multi-mode approach for wind direction orthogonal to

the deck axis was modified to account for directionality (θ), as originally proposed in

(Scanlan 1993). Background theory is described in Chapter 2. The latter approach enabled

the analysis of the combined influence of wind directionality and velocity, as these should

be more realistically used in serviceability analysis. This observation led to the derivation of

79

TEP surfaces (Filliben et al. 2002; Grigoriu 2002) as a function of various pre-selected

threshold levels.

Figure 4.18 shows the flutter derivative surfaces corresponding to T2 threshold in

Table 4.5 that include wind directionality affecting buffeting response of a bridge deck. In

Fig. 4.19, structural properties of the bridge were combined with NEU’s experimental data

in Fig. 3.5. A skew wind angle, varying from θ = –40° to θ = 40°, was considered in this

analysis.

Some influence on directionality can be observed in Fig. 4.18, especially for σαα.

Figures 4.18 and 4.19 confirm that the skew wind angle equal to zero degrees is the most

conservative assumption for both vertical and torsional buffeting responses.

4.5 Exploratory Performance Analysis on a Full-Scale Structure

In this section, the results in Section 4.4 were applied to the serviceability analysis of the

actual bridge structure, from which the structural and aeroelastic model was originally

derived. The structural performance was derived by extension of Eq. (4.2) to the case of

combined dependency on mean wind speed and direction. The exceedance probability and

“loss of performance” (PT) of the full-scale bridge structure was referred to TEP surfaces in

Fig. 4.18 and threshold T2. Equation (4.2) was re-written as follows:

180

180 0

| , , .T T UP F Y T U u f u dud

(4.5)

80

In Eq. (4.5), the conditional probability function, denoted as FT[], is the “TEP surface”

and fUΘ is the joint probability density function between mean wind velocity (U) and mean

wind incidence angle with respect to the axis of the bridge, which must be defined in the

interval –180° ≤ θ ≤ 180° (θ = 0° orthogonal to the bridge, θ = ±90° parallel to the bridge).

The wind data, used to estimate fU(U) and fθ(θ) given the assumption fUθ = fU·fθ, was

extracted from the historical records of a meteorological buoy close to the coast of California

in San Francisco (USA), which is part of the NOAA (National Oceanic and Atmospheric

Administration) system and National Data Buoy Center (NOAA Station 9414290, Latitude:

37.807 N, Longitude: 122.465 W, Shown in Fig. 4.20). This particular meteorological station

was selected because the prototype full-scale application (Golden Gate Bridge) is located in

proximity of the station. Data are available on-line from http://www.ndbc.noaa.gov/. Wind

speeds and directions at this station are measured using an anemometer located 7.3 meters

above mean sea level.

The histogram of annual maxima (“peak data” shown in Fig. 4.21) was derived from

the NOAA annual wind speed maxima over a 16-year period (1996-2011) and was employed

to evaluate fU(U). The annual wind speed maxima, obtained from the 16-year period, was

fitted to an Extreme Value Type-I distribution with two parameters, i.e., scale equal to 30.0

and location equal to 9.6 (Simiu and Scanlan 1996). The distribution of the mean yaw-wind

angle was based on a non-parametric model for the statistical distribution and derived from

the histogram of the mean wind direction (azimuthal) recorded by the sensor at elevation 7.3

m, postulating little effects of elevation on directionality. The azimuthal direction was later

converted to yaw-wind angle by using actual orientation of the bridge axis.

81

Table 4.6 summarizes the results of the performance analysis and shows the

estimation of the PT for both RMS vertical and torsional responses (dynamic component

only) for the bridge, evaluated at deck section x = l/4. These are respectively labeled as case

1 and case 2 in the table. In both cases the total probabilities of exceeding threshold T2 in

Table 4.5(a) are of the order of 2% by considering both wind speed and directionality.

Overall, the performance of the bridge is good from the point of view of the serviceability

standards. This analysis was completed by estimation of a “generalized safety index” for the

prototype application, defined as β = Ф-1(1 – PT) (Haldar and Mahadevan 2000) with Ф being

the standard normal cumulative density function.

The safety indices are also shown in Table 4.6; these should be interpreted as being

related to serviceability buffeting limit states. Reliability is acceptable, with safety indices

above 2 in both cases. Since effect of chord-wise aerodynamic admittance was not

considered in the model, a further reduction in the buffeting loading may possibly lead to a

decrement in PT; therefore, the results of Table 4.6 can be interpreted as a conservative

estimation of reliability and β.

Reliability against collapse (e.g., flutter) is not contemplated in this chapter, even

though it was indirectly included in the assessment of by Eq. (4.3).

4.6 Summary

The Monte-Carlo-based methodology for predicting the buffeting response of a long-span

bridge, including the uncertainty in the flutter derivative, was developed. The model curve

was utilized in the MC-based methodology; this curve is a second order polynomial

2 | ,TP Y Y U u

82

description of the FDs where uncertainty is associated with coefficient of the polynomial.

Statistical properties of the polynomial in the flutter derivatives were estimated both from

the literature and a set of measurements. Numerical procedures were coded and implemented

in MATLAB, using the methodology described in this chapter; the numerical program was

employed, to derive the dynamic response of two bridge structures at full scale.

The numerical results show that the RMS buffeting response of a long-span cable

supported bridge can be estimated (with sufficient accuracy to be practically useful) using a

second order polynomial description of the flutter derivatives. The uncertainty in these

derivatives in the model is captured by specifying the coefficients of the polynomial as a

vector of random variables having a specified mean and covariance (the values of these

properties used in the numerical analyses were obtained experimentally).

It is found that there is a significant computational advantage in using Monte-Carlo

methods for calculating a double integral that arises in the estimation of the generalized

buffeting loading, needed by the multi-mode buffeting analysis. The computational time is

approximately one order of magnitude smaller than that of using a standard integration

algorithm (repeated application of trapezoidal rule).

The results also suggest that the proposed MC-based methodology for buffeting

analysis might be used in conjunction with other dynamic indicators for analyzing the

serviceability in a full-scale structure (e.g., for comfort of the bridge users).

83

Table 4.1 Natural frequencies and mode types of Golden Gate Bridge (Jain 1996).

Note: L is Lateral, V is Vertical, T is Torsional, S is Symmetric, AS is Anti-symmetric. Note [1]: Modal inertia normalized to one.

Table 4.2 Comparison of closed-form numerical solution with literature results at l/4.

G hihi G pipi G αiαi

1 0.0490 LS 2.63E-16 3.30E-01 8.03E-05

2 0.0870 VAS 3.20E-01 7.65E-15 1.90E-15

3 0.1124 LAS 1.72E-14 3.20E-01 1.24E-02

4 0.1285 VS 1.90E-01 8.28E-14 1.37E-14

5 0.1340 VAS 3.40E-01 5.97E-14 2.66E-14

6 0.1638 VS 3.40E-01 4.00E-13 1.84E-14

7 0.1916 TAS 6.67E-12 3.32E-02 3.20E-01

8 0.1972 TS 2.49E-12 2.50E-01 1.80E-01

9 0.1988 VAS 1.80E-01 4.60E-13 9.61E-12

10 0.2021 VS 2.60E-01 1.49E-15 8.00E-15

NEU(Hz)

ModeNo.

Modetype

Modal Integrals [1]

σ hh (m) 0.13779 0.13235 4.1

σ αα (rad) 0.0060 0.0021 182.1

σαα (deg) 0.34° 0.12°

RMS Response

From Jain(1996)

Closed-Form(CF)

Diff.(%)

84

Table 4.3 Bias and relative errors in the MC case: (a) for heave σhh; (b) for torsion σαα.

(a)

(b)

N MC Bias × 10-4

(σy,target‐Yσ,mean)

Relative Error (%)

1,000 -0.23 -1.645,000 -0.27 -1.8710,000 -0.24 -1.6750,000 -0.24 -1.71100,000 -0.21 -1.49Note: y denotes either h or α

N MC Bias × 10-3

(σy,target‐Yσ,mean)

Relative Error (%)

1,000 0.52 0.395,000 0.62 0.4710,000 0.78 0.5950,000 0.82 0.62100,000 0.89 0.67Note: y denotes either h or α

85

Table 4.4 Bias and relative errors in the QMC case: (a) for heave σhh; (b) for torsion σαα.

(a)

(b)

N QMC Bias × 10-3

(σy,target‐Yσ,mean)

Relative Error (%)

1,000 -0.69 -0.525,000 -0.69 -0.5210,000 -0.82 -0.6250,000 N.A. N.A.100,000 N.A. N.A.

Note: y denotes either h or α

N Q MC Bias × 10-4

(σx,target‐xσ,mean)

Relative Error (%)

1,000 0.36 2.555,000 0.24 1.6610,000 0.24 1.6850,000 N.A. N.A.100,000 N.A. N.A.Note: y denotes either h or α

86

Table 4.5 Threshold values for σhh and σαα, employed in the TEP analysis with flutter

derivatives from the literature.

MD: median value. D: deck depth

Table 4.6 Study cases used for serviceability on full-scale structure.

Vertical (m) Torsional (rad)

T 1 = 0.5 MD 0.008D 4.06×10-4

T 2 = 1.0 MD 0.016D 8.11×10-4

T 3 = 1.5 MD 0.024D 1.22×10-3

ThresholdLabel

Response Threshold

Case Type of Response Wind Velocity and Direction Data(NOAA Station 9414290)

Exceedance Probability

for Threshold T 2, P T

Generalized Safety Index, β

1 RMS of vertical response at l /4

1-year continuous data 0.0126 2.2379

2 RMS of torsionalresponse at l /4

1-year continuous data 0.0088 2.3738

87

Figure 4.1 Flowchart describing the MC-based methodology for buffeting analysis.

