Dynamic performance of bridges and vehicles under strong wind

209
Louisiana State University LSU Digital Commons LSU Doctoral Dissertations Graduate School 2004 Dynamic performance of bridges and vehicles under strong wind Suren Chen Louisiana State University and Agricultural and Mechanical College Follow this and additional works at: hps://digitalcommons.lsu.edu/gradschool_dissertations Part of the Civil and Environmental Engineering Commons is Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please contact[email protected]. Recommended Citation Chen, Suren, "Dynamic performance of bridges and vehicles under strong wind" (2004). LSU Doctoral Dissertations. 1949. hps://digitalcommons.lsu.edu/gradschool_dissertations/1949

Transcript of Dynamic performance of bridges and vehicles under strong wind

Page 1: Dynamic performance of bridges and vehicles under strong wind

Louisiana State UniversityLSU Digital Commons

LSU Doctoral Dissertations Graduate School

2004

Dynamic performance of bridges and vehiclesunder strong windSuren ChenLouisiana State University and Agricultural and Mechanical College

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_dissertations

Part of the Civil and Environmental Engineering Commons

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion inLSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please [email protected].

Recommended CitationChen, Suren, "Dynamic performance of bridges and vehicles under strong wind" (2004). LSU Doctoral Dissertations. 1949.https://digitalcommons.lsu.edu/gradschool_dissertations/1949

Page 2: Dynamic performance of bridges and vehicles under strong wind

DYNAMIC PERFORMANCE OF BRIDGES AND VEHICLES

UNDER STRONG WIND

A Dissertation

Submitted to the Graduate Faculty of the Louisiana State University and

Agricultural and Mechanical College in partial fulfillment of the

requirements for the degree of Doctor of Philosophy

in

The Department of Civil and Environmental Engineering

By Suren Chen

B.S., Tongji University, 1994 M.S., Tongji University, 1997

May 2004

Page 3: Dynamic performance of bridges and vehicles under strong wind

DEDICATION

To my parents, my wife and my son

ii

Page 4: Dynamic performance of bridges and vehicles under strong wind

ACKNOWLEDGMENTS

I am indebted to Professor Steve Cai, my advisor, for his active mentorship, constant

encouragement, and support during my Ph. D study at LSU and KSU. It has been my greatest pleasure to work with such a brilliant, considerate and friendly scholar.

I also want to express my sincere gratitude to Professor Christopher J. Baker of The

University of Birmingham. The advice obtained from him on the vehicle accident assessment was very helpful and encouraging. The advice and help given by Dr. John D. Holmes on the time-history simulations are particularly appreciated. I also want to thank Professor Marc L. Levitan, the director of the Hurricane Center at LSU, for his very helpful courses on hurricane engineering and his great work as a member of my committee. Gratitude is also extended to Professor M. Gu at Tongji University and Professor C. C. Chang at Hong Kong University of Science and Technology for their continuous encouragement and support.

Thanks are also extended to my other committee members: Professor Dimitris E.

Nikitopoulos of Mechanical Engineering, Professor Jannette Frandsen of Civil Engineering, and Professor Jaye E. Cable of Oceanography & Coastal Sciences for very helpful suggestions in the dissertation.

The Graduate Assistantship offered by Louisiana State University and the National Science

Foundation (NSF) made it possible for me to proceed with my study. Last but not the least, I would like to thank my beloved wife and my son for their strong

support. The dissertation could not have been completed without their encouragement, their love and their patience.

iii

Page 5: Dynamic performance of bridges and vehicles under strong wind

TABLE OF CONTENTS

DEDICATION.................................................................................................................................ii ACKNOWLEDGMENTS ............................................................................................................ iii ABSTRACT.................................................................................................................................. vi CHAPTER 1.INTRODUCTION ................................................................................................... 1

1.1 Wind Hazard .......................................................................................................................1 1.2 Bridge Aerodynamics .........................................................................................................3 1.3 Vehicle Dynamic Performance on the Bridge under Wind.................................................5 1.4 Structural Control on Wind-induced Vibration of Bridges.................................................7 1.5 Present Research.................................................................................................................8

CHAPTER 2. MODAL COUPLING ASSESSMENTS AND APPROXIMATED PREDICTION OF COUPLED MULTIMODE WIND VIBRATION OF LONG-SPAN BRIDGES.................. 10

2.1 Introduction........................................................................................................................10 2.2 Mathematical Formulations...............................................................................................11 2.3 Approximated Prediction of Coupled Buffeting Response...............................................18 2.4 Numerical Example...........................................................................................................19 2.5 Concluding Remarks..........................................................................................................31

CHAPTER 3. EVOLUTION OF LONG-SPAN BRIDGE RESPONSE TO WIND- NUMERICAL SIMULATION AND DISCUSSION.................................................................. 33

3.1 Introduction .......................................................................................................................33 3.2 Motivation of Present Research ....................................................................................... 33 3.3 Analytical Approach......................................................................................................... 34 3.4 Numerical Procedure.........................................................................................................38 3.5 Numerical Example...........................................................................................................40 3.6 Concluding Remarks .........................................................................................................56

CHAPTER 4. DYNAMIC ANALYSIS OF VEHICLE-BRIDGE-WIND DYNAMIC SYSTEM..................................................................................................................................... . 58

4.1 Introduction....................................................................................................................... 58 4.2 Equations of Motion for 3-D Vehicle-Bridge-Wind System............................................ 59 4.3 Dynamic Analysis of Vehicle-Bridge System under Strong Wind...................................68 4.4 Numerical Example...........................................................................................................70 4.5 Concluding Remarks..........................................................................................................93 4.6 Matrix Details of the Coupled System…………………………………………………...94

CHAPTER 5. ACCIDENT ASSESSMENT OF VEHICLES ON LONG-SPAN BRIDGES IN WINDY ENVIRONMENTS..................................................................................................... 101

5.1 Introduction......................................................................................................................101 5.2 Dynamic Interaction of Non-Articulated Vehicles on Bridges....................................... 102

iv

Page 6: Dynamic performance of bridges and vehicles under strong wind

5.3 Accident Analysis Model for Vehicles on Bridges..........................................................105 5.4 Numerical Example.........................................................................................................113 5.5 Concluding Remarks....................................................................................................... 129

CHAPTER 6. STRONG WIND-INDUCED COUPLED VIBRATION AND CONTROL WITH TUNED MASS DAMPER FOR LONG-SPAN BRIDGES. .................................................... 131

6.1 Introduction......................................................................................................................131 6.2 Closed-Form Solution of Bridge-TMD System........................................................... .. 132 6.3 Coupled Vibration Control with a Typical 2DOF Model.............................................. .138 6.4 Analysis of a Prototype Bridge ...................................................................................... 143 6.5 Concluding Remarks....................................................................................................... 154

CHAPTER 7. OPTIMAL VARIABLES OF TMDS FOR MULTI-MODE BUFFETING CONTROL OF LONG-SPAN BRDGES................................................................................... 156

7.1 Introduction......................................................................................................................156 7.2 Formulations of Multi-mode Coupled Vibration Control with TMDs .......................... 157 7.3 Parametrical Studies on “Three-row” TMD Control ......................................................161 7.4 Concluding Remarks....................................................................................................... 177

CHAPTER 8. WIND VIBRATION MITIGATION OF LONG-SPAN BRIDGES IN HURRICANES.......................................................................................................................... 178

8.1 Introduction......................................................................................................................178 8.2 Equations of Motion of Bridge-SDS System ................................................................. 179 8.3 Solution of Flutter and Buffeting Response.....................................................................181 8.4 Numerical Example: Humen Bridge-SDS system.......................................................... 182 8.5 Concluding Remarks....................................................................................................... 187

CHAPTER 9. CONCLUSIONS AND FURTHER CONSIDERATIONS ................................189

9.1 Summary and Conclusions..............................................................................................189 9.2 Future Work ....................................................................................................................191

REFERENCES…………………………………………………………………………………193 VITA………................................................................................................................................201

v

Page 7: Dynamic performance of bridges and vehicles under strong wind

vi

ABSTRACT

The record of span length for flexible bridges has been broken with the development of

modern materials and construction techniques. With the increase of bridge span, the dynamic response of the bridge becomes more significant under external wind action and traffic loads. The present research targets specifically on dynamic performance of bridges as well as the transportation under strong wind.

The dissertation studied the coupled vibration features of bridges under strong wind. The current research proposed the modal coupling assessment technique for bridges. A closed-form spectral solution and a practical methodology are provided to predict coupled multimode vibration without actually solving the coupled equations. The modal coupling effect was then quantified using a so-called modal coupling factor (MCF). Based on the modal coupling analysis techniques, the mechanism of transition from multi-frequency type of buffeting to single-frequency type of flutter was numerically demonstrated. As a result, the transition phenomena observed from wind tunnel tests can be better understood and some confusing concepts in flutter vibrations are clarified.

The framework of vehicle-bridge-wind interaction analysis model was then built. With the interaction model, the dynamic performance of vehicles and bridges under wind and road roughness input can be assessed for different vehicle numbers and different vehicle types. Based on interaction analysis results, the framework of vehicle accident analysis model was introduced. As a result, the safer vehicle transportation under wind can be expected and the service capabilities of those transportation infrastructures can be maximized. Such result is especially important for evacuation planning to potentially save lives during evacuation in hurricane-prone area.

The dissertation finally studied how to improve the dynamic performance of bridges under wind. The special features of structural control with Tuned Mass Dampers (TMD) on the buffeting response under strong wind were studied. It was found that TMD can also be very efficient when wind speed is high through attenuating modal coupling effects among modes. A 3-row TMD control strategy and a moveable control strategy under hurricane conditions were then proposed to achieve better control performance.

Page 8: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 1. INTRODUCTION

The dissertation is made up of nine chapters based on papers that have either been accepted, or are under review, or are to be submitted to peer-reviewed journals, using the technical paper format that is approved by the Graduate School.