F.E.MK, M, C(Deterministic)

Build deck section

Spectrum analysis of buffeting loads, SQQ

Formulate the Equation of Motion in the frequency domain and

incorporate the FD information

Estimate the RMS responseor

Test to obtain Data

Use Data to formulate a model for the FD that includes variability

Estimate the PSDof the response Sξξ

MC sampling of FD simulation

hh

Note: The numerical procedure depicted in the box is repeated for NMC times at various wind speeds and skew wind angles

Select wind speed (U) and skew angle (θ)

Model curves:Hi* = CjUR2+Cj+1URAi* = DjUR2+Dj+1UR

Calculate aeroelastic loads

88

(a)

(b)

Figure 4.2 Two-dimensional sample points = 1,000: (a) MC with uniform distribution, (b)

QMC with Halton sequence.

0 0.2 0.4 0.6 0.8 10

0.2

0.4

0.6

0.8

1

xB

xA

0 0.2 0.4 0.6 0.8 10

0.2

0.4

0.6

0.8

1

xA

xB

89

(a)

(b)

(c)

0 200 400 600 800 1000 12000

0.2

0.4

0.6

0.8

1Mode n. 1

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

1Mode n. 2

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

1Mode n. 3

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

90

(d)

(e)

(f)

0 200 400 600 800 1000 12000

0.2

0.4

0.6

0.8

1Mode n. 4

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

1Mode n. 5

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

Mode n. 6

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

91

(g)

(h)

(i)

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

1Mode n. 7

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

0 200 400 600 800 1000 12000

0.2

0.4

0.6

0.8

1Mode n. 8

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

1Mode n. 9

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

92

(j)

Figure 4.3 Ten simplified (sinusoidal-like) mode shapes used in the multi-mode buffeting

analysis: (a) LS, 0.049 Hz; (b) VAS, 0.087Hz; (c) LAS, 0.112 Hz; (d) VS, 0.129 Hz; (e)

VAS, 0.134 Hz; (f) VS, 0.164 Hz; (g) TAS, 0.192 Hz; (h) TS, 0.197 Hz; (i) VAS, 0.199 Hz;

(j) VS, 0.202 Hz. (Note: L is Lateral, V is Vertical, T is Torsional, S is Symmetric, AS is

Anti-symmetric).

0 200 400 600 800 1000 1200-1

-0.5

0

0.5

1Mode n. 10

Mod

al A

mpl

itud

e

Longitudinal Abscissa

ph

93

(a)

(b)

Figure 4.4 MC-based scatter plots of RMS response for deck section at x= l/4 and for U =

22.2 m/s as a function of NMC: (a) heave σhh; (b) torsion σαα.

0.126

0.128

0.130

0.132

0.134

0.136

0.138

RM

S R

espo

nse

, σhh

(m)

NMC= 1,000NMC = 5,000NMC = 10,000NMC = 50,000NMC = 100,000Target Value from CF (RT: 1.46)

RT: 0.05

RT: 0.13RT: 0.21

RT: 1.00 RT: 2.26

Note: Relative Run Time (RT) to NMC = 50,000

NMC=1,000NMC=5,000NMC=10,000NMC=50,000NMC=100,000

0.0011

0.0012

0.0013

0.0014

0.0015

0.0016

0.0017

RM

S R

espo

nse

, σαα

(rad

)

NMC = 1,000NMC = 5,000NMC = 10,000NMC = 50,000NMC = 100,000Target Value from CF (RT: 1.46)

RT: 0.05

RT: 0.13 RT: 0.21RT: 1.00 RT: 2.26

Note: Relative Run Time (RT) to NMC = 50,000

NMC=1,000NMC=5,000NMC=10,000NMC=50,000NMC=100,000

94

(a)

(b)

Figure 4.5 Coefficient of variation of the RMS response at x = l/4 for U = 22.2 m/s,

computed by MC algorithm: (a) heave σhh; (b) torsion σαα.

0.0%

0.1%

1.0%

10.0%

1,000 10,000 100,000

CoV

, σhh

NMC (No. of MC Integration Points)

0.0%

0.1%

1.0%

10.0%

1,000 10,000 100,000

CoV

, σαα

NMC (No. of MC Integration Points)

95

(a)

(b)

0.126

0.128

0.130

0.132

0.134

0.136

0.138

0.140

0 25 50 75 100

RM

S R

esp

onse

, σhh

(m)

MC Simulation Index

NMC = 1,000 Target Value from CF Tolerance LimitsNMC = 1,000

0.130

0.131

0.132

0.133

0.134

0.135

0.136

0 25 50 75 100

RM

S R

esp

onse

, σhh

(m)

MC Simulation Index

NMC = 5,000 Target Value from CF Tolerance LimitsNMC = 5,000

96

(c)

(d)

0.131

0.132

0.133

0.134

0.135

0.136

0 25 50 75 100

RM

S R

espo

nse,

σhh

(m)

MC Simulation Index

NMC = 10,000 Target Value from CF Tolerance LimitsNMC = 10,000

0.1320

0.1325

0.1330

0.1335

0.1340

0.1345

0 25 50 75 100

RM

S R

esp

onse

, σhh

(m)

MC Simulation Index

NMC = 50,000 Target Value from CF Tolerance LimitsNMC = 50,000

97

(e)

Figure 4.6 Tolerance intervals for vertical RMS response (σhh) of 100 MC simulations: (a)

NMC = 1,000; (b) NMC = 5,000; (c) NMC = 10,000; (d) NMC = 50,000; (e) NMC = 100,000.

0.1320

0.1325

0.1330

0.1335

0.1340

0.1345

0 25 50 75 100

RM

S R

espo

nse,

σhh

(m)

MC Simulation Index

NMC = 100,000 Target Value from CF Tolerance LimitsNMC = 100,000

98

(a)

(b)

0.00110

0.00120

0.00130

0.00140

0.00150

0.00160

0.00170

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

.)

MC Simulation Index

NMC = 1,000 Target Value from CF Tolerance LimitsNMC = 1,000

0.00125

0.00130

0.00135

0.00140

0.00145

0.00150

0.00155

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

)

MC Simulation Index

NMC = 5,000 Target Value from CF Tolerance LimitsNMC = 5,000

99

(c)

(d)

0.00130

0.00133

0.00136

0.00139

0.00142

0.00145

0.00148

0.00151

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

.)

MC Simulation Index

NMC = 10,000 Target Value from CF Tolerance LimitsNMC = 10,000

0.00134

0.00136

0.00138

0.00140

0.00142

0.00144

0.00146

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

.)

MC Simulation Index

NMC = 50,000 Target Value from CF Tolerance LimitsNMC = 50,000

100

(e)

Figure 4.7 Tolerance intervals for torsional RMS response (σαα) of 100 MC simulations: (a)

NMC = 1,000; (b) NMC = 5,000; (c) NMC = 10,000; (d) NMC = 50,000; (e) NMC = 100,000.

0.00137

0.00138

0.00139

0.00140

0.00141

0.00142

0.00143

0.00144

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

.)

MC Simulation Index

NMC = 100,000 Target Value from CF Tolerance LimitsNMC = 100,000

101

(a)

(b)

Figure 4.8 QMC-based scatter plots of RMS response for deck section at x= l/4 and for U =

22.2 m/s as a function of NQMC: (a) heave σhh; (b) torsion σαα.

0.126

0.128

0.130

0.132

0.134

0.136

0.138

RM

S R

esp

onse

, σhh

(m)

NQMC = 1,000

NQMC = 5,000

NQMC = 10,000

Target Value from CF (RT: 1.46)

RT: 0.57RT: 5.05 RT: 19.27

Note: Relative Run Time (RT) to NMC = 50,000

NQMC = 1,000

NQMC = 5,000

NQMC = 10,000

0.0011

0.0012

0.0013

0.0014

0.0015

0.0016

0.0017

RM

S R

esp

onse

, σαα

(rad

.)

NQMC = 1,000

NQMC = 5,000

NQMC = 10,000

Target Value from NI

RT: 0.57

RT: 5.05 RT: 19.27

Note: Relative run time (RT) to NMC = 50,000

NQMC = 1,000

NQMC = 5,000

NQMC = 10,000

102

(a)

(b)

Figure 4.9 Coefficient of variation of the RMS response at x = l/4 for U = 22.2 m/s,

computed by QMC algorithm: (a) heave σhh; (b) torsion σαα.

0.0%

0.1%

1.0%

10.0%

1,000 10,000 100,000

Co

V, σhh

NQMC (No. of QMC Integration Points)

0.1%

1.0%

10.0%

1,000 10,000 100,000

CoV

, σαα

NQMC (No. of QMC Integration Points)

103

(a)

(b)

0.126

0.128

0.130

0.132

0.134

0.136

0.138

0.140

0 25 50 75 100

RM

S R

esp

onse

, σhh

(m)

QMC Simulation Index

NMC = 1,000 Target Value from CF Tolerance LimitsNQMC = 1,000

0.130

0.131

0.132

0.133

0.134

0.135

0.136

0 25 50 75 100

RM

S R

esp

onse

, σhh

(m)

QMC Simulation Index

NMC = 5,000 Target Value from CF Tolerance LimitsNQMC = 5,000

104

(c)

Figure 4.10 Tolerance intervals for vertical RMS response of 100 MC simulations (σhh): (a)

NQMC = 1,000; (b) NQMC = 5,000; (a) NQMC = 10,000.

0.131

0.132

0.133

0.134

0.135

0.136

0 25 50 75 100

RM

S R

esp

onse

, σhh

(m)

QMC Simulation Index

NMC = 10,000 Target Value from CF Tolerance LimitsNQMC = 10,000

105

(a)

(b)

0.00110

0.00120

0.00130

0.00140

0.00150

0.00160

0.00170

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

.)

QMC Simulation Index

NMC = 1,000 Target Value from CF Tolerance LimitsNQMC = 1,000

0.00125

0.00130

0.00135

0.00140

0.00145

0.00150

0.00155

0 25 50 75 100

RM

S R

espo

nse

, σαα

(rad

.)

QMC Simulation Index

NMC = 5,000 Target Value from CF Tolerance LimitsNQMC = 5,000

106

(c)

Figure 4.11 Tolerance intervals for RMS response of 100 MC simulations (σαα): (a) NQMC =

1,000; (b) NQMC = 5,000; (a) NQMC = 10,000.

0.00130

0.00133

0.00136

0.00139

0.00142

0.00145

0.00148

0.00151

0 25 50 75 100

RM

S R

esp

onse

, σαα

(rad

.)

QMC Simulation Index

NMC = 10,000 Target Value from CF Tolerance LimitsNQMC = 10,000

107

Figure 4.12 RMS values for vertical buffeting response as a function of wind speed U

corresponding to a given confidence level.

10 15 20 25 30 350

0.5

1

1.5

2

RM

S V

erti

cal R

espo

nse

(m)

U (m/s)

95% Confidence Level98% Confidence Level

108

(a)

(b)

Figure 4.13 Flutter derivatives H1* (a) and A2

* (b) of the Golden-Gate Bridge girder with

aspect ratio B/D = 3.5:1. Data sets were reproduced from (Jain 1996; Jain et al. 1996) with

α0=0°. The (“reference”) coefficients of the “Polynomial Model” were derived by

regression of the data sets, according to Eqs. (3.1) and (3.2). Tolerance limits (dashed lines)

were based on approximate evaluation of one standard deviation.

-7.0

-6.0

-5.0

-4.0

-3.0

-2.0

-1.0

0.0

0 2 4 6 8 10 12

Flu

tter

Der

ivat

ive,

Lif

t

Reduced Velocity

G.Gate H1*Reference

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 2 4 6 8 10 12

Flu

tter

Der

ivat

ive,

Mom

ent

Reduced Velocity

G.Gate A2*

Reference

109

(a)

(b)

Figure 4.14 TEP curves of RMS response with respect to thresholds T1 to T3 at the deck

section l/4: (a) σhh; (b) σαα (DFV: “Deterministic” Flutter Velocity).

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Pro

babi

lity

of E

xcee

danc

e

U (m/s)

DF

V T1

T2

T3

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Pro

babi

lity

of E

xcee

danc

e

U (m/s)

DF

V

T1

T2

T3

110

(a)

(b)

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

P[(Y

>T

)]

U (m/s)

DF

V

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

P[~

Flu

tter

]

U (m/s)

DF

V

111

(c)

Figure 4.15 Procedure for rescaling the TEP curves in Fig. 6.11(c) based on Eq. (6.3): (a)

prior probability or TEP; (b) marginal likelihood function; (c) Posterior probability or TEP

(DFV: “Deterministic” Flutter Velocity).

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

P[(Y

>T

)|~F

lutt

er]

U (m/s)

DF

V

112

(a)

(b)

Figure 4.16 TEP curves of RMS response at deck section l/4 (T2 threshold only) before (T2)

and after rescaling (T2M): (a) σhh; (b) σαα (DFV: “Deterministic” Flutter Velocity).

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Pro

babi

lity

of E

xcee

danc

e

U (m/s)

DF

VT

2

T2M

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Pro

babi

lity

of E

xcee

danc

e

U (m/s)

DF

V

T2

T2M

113

(a)

(b)

Figure 4.17 TEP curves of RMS responses with thresholds based on the RMS

displacement, deck section at l/4 and NEU’s flutter derivatives: (a) σhh; (b) σαα.

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

T2M

10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

T2M

114

(a)

(b)

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

115

(c)

(d)

Figure 4.18 TEP surfaces of RMS displacement for T2M threshold as a function of wind

accounting for effects of skew winds at l/4 with literature flutter derivatives: (a) σhh at l/4;

(b) σαα at l/4; (c) σhh at l/2; (d) σαα at l/2.

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

116

(a)

(b)

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

117

(c)

(d)

Figure 4.19 TEP surfaces of RMS displacement for T2 threshold as a function of wind

accounting for effects of skew winds at l/4 with NEU’s flutter derivatives: (a) σhh at l/4; (b)

σαα at l/4; (c) σhh at l/2; (d) σαα at l/2.

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

1020

3040

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ility

of

Exc

eed

ance

118

Figure 4.20 National Data Buoy Center (NOAA Station 9414290, Latitude: 37.807 N,

Longitude: 122.465 W) (Photo reproduced from NOAA, http://www.ndbc.noaa.gov/).

119

Figure 4.21 PDFs of “parent” (continuous time) mean wind velocity and annual maxima of

mean wind velocity, data from NOAA (NOAA).

0 20 40 60 80 1000

0.02

0.04

0.06

0.08

0.1

0.12

U (m/s)

PD

F

Parent wind velocityAnnual peak wind velocity

120

Chapter 5

Lifetime Cost Analysis due to Buffeting Response on a

Long-Span Bridge, accounting for Variability in Flutter

Derivatives

5.1 Introduction

This chapter discusses the estimation of lifetime cost analysis for a long-span bridge due to

wind-induced damages (i.e., monetary losses which can be associated with interventions

needed to ensure safety). In the estimation of the costs the variability in the flutter derivative

(FDs) is included. The focus of the chapter is on the buffeting bridge response.

The cost analysis employs the Monte-Carlo-based methodology for buffeting

response, described in Chapter 4. In the proposed probabilistic setting, “threshold

exceedance probability” (TEP) curves and surfaces are used to estimate the probability that

a given threshold for the variance of the response is exceeded. These probabilities are used,

together with information on the probability of the wind velocity at a given site, to predict

121

the expected value of a “lifetime monetary loss function”, derived for wind-induced

buffeting response.

The set of threshold levels, used in this chapter to derive these graphs, is different

from the ones used in Chapter 4. Thresholds are re-selected and based on both peak deck

acceleration (for user comfort) and peak dynamic displacements (e.g., for deformation of the

bridge deck or superstructure). Peak responses were directly estimated from the RMS

displacements through the gust effect factor (e.g., (Scanlan and Jones 1990a)). This choice

is dictated by the need for more realistically replicating actual damage scenarios on a large

bridge structure. The proposed model for “cost analysis” was adapted from an existing life-

cycle simulation algorithm for earthquake hazards (Wen and Kang 2001). The same truss-

type 1200m truss-type bridge model, analyzed in the previous chapters, was used in this

chapter.

5.2 Peak Estimation via RMS Response

The peak value of the dynamic displacement at any given deck section, either vertical

displacement h or torsional rotation α, can be determined by multiplying the RMS response

value by the peak effect factor (g). This factor was calculated through the equation proposed

by Davenport (Davenport 1964), which is strictly valid for a stationary random processes,

as follows:

(5.1) 0, 0

0, 0

0.577ˆ2log ,ˆ2log

e y

e y

g E g TT

122

(5.2)

where = generic RMS displacement (e.g., σhh for vertical response in Eq. 5.2), = the

generic RMS velocity corresponding to y, = reference temporal duration (averaging wind

time) for estimation of peak crossings, usually equal to 10 minutes.

In order to evaluate the effect of wind-induced vibration on user comfort and human

perception (e.g., Simiu and Scanlan 1996), RMS acceleration of the response and its peak

value are more relevant than dynamic displacements. These quantities were calculated and

applied in the cost analysis, later discussed. For example, the cross-RMS acceleration hh

can be derived from Eq. (5.2) and the corresponding peak vertical deck acceleration of

the generic section x can be found as follows:

(5.3)

(5.4)

5.3 Lifetime Expected Cost Analysis

Over a time period (t, in years), which may be the design life of a new bridge or the remaining

life of an existing bridge, the expected total cost can be expressed as a function of t as follows

(Wen and Kang 2001):

0,

1,

2yy

yyy

yy yy

0̂T

ˆ( )h x

4

0

, ( , , ) ,A B hh A Bhhx x K S x x K dK

ˆ( ) ( , ) .hh

h x x x g

123

(5.5)

In the previous equation, E[] denotes expected value; C0 is the initial construction

cost of the structure; i is an index describing each severe loading occurrence; ti ≤ t is the

loading occurrence time of event “i”, a random variable. Moreover, N(t) is the total number

of wind damaging events over time t (another random variable, described by a Poisson

counting process as suggested in Wen and Kang, 2001). The quantity Cj is the cost in present

dollar value of j-th limit state being reached at time ti of the loading occurrence; e-λti is the

“discounted factor” (Wen and Kang 2001) of the repair cost Cj over time t; λ is a constant

discount rate per year; Pj is the probability of occurrence for limit state j, which is assumed

as a constant over time since no structural deterioration is anticipated (usually acceptable in

the context of serviceability limit states); the integer index k is the total number of limit states

under consideration. The Pj probabilities must be found from the TEP curves and surfaces,

and are estimated from Eqs. (4.2) and (4.5) through the MC-based methodology presented

in Chapter 4; the actual procedure is later described. As suggested in (Wen and Kang 2001)

the integer variable N(t) can be approximated as a Poisson process.

In the specific prototype application under investigation four limit states were

separately considered. As a consequence, Eq. (5.5) was modified and adapted from its

original form to obtain the expected value of maintenance and repair cost CE, normalized

with respect to the initial construction cost C0 at t=0, as

( )

01 1

.i

N t kt

j ji j

E C t C E C e P

124

'

( )

1 11

( )

2 210

( )0

3 31

( )

4 41

i

i

i

i

N tt

i

N tt

i

E N tt

i

N tt

i

E e P

E e PE C t C

CC

E e P

E e P

'

( )

1

( )

1

( )

1

( )

1

=1 only,User comfort

=2 only, Deformation, lower tolerance

=3 only, Deformation, higher tolerance

=4 only, Collapse

N t

i

N t

i

N t

i

N t

i

j

j

j

j

(5.6)

In Eq. (5.6) the cost, associated with each limit state, is expressed as εjC0 (j=1,2,3,4)

with εj being the ratio of this cost relative to C0. The expected value of intervention cost

normalized to the initial construction cost in Eq. (5.6), εj (j=1,2,3,4) will be discussed in

Section 5.8 along with the numerical example. In Eq. (5.6) the expected value of the cost

was found by categorizing the number of occurrences for wind events into “moderate wind

storm” (serviceability and buffeting limit states from j=1 to j=3) and “severe wind storm”

(j=4, collapse induced by the onset of flutter). This selection was dictated by the need for

separating the frequent occurrences of a moderate wind-speed storm (based on a Poisson

counting process N(t)) with extremely rare and severe events (with N’(t)<N(t)). Therefore,

the arrival rate of each Poisson Process, N(t) and N’(t), was taken as distinct.