Chapter 1 introduces the related background knowledge of the dissertation, the research scope and structure of the dissertation. Chapter 2 discusses the modal coupling effect on bridge aerodynamic performances (Chen et al. 2004). Chapter 3 covers the evolution of the long-span bridge response to the wind (Chen and Cai 2003a). Chapter 4 discusses the dynamic analysis of the vehicles-bridge-wind system (Cai and Chen 2004a). Chapter 5 discusses the vehicle safety assessment of vehicles on long-span bridges under wind (Chen and Cai 2004a). Chapter 6 investigates the new features of strong-wind induced vibration control with Tuned Mass Dampers on long-span bridges (Chen and Cai 2004b). Chapter 7 studies the optimal variables of Tuned Mass Dampers on multiple-mode buffeting control (Chen et al. 2003). Chapter 8 investigates the wind vibration mitigation on long-span bridges in hurricane conditions (Cai and Chen 2004b). Chapter 9 summarizes the dissertation and gives some suggestions for future research.

This introductory chapter gives a general background related to the present research. More detailed information can be seen in each individual chapter.

1.1 Wind Hazard

Wind is about air movement relative to the earth, driven by different forces caused by pressure differences of the atmosphere, by different solar heating on the earth’s surface, and by the rotation of the earth. It is also possible for local severe winds to be originated from local convective effects and the uplift of air masses. Wind loading competes with seismic loading as the dominant environmental loading for modern structures. Compared with earthquakes, wind loading produces roughly equal amounts of damage over a long time period (Holmes, 2001). The major wind storms are usually classified as follows:

Tropical cyclones: Tropical cyclones belong to intense cyclonic storms which usually occur over the tropical oceans. Driven by the latent heat of the oceans, tropical cyclones usually will not form within about 5 degrees of the Equator. Tropical cyclones are called in different names around the world. They are named hurricanes in the Caribbean and typhoons in the South China Sea and off the northwest coast of Australia (Holmes, 2001).

Thunderstorm: Thunderstorms are capable of generating severe winds, through tornadoes and downbursts. They contribute significantly to the strong gusts recorded in many countries, including the United States, Australia and South Africa. They are also the main source of high winds in the equatorial regions (within about 10 degrees of the Equator), although their strength is not high in these regions (Holmes, 2001; Simiu and Scanlan, 1986).

Tornadoes: These are larger and last longer than “ordinary” convection cells. The tornado, a vertical, funnel-shaped vortex created in thunderclouds, is the most destructive of wind storms. They are quite small in their horizontal extent-of the order of 100 m. However, they

1

Page 9: Dynamic performance of bridges and vehicles under strong wind

can travel for quite a long distance, up to 50 km, before dissipating, producing a long narrow path of destruction. They occur mainly in large continental plains, and they have very rarely passed over a weather recording station because of their small size (Holmes, 2001).

Downbursts: Downbursts have a short duration and also a rapid change of wind direction during their passage across the measurement station. The horizontal wind speed in a thunderstorm downburst, with respect to the moving storm, is similar to that in a jet of fluid impinging on a plain surface (Holmes, 2001).

Damage to buildings and other structures caused by wind storm has been a fact of life for human beings since these structures appeared. In nineteenth century, steel and reinforcement were introduced as construction materials. During the last two centuries, major structural failures due to wind action have occurred periodically and provoked much interest in wind loadings by engineers. Long-span bridges often produced the most spectacular of these failures, such as the Brighton Chain Pier Bridge in England in 1836, the Tay Bridge in Scotland in 1879, and the Tacoma Narrows Bridge in Washington State in 1940. Besides, other large structures have experienced failures as well, such as the collapse of the Ferrybridge cooling tower in the U. K. in 1965, and the permanent deformation of the columns of the Great Plains Life Building in Lubbock, Texas, during a tornado in 1970. Based on annual insured losses in billions of US dollars from all major natural disasters, from 1970 to 1999, wind storms account for about 70% of total insured losses (Holmes, 2001). This research addresses transportation-related issues due to hurricane-induced winds.

Hurricanes and hurricane-induced strong wind are, by many measures, the most devastating of all catastrophic natural hazards that affect the United States. The past two decades have witnessed exponential growth in damage due to hurricanes, and the situation continues to deteriorate. The most vulnerable areas, coastal countries along the Gulf and Atlantic seaboards, are experiencing greater population growth and development than anywhere else in the country. In the United States, annual monetary losses due to tropical cyclones and other natural hazards have been increasing at an exponential pace, now averaging up to $1 billion a week (Mileti, 1999). Large hurricanes can have impacts that are national or even international in scope. Damage from Hurricane Andrew was so extensive (total loss approximately $25 billion) that it caused building materials shortages nationwide and bankrupted many Florida insurance companies. Had Andrew’s track shifted just a few miles, it could have gone through downtown Miami, hit Naples on the west coast of Florida, and then devastated New Orleans. Projections for the total losses in this scenario are several times greater than the $25 billion in damages caused by Andrew. Losses of this magnitude threaten the stability of national and international reinsurance markets, with potentially global economic consequences. When a hurricane or tropical storm does strike the gulf coast, the results are generally devastating.

In additional to huge loss of property, loss of life is even more stunning. Compared to the U. S., developing countries which lack predicting and warning systems are suffering even more from hurricane-associated hazards. The cyclone in October 1999 killed tens of thousands in India, and Hurricane Mitch killed thousands in Honduras in 1998. Even as storm prediction and tracking technologies improve, providing greater warning times, the U. S. is still becoming ever more susceptible to the effects of hurricanes, due to the massive population growth in the South and Southeast along the hurricane coast from Texas to Florida to the Carolinas. This growth has

2

Page 10: Dynamic performance of bridges and vehicles under strong wind

spurred tremendous investments in areas of greatest risk. The transportation infrastructure has not increased capacity at anything like a similar pace, necessitating longer lead times for evacuations and forcing some communities to adopt a shelter-in-place concept. This concept recognizes that it will not be possible for everyone to evacuate, so only those in areas of greatest risk from storm-surge are given evacuation orders.

New Orleans is a typical example of the hurricane-prone cities in the United States. Due to the fact that most of the city is at or below sea level, protected only by levees, it has been estimated that a direct hit by a Category 3 or larger hurricane will “fill the bowl”, submerging most of the city in 20 feet or more of water (Fischetti 2001). In extreme cases, evacuations are essential to minimize the loss of lives and properties. In New Orleans, four of the five major evacuation routes out of the city include highway bridges over open water. The Louisiana Office of Emergency Preparedness estimates that under current conditions, there will be time to evacuate only 60-65% of the 1.3 million Metro area populations in the best-case scenario, with a 10% casualty rate for those remaining in the city.

To ensure a successful evacuation, smooth transportation is the key to the whole evacuation process. There are two categories of problems to be dealt with: the safety and efficient service of the transportation infrastructures, such as bridges and highways; the safe operation of vehicles on those transportation infrastructures (Baker 1994; Baker and Reynolds 1992). It is very obvious that maximizing the opening time of the evacuation routes as the storm approaches is very important. The present study investigates these two kinds of problems.

1.2 Bridge Aerodynamics

The record of span length for flexible structures, such as suspension and cable-stayed bridges, has been broken with the development of modern materials and construction techniques. The susceptibility to wind actions of these large bridges is increasing accordingly. The well-known failure of the Tacoma Narrows Bridge due to the wind shocked and intrigued bridge engineers to conduct various scientific investigations on bridge aerodynamics (Davenport et al. 1971, Scanlan and Tomko 1977, Simiu and Scanlan 1996, Bucher and Lin 1988). In addition to the Tacoma Narrows Bridge, some existing bridges, such as the Golden Gate Bridge, have also experienced large, wind-induced oscillations and were stiffened against aerodynamic actions (Cai 1993). Basically, three approaches are currently used in the investigation of bridge aerodynamics: the wind tunnel experiment approach, the analytical approach and the computational fluid dynamics approach.

Wind Tunnel Experiment Approach: The wind tunnel experiment approach tests the scaled model of the structure in the wind tunnel laboratory to simulate and reproduce the real world. Wind tunnel tests can either be used to predict the performance of structures in the wind or be used to verify the results from other approaches. The wind tunnel experiment approach is designed to obtain all the dynamic information of the structure with wind tunnel experiments. Bluff body aerodynamics emphasizes on flows around sharp corners, or separate flows. Simulating the atmospheric flows with characteristics in the wind tunnel similar to those of natural wind is usually required in order to investigate the wind effect on the structures. For such purposes, the wind environment should be reproduced in a similar manner, and the structures should be modeled with similarity criteria (Simiu and Scanlan 1986). To achieve similarity between the model and the prototype, it is desirable to reproduce at the requisite scale the

3

Page 11: Dynamic performance of bridges and vehicles under strong wind

characteristics of atmospheric flows expected to affect the structure of concern. These characteristics include: (1) the variation of the mean wind speed with height; (2) the variation of turbulence intensities and integral scales with height; and (3) the spectra and cross-spectra of turbulence in the along-wind, across-wind, and vertical directions.

Wind tunnels used for civil engineering are referred to as long tunnels, short wind tunnels and tunnels with active devices. The long wind tunnels, a boundary layer with a typical depth of 0.5 m to 1 m, develop naturally over a rough floor of the order of 20 m to 30 m in length. The depth of the boundary layer can be increased by placing passive devices at the test section entrance. Atmospheric turbulence simulations in long wind tunnels are probably the best that can be achieved currently. The short wind tunnel has the short test section, and is ideal for tests under smooth flow, as in aeronautical engineering. To be used in civil engineering applications, passive devices, such as grids, barriers, fences and spires usually should be added in the test section entrance to generate a thick boundary layer (Simiu and Scanlan 1986). The wind tunnel approach totally relies on the experiments in the laboratory and may be very expensive and time-consuming.

Analytical Approach: Another way is to build up analytical models based on the insight of aerodynamic aspects of the structure obtained from the wind tunnel tests, as well as knowledge of structural dynamics and fluid mechanics. With the models, the dynamic performance of the structure can be predicted numerically. However, although the science of theoretical fluid mechanics is well developed and computational methods are experiencing rapid growth in the area, it still remains necessary to perform physical wind tunnel experiments to gain necessary insights into many aspects associated with fluid. So the analytical approach is actually a hybrid approach of numerical analysis and wind tunnel tests. Due to its convenient and inexpensive nature, the analytical approach is adopted in most cases. The dissertation also uses the analytical approach to carry out all the research.

Computational Fluid Dynamics (CFD): Computational fluid dynamics (CFD) techniques have been under development in wind engineering for several years. Since this topic is out of the scope for the dissertation, no comprehensive review is intended here.