In particular, the number of occurrences for moderate wind storms N(t) at time t was

based on a Poisson Process with an annual occurrence rate, derived from the probability of

annual wind speed maxima exceeding U=20 m/s, based on actual site winds in the proximity

of the actual bridge; this is equal to 0.94 independent of direction θ. This wind speed value

was chosen to reflect a minimum demand level, corresponding to a small but non-negligible

125

deck vibration (e.g., vertical vibrations exceeding 0.27 m or 1/30 of the deck height D). The

j=4 limit state employs a second Poisson counting process N’(t) with a much smaller

occurrence rate; the latter rate was determined as the probability of annual mean wind speed

maxima exceeding a wind speed (U) equal to the median of the critical flutter velocity

distribution for this bridge, considering variability and uncertainty in flutter derivatives, and

for wind direction orthogonal to the bridge axis (θ=0°). The median of the critical velocity

was estimated as U = 60 m/s, which leads to 0.01 as the mean occurrence rate of N’(t).

Table 5.1 shows the limit states employed in the cost analysis for the case study.

Three thresholds for limit states (T1 to T3) were used in the cost analysis under buffeting

response. The level T1, associated with limit state j=1 and maximum vertical deck

accelerations, was selected as a user comfort level based on human perception (e.g., Simiu

and Scanlan, 1996). The thresholds associated with maximum vertical dynamic

displacements was sub-categorized as T2 (limit state j=2) and T3 (limit state j=3); T2 is a more

stringent level (lower tolerance) and T3 is a less restrictive level (higher tolerance). In Table

5.1 the threshold T1 was taken as 20 milli-g of peak deck dynamic acceleration; T2 and T3

were taken as 1 m and 2 m of peak vertical deck oscillation, respectively (simulating

unacceptable deformation in the bridge deck or superstructure during serviceability). They

are all applied to buffeting response analysis and response. On the contrary, the quantity T4

was applied to the onset of flutter and collapse; and T4 was separately analyzed. The

thresholds listed in Table 5.1 were employed in the TEP analysis for vertical deck vibrations;

the same thresholds were used for peak torsional acceleration and deck rotation and

normalized with respect to B in order to be compatible with an equivalent vertical oscillation

126

of the same order of magnitude (for example with T1 being equal to 20 milli-g, T2 equal to 1

m and T3 equal to 2 m).

5.4 Description of the Structural and Aeroelastic Model

For the cost analysis a four-mode buffeting and flutter analysis of the Golden Gate Bridge

model was carried out by considering the first two vertical (v1 and v2) and torsional modes

(t1 and t2) with frequencies nv1 = 0.087 Hz and nv2 = 0.129 Hz, nt1 = 0.192 Hz and nt2 =

0.197 Hz. Simplified mode shapes were used to describe the deck and cable vibration, as

assumed in Chapters 3, 4 and in Figs. 3.5(b,d,g,h).

The TEP curves for T1 to T3 in Table 5.1 were first calculated using the model curves

based on the literature data (Jain et al. 1998 and Section 4.3.2); as described in previous

chapters the statistical properties of the coefficients of the model curves were synthetically

reproduced since no FD error analysis was available.

As a second example, TEP curves were estimated from the results from the tests

conducted at NEU, in which the eight flutter derivatives H1*,…, H4

* and A1*,…, A4

* of the

section model in Fig. 3.5 were found. The first- and second-moment properties of the

experimental errors were evaluated by repeating the measurements numerous times, as

described in Chapter 3.

For both sets of flutter derivatives the “randomization” of H1*,…, H4

* and A1*,…,

A4* was carried out by means of the model curves introduced in Chapter 3.

127

5.5 Estimation of Peak Dynamic Response during Buffeting

Peak estimation via RMS dynamic response with a zero skew wind angle is discussed in this

section. As a first implementation of the model for cost analysis, both peak dynamic

displacements and accelerations were restricted to the deck section at quarter span

(x=l/4=400 m) (i.e., cross-section at the anti-nodal point of the fundamental modes “v1” and

“t1”).

Initially, a “deterministic buffeting” analysis without considering variability in the

FDs was carried out to estimate the reference values of the peak vibration at the quarter-span

section of the deck and for verification that the oscillation magnitudes were indeed

realistically compatible and consistent with an increment of mean wind speed. An example

of peak estimation for wind orthogonal to the bridge is shown in Fig. 5.1 for both Jain’s data

and NEU’s data. For example, the comparison of peak vertical displacements is shown in

Fig. 5.1(a); vertical accelerations were calculated in terms of milli-g (mg) in Fig. 5.1(b). The

effect of the static wind loads on deck vertical displacements and rotations were also

estimated but were not incorporated in the figure due to their negligible influence in

comparison with dynamic effects. Both graphs clearly show that the peak values increase

non-linearly with increasing mean wind velocity U. Furthermore, peak vertical-response

displacements (Fig. 5.1a), using Jain’s flutter derivatives, are lower than the ones based on

flutter derivatives from NEU tests.

128

5.6 Monte-Carlo-based Methodology for “Statistical Buffeting” Analysis

considering Uncertainty in the FD

5.6.1 Wind-Direction-Independent Analysis

In Fig. 5.2 three “threshold exceedance threshold” (TEP) curves, associated with the bridge

model and based on flutter derivatives provided in (Jain et al. 1998) are presented for relative

wind direction orthogonal to the bridge axis (θ=0°). The TEP curves were calculated for both

vertical and torsional response at the quarter span of the simulated bridge. The curves in Fig.

5.2 were numerically computed by exclusively considering the buffeting response without

the analysis of collapse probability influenced by the onset of flutter. The MC-based

methodology for buffeting analysis, presented in Chapter 4, numerically calculates the peak

acceleration (T1) and the peak displacements (T2, T3) of both vertical and torsional responses.

At a given wind speed, since flutter has occurred on a sub-set of events only (statistically),

the procedure can still find a finite peak displacement if one restricts the attention to the sub-

set of events for which flutter has not occurred. Since the cost analysis (Eq. 5.6) requires the

separation between events leading to buffeting vibration and those leading to flutter, it is

necessary to “exclude” the events leading to flutter in the estimation of buffeting vibration

TEP. Therefore, the approximate “rescaling” described in Section 4.3.2 was utilized for

buffeting response; the “prior” TEP curves (which are still including events for which flutter

has occurred) were modified using Eqs. (4.3) and (4.4). The rescaled curves for thresholds

T1, T2, T3 are shown in Fig. 5.3. The scaling requires the flutter probability to be evaluated;

this was carried out numerically, as described in a later section (Section 5.7).

129

Figure 5.4 shows the TEP curves obtained with NEU’s experimental data, shown in

Fig. 3.5, and for a mean wind direction orthogonal to the bridge axis (θ=0°). These curves

were also based on the same threshold levels as in Table 5.1 and after rescaling (Eq. 4.3).

The comparison of TEP curves considers both Jain’s data and NEU’s data; the curves

are shown in Figs 5.5 and 5.6. The TEP curves using NEU’s data for vertical response are

shifted to the left of the graph, i.e., toward a lower wind velocity region in comparison with

Jain’s data for T1 T2 and T3. In contrast, the situation seems opposite for torsional response

since the curves using NEU’s data are shifted to the right of the graphs.

The comparison between the two curves for each threshold level in Figs. 5.5 and 5.6,

based on Jain’s FD data and NEU’s FD data, also suggests that the variability in the

aeroelastic loads plays a significant role in the input-to-output uncertainly propagation for

the analysis of the buffeting response for this bridge example. Exceedance probabilities are

crucially influenced by the flutter derivatives which are employed. A reduction in

exceedance probability is observed if the flutter derivatives from Jain’s data were used to

estimate the wind loading and deck response. This observation is valid for the vertical

dynamic deck response only and the deck section at the quarter span: in Fig. 5.5 for U=25

m/s the exceedance probability of the T2 threshold is less than 15% with Jain’s derivatives

whereas it is almost 86% for NEU’s data. In contrast, the torsional response appears to be

opposite.

5.6.2 Wind-Direction-Dependent Analysis

Figures 5.7 and 5.8 shows the TEP surfaces for vertical and torsional response, based on

flutter derivatives from Jain’s data (Fig 3.6) and corresponding to the thresholds in Table 5.1

130

(for vertical response thresholds are: T1 = 20 milli-g, T2 = 1 m and T3 = 2 m). A relative skew

wind angle, varying from θ = –40° to θ = 40°, was exclusively considered. Figures 5.9 and

5.10 show the TEP surfaces as a function of U and θ for vertical and torsional response,

based on flutter derivatives from NEU, corresponding to the same thresholds as in Table 5.1.

The analysis of all TEP surfaces, Figs. 5.7 through 5.10, confirms that the relative

skew wind angle equal to zero degrees is the most critical direction for both vertical and

torsional buffeting responses.

5.7 Flutter Analysis: Numerical Results

The flutter velocity in the absence of flutter derivative errors (i.e., “deterministic”) is equal

to 19.7 m/s for Jain’s derivatives, whereas is larger than 80.0 m/s for NEU’s data. The

“statistical” flutter velocity accounting for FD uncertainty was investigated and used to

determine the limit-state probability P4 in Eq. (5.6).

The Probability Density Function (PDF) and Cumulative Density Function (CDF) of

flutter velocity Ucr were numerically calculated from the histograms of occurrences, derived

through a MC-based sample at various skew wind angles (–40° ≤ θ ≤40°) exclusively using

NEU’s flutter derivatives; three cases (θ = 0°, 20°, and 40°) are compared and shown in Fig.