Long cable-stayed and suspension bridges must be designed to withstand the drag forces induced by the mean wind. In addition, such bridges are susceptible to aeroelastic effects, which include torsion divergence (or lateral buckling), vortex-induced oscillation, flutter, galloping, and buffeting in the presence of self-excited forces (Simiu and Scanlan 1986). The aeroelastic effects between the bridge deck and the moving air are deformation dependent, while the aerodynamic effects are induced by the forced vibration from the turbulence of the air. Usually divergence, galloping and flutter are classified as aerodynamic instability problems, while vortex shedding and buffeting are classified as wind-induced vibration problems. All these phenomena may occur alone or in combination. For example, both galloping and flutter only happen under certain conditions. At the mean time, the wind-induced vibrations, like vortex shedding and buffeting may exist. The main categories of wind effects on bridges with boundary layer flow theory are flutter and buffeting. While flutter may result in dynamic instability and the collapse of the whole structures, large buffeting amplitude may cause serious fatigue damage to structural members or noticeable serviceability problems.

4

Page 12: Dynamic performance of bridges and vehicles under strong wind

Typically, to deal with these wind-induced problems, a scaled-down bridge model is tested in a wind tunnel for two purposes. First is to observe the aerodynamic behavior and then to develop some experimentally-based countermeasures (Huston 1986). Second is to measure some aerodynamic coefficients, such as flutter derivatives and static force coefficients, in order to establish reasonable analytical prediction models (Tsiatas and Sarkar 1988, Scanlan and Jones 1990, Namini et al. 1992).

Long-span bridges are often the backbones of the transportation lines in coastal areas and are vulnerable to wind loads. Maintaining the highest transportation capacity of these long-span bridges, such as the Luling Bridge near New Orleans, Louisiana, and the Sunshine Skyway Bridge near Tampa, Florida, is vital to supporting hurricane evacuations. Due to the aeroelastic and aerodynamic effects from high winds on long-span bridges, strong dynamic vibrations will be expected. Excessive vibrations will cause the service and safety problems of bridges (Conti et al. 1996; Gu et al. 1998, 2001, 2002). Stress induced from dynamic response may also cause fatigue accumulation on some local members and damages to some connections. With the increase of wind speed, the aerodynamic stability of the bridge may also become a problem. In extreme high wind speed, the aerodynamic instability phenomenon, flutter, may happen. As a result of flutter, the bridge may collapse catastrophically (Amann et al. 1941).

1.3 Vehicle Dynamic Performance on the Bridge under Wind

Economic and social developments increase tremendously the traffic volume over bridges and roads. Heavy vehicles on bridges may significantly change the local dynamic behavior and affect the fatigue life of the bridge. On the other hand, the vibrations of the bridge under wind loads also in turn affect the safety of the vehicles. For vehicles running on highway roads, the wind loading on the vehicle, as well as grade and curvature of the road, may cause safety and comfort problems (Baker 1991a-c, 1994).

Interaction analysis between moving vehicles and continuum structures originated in the middle of 20th century. From an initial moving load simplification (Timishenko et al. 1974), to a moving-mass model (Blejwas et al. 1979) to full-interaction analysis (Yang and Yau 1997; Pan and Li 2002; Guo and Xu 2001), the interaction analysis of vehicles and continuum structures (e.g. bridges) has been investigated by many people for a long time. In these studies, road roughness was treated as the sole excitation source of the coupled system.

Recently, the dynamic response of suspension bridges to high wind and a moving train has been investigated (Xu et al. 2003), while no wind loading on the train was considered since the train was moving inside the suspension bridge deck. It was found that the suspension bridge response is dominated by wind force in high wind speed. The bridge motions due to high winds considerably affected the safety of the train and the comfort of passengers (Xu et al. 2003). The coupled dynamic analysis of vehicle and cable-stayed bridge system under turbulent wind was also conducted recently (Xu and Guo 2003). However, only vehicles under low wind speed were explored, and the study did not consider many important factors, such as vehicle number, and driving speeds.

Studied on the wind effects on ground vehicles were mainly focused on cross wind. The cross wind effect can be broadly considered as two types: (1) low wind speed effects, such as an increase of drag coefficient and vehicle aerodynamic stability considerations; and (2) high wind

5

Page 13: Dynamic performance of bridges and vehicles under strong wind

speed effects (Baker 1991a). The latter is, of most concern to researchers, is composed of variety of forms. For example, the suspension modes of vehicles may be excited by the strong cross-wind if the wind energy is enough around the modal frequency of the suspension system. It is believed that there are three major types of accidents for wind-induced vehicle accidents in high wind: overturning accidents, sideslip accidents and rotation accidents. It is possible for high-sided vehicles to be overturned, especially where some grade on section and road curvature exists (Coleman and Baker 1990; Sigbjornsson and Snajornsson 1998; Baker 1991a-c, 1994). Even for small vehicles, like vans and cars, severe course deviation at gusty sites may occur (Baker 1991 a-b). For vehicles traveling on bridges, the problems are even more complicated because the bridge itself is a kind of dynamic-sensitive structure in a strong wind (Bucher and Lin 1988; Cai and Albrecht 2000; Cai et al. 1999a). The interactive effect between the bridge and vehicles makes the assessment of the vehicle performance on bridges more difficult.

In the 1990’s, many vehicle accidents were reported around the world (Baker and Reynolds 1992; Coleman and Baker 1990; Sigbjornsson and Snajornsson 1998). From the statistics of a great many accidents which occurred during the severe storm on Jan 25 1990 in British, it is reported that overturning accidents were the most common type of wind-induced accidents, accounting for 47% of all accidents. Course deviation accidents made up 19% of the total accidents, and accidents involving trees made up 16%. Among all accidents, 66% involved high-sided lorries and vans, and only 27% involved cars (Baker and Reynolds 1992).

Safety study of vehicles under wind on highways began in the 1980s. In his representative work, Baker (Baker 1991 a-c) proposed the fundamental equations for wind action on vehicles. Without considering the driver’s performance, the wind speed at which these accident criteria are exceeded (the so-called accident wind speed) was found to be a function of vehicle speed and wind direction. Using meteorological information, the percentage of the total time for which this wind speed is exceeded can be found. Some quantification of accident risk has been made. Based on Baker’s model, Sigbjornsson and Snajornsson (1998) tried the risk assessment of an accident which happened in Iceland through a reliability approach. In the approach, the so-called “safety index approach” was adopted to describe the risk of the accident and some valuable results were given. However, the driver’s behavior was not included in this analysis, which made the analysis not general enough for further application. As in Baker’s study, Sigbjornsson did not consider road roughness, which could be a very important factor affecting the dynamic behavior of vehicles on highway roads. In addition, no study has included vehicle performance on a bridge under strong wind.

During periods of high winds, it is a common practice to either slow down traffic or stop vehicle movement altogether at exposed windy sites such as long span bridges (Baker 1987). Various criteria used to initiate traffic control differ substantially between sites. In some places, the “two-level” system is operated; vehicles are slowed down, and warning signs are activated when the wind speed exceeds a certain level. It seems that the various criteria for imposing traffic restrictions are, to an extent, rather arbitrary and ill-defined (Baker 1987). Baker (Baker 1986) suggested that an overturning accident may occur if, within the 0.5 s of the vehicle entering the gust, one of the tire reactions fell to zero. A side-slipping accident happens when the lateral displacement exceeded 0.5m, and a rotational accident happens when the rotational displacement exceeds 0.2 radians. These definitions are based on some assumptions, especially that within 0.5s, the driver of the vehicle would not react to correct the lateral and rotation

6

Page 14: Dynamic performance of bridges and vehicles under strong wind

displacement. With the introduction of the driver behavior model, the assumption of 0.5 second can be eliminated (Baker 1994).

1.4 Structural Control on Wind-induced Vibration of Bridges

When extreme wind such as a hurricane attacks the long span bridges, large vibration response may force the bridge to be closed to transportation for the safety of the bridge and of vehicles. An efficient control system is needed to guarantee the safety of the bridge, to prolong the opening time for traffic under mitigation conditions, and to reduce direct and indirect financial loss as well as civilian lives. In the past few decades, bridge engineers have had modest success in dealing with wind-induced vibrations by structural strengthening or streamlining/modifying the bridge geometry based on the results of experimental /analytical studies. However, with the increase in bridge-span length, wind-structure interaction is becoming increasingly important. The Akashi Kaikyo suspension bridge in Japan has set a new record of 1900 m in span length. An even longer span length of 5000 m for the Gibraltar Strait Bridge is under discussion (Wilde et. al 2001). Wind-induced vibrations become one of the major controlling factors in long-span bridge design. To meet the serviceability and strength requirements under wind actions, structural control will be the most economical and the only feasible alternative for these ultra-long-span bridges (Anderson and Pederson 1994).

The control strategies for wind-induced vibration of long-span bridges can be classified as structural (passive), aerodynamic (passive or active), or mechanical (passive or active) countermeasures. Passive structural countermeasures aim at increasing the stiffness of the structures by increasing member size, adding additional members, or changing the arrangement of structural members.

Passive aerodynamic countermeasures focus on selecting bridge deck shapes and details to satisfy aerodynamic behaviors. Examples are using shallow sections, closed sections, edge streamlining and other minor or subtle changes to the cross-section geometry (Wardlaw 1992). It was found that the aerodynamic countermeasures are more efficient than structural strengthening (Cai et al. 1999b). However, these passive aerodynamic countermeasures are not adequate for ultra long-span bridges and in extreme high wind, such as a hurricane.

Active aerodynamic countermeasures use adjustable control surfaces for increasing critical flutter wind velocity (Ostenfeld and Larsen 1992, Predikman and Mook 1997, Wilde and Fujino 1998, 2001). By adjusting the rotation of these control surfaces to a predetermined angle, stabilizing aerodynamic forces are generated. One of the disadvantages is that the control efficiency is sensitive to the rotation angles. A wrong direction of rotation, due to either the failure of control system or inaccurate theoretical predictions, will have detrimental effects on the bridge. Predicting a bridge’s performance under hurricane wind and designing a reliable/efficient control mechanism is still extremely difficult.

Mechanical countermeasures focus mainly on the flutter and buffeting controls with passive devices, such as tuned mass dampers (TMD). A TMD consists of a spring, a damper and a mass. It is easy to design and install and has been used in the vibration control of some buildings and bridges, such as Citicorp Center in New York, John Hancock tower in Boston, and the Normandy Bridge in France. In a typical passive TMD system, the natural frequency of the TMD is tuned to a pre-determined optimal frequency that is dependent on the dynamic

7

Page 15: Dynamic performance of bridges and vehicles under strong wind

characteristics of the bridge system and wind characteristics (Gu et al. 1998, 2001, Gu and Xiang 1992).

1.5 Present Research

The present research discusses the safety issues of long-span bridges and transportation under wind action. It covers three interrelated parts: Part I - multimode coupled vibration of long-span bridges in strong wind; Part II – vehicle-bridge-wind interaction and vehicle safety; and Part III - bridge vibration control under strong wind.

Chapters 2 and 3 are devoted to Part I. With the increase of bridge span, the dynamic response of the bridge becomes more significant under external wind action and traffic loads. Longer bridges usually have closer mode frequencies than those of short-span bridges. Under the action of aeroelastic and aerodynamic forces, the response component contributed by one mode may affect the aeroelastic effects on another mode when their frequencies are close, due to aerodynamic coupling. Such coupling effects among modes are usually gradually strengthened when wind speed is high. As an important phenomenon for long-span bridges under strong wind, modal coupling of a bridge under wind action is assessed in Chapter 2. A practical approximation approach of predicting the coupled response of the bridge under wind action is also introduced. Another part of research in the bridge aerodynamic is to investigate the phenomena of buffeting and flutter of bridges. Buffeting and flutter are usually treated as two different phenomena. Buffeting is believed to be a forced vibration, and flutter is usually treated as a system instability phenomena. Chapter 3 aims at establishing the connections between buffeting and flutter phenomena. It is believed in the current work that the two phenomena are continuous, and the evolution process from buffeting to flutter is investigated. In the meantime, the hybrid analysis approach introduced will also benefit the following analysis of vehicle-bridge-wind interaction analysis.

Based on the work of Part I, Chapters 4 and 5 develop the analytical framework of vehicle-bridge-wind interaction (Part II). The previous works are mainly focused on the dynamic performance of vehicles on the road under wind action, or vehicle dynamic performance on the bridge without wind or with slight wind. In Chapter 4, a general analysis model is built for dynamic coupling analysis of a bridge and vehicles. With finite-element based dynamic analysis results, modal coupling assessment techniques introduced in Chapter 2 are adopted to choose those important modes to be included in the analysis. With limited key modes and the vehicle dynamic model, the dynamic response of the moving vehicle can be predicted at any time on the bridge under wind action. Part of the obtained dynamic response of vehicles is to be used in the following vehicle safety assessment part. Regarding vehicle safety assessment, all previous works are only for vehicles on the road. In Chapter 5, the general model of vehicle safety assessment on the bridge is introduced. With the safety assessment model, the accident risks corresponding to different accident types are assessed under different situations.

Chapters 6, 7, and 8 form Part III - vibration control. Bridges exhibit large dynamic response under strong wind. Excessive response may cause the safety problem of the bridge known as flutter. It may also cause serviceability problems and fatigue accumulation. So, in some circumstances, structural control is an important way to enhance the safety, serviceability, and durability of a bridge. In Chapter 6, the special features of structural control with Tuned Mass Dampers (TMD) on the buffeting response under strong wind are studied. In addition to

8

Page 16: Dynamic performance of bridges and vehicles under strong wind

9

the well-known resonant suppression mechanism of TMD, it is also found that the TMD can be used to suppress the strong coupling effects among modes of a bridge when the wind speed is high. Such a new mechanism enables TMD to control the bridge buffeting efficiently even when wind speed is high. Following the work of Chapter 6, a 3-row TMD control strategy is proposed to achieve better control performance in Chapter 7. Finally, in Chapter 8, a moveable control strategy is introduced to facilitate the vibration control of long-span bridges under hurricane.

Page 17: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 2. MODAL COUPLING ASSESSMENTS AND APPROXIMATED PREDICTION OF COUPLED MULTIMODE WIND VIBRATION OF LONG-SPAN

BRIDGES

2.1 Introduction

Bridges with record-breaking span lengths are currently being designed or expected worldwide in the future. For example, Messina Straits Bridge with a span length of 3,300 m is under design and Gibraltar Straits Crossing with a span length of 5,000 m is under discussion. Lighter and more aerodynamically profiled cross sections of decks are most commonly used for these increasingly more flexible long-span bridges. In these circumstances, the structural characteristics of the bridges tend to result in closer modal frequencies. As a result, the possibilities of modal coupling through aerodynamic and aeroelastic effects increase (Chen et al. 2000).

The mechanisms of modal coupling in the wind-induced vibrations of bridges have been studied (Cai et al. 1999; Namini et al. 1992; Thorbek and Hansen 1998; Katsuchi et al. 1998; Bucher and Lin 1988). Thorbek and Hansen (1998) once focused on the effect of modal coupling between vertical and torsional modes on the buffeting response and gave a general guideline regarding the need of including modal coupling in the calculations. They suggested that for suspension bridges with streamlined bridge deck sections and a ratio of 2-3 between the torsional and vertical natural frequencies in still air, the effect of modal coupling be taken into account in the calculation if the mean wind velocity exceeds approximately 60% of the critical flutter wind velocity. D'Asdia and Sepe (1998) and Brancaleoni and Diana (1993) also highlighted the importance of aeroelastic effects in the analysis of Messina Bridge and modal coupling effects were studied. Other previous works (Bucher and Lin 1988) also recognized the importance of conducting coupled multimode analysis.

For a long time, wind-induced buffeting response of bridges is obtained from the square root of the sum of the squares (SRSS) of single-mode responses (Simiu and Scanlan 1996), which is called “single-mode approach” hereafter. Single-mode response has been very attractive in engineering practice and preliminary analysis due to its convenience. However, since single-mode approach completely neglects the modal coupling effects among different modes, it sacrifices significantly the accuracy when modal coupling is not weak (Jain et al. 1996; Tanaka et al. 1994). In contrast, coupled multimode procedures take into account all the aeroelastic and aerodynamic coupling effects by solving simultaneous equations (Chen et al. 2000; Bucher and Lin 1988; Jain et al. 1996; Cai et al. 1999b). The calculation efforts vary with the total number of modes being included in the simultaneous equations.

Katsuchi et al. (1998, 1999) investigated the multimode behavior of the Akashi-Kaikyo Bridge with a span length of 1990 m by using the numerical procedures based on (Jain et al. 1996). The analytical results showed some strongly coupled aeroelastic behaviors that are consistent with the wind tunnel observations (Katsuchi et al. 1998, 1999). In their study, up to 25 and 17 modes were analyzed for the flutter and buffeting response, respectively. Through the comparison between the coupled multimode and single-mode analyses, it was found that only 6 key modes are important to the flutter analysis. Meanwhile, no significant difference

10

Page 18: Dynamic performance of bridges and vehicles under strong wind

appeared between the buffeting results of coupled multimode analysis and single-mode analysis until the 10th mode was included (Katsuchi et al. 1999). This implied that it is only a few key modes that are crucial for the accuracy of coupled multimode analysis.

However, a rational method of assessing the modal coupling effect and thus identifying these key modes is still absent. One does not know if a coupled multimode analysis is needed, or how many key modes should be included before an actual coupled multimode analysis is conducted. As a result, the modes that are included in the multimode analysis can only be selected based on subjective judgment of modal properties and flutter derivatives. Consequently, one may choose either too many or too few modes for the coupled analysis. The former may include many unessential modes while the latter may miss some important ones. In recent years, finite element method (FEM) has become more and more popular for the aerodynamic analysis of bridges (Cai et al. 1999; Namini et al. 1992). Assessing the necessity of coupled multimode analysis and then including the key modes in the analysis will achieve better computation efficiency.

While coupled multimode analysis is more accurate in predicting the wind-induced vibration than the single-mode analysis, it is still advantageous and desirable to develop a more convenient method that can balance the simplicity of single-mode analysis and the accuracy of coupled multimode analysis. This method can be used for practical applications or in cases when a complicated coupled multimode analysis is not desirable, such as in a preliminary analysis.

In the present study, firstly, a closed-form formulation of coupled multimode response is derived where only the primary modal coupling effects are considered while the trivial secondary ones are ignored. As will be seen later, this approximate method gives much more accurate results than the traditional single-mode analysis and agrees well with the coupled multimode analysis results. Secondly, the tendency of modal coupling effect is quantitatively assessed by using a modal coupling factor (MCF). Though the derivation of MCF is based on buffeting analysis, it can also disclose the nature of modal coupling mechanism that will govern the flutter behavior of long-span bridges. Lastly, another important application of the MCF is in the design of vibration control strategies (Gu et al. 2002a). An optimal control design may call for suppressing the response induced by modal coupling effect other than traditional resonant component in high wind speed. To develop such control strategy, it is essential to know the coupling characteristics of modes in advance.

2.2 Mathematical Formulations

2.2.1 Coupled Multimode Buffeting Analysis

For the convenience of discussion, the multimode analysis procedure (Jain et al. 1996) is briefly reviewed below. Deflection components of a bridge can be represented in terms of

the generalized coordinate as )t(iξ

(2.1) n

r i ii 1

r(x, t) r (x) (t)=

= δ ξ∑ 11

Page 19: Dynamic performance of bridges and vehicles under strong wind

where ri(x) = hi(x), pi(x) or αi(x); ,)x(h i )x(iα and = modal shape functions in vertical,

torsional and lateral direction, respectively; i = 1 to n; n = total number of modes considered in the calculation, and

)x(pi

i ir

i i i

1, if r (x) (x)b (deck width), if r (x) h (x),or p (x)

= αδ = =

(2.2)

The equations of motion for a bridge under wind action can be expressed as

'' '+ + = bIξ Aξ Bξ Q (2.3)

where = vector of generalized coordinates; the superscript prime represents a derivative

with respect to dimensionless time parameter

ξ

( )t Ut / b= ; = identity matrix; Q = vector of

normalized buffeting force (Jain et al. 1996); and the general terms of matrices A and B are

I b

ij i i ij i ijA (K) 2 K J KZ= ζ δ − (2.4)

2ij i ij i ijB (K) K J K T= δ − 2 (2.5)

where

4

ii

bJ2I

lρ= (2.6)

i j i j i j

i j i j i j i j i j i j

* * *ij 1 h h 2 h 5 h p

* * * * * *1 p p 2 p 5 p h 1 h 2 5 p

Z H G H G H G

P G P G P G A G A G A Gα

α α α α

= + +

+ + + + + + α

α

(2.7)

i j i j i j

i j i j i j i j i j i j

* * *ij 4 h h 3 h 6 h p

* * * * * *4 p p 3 p 6 p h 3 4 h 6 p

T H G H G H G

P G P G P G A G A G A G

α

α α α α

= + +

+ + + + + + (2.8)

The modal integrals ( G ) are calculated as jisr

i jr s i j0

dxG r (x)s (x)= ∫l

l (2.9)

12

Page 20: Dynamic performance of bridges and vehicles under strong wind

where ; ; ; i i i ir (x) h (x), p (x) or (x)= α j j j js (x) h (x),p (x) or (x)= α i iK b / U= ω ijδ =

Kronecker delta function (=1 if i = j or = 0 if ji ≠ ); = circular natural

frequency, damping ratio and generalized modal mass of i-th mode, respectively; ρ = air

density; U = mean velocity of the oncoming wind; K(

i i, aω ζ iInd

b / U 2 f b / U= ω =

*iA (i 1 6)

π ) = reduced

frequency; and f are vibrational circular frequency and frequency, respectively; b = bridge

width; l = bridge length; and H ,

ω* *i iP , = − = experimentally determined flutter

derivatives.