5.11. In order to increase accuracy in the Monte-Carlo estimation of low probabilities, a total

of 50,000 Monte-Carlo realizations were employed to synthetically generate the random set

of flutter derivatives for stochastic flutter analysis (Seo and Caracoglia 2011). As for the

buffeting analysis, random flutter derivatives were generated from the polynomial model

curves with random coefficients described in Section 3.6.

131

The accuracy of the Monte-Carlo estimation in relation to the number of realizations

(NMC) was also analyzed by comparing the numerical estimation of the CDF with the

“standard error measure”. The standard error can be defined as (Bucher 2009)

2 .E MCS N (5.7)

According to Bucher (2009), the variance of the CDF estimator was not derived by

repeated evaluation of the flutter probability (PUcr) but was approximately assessed along

the Ucr axis from the curve corresponding to case with θ=0° in Fig. 5.11(b) as

.U crcr

P U MCP N (5.8)

Figure 5.12 was also used for “guidance” to provide indications for sufficient or

insufficient resolution of the stochastic flutter analysis, since it compares the absolute

differences between any two contiguous numerical flutter CDFs at various θ (i.e., the curves

shown in Fig. 5.11) to the standard error (Eq. 5.6). For example, the blue dotted line in Fig.

5.12 is the absolute difference between the two cases for mean wind incident angles θ=0°

and θ=20°; this difference is greater than the standard error (red solid line) when flutter

velocity is approximately less than 85 m/s. The black dashed-dotted line is the absolute

difference between the two cases with θ=20° and θ=40°; this quantity is greater than the

standard error (red solid line) when Ucr ≤ 90 m/s. The interval of Ucr where the relative

132

difference between any two scenarios (absolute variations due to change in skew wind angle

θ) is greater than the standard error can be used as a “confidence region” with acceptable

resolution.

As a result of the comparisons in Fig. 5.12, it was suggested that an “acceptance

region” in the MC-based methodology for stochastic flutter speed with skew wind angle can

be approximately found for mean wind speed less than 80 m/s; this velocity is usually lower

that the upper limits used to estimate the TEP curves and surfaces presented in this chapter.

The probability of the onset of flutter (PUcr) also coincides with the limit-state

probability P4 in Eq. (5.6). As a result of numerical flutter analysis, the limit-state probability

P4 was numerically found as 0.0029 for the case using Jain’s FD data and 0.0964 with NEU’s

FD data. The flutter probabilities P4 for both cases (i.e., using Jain’s FDs and NEU’s FDs)

will not be considered in the cost analysis.

5.8 Lifetime Expected Intervention Cost Analysis - Numerical Results

In this section, an example of expected lifetime cost analysis is described. NEU’s FD data

was exclusively used for buffeting simulations considering the uncertainty in the flutter

derivatives.

5.8.1 Estimation of the Limit-State Probabilities Pj from TEP Analysis

Table 5.2 shows the probabilities of each limit state (Pj) based on the buffeting intervention

thresholds described in Table 5.1 for both the case with considering the variability in the FD

and the case in the absence of the variability. As outlined earlier, the threshold level T1 for

user comfort and two levels for deck deformation were investigated: T2 a more conservative

133

intervention threshold level and T3 a less restrictive level. However, inspection revealed that

the limit state probabilities P2 and P3 for torsional response are zero values for 0<U<80 m/s

when variability in FD is neglected (i.e., for “deterministic buffeting”). This remark confirms

that “deterministic buffeting” analysis underestimates the loss function significantly without

further investigation. Therefore, the results of the TEP analysis for vertical deck response

were exclusively used in the cost analysis.

As described in previous sections and in Table 5.2 the threshold T1 used in the cost

analysis corresponds to peak deck dynamic acceleration (comfort of users); T2 and T3

correspond to the peak vertical deck oscillation requiring intervention for serviceability.

These three thresholds (i.e., T1, T2 and T3) are estimated from dynamic buffeting analysis.

The probabilities in Table 5.2 were computed from the TEP surfaces in Fig. 5.9

through Fig. 5.10. As originally introduced in Eq. (4.5), if the TEP surface is derived from

the equation below

FTj(U,θ) = P[Y>Tj|U = u,Θ = θ], (5.9)

the probability for limit states j=1,2,3 is

Pj = ∫∫FTj(U,θ)fU(U)fθ(θ)dθdU. (5.10)

In Eq. (5.10) the probability density functions of the wind speed fU(U) and mean

wind direction fθ(θ) were obtained from the “raw” wind data, assuming U and θ as

independent random variables, recorded at the meteorological station close to the benchmark

134

bridge, as described in Section 4.5. The histogram of annual 10-minute averaged wind speed

maxima shown in Fig. 4.21 was derived from the NOAA annual wind speed maxima over a

16-year period (1996-2011) and was employed to evaluate fU(U). A detailed description of

the station and data resources can also be found in (Seo and Caracoglia 2012).

Derivation of the probability density functions of the annual wind speed maxima (fU)

and mean skew wind angle (fθ) enabled the estimation of Pj (j=1,2,3) in Eq. (5.10) after TEP

analysis (Section 5.5).

5.8.2 Expected Intervention Cost - Description of the Simulations

Probability mass function (PMF) was utilized to investigate how the cost varies with a

random arrival time of each wind event (ti and ti’ in Eq. 5.6). The arrival time of the events

was modeled as a uniform random variable between 0 and t (or 0 and t’) with the counting

of the total number of Poisson’s events, N(t) or N’(t). Monte-Carlo sampling was employed

to estimate the PMF of the relative maintenance and repair cost, (C(t)-C0)/C0, using 500

repeated Monte-Carlo realizations and five percent discount rate/year (λ). Intervention

criteria can be found in Table 5.1 (Tj, j=1,2,3,4); a time projection of 80 years was initially

used; probabilities of exceeding the limit states were taken from the Pj values in Table 5.2

(j=1,2,3). The ratio of this cost relative to construction cost C0 was chosen to be 20% for all

intervention levels ε1, ε2, ε3. The results of the cost analysis are summarized in Figs. 5.13 and

5.14. These figures were compiled based on two different intervention criteria: Fig. 5.13 for

user comfort (i.e., P1), and Fig. 5.14 due to deformation in the deck (i.e., P2 or P3). Five

markers (black +, blue ◊, cyan ○, magenta □, and green Δ) and one red solid line are shown

in the stem plots (Fig. 5.13). Each point on the stem plots represents a probability range: +

135

probability between 0 and 0.2, ◊ between 0.2 and 0.4, ○ between 0.4 and 0.6, □ between 0.6

and 0.8, Δ less than 1.0. The solid red line describes the evolution of the CE (Eq. 5.6) expected

relative cost, as a function of time, derived from the PMF stem plots.

5.8.3 Expected Intervention Cost - Numerical Results using NEU’s FD Data

Figure 5.14 shows the results of the cost analysis due to buffeting response using NEU’s FD

data. For an intervention level threshold T1 of 20 milli-g of maximum vertical acceleration

(for user comfort) the loss function estimated without considering the uncertainty in the FDs

is approximately 14% lower for exposure times of 40 years or more, as indicated in Fig.

5.13(b). For an intervention level threshold T2 of 1m of maximum vertical deformation

(lower tolerance), the underestimation was approximately 19% for 40 years of exposure or

longer, as in Fig. 5.14(a). For the higher tolerance case T3 of 2 m of maximum vertical

deformation the underestimation proved to be 27% for 40 years of exposure or longer, with

a progressive reduction trend with lower exposure times similar to that for the lower

threshold.

5.9 Discussion and Remarks

The cost analysis result shows that the expected value of the loss function due to the buffeting

response of a suspended bridge (a function proportional to the cost associated with

interventions needed to ensure safety) is affected by the variability of the FDs in a manner

that depends on the time of exposure and one the threshold used to decide on the need for

intervention (maintenance or repair).

136

For user comfort level threshold (T1) of 20 milli-g of maximum vertical acceleration

the loss function estimated without considering the uncertainty in the FDs is approximately

14% low for exposure times of 40 years or more. For exposure times less than 40 years the

underestimation decreases (approximately) linearly for both cases, up to zero for zero

exposure time.

For the intervention level threshold with lower tolerance (T2) of 1m of maximum

vertical deformation the underestimation was 19% for 40 years of exposure or longer. For

higher tolerance case, T3 of 2m of maximum vertical deformation, the underestimation

proved to be 27% for NEU’s case for 40 years of exposure or longer, with a reduction trend

with exposure similar to that for the lower threshold. This result shows that consideration of

the FD variability in the estimation of the loss function is important when the threshold for

intervention is relatively high for this benchmark bridge.

137

Table 5.1 Structural performance thresholds for vertical deck response (Tj).

Table 5.2 Probabilities of each damage state (Pj) due to buffeting response based on the

structural performance thresholds (T = Tj) using NEU’s FD data.

Flutter ThresholdHuman comfort

(T 1, peak acc.)

Lower ToleranceDeformation

(T 2, peak disp.)

Higher ToleranceDeformation

(T 3 peak disp.)

Instability

(T 4, collapse)

20 milli-g 1 m 2 m N.A.

Buffeting Thresholds (deck)

P 1 P 2 P 3

Without variability in FD 3.48E-03 3.20E-03 2.62E-03

With variability in FD 4.04E-03 3.94E-03 3.56E-03

Without variability in FD 5.96E-04 0.00E+00 0.00E+00

With variability in FD 3.58E-03 3.04E-03 2.70E-03

Probabilities of Damage State, P [Y > T ]

TorsionalResponse

VerticalResponse

Using NEU's FD Data

138

(a)

(b)

Figure 5.1 Reference peak vertical dynamic response, (“deterministic” without variability

in FD) as a function of wind velocity at θ = 0° with both Jain’s flutter derivatives and

NEU’s flutter derivatives at l/4: (a) displacement; (b) acceleration.

10 12 14 16 18 200

0.1

0.2

0.3

0.4

U (m/s)

Dis

pla

cem

ent

at l

/4, (

m)

FDs from JainFDs from NEU

10 12 14 16 18 200

2

4

6

8

10

U (m/s)

Acc

eler

atio

n a

t l/4

, (m

g)

FDs from JainFDs from NEU

139

(a)

(b)

Figure 5.2 TEP curves of the peak dynamic response with respect to thresholds T1, T2, T3

deck section at l/4 using Jain’s FD data: (a) vertical response; (b) torsional response.

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

140

(a)

(b)

Figure 5.3 Recaled TEP curves (modified by Eq. 4.3) of the peak dynamic response with

respect to thresholds T1, T2, T3 deck section at l/4 using Jain’s FD data: (a) vertical

response; (b) torsional response.

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

141

(a)

(b)

Figure 5.4 Rescale TEP curves (modified by Eq. 4.3) of the peak dynamic response with

respect to thresholds T1, T2, T3 deck section at l/4 using NEU’s FD data: (a) vertical

response; (b) torsional esponse.

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1

T2

T3

142

(a)

(b)

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1 (NEU)

T1 (Jain)

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T2 (NEU)

T2 (Jain)

143

(c)

Figure 5.5 A comparison between two sets of curves for vertical response based on NEU’s

FD data (continuous lines) and Jain’s FD data (dotted lines): (a) T1=20 milli-g; (b) T2=1m;

(c) T3=2m.

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T3 (NEU)

T3 (Jain)

144

(a)

(b)

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T1 (NEU)

T1 (Jain)

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T2 (NEU)

T2 (Jain)

145

(c)

Figure 5.6 A comparison between two sets of curves for torsional response based on

NEU’s FD data (continuous lines) and Jain’s FD data (dotted lines): (a) T1=20 milli-g; (b)

T2=1m; (c) T3=2m.

10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Pro

bab

ilit

y of

Exc

eed

ance

U (m/s)

T3 (NEU)

T3 (Jain)

146

(a)

(b)

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

147

(c)

Figure 5.7 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic vertical

response with various intervention levels at l/4 using Jain’s FD data: (a) T1=20 milli-g; (b)

T2=1m; (c) T3=2m.

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

148

(a)

(b)

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

149

(c)

Figure 5.8 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic (normalized to

B/2) torsional response with various intervention levels at l/4 using Jain’s FD data: (a)

T1=20 milli-g; (b) T2=1m; (c) T3=2m.

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

150

(a)

(b)

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

151

(c)

Figure 5.9 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic vertical

response with various intervention levels at l/4 using NEU’s FD data: (a) T1=20 milli-g; (b)

T2=1m; (c) T3=2m.

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

152

(a)

(b)

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

153

(c)

Figure 5.10 Recaled TEP surfaces (modified by Eq. 4.3) of the peak dynamic (normalized

to B/2) torsional response with various intervention levels at l/4 using NEU’s FD data: (a)

T1=20 milli-g; (b) T2=1m; (c) T3=2m.

2040

6080

-40-20

020

400

0.5

1

U (m/s) (deg)

Pro

bab

ilit

y of

Exc

eed

ance

154

(a)

(b)

Figure 5.11 Probability distributions of flutter velocity, Ucr, using NEU’s flutter

derivatives: a) PDF (probability density function); b) CDF (cumulative density function).

0 50 100 150 2000

0.005

0.01

0.015

0.02

Ucr (m/s)

Nu

mer

ical

PD

F

PDF(Ucr|=0)

PDF(Ucr|=20)

PDF(Ucr|=40)

0 50 100 150 2000

0.2

0.4

0.6

0.8

1

Ucr (m/s)

Nu

mer

ical

CD

F

CDF(Ucr|=0)

CDF(Ucr|=20)

CDF(Ucr|=40)

155

Figure 5.12 Resolution of the Monte-Carlo-based flutter procedure vs. standard error.

0 50 100 150 2000

0.02

0.04

0.06

0.08

Ucr (m/s)

Res

olu

tion

CDF(Ucr|=0) - CDF(Ucr|=20)

CDF(Ucr|=20) - CDF(Ucr|=40)

Standard error

156

(a)

(b)

Figure 5.13 Intervention costs normalized to the initial construction cost for user comfort

level threshold T1=20 milli-g over time based on NEU’s FD data :

(a) 3D PMF (probability mass function stem plot); (b) 2D expected normalized cost, -

discount rate/year λ=0.05.

0

0.2

0.4

020406080

0

0.2

0.4

0.6

0.8

1

Time (yrs)Relative Cost

PM

F

0 20 40 60 800

0.02

0.04

0.06

0.08

0.1

Time (yrs)

Exp

ecte

d N

orm

aliz

ed C

ost,

CE

With Variability in FDWithout Variability in FD

0 0EC E C t C C

157

(a)

(b)

Figure 5.14 Expected intervention costs normalized to the initial construction cost due to

deformation in deck based on NEU’s FD data : (a) lower

tolerance case (T2=1m); (b) higher tolerance case (T3=2m), - discount rate/year λ=0.05.

0 20 40 60 800

0.02

0.04

0.06

0.08

0.1

Time (yrs)

Exp

ecte

d N

orm

aliz

ed C

ost,

CE

With Variability in FDWithout Variability in FD

0 20 40 60 800

0.02

0.04

0.06

0.08

0.1

Time (yrs)

Exp

ecte

d N

orm

aliz

ed C

ost,

CE

With Variability in FDWithout Variability in FD

0 0EC E C t C C

158

Chapter 6

Summary and Conclusions

6.1 Summary

The primary intent of the research, described in this dissertation, was to implement a

methodology for analyzing the buffeting response of cable-supported bridges, including

uncertainty in the aeroelastic input (flutter derivatives). Flutter derivatives are the most

important part of wind loading and are estimated in a wind tunnel experiment. A second

order polynomial model for the flutter derivatives was used in the research and labeled as

“model curve”.

A general discussion on the state of the art regarding the general introduction of

wind-induced phenomena of long-span bridges is carried out in Chapter 1. Also the

motivation of this research is presented. Fast and efficient integration procedures of the wind

loading acting over the deck span are also necessary for structural performance.

Chapter 2 summarizes the background theory of wind-induced response of long-span

bridges and reviews the fundamental aspects of aerodynamics and aeroelasticity of long-

span suspended bridge decks. The general approach for flutter instability and buffeting

159

response is discussed in this chapter. The effect of wind directionality on the bridge deck

response is also presented.

Chapter 3 describes the experimental setup, measurements and experimental results,

used in this research. This chapter also describes the development of a “model curve” for

representation of the behavior of flutter derivatives as a function of reduced wind velocity.

Flutter derivatives were measured in the wind tunnel at Northeastern University.

Chapter 4 discusses the buffeting analysis of long-span bridges including uncertainty

in the FD. In the first part of this chapter, “deterministic buffeting” analysis is discussed

using the model curve that the result is the value of the RMS dynamic response at a given

wind speed as the standard approach. In the second part of this chapter, “statistical buffeting”

analysis is discussed, which includes uncertainty in the FDs; the coefficients of the model

curve are treated as random variables, whose probability distribution is conditional on

reduced wind speed. The “statistical buffeting” response was evaluated by adopting the

concept of “fragility”; this was also employed in the calculation of the exceedance

probability of pre-selected vibration thresholds, conditional on mean wind speed and

direction at the deck level. Threshold exceedance probability curves and surfaces, associated

with RMS dynamic buffeting response, were numerically derived.

Finally, the expected value of the loss function due to the buffeting response of a

1200-meter suspension bridge is evaluated by considering the uncertainty in the FDs.

160

6.2 Conclusions

The findings of this research can be summarized as follows:

Regarding Performance

It is shown that the expected value of the loss function due to the buffeting response

of a suspended bridge (a function proportional to the cost associated with

interventions needed to ensure safety) is affected by the variability of the FDs in a

manner that depends on the time of exposure and one the threshold used to decide on

the need for intervention.

For an intervention level threshold of 1m of maximum vertical deformation the loss

function estimated without considering the uncertainty in the FDs is approximately

19% low for exposure times of 40 years or more. For exposure times less than 40

years the underestimation decreases (approximately) linearly, up to zero for zero

exposure time.

For an intervention level threshold of 2m of maximum vertical deformation (higher

tolerance in allowing the bridge to operate) the underestimation proved to be 27%

for 40 years of exposure or longer, with a reduction trend with exposure similar to

that for the lower threshold. This result shows that consideration of the FD variability

in the estimation of the loss function is important when the threshold for intervention

is relatively high.

161

On Computational Issues

It is shown that the RMS buffeting response of a long-span cable supported bridge

can be estimated (with sufficient accuracy to be practically useful) using a second

order polynomial description of the flutter derivatives (FDs). The uncertainty in these

derivatives in the model is captured by specifying the coefficients of the polynomial

as a vector of random variables having a specified mean and covariance (the values

of these properties used in the numerical analyses were obtained experimentally).

It is found that there is a significant computational advantage in using Monte Carlo

methods for calculating a double integral that arises on the estimating of the

generalized buffeting loading used in a multi-mode buffeting analysis. The

computational time is approximately one order of magnitude smaller than that of

using a standard integration algorithm (repeated trapezoidal rule).

It is found that the polynomial description of the FD can also be used with

computational advantages to estimate the onset of flutter.

6.3 Recommendations for Future Research

Expand the treatment of the uncertainties to include structural properties and wind

directionality.

More detailed investigation of Quasi-Monte-Carlo methods should be taken,

especially for calculating the double integral in the analysis of multi-variable

“random domains”.

162

More detailed investigation in the estimation of the expected value of the loss

function due to the buffeting response of a suspended, affected by the variability of

the FDs, should be taken.

Account for lateral dynamic bridge response in the buffeting analysis.