Fourier transformation of Eq. (2.3) gives

ξ = bE Q (2.10)

where ξ and bQ = Fourier transformation of ξ and , respectively; E = impedance matrix

with and

bQ

2ijδij ij ijE K K A (K) B (K)i= − + ⋅ ⋅ + 1i = − .

The power spectral density (PSD) matrix for the generalized displacements ξ is derived

using Eq. (2.10) as

b b

11 *Q Q(K)

−−ξξ = S E S E (2.11)

where = complex conjugate transpose of ; and S is the spectrum of buffeting force

(Jain et al. 1996).

*E EbbQQ

The mean-square values of physical displacements can be obtained as

( ) ( ) ( ) ( )i j

n n2 2r r i j 0

i 1 j 1

x r x r x S K∞

ξ ξ= =

σ = δ∑∑ ∫ dK (2.12)

2.2.2 Simplification of Coupled Multimode Buffeting Analysis

Generally, the equations of motion, Eq. (2.3), are coupled and can only be solved simultaneously. By ignoring modal coupling effects among modes, i.e., ignoring the off-diagonal elements in matrices A and B of Eq. (2.3), Eq. (2.12) reduces to

13

Page 21: Dynamic performance of bridges and vehicles under strong wind

( )2 2

i i

n nun un 2 2 unr ri r i 0

i 1 i 1

(x) (x) r (x) S K dK∞

ξ ξ= =

σ = σ = δ∑ ∑ ∫ (2.13)

( )( )

( )( ) ( )( )b bi i

i i

Q Qun2 22

ii ii

S KS K

B K K K A Kξ ξ =

− + ⋅

(2.14)

where (and hereafter) the superscript ‘un’ stands for the single-mode result from uncoupled single-mode solution.

As the combination of the resonant response and background response, the single-mode uncoupled response can be written in the approximated closed-form as (Simiu and Scanlan 1996).

( )( ) ( )

( )b bi i

b bi i

nQ Q iun 2 2

r r i Q Q220 02 2i 1i i i

S K(x) r (x) dK S K dK

K K 2K K

∞ ∞

=

σ ≈ δ + − + ⋅ ⋅ζ

∑ ∫ ∫ (2.15)

where the first and second terms of Eq. (2.15) represent resonant and background response components, respectively, and

i i iK K 1 J T= − ii (2.16)

( ) ( )i i i i i i iK / K J K / K Zζ = ζ − ii

K

(2.17)

After simplifying the vibration frequency K with K in the spatial correlation function

of , the background response can be given as a closed-form, and the total

response is derived as (Simiu and Scanlan 1996; Gu et al. 2002b):

i

( )b bi iQ Q0

S K d∞

( ) ( )( )b bi i

nun 2 2 i rr r i Q Q i

ri 1 i 5/3 5/3ii

UK U 6A 1.67(x) r (x) S K50A4z8b 0.841 10f1 50f

=

+ σ ≈ δ ⋅ + ζ + + +

∑ (2.18)

where 2 2

M L' 'rM L

2C 2CA ,C C = + DC

or 2

D'D

2CC

when r = α, h or p, respectively;

ii

K zf2 b

; z = elevation of the bridge deck above water; CL, CD and CM are the static wind

14

Page 22: Dynamic performance of bridges and vehicles under strong wind

coefficients of lift, drag and moment of the bridge deck, respectively; and a prime over the coefficients represents a derivative with respect to attack angle.

Eq. (2.13) is usually called SRSS single-mode method based on traditional mode-by-mode single-mode buffeting analysis procedure. Eqs. (2.15) and (2.18) are practical formulas after fair approximations are made on Eq. (2.13). As will be seen in the numerical example, the accuracy of such results at high wind velocity is significantly scarified due to ignoring modal coupling effects. To consider modal coupling effect in the buffeting analysis, coupled multimode simultaneous equations need to be solved (Chen et al. 2000). In the following part, a new approach to decoupling those equations will be developed. The new approximate solution will not only lay a foundation to develop the MCF for modal coupling quantification but also provide a method to predict the coupled multimode response based on SRSS results without solving complicated simultaneous equations.

To continue the derivation, the i-th equation is extracted from Eq. (2.10) and is rewritten as

i

n2 2i i j ij j

j 1(K K ) (K) K K D (K) (K) Qi

=

− ξ + ⋅ ⋅ ξ =∑ b (2.19)

where

ij i ijij

j j

A J K TD (K)

K Ki⋅ ⋅ ⋅

= + (2.20)

When off-diagonal terms in matrices A and B are dropped, Eq. (2.3) is decoupled. Correspondingly, Dij = 0 (if i ≠ j) and the uncoupled single-mode response of mode i can be easily derived from Eq. (2.19) as

ibuni 2 2

i i

Q(K)

K K i K K Dξ =

− + ⋅ ⋅ ii

(2.21)

where the superscript ‘un’ in )K(uniξ stands for the result from uncoupled single-mode

solution.

Defining a non-dimensional parameter

ib

iiii22

iuni

ii Q

)K()DKKiKK()K(

)K()K(~ ξ⋅⋅+−

=ξξ

=ξ (2.22)

and then introducing it into Eq. (2.19) gives

15

Page 23: Dynamic performance of bridges and vehicles under strong wind

nij

ijj i

jjj

i D(K) (K) 1K K i D

K K≠

⋅ εξ + ξ =

− + ⋅ε∑ j (2.23)

where ijij ij QD D Rε = ; j

i

bQij

b

Q (K)R (K)

Q (K)= ; ε is introduced to classify the coupling coefficients;

for weak coupling 1<<ε . The chosen magnitude of ε is to make the normalized coupling

coefficients Dij~ of the order of 1, denoted as . [1]O

Since the term on the right hand side of Eq. (2.23) equals to 1, there exist the following results when ε << 1 (Linda and Donald, 1998):

jij

j jjj

j

[ ],K Ki DK [1],K KK i DK K

ε ≠⋅ε = ≈− + ⋅ε

OO

(2.24)

According to the definition in Eq. (2.22), and equal to 1 if the modal

coupling effect is entirely omitted. Otherwise, there exists the condition that . When the i-th mode is under study, the resonant reduced frequency equals to K

i (K)ξ j (K)ξ

1<<εi that is usually different

from Kj. For satisfying the conditionε 1<<ε , following equation can be approximately derived (Linda and Donald, 1998):

j (K) 1 ( ) 1ξ = − ε ≈O (2.25)

With Eqs. (2.24) and (2.25), Eq. (2.23) can be simplified by ignoring the term of O(ε2) as:

nij

ijj i

jjj

i D(K) 1 1K K i D

K K≠

⋅ εξ + ⋅ =

− + ⋅ε∑ (2.26)

In the above process, coupling effects with the order of O(ε2) are ignored. When the response of the i-th mode is to be determined, the coupling between the i-th mode and any other mode is defined here as the primary modal coupling; the coupling effects between arbitrary j-th mode and any other mode (except for i-th mode) is defined as secondary modal coupling. Physically, the above simplification process ignores the secondary coupling effect while including the primary coupling effect. Since only the high order small terms are ignored, the accuracy of the solution should not be significantly scarified. The corresponding solution is much more accurate, as will be seen in the numerical example, than the traditional uncoupled single-mode analysis with Eqs. (2.13) and (2.18).

16

Page 24: Dynamic performance of bridges and vehicles under strong wind

Restoring Eq. (2.26) into the original form like Eq. (2.19) gives

i j

2ni ij ij

i b 2 2 22 2 2j i j j jj jji i ii ii

J K T i K A1(K) Q (K) Q (K)(K K J K T ) i K A(K K J K T ) i K A ≠

b

− + ⋅ ⋅ξ = × −

− − + ⋅ ⋅ − − + ⋅ ⋅ ∑ (2.27)

Rewriting Eq. (2.27) in a matrix form gives

n n×= bH Qξ (2.28)

where

2i ij ij

2 2 22 2 2j j jj jji i ii ii

ij

2 2 2i i ii ii

J K T i K A1 ,(i j)(K K J K T ) i K A(K K J K T ) i K A

H1 ,(i j)

(K K J K T ) i K A

− ⋅ ⋅× ≠

− − + ⋅ ⋅ − − + ⋅ ⋅ = = − − + ⋅ ⋅

(2.29)

The physical mean square response in the r direction can be expressed as

( ) ( ) ( ) ( ) ( )ii ij i i i j

n n n n n2 2 2 2 2r r r r i r i r j0 0

i 1 i 1 i 1 i 1 j i

r x S K dK r x r x S K dK∞ ∞

ξ ξ ξ ξ= = = = ≠

σ = σ + σ = δ + δ δ∑ ∑ ∑ ∑ ∑∫ ∫ (2.30)

where the first part of Eq. (2.30) represents the mean square of primary response which is the summation of each single-mode response in r direction and the second part represents the secondary response which is the summation of cross-modal response in r direction.

The power spectral density (PSD) for the generalized i-th mode displacement iξ can be

derived from Eq. (2.27) as:

i i i i j j b bi j

n nun un *

ij ii ij Q Qj i j i

S (K) S (K) (K) S (K) 2 Re(H H ) Sξ ξ ξ ξ ξ ξ≠ ≠

= + β ⋅ + ⋅ ⋅∑ ∑ (2.31)

2 2ij ij2

ij i 2 22i i

i ii i i ii2

Z T(K) J

K K1 J T 2 J ZK K

+β =

− − × + ζ × − ×

(2.32)

Similarly, cross modal PSD between mode i and mode j for the generalized displacements can be written as

i j 1 2 b bL L1 21 2

n n*

iL jL Q QL 1 L 1

S (K) H H Sξ ξ= =

= ⋅∑∑ (i ≠ j) (2.33)

17

Page 25: Dynamic performance of bridges and vehicles under strong wind

2.3 Approximated Prediction of Coupled Buffeting Response

It can be seen from Eq. (2.31) that the approximated response spectrum of mode i consists of three parts. The first part is the response spectrum of mode i from the traditional uncoupled single-mode analysis, in which the modal coupling effect with other modes is completely neglected. The second part is the response contributed by other modes due to modal coupling and is written as a linear combination of the single-mode response spectra of each mode. The third part is related to the cross-modal buffeting force spectrum between mode i and other modes. Since cross-modal buffeting force spectrum has small effect on aeroelastic coupling, it is negligible (Jain et al. 1996). Similarly, cross-modal response shown in Eq. (2.33), namely the second part of Eq. (2.30), is usually also omitted since the contribution to the total response is normally insignificant (Katsuchi et al. 1999). The numerical verification on such approximations will be made in the example described later.

2.3.1 RMS Response of Coupled Analysis

After omitting trivial terms as discussed above, the mean-square values of physical displacements can be obtained using Eqs. (2.30) and (2.31).

( ) ( ) ( )i i i i j j

n n n2 2 2 2 2 un unr r i r i ij0 0 0

i 1 i 1 j i

r x S K dK r x S (K)dK (K) S (K)dK∞ ∞ ∞

ξ ξ ξ ξ ξ ξ= = ≠

σ ≈ δ ≈ δ + β ⋅

∑ ∑ ∑∫ ∫ ∫ (2.34)

An examination of βij (K) shows that there exists a peak value in the curve of βij (K) when K = Ki. However, βij (K) decays fast when K is away from Ki. Since the frequencies of bridge modes are usually distinctly separated, βij (K) in Eqs. (2.31) and (2.34) can be approximated as βij (Kj), i.e., using the modal natural frequency to replace the oscillation frequency. This is similar to the approach used by Simiu and Scanlan (1996) and Cai et al. (1999a).

In the premise of neglecting the cross-modal buffeting spectrum, namely, the third part of Eq. (2.31), the physical root-mean-square (RMS) response in r motion direction can be finally expressed from Eq. (2.34) as

( ) ( ) ( )2

i i i

n n n2 2 un un

r r i ij r0i 1 j i i 1 j i

x r x S (K)dK 1 x 1∞

ξ ξ= ≠ =

σ ≈ δ + ϑ = σ + ϑ

∑ ∑ ∑∫

n

ij≠

∑ (2.35)

where

j j

i i

un

0ij ij j un

0

S (K)dK(K )

S (K)dK

ξ ξ

ξ ξ

ϑ = β ⋅ ∫∫

( i j≠ ) (2.36)

Above formula can be used for engineers to conveniently, yet reliably, calculate the coupled buffeting response. It will be seen in the following example that Eq. (2.35) is much more accurate than the previous traditional mode-by-mode single-mode analysis shown in

18

Page 26: Dynamic performance of bridges and vehicles under strong wind

Eqs. (2.13) or (2.18), especially in the case of high wind velocity. If the coupling effect is

entirely ignored, i.e., ijϑ = 0, then Eq. (2.35) reduces to Eq. (2.13).

2.3.2 Modal Coupling Factor (MCF)

When the background response is omitted for simplicity to assess MCF, the MCF in Eq. (2.36) can be simplified as a simple closed-form like

( ) ( )( )( )

b bj j

b bi i

2 2Q Q jij ij j i2

ij i 2 22 2i j Q Q ii j i ii i i j i ii

S KZ T KJ

K S KK / K 1 J T 2 K / K J Z

+ ζϑ = × ζ− − × + ζ × − ×

(2.37)

The coefficient ϑij indicates the relative contribution of mode j to mode i responsei i

2ξ ξσ

due to modal coupling effect; it is here named Modal Coupling Factor (MCF). Since MCF represents the relative significance of modal coupling, this information gives a convenient way to quantitatively assess the degree and prone of modal coupling between any pair of modes under study.

2.4 Numerical Example

2.4.1 Prototype Bridge

The developed procedure was applied to the analysis of Yichang Suspension Bridge that is located in the south of China with a main span length of 960m and two side spans of 245m each. The deck elevation is 50m above the sea level, the design wind speed is 29 m/s, and the predicted critical flutter wind velocity Ucr is 73 m/s (Lin et al. 1998). The major parameters along with some other information are summarized in Table 2.1.

Table 2.1 Main parameters of Yichang Suspension Bridge

Main span (m) 960 Lift coefficient at 0o attack angle -0.12

Width of the deck (m) 30 Drag coefficient at 0o attack angle 0.858

Clearance above water (m) 50 Pitching coefficient at 0o attack angle 0.023

Equivalent mass per length

( 10× 3 kg/m)

15.07 ( ) o/CL 0=∂∂

αα 4.43

Equivalent inertial moment of mass per length (× 103 kg⋅ m)

1111 ( ) o/CM 0=∂∂

αα 1.018

Design wind speed (m/s) 29 Structural damping ratio 0.005

19

Page 27: Dynamic performance of bridges and vehicles under strong wind

0 200 400 600 800 1000

-0.5

0.0

0.5

1.0

1st vertical symmetric 1st vertical asymmetric 1st lateral symmetric

h(x)

,y(x

)

x (m)

0 200 400 600 800 1000-1.0

-0.5

0.0

0.5

1.0

1st symmetric torsion 1st asymmetric torsion

α(x

)

x (m)

Fig. 2.1 Major mode shapes of main-span for Yichang Bridge

20

Page 28: Dynamic performance of bridges and vehicles under strong wind

0 2 4 6 8 10-8

-6

-4

-2

0

2

Fl

utte

r der

ivat

ives

Hi*

U/fB

H1* H3

*

H2* H4

*

Fig. 2.2 Flutter derivatives H*i of Yichang Bridge

The characteristics of natural modes were predicted with finite element methods. To facilitate the discussion, only the seven important modes are presented here for the wind-induced vibration analysis. These seven modes are numbered from 1 to 7 in sequence corresponding to one lateral bending, two vertical symmetric bending, one vertical asymmetric bending, two torsional symmetric, and one torsional asymmetric modes as given in Table 2.2. These modes are plotted in Fig. 2.1 for the main span. Flutter derivatives are the other important information for aeroelastic analysis. Eight flutter derivatives of Yichang

Bridge are shown in Figs. 2.2 and 2.3 (Lin et al. 1998). The relatively large values of *1A

(representing the effect of vertical vibration on torsional vibration) and *2H (representing the

effect of torsional vibration on vertical vibration) is an indication of possible strong coupling between vertical and torsional modes. However, this kind of observation is very preliminary and many other factors, such as modal characteristics, wind velocities, and their combinations, will affect the modal coupling. A more rational quantification method for modal coupling effect, as has been developed in the present study, is necessary.

21

Page 29: Dynamic performance of bridges and vehicles under strong wind

0 2 4 6 8 10

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

Fl

utte

r der

ivat

ives

Ai*

U/fB

A1* A3

*

A2* A4

*

Fig. 2.3 Flutter derivatives A*i of Yichang Bridge

2.4.2 Summary of Numerical Procedure

The whole procedure of assessing modal coupling and predicting the coupled buffeting response with the proposed approximate method can be described as:

Firstly, the mean square response for mode i is calculated using traditional single-mode analysis method, i.e., Eqs. (2.13) or (2.18). From these results, it can be decided preliminarily which modes are the main contributors to the physical response. These mode numbers are recorded as the selected i mode for following response analysis in each motion direction.

Secondly, for each selected mode number i in any r motion direction, Zij and Tij are

decided for any mode i and j combination with Eqs. (2.7) and (2.8); ij j(K )β in Eq. (2.36) is

then obtained. Since j j i i

un un

0 0S (K)dK S (K)dK

∞ ∞

ξ ξ ξ ξ∫ ∫

ij

is obtained in the first step, one can finally

get using Eq. (2.36) or Eq. (2.37); ijϑ ϑ is used to assess the tendency of coupling between

mode j and mode i. Since ϑ is directly related to the contribution to the response of mode i

due to modal coupling effect of mode j, a preset threshold value of

ij

ijϑ can be used to select

22

Page 30: Dynamic performance of bridges and vehicles under strong wind

only those key modes j. The choice of threshold value totally depends on the required

accuracy. Meanwhile, since describes aeroelastic coupling effect, which is one of the

major sources of flutter instability, it can also be used to decide which modes are necessary to be included into flutter analysis.

ijϑ

Lastly, for any motion in r direction, with selected mode i in step one and corresponding

obtained in step two, approximate RMS response can be easily derived with Eq. (2.35). ijϑ

Table 2.2 Modal properties of Yichang Suspension Bridge

Mode number Natural frequency (Hz) Mode type

1 0.071 1st Lateral Symmetric Mode

2 0.105 1st Vertical Asymmetric Mode

3 0.161 1st Vertical Symmetric Mode

4 0.233 2nd Vertical Symmetric Mode

5 0.337 1st Torsional Symmetric Mode

6 0.423 1st Torsional Asymmetric Mode

7 0.493 2nd Torsional Symmetric Mode

2.4.3 Assessment of Modal Coupling Effect Using MCF ijϑ

Only the results for the mid-point of the central span are presented here since this location is usually the most important one to study the vibration of the bridge. The MCF

values, , were calculated with Eq. (2.36) for different wind velocities. For the sake of

brevity, only the coupling effects between the four typical modes (Modes 3 to 6) and the other modes (Modes 1 to 7) are shown in Figs. 2.4 to 2.7, respectively. In these figures, the x-axis represents the ratio between the wind speed and flutter critical wind speed (U

ijϑ

cr = 73 m/s). The log scale is used for the y-axis to fit all curves in the figures.