Include chord-wise admittance effects in the formulation of the buffeting loading.

Even though the use of chord-wise admittance can lead to a reduction of magnitude

in the buffeting loads, this effect should be included in the future.

6.4 Outcome of the PhD Studies: List of Deliverables

As of February 2013, the work, described in this dissertation, has been published or is in the

process of being published in a series of journal articles (three) and conference proceedings

(four). A description of these items is provided below for completeness.

6.4.1 Journal Publications (Published/under review)

A. Seo, D.-W. and Caracoglia, L., “Estimating Life-Cycle Monetary Losses due to Wind

Hazards: Fragility Analysis of Long-Span Bridges,” Engineering Structures, Submitted

for review, 2012.

B. Seo, D.-W. and Caracoglia, L., “Statistical Buffeting Response of Long-Span Bridges

Influenced by Errors in Aeroelastic Loading Estimation,” Journal of Wind Engineering

and Industrial Aerodynamics, Vol. 104-106, 2012, pp. 129-140. Note: The paper was

selected for potential publications in a special issue of the journal, as an extended

version of ICWE-13 Conference paper.

C. Seo, D.-W. and Caracoglia, L., “Estimation of Torsional-flutter Probability in Flexible

Bridges Considering Randomness in Flutter Derivatives,” Engineering Structures, Vol.

33, No. 8, 2011, pp. 2284–2296.

163

6.4.2 Other Journal Publications (not related to the main topic of this Dissertation)

D. Seo, D.-W. and Caracoglia, L., “Derivation of Equivalent Gust Effect Factors for Wind

Loading on Low-Rise Buildings through Database-Assisted-Design Approach,”

Engineering Structures, Vol. 32, No. 1, 2010, pp. 328-336.

6.4.3 Full Papers in Conference Proceedings

E. Seo, D.-W. and Caracoglia, L., “A Numerical Algorithm for Predicting Life-Cycle

Maintenance Costs for Slender Bridges under Wind Hazards,” Proceedings of the 3rd

Workshop of the American Association for Wind Engineering (AAWE), Hyannis,

Massachusetts, August 12-14, 2012, CD-ROM.

F. Seo, D.-W. and Caracoglia, L., “Monte-Carlo Methods for Estimating the Buffeting

Response of a Bridge Contaminated by Flutter-Derivative Errors,” Proceedings of 13th

International Conference on Wind Engineering (ICWE-13), Amsterdam, NL, July 10-

15, 2011, paper No. 214, CD-ROM.

G. Seo, D.-W. and Caracoglia, L., “Flutter Velocity Estimation using Experimentally-

Derived (Co)-Variances of Aeroelastic Coefficients,” Proceedings of the International

Conference of the Engineering Mechanics Institute (EMI2011), American Society of

Civil Engineers, Northeastern University, Boston, Massachusetts, USA, June 02-04,

2011, CD-ROM (full paper). Note: The paper was selected among the finalists for the

Probabilistic Methods Student Paper Award.

H. Seo, D.-W. and Caracoglia, L., “Quasi- and Monte-Carlo-Based Methods for Statistical

Buffeting Analysis of Long-Span Bridges under the Effects of Turbulent Wind,”

Proceedings of the 2nd Workshop of the American Association for Wind Engineering

(AAWE), Marco Island, Florida, August 18-20, 2010.

6.4.4 Other Papers Published as Conference Proceedings (not related to the main

topic of the Dissertation)

I. Seo, D.-W., Moghim, F., and Bernal, D., “Normalization of Complex Modes from Mass

Perturbations,” Proceeding of A Conference and Exposition on Structural Dynamics

164

(IMAC XXIX), Society for Experimental Mechanics, Jacksonville, FL, January 04-07,

2011, PAPER No. 41.

6.4.5 Poster Presentations

J. Seo, D.-W. and Caracoglia, L., “A Life-Cycle Cost Analysis for Structural Maintenance

of Flexible Bridges under Wind Hazards,” 2012 NSF-CMMI Engineering Research and

Innovation Conference, Boston, Massachusetts, USA, July 9–12 2012 (student poster

presentation).

K. Seo, D.-W. and Caracoglia L., “A Life-Cycle Cost Model for Structural Maintenance of

Long-Span Bridges Under Wind Hazards,” NU Research Exposition 2012, Boston,

Massachusetts, USA, April 6, 20112, Research poster No. 67.

L. Seo, D.-W. and Caracoglia, L., “Statistical Buffeting Simulations of Long-Span Bridge

Response under Wind Hazards: Fragility Curves and Surfaces,” Second US-Japan Mini

Workshop on Structural Dynamics and Monitoring of Bridges and Flexible Structures

against Wind Hazards, Boston, Massachusetts, USA, November 12-14, 2011.

M. Seo, D.-W. and Caracoglia L., “Statistical Buffeting Simulations of Long-span Bridge

Response under Wind Hazards: Recent Case Studies,” NU Research Exposition 2011,

Boston, Massachusetts, USA, April 6, 2011, No. 1968.

N. Seo, D.-W. and Caracoglia L., “Estimation of Torsional Bridge Flutter Collapse by

Numerical Statistical Methods,” NU Research Exposition 2010, Boston, Massachusetts,

USA, March 24, 2010, No. 1410. O. Seo, D.-W. and Caracoglia L., “Assessment of Gust Effect Factors for Wind Loading on

Low-rise Buildings through Database-Assisted-Design Method and Current Structural

Design Standards,” NU Research Exposition 2009, Boston, Massachusetts, USA,

March 26, 2009, No. 1294.

165

References

Barlow, J. B., Rae, W. H., and Pope, A. (1999). Low-speed wind tunnel testing (third edition),

John Wiley and sons, New York, New York, USA.

Bartoli, G., Contri, S., Mannini, C., and Righi, M. (2009). "Toward an Improvement in the

Identification of Bridge Deck Flutter Derivatives." Journal of Engineering

Mechanics, 135(8), 771-785.

Bashor, R., and Kareem, A. (2007). "Probabilistic performance evaluation of buildings: An

occupant comfort perspective." 12th International Conference on Wind Engineering

(12-ICWE), Cairns, Australia, July 1-6, 2007, 1335-1342.

Bienkiewicz, B. (1987). "Wind-tunnel study of effects of geometry modification on

aerodynamics of a cable-stayed bridge deck." Journal of Wind Engineering and

Industrial Aerodynamics, 26(3), 325-339.

Brito, R. (2008). "Extraction of aeroelastic coefficients for bridge decks from small-scale

wind tunnel tests," Northeastern Unviersity, Boston.

Brito, R., and Caracoglia, L. (2009). "Extraction of flutter derivatives from small scale wind

tunnel experiments." 11th Americas Conference on Wind Engineering, American

Association for Wind Engineering (AAWE), San Juan, Puerto Rico, CD-Rom.

Bucher, C. (2009). Computational analysis of randomness in structural mechanics, Taylor

Francis Group, London, UK.

Caracoglia, L. (2008a). "Influence of uncertainty in selected aerodynamic and structural

parameters on the buffeting response of long-span bridges." Journal of Wind

Engineering and Industrial Aerodynamics, 96(3), 327-344.

Caracoglia, L. (2008b). "Recent investigations on long-span bridge aeroelasticity in the

presence of turbulence fields with uncertain span-wise correlation." Sixth

International Colloquium on Bluff Bodies Aerodynamics & Applications (BBAA VI),

Polytechnic University of Milan, Milan, Italy, ISBN 88-901916-3-5, 152-155.

166

Caracoglia, L. (2011). "Simulation of linear and non-linear propagation effects of a random

turbulence field on bridge flutter instability." Journal of Wind Engineering and

Industrial Aerodynamics, 99(9), 945-954.

Caracoglia, L., and Jones, N. P. (2003). "Time domain vs. frequency domain characterization

of aeroelastic forces for bridge deck sections." Journal of Wind Engineering and

Industrial Aerodynamics, 91(3), 371-402.

Caracoglia, L., and Velazquez, A. (2007). "Sensitivity of long-span bridge buffeting response

predictions to uncertainty in the definition of selected flutter derivatives." 12th

International Conference on Wind Engineering (12-ICWE), Cairns, Australia, July

1-6, 2007, 135-142.

Chowdhury, A. G., and Sarkar, P. P. (2003). "A new technique for identification of eighteen

flutter derivatives using a three-degree-of-freedom section model." Engineering

Structures, 25(12), 1763–1772.

Davenport, A. G. (1962). "Buffeting of a suspension bridge by storm winds." Journal of

Structural Division, ASCE, 88(ST3), 233-264.

Davenport, A. G. (1964). "Note on the distribution of the largest value of a random function

with application to gust loading." Journal of the Institution of Civil Engineers, 24,

187-196.

Davenport, A. G. (1967). "Gust loading factors." Journal of Structural Division, ASCE,

93(ST3), 11-34.

Denoël, V. (2009). "Estimation of modal correlation coefficients from background and

resonant responses." Structural Engineering and Mechanics, 32(6), 725-740.

Dragomirescu, E., Miyata, T., Yamada, H., and Katsuchi, H. (2003). "Probabilistic approach

of structural reliability applied to bridge components." 11th International Conference

on Wind Engineering (11-ICWE), Taxas tech University, Lubbock, Texas, USA, 789-

796.

Ehsan, F., Jones, N. P., and Scanlan, R. H. (1993). "Effect of sidewalk vents on bridge

response to wind." Journal of Structural Engineering, 119(2), 484-504.

Ellingwood, B. R. (2000). "LRFD: Implementing structural reliability in professional

practice." Engineering Structures, 22(2), 106-115.

167

Ellingwood, B. R., Rosowsky, D. V., Li, Y., and Kim, J. H. (2004). "Fragility assessment of

light-frame wood construction subjected to wind and earthquake hazards." Journal

of Structural Engineering, 130(12), 1921-1930.