23

Page 31: Dynamic performance of bridges and vehicles under strong wind

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.11E-8

1E-7

1E-6

1E-5

1E-4

1E-3

0.01

0.1

1

M

CF ϑ

3j

U/Ucr

ϑ31 ϑ32 ϑ34 ϑ35 ϑ36 ϑ37

Fig. 2.4 Modal Coupling Factor between mode 3 and other modes

Fig. 2.4 shows the coupling effects between Mode 3 (the 1st symmetric vertical mode)

and the other modes, denoted as (j = 1 to 7). It was found that the values of are all

very small when the wind velocity is lower than 10 m/s, but goes higher when the wind

velocity increases. The MCF between Modes 3 and 5,

3 jϑ 3 jϑ

35ϑ , is significantly larger than other

. When wind velocity is close to 70 m/s, 3 jϑ 35ϑ is about three orders larger than the second

largest value ϑ34. The results conclude that Mode 5 (the 1st symmetric torsion mode) contributes most to the buffeting response of the coupling part of Mode 3 (Again, each mode’s vibration consists of one part from single-mode vibration and another part from mode coupling, as discussed earlier). Other modes contribute insignificantly (10-5-10-4 level) to the response of Mode 3 and can thus be ignored in calculating the buffeting response of the coupling part of Mode 3.

24

Page 32: Dynamic performance of bridges and vehicles under strong wind

Fig. 2.5 shows the MCF values between Mode 4 (the 2nd vertical symmetric mode) and the other modes. It can be seen that Mode 3 (the 1st vertical symmetric mode) contributes relatively large to the buffeting response of the coupling part of Mode 4. However, the value

of is very small for all the modes, indicating a weak coupling between Mode 4 and the

other modes.

4 jϑ

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1

1E-8

1E-7

1E-6

1E-5

1E-4

1E-3

MCF

ϑ4j

U/Ucr

ϑ41 ϑ42 ϑ43 ϑ45 ϑ46 ϑ47

Fig. 2.5 Modal Coupling Factor between mode 4 and other modes

Similarly, the MCF values between Mode 5 (the 1st symmetric torsional mode) and the other modes are shown Fig. 2.6. It can be seen that Mode 3 (the 1st vertical symmetric mode) makes the largest contribution to the buffeting response of Mode 5. However, ϑ53 is significantly less than ϑ35 (see Fig. 2.4), suggesting that the aeroelastic modal coupling effects among modes may not have the property of reciprocity, i.e., ϑij may not be equal to ϑji. This is due to the well-known fact that the aeroelastic matrix is not symmetric.

25

Page 33: Dynamic performance of bridges and vehicles under strong wind

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1

1E-6

1E-5

1E-4

1E-3

0.01

M

CF ϑ

5j

U/Ucr

ϑ51 ϑ52 ϑ53 ϑ54 ϑ56 ϑ57

Fig. 2.6 Modal Coupling Factor between mode 5 and other modes

The MCF values between Mode 6 (the 2nd symmetric torsional mode) and the other modes are shown in Fig. 2.7. While both Modes 3 and 5 make more significant contributions than the other modes, the absolute MCF values are very small, indicating that modal coupling between Mode 6 and the other modes are very weak and can thus be ignored.

As observed above, the MCF values for all the modes are relatively small in low wind velocity and increase with the wind velocity. This indicates that modal coupling effect is mostly due to the aeroelastic modal coupling effect since the structural coupling has nothing to do with wind velocity.

26

Page 34: Dynamic performance of bridges and vehicles under strong wind

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1

1E-8

1E-7

1E-6

1E-5

1E-4

1E-3

M

CF ϑ

6j

U/Ucr

ϑ61 ϑ62 ϑ63 ϑ64 ϑ65 ϑ66

Fig. 2.7 Modal Coupling Factor between the mode 6 and other modes

For the Yichang Suspension Bridge, the calculated MCF can clearly disclose the nature of modal coupling as well as the contribution of other modes to a given mode. Furthermore, the necessity to include a specific mode in the multimode analysis can be judged through the MCF values. For engineering practice, it can be concluded that Mode 3 (the 1st symmetric vertical bending mode) and Mode 5 (the 1st symmetric torsional mode) should be included in the coupled analysis, while the response of other modes can be solved in a mode-by-mode manner without considering modal coupling. In other words, for any other mode i except for modes 3 &5, ϑij = 0 can be used in Eq. (2.35). The strong coupling effect of buffeting response between Modes 3 and 5 indicates also a strong coupling tendency of flutter behavior between these two modes. This was verified in wind tunnel test showing strong coupling between these two modes for both buffeting and flutter behaviors (Lin et al. 1998).

2.4.4 Buffeting Prediction Using the Proposed Approximate Method

Considering only a limited number of modes, the approximate response spectrum for each mode of coupled buffeting can be obtained through Eq. (2.31) after omitting the third part of the formula. To verify the accuracy of mode selection using the MCF method discussed above, Figs. 2.8 to 10 show the normalized spectral response for Mode 3 (the 1st 27

Page 35: Dynamic performance of bridges and vehicles under strong wind

symmetric vertical mode) and Mode 5 (the 1st symmetric torsional mode) under the wind velocity of 20, 40 and 70 m/s, respectively. The curve labelled as Coupled Multimode Analysis corresponds to a fully coupled analysis of all the seven modes. The curve labelled as Proposed Approximated Method corresponds to the MCF method, considering only the coupling effect between Modes 3 and 5. Finally, Uncoupled Single-mode Analysis corresponds to traditional mode-by-mode single-mode analysis.

0.08 0.14 0.20 0.26 0.32 0.38.0001

0.001

0.01

0.1

1

10

0.08 0.14 0.20 0.26 0.32 0.380.0001

0.001

0.01

0.1

1

10

Coupled multi-mode analysis

Proposed Approximated method

Uncoupled single-mode analysis

Nor

mal

ized

PSD

of M

ode

0Nor

mal

ized

PSD

of M

ode

f (Hz)

Fig. 2.8 Normalized PSD of modes 3&5 when U=20 m/s

28

Page 36: Dynamic performance of bridges and vehicles under strong wind

Fig. 2.9. Normalized PSD of modes 3&5 when U=40 m/s Proposed Approximated method

0.08 0.14 0.20 0.26 0.32 0.38.001

0.01

0.1

1

10

0.08 0.14 0.20 0.26 0.32 0.38

0.001

0.01

0.1

1

10

Coupled multi-mode analysis

Uncoupled single-mode analysis

0.0001Nor

mal

ized

PSD

of M

ode

5

0Nor

mal

ized

PSD

of M

ode

3

f (Hz)

Fig. 2.9 Normalized PSD of modes 3&5 when U=40 m/s

29

Page 37: Dynamic performance of bridges and vehicles under strong wind

0.08 0.14 0.20 0.26 0.32 0.380.0001

0.001

0.01

0.1

1

10

Coupled multi-mode analysis

Proposed Approximated method

Uncoupled single-mode analysis

0.08 0.14 0.20 0.26 0.32 0.380.0001

0.001

0.01

0.1

1

10N

orm

aliz

ed P

SD o

f Mod

e 3

Nor

mal

ized

PSD

of M

ode

5

f (Hz)Fig. 2.10 Normalized PSD of modes 3&5 when U=70 m/s

It is found in these figures that the Proposed Approximated Method predicts very close results to those of Coupled Multimode Analysis except for a shifting of peak frequency shown in Fig. 2.10, due to the approximation of K ≈ Kj discussed earlier. In this figure, the investigated wind velocity 70 m/s is much higher than the design wind velocity of 29 m/s. Even though this observed shifting of the peak value frequency, the calculated RMS values of buffeting responses from these two methods are very close. This is because that the total areas under these two curves are about the same. The RMS errors for the vertical displacement of Mode 3 in terms of the generalized coordinate under wind velocities of 20 m/s, 40 m/s, and 70 m/s are 0 %, 1.1 %, and 3.1 %, respectively.

30

Page 38: Dynamic performance of bridges and vehicles under strong wind

Results from Figs. 2.8 to 2.10 show two peak values for Mode 3 (top half of the figure). The first one, corresponding to its own natural frequency, represents the resonant vibration of Mode 3. The second peak, corresponding to the natural frequency of Mode 5, represents the contribution of Mode 5 to Mode 3 vibration due to modal coupling. Comparison of these figures shows that the second peak value of Mode 3 increases with the wind velocity. At the wind velocity of 70 m/s shown in Fig. 2.10, the second peak value is even larger than the first one, indicating that the coupling effect is a significant, or even become a major, contributor to the buffeting response. In this case, ignoring the coupling effect of Mode 5 on Mode 3 would result in a significant error by comparing the curve labelled “Uncoupled Single-mode Analysis” and the curve labelled “Proposed Approximated Method.” This is due to the strong modal coupling effect as indicated by a large value of ϑ35 (see Fig. 2.4). In comparison, for the Mode 5 (bottom half of the figure), the difference between these two curves (and the other one) is trivial since the modal coupling effect is weak as indicated by a small value of ϑ53 (see Fig. 2.6).

The above calculation is based on the spectrum of individual mode in terms of generalized coordinates. To study the accuracy of the proposed MCF method in terms of the physical displacement, the total RMS values of buffeting response were calculated using Eq. (2.35) under different wind velocities. The results of the MCF method (considering coupling effect between modes 3 and 5) are compared in Table 2.3 with that of fully coupled analysis (considering coupling effect among all of the 7 modes) and the uncoupled single-mode calculation (ignoring coupling effect among all of the 7 modes). Comparison of the results suggests that the proposed method results in an error of less than 5% in terms of the RMS of the total buffeting response, an accuracy good enough for engineering application.

2.5 Concluding Remarks

In the present study, a general modal coupling quantification method is introduced through analytical derivations of coupled multimode buffeting analysis. With the proposed Modal Coupling Factor, modal coupling effect between any two modes can be quantitatively assessed, which will help better understand modal coupling behavior of long-span bridges under wind action. Such assessment procedure will also help providing a quantitative guideline in selecting key modes that need to be included in coupled buffeting and flutter analyses. As seen in the numerical example, only as few as two key modes are necessary to be included into modal coupling analysis for the engineering practice. While the example does not necessarily represent the most common cases for long span bridges, it demonstrates that only the coupling effect among limited modes are really necessary to be considered in coupled analysis and selecting only the necessary modes will significantly reduce the calculation effort.