Filiben, J. J., Gurley, K., Pinelli, J.-P., and Simiu, E. (2002). "Fragility curves, damage

matrices, and wind induced loss estimation." Third International Conference on

Computer Simulation in Risk Analysis and Hazard Mitigation, Sintra, Portugal, 119-

126.

Filliben, J. J., Gurley, K., Pinelli, J.-P., and Simiu, E. (2002). "Fragility curves, damage

matrices, and wind induced loss estmation." Third International Conference on

'Computer Simulation in Risk Analysis and Hazard Mitigation', Sintra, Portugal, 119-

126.

Ge, Y. J., Xiang, H. F., and Tanaka, H. (2000). "Application of a reliability analysis model

to bridge flutter under extreme winds." Journal of Wind Engineering and Industrial

Aerodynamics, 86(2-3), 155-167.

Grigoriu, M. (2002). Stochastic calculus. Applications in Science and Engineering,

Birkhäuser, Boston, MA, USA.

Gu, M., and Zhou, X.-Y. (2009). "An approximation method for resonant response with

coupling modes of structures under wind action." Journal of Wind Engineering and

Industrial Aerodynamics, 97(11-12), 573-580.

Haldar, A., and Mahadevan, S. (2000). Reliability assessment using stochastic finite-element

analysis, John Wiley and Sons, New York, NY, USA.

Huston, D. R., Bosch, H. R., and Scanlan, R. H. (1988). "The effects of fairings and of

turbulence on the flutter derivatives of a notably unstable bridge deck." Journal of

Wind Engineering and Industrial Aerodynamics, 29(1-3), 339-349.

Ibrahim, S. R., and Mikulcik, E. C. (1977). "Method for the direct identification of vibration

parameters from the free response." Shock and Vibration Bulletin(47).

Jain, A. (1996). "Multi-mode aeroelastic and aerodynamic analysis of long-span bridges,"

Dissertation, The Johns Hopkins University, Baltimore.

Jain, A., Jones, N. P., and Scanlan, R. H. (1996). "Coupled aeroelastic and aerodynamic

response analysis of long-span bridges." Journal of Wind Engineering and Industrial

Aerodynamics, 60(1-3), 69-80.

168

Jain, A., Jones, N. P., and Scanlan, R. H. (1998). "Effect of modal damping on bridge

aeroelasticity." Journal of Wind Engineering and Industrial Aerodynamics, 77-78,

421-430.

Jones, N. P., and Scanlan, R. H. (2001). "Theory and full-bridge modeling of wind response

of cable-supported bridges." Journal of Bridge Engineering, ASCE, 6(6), 365-375.

Katsuchi, H. (1997). "An analytical study on flutter and buffeting of the Akashi-Kaikyo

bridge," The Johns Hopkins University, Baltimore.

Katsuchi, H., Jones, N. P., and Scanlan, R. H. (1999). "Multimode coupled buffeting and

flutter analysis of the Akashi-Kaikyo Bridge." Journal of Structural Engineering,

ASCE, 120(8), 60-70.

Katsuchi, H., Jones, N. P., Scanlan, R. H., and Akiyama, H. (1998). "Multi-mode flutter and

buffeting analysis of the Akashi-Kaikyo bridge." Journal of Wind Engineering and

Industrial Aerodynamics, 77-78, 431-441.

Kwon, S.-D. (2010). "Uncertainty of bridge flutter velocity measured at wind tunnel tests."

The Fifth International Symposium on Computational Wind Engineering (CWE2010),

Chapel Hill, North Carolina, USA, CD-Rom.

Mannini, C., and Bartoli, G. (2007). "A probabilistic approach to bridge deck flutter."

ICWE12 CAIRNS, Australia, 2351-2358.

Mannini, C., Bartoli, G., and Borri, C. (2012). "New developments in bridge flutter

analysis." Proceedings of the Institution of Civil Engineers: Structures and Buildings,

165(3), 139-159.

Meirovitch, L. (1970). Analytical methods in vibrations, MCGraw-Hill, New York, NY,

USA.

Miyata, T., Yamada, H., Katsuchi, H., and Kitagawa, M. (2002). "Full-scale measurements

of Akashi-Kaikyo Bridge during typhoon." Journal of Wind Engineering and

Industrial Aerodynamics, 90, 1517-1527.

Namini, A., Albrecht, P., and Bosch, H. (1992). "Finite element-based flutter analysis of

cable-suspended bridges." Journal of Structural Engineering, 118(6), 1509-1526.

Newland, D. E. (1993). Random vibrations, spectral & wavelet analysis, 3rd edition Ed.,

Longmann, Edinburgh, UK.

169

NOAA. (http://www.ndbc.noaa.gov/). National Data Buoy Center.

Norton, T., Abdullah, M., M. , and Stephens, D. (2008). "Proposed methodology for

performance-based vulnerability assessment of wind-excited tall buildings." Fourth

International Conference on 'Advances in Wind and Structures (AWAS'08)', Techno-

Press, Korea, ISBN 978-89-89693-23-9-98530, Jeju, South Korea, 1228-1246.

Ostenfeld-Rosenthal, P., Madsen, H. O., and Larsen, A. (1992). "Probabilistic flutter criteria

for long span bridges." Journal of Wind Engineering and Industrial Aerodynamics,

42(1-3), 1265-1276.

Ozkan, E. (2003). "Evaluation of response prediction methodology for long-span bridges

using full-scale measurements," Ph.D. Dissertation, The Johns Hopkins University,

Baltimore, MD, USA.

Ozkan, E., and Jones, N. P. (2003). "Evaluation of response prediction methodology for

long-span bridges using full-scale measurements." 11th Int. Conference on Wind

Engineering, Lubbock, TX, USA, 1407-1414.

Pourzeynali, S., and Datta, T. K. (2002). "Control of flutter of suspension bridge using

TMD." Wind and Structures, 5(5), 407-422.

Robert, C. P., and Casella, G. (2004). Monte Carlo statistical methods (2nd ed.), Springer

Science, New York, New York, USA.

Sarkar, P., Caracoglia, L., Haan, F. L., Sato, H., and Murakoshi, J. (2009). "Comparative and

sensitivity study of flutter derivatives of selected bridge deck sections. Part 1:

Analysis of inter-laboratory experimental data." Engineering Structures, 31(1), 158-

169.

Saxena, V., Deodatis, G., Shinozuka, M., and Feng, M. Q. (2000). "Development of fragility

curves for multi-span reinforced concrete bridges." International Conference on

Monte Carlo Simulation, Balkema, Principality of Monaco.

Scanlan, R. H. (1987). "Interpreting aeroelastic models of cable-stayed bridges." Journal of

Engineering Mechanics, 113(4), 555-575.

Scanlan, R. H. (1993). "Bridge buffeting by skew winds in erection stages." Journal of

Engineering Mechanics, 119(2), 251-269.

170

Scanlan, R. H. (1999). "Estimates of skew wind speeds for bridge flutter." Journal of Bridge

Engineering, 4(Compendex), 95-98.

Scanlan, R. H., Béliveau, J. G., and Budlong, K. S. (1974). "Indicial Aerodynamic Functions

for bridge decks." Journal of Engineering Mechanics, ASCE, 100(EM4), 657-673.

Scanlan, R. H., and Jones, N. P. (1990a). "Aeroelastic analysis of cable-stayed bridges."

Journal of structural Engineering, ASCE, 116(2), 279-297.

Scanlan, R. H., and Jones, N. P. (1990b). "A minimum design methodology for evaluating

bridge flutter and buffeting response." Journal of Wind Engineering and Industrial

Aerodynamics, 36(1-3), 1341-1353.

Scanlan, R. H., and Tomko, J. J. (1971). "Airfoil and bridge deck flutter derivatives." Journal

of Engineering Mechanics, ASCE, 97(EM6), 1717-1737.

Seo, D.-W., and Caracoglia, L. (2011). "Estimation of torsional-flutter probability in flexible

bridges considering randomness in flutter derivatives." Engineering Structures, 33(8),

2284-2296.

Seo, D.-W., and Caracoglia, L. (2012). "Statistical buffeting response of flexible bridges

influenced by errors in aeroelastic loading estimation." Journal of Wind Engineering

and Industrial Aerodynamics, 104-106, 129-140.

Simiu, E., and Scanlan, R. H. (1996). Wind effects on structures, 3rd edition Ed., John Wiley

and Sons, New York, NY, USA.

Singh, L. (1997). "Experimental determination of aeroelastic parameters of long-span

bridges," The Johns Hopkins University, Baltimore, USA.

Smith, M. A., and Caracoglia, L. (2011). "A Monte Carlo based method for the dynamic

“fragility analysis” of tall buildings under turbulent wind loading." Engineering

Structures, 33(2), 410-420.

Strømmen, E. N. (2006). Theory of bridge aerodynamics, Springer-Verlag, Berlin-

Heidelberg, D.

Tempo, R., Calafiore, G., and Dabbene, F. (2005). Randomized algorithms for analysis and

control of uncertain systems, Springer-Verlag London Limited, London.

171

Theodorsen, T. (1935). "General theory of aerodynamic instability and the mechanism of

flutter." Technical Report 496, National Advisory Committee for Aeronautics,

Washington, DC, USA.

Walpole, R., Myers, R., Myers, S., and Ye, K. (2002). Probability and Statistics for

Engineers and Scientists, Prentice Hall, Upper Saddle River, NJ USA.

Wen, Y. K., and Kang, Y. J. (2001). "Minimum building life-cycle cost design criteria. I:

Methodology." Journal of Structural Engineering, 127(3), 330-337.

Xu, Y. L., Guo, W. W., Chen, J., Shum, K. M., and Xia, H. (2007). "Dynamic response of

suspension bridge to typhoon and trains. I: Field measurement results." Journal of

Structural Engineering, 133(1), 3-11.

Xu, Y. L., and Zhu, L. D. (2005). "Buffeting response of long-span cable-supported bridges

under skew winds. Part 2: case study." Journal of Sound and Vibration, 281(3-5),

675-697.