Since the MCF represents the inherent characteristics of modal coupling among modes, it can also provide useful information for flutter analysis. For modern bridges with streamlined section profiles, coupled modes instead of single-mode usually control flutter behaviors. Therefore, knowing the coupling characteristics among modes is extremely important in order to select appropriate modes for coupled flutter analysis and to better understand the aerodynamic behavior.

31

Page 39: Dynamic performance of bridges and vehicles under strong wind

Table 2.3 Comparison between fully-coupled method and proposed approximate method

RMS of vertical displacement

hσ (m)

RMS of torsion displacement

ασ (rad)

A B C A B C

U

(m/s)

Uncoupled

single-mode

analysis

(Eq. 2.18)

Fully Coupl

ed analys

is

(Eq, 2.12)

Proposed MCF analysis

(Eq. 2.35)

Error =

|C/B-1|*100

(%)

Uncoupled

single-mode

analysis

(Eq. 2.18)

Fully Coupled analysis

(Eq. 2.12)

Proposed

MCF analysis (Eq. 2.35)

Error=

|C/B-1|*100

(%)

20 0.12 0.13 0.13 0 0.003 0.003 0.003 0

40 0.30 0.40 0.39 2.6 0.01 0.011 0.011 0

70 0.48 0.99 0.95 4.2 0.025 0.030 0.029 3.5

(RMS displacement at the mid-point of main span for Yichang Bridge)

By using the MCF, an approximate method for predicting the coupled multimode buffeting response was derived through a closed-form formula. Numerical results of a prototype bridge have proven that the proposed method is much more accurate than the traditional uncoupled single-mode method. That is especially true when the coupling effect is significant at high wind velocity. Difference of the predicted buffeting responses between the approximate and fully coupled analysis methods (considering the coupling of all modes) is less than 5%.

Another important potential application of the MCF values is in the design of adaptive control strategies. To achieve the optimal control efficiency, an adaptive control may not only aim at controlling a single-mode resonant vibration, but also at reducing coupled vibration by breaking the coupling mechanism. For this purpose, the coupling characteristics of modes need to be known in advance.

32

Page 40: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 3 EVOLUTION OF LONG-SPAN BRIDGE RESPONSE TO WIND- NUMERICAL SIMULATION AND DISCUSSION

3.1 Introduction

Long-span bridges are susceptible to wind actions and flutter and buffeting are their two common wind-induced phenomena. Buffeting is a random vibration caused by wind turbulence in a wide range of wind speeds. With the increase of wind speed to a critical one, the bridge vibration may become unstable or divergent - flutter (Scanlan 1978). This critical wind speed is called flutter wind speed. Flutter can occur in both laminar and turbulent winds.

In laminar flow, the bridge vibration prior to flutter is essentially a damped multi-frequency free vibration, namely, a given initial vibration will decay to zero. When the wind speed increases to the flutter wind speed, the initial vibration (or self-excited vibration) would be amplified to become unstable. When the bridge starts to flutter due to the increased wind speed, it was observed that all modes respond to a single frequency that is called flutter frequency (Scanlan and Jones, 1990).

In turbulent flow, random buffeting response occurs before flutter. When wind speed is low, each individual mode vibrates mainly in a frequency around its natural frequency and the buffeting vibration is a multi-frequency vibration in nature. Keeping increase of wind speed will lead to a single-frequency dominated divergent buffeting response near the flutter velocity. The divergent buffeting response represents the instability of the bridge-flow system, which can also be interpreted as the occurrence of flutter (Cai et al. 1999). Therefore, physically, buffeting and flutter are two continuous dynamic phenomena induced by the same incoming wind flow. It is a continuous evolution process where a multi-frequency buffeting response develops into single-frequency flutter instability.

Previous flutter analysis usually focused only on finding the flutter wind speed and the corresponding flutter frequency. Except for some generic statements, how the multi-frequency pre-flutter vibration turns into a single-frequency oscillatory vibration at the onset of flutter has not yet been well demonstrated numerically. The present study will simulate and discuss the two divergent vibration processes near flutter wind speed. The first case is from self-excited flutter and the second one is from random buffeting vibration. The simulated process will clearly demonstrate how the multi-frequency vibration process evolves into a single-frequency vibration, which will help clarify some confusing statements made in the literature and help engineers better understand the flutter mechanism.

3.2 Motivation of Present Research

As discussed above, the bridge vibration is a multi-frequency vibration at low wind speeds. However, when the wind speed approaches the flutter wind speed, the multi-frequency vibration merges into a single-frequency dominated flutter vibration. Numerical simulations of the transition phenomenon from multi-frequency buffeting to single-frequency flutter have not been well introduced.

Flutter was classified by Scanlan (1987) as “stiffness-driven type” and “damping-driven type”. Classical aircraft-type flutter, called “stiffness-driven type” was believed to have typically two coupling modes coalesce to a single flutter frequency (Namini et al. 1992). On the other hand, single-degree-of-freedom, which was also called damping-driven flutter, has

33

Page 41: Dynamic performance of bridges and vehicles under strong wind

different scenario (frequencies does not coalesce) (Scanlan 1987). These statements imply that the frequency of each mode changes to a single value at flutter for stiffness-driven flutter. However, based on the results of eigenvalue analysis that will be shown later, the predicted modal oscillation frequencies of different modes are not necessarily the same when flutter occurs. Similar observation (oscillation frequencies of the modes are not the same at flutter) can also be made from the numerical results of other investigators (Namini et al. 1992). Chen et al. (2001) analyzed examples with different pair of frequencies for coupled modes. “Veering” phenomena was observed when the frequencies are very close (extreme case). In the examples with not too close frequencies (like most realistic bridges), the frequencies of the coupled modes did not have the chance to coalesce. However, in the wind tunnel test as well as the observations of the Tacoma Narrows’ failure, the vibration is known to usually exhibit dominant torsion vibration with a single frequency right before the occurrence of flutter instability. The predicted different oscillation frequencies among modes seem to contradict with the observed “single-frequency” flutter vibration. As will be seen later from the numerical example, flutter is observed as a single-frequency vibration because that the modal coupling effects force all modes to respond to the oscillation frequency of the critical mode. The oscillation frequency of each mode does not necessarily merge or change to a single value. The “meeting” of the vertical and torsion frequencies may occur beyond the flutter velocity, not necessarily at the onset point of flutter as previously stated in some papers (Scanlan, 1978).

So far, the nature of transition of frequencies during the flutter initiation process has not been satisfactorily explained. The writers believe that more specific numerical simulations and discussions are necessary to understand how the multi-frequency vibration always turns into a single-frequency one at flutter. The present work will help explain this phenomenon and better understand the evolution process from buffeting vibrations under strong wind to flutter occurrence. For this purpose, an examination of the frequency characteristics of bridge vibration in a full range of wind speeds is necessary.

h

Incoming flow

α

Fig.3.1 2-D model of bridge section in incoming wind

3.3 Analytical Approach

Flutter and buffeting analyses were carried out either in a time domain (Ding et al. 2000; Boonyapinyo et al. 1999) or in a frequency domain (Scanlan and Jones 1990). In both methods, aerodynamic forces are defined (directly or indirectly) with the so-called flutter

34

Page 42: Dynamic performance of bridges and vehicles under strong wind

derivatives that can be determined for each type of bridge deck through a specially designed wind tunnel experiment. To facilitate the presentation and discussion, a brief description of the related theoretical aspects of flutter and buffeting analyses is given below.

3.3.1 Equations of Motion

For the convenience of theoretical presentation, a typical 2-dimensional section model (with a unit length of bridge deck) shown in Fig. 3.1 is used in the formulation of flutter and buffeting analyses. Two degree-of-freedom, i.e., vertical displacement h(x, t) and torsion displacement α(x, t), are considered here. The equations of motion for the section model can be expressed as (Simiu and Scanlan 1986):

( ) ( ) ( ) ( )( ) ( ) ( )( ) ( ) ( ) ( )( ) ( ) (

2h h h ae b

2ae b

m x h x, t 2 h x, t h x, t L x, t L x, t

I x x, t 2 x, t x, t M x, t M x, tα α α

+ ζ ω + ω = +

α + ζ ω α + ω α = + ) (3.1)

where the “dot” above the h and α represents a derivative with respect to time, m and I = generalized mass and generalized mass moment of inertia of a unit length of deck, respectively; ζα and ζh = structural damping ratios for torsion and vertical modes, respectively; ωα and ωh = natural circular frequencies for torsion and vertical modes, respectively; and = unit length self-excited lift force and pitch moment, respectively; and and M = unit length buffeting lift force and pitch moment, respectively. These aerodynamic forces are expressed as

aeL

b

aeM

bL

( ) ( ) ( ) ( )2 2h x, t x, tL U B KH KH K H x, t K H h x, t∗ ∗ ∗ ∗ α

= ρ + + α + 2ae 1 2 3 4U U (3.2)

( ) ( ) ( ) ( )2 2 2 2h x, t x, tM U B KA KA K A x, t K A h x, t∗ ∗ ∗ ∗ α

= ρ + + α + ae 1 2 3 4U U (3.3)

( )2 '1 u(t) w(t)b L L DL U B C 2 C C

2 U U

(3.4) = ρ + +

2 2 '1 u(t) w(t)b M MM U B C 2 C

2 U U

(3.5) = ρ +

where reduced frequency BKUω

= ; B = bridge width; U = average wind speed; ω =

oscillation circular frequency; Hi* and Ai

* = flutter derivatives from wind tunnel test; CL, CD

35

Page 43: Dynamic performance of bridges and vehicles under strong wind

and CM = static coefficients for lift force, drag force and pitch moment of bridge deck, respectively; and u(t) and w(t) = turbulent wind components in the lateral and vertical directions, respectively. A “prime” over the coefficients represents a derivative with respect to the wind attack angle.

The two vibration displacements can be expressed using modal superposition as:

( ) ( ) (n

i ii 1

h x, t h x t=

= ξ∑ )

)

(3.6)

( ) ( ) (n

i ii 1

x, t x t=

α = α γ∑ (3.7)

where n = number of mode shapes; hi(x) and αi(x) = mode shapes in vertical and torsion directions, respectively; and ξi(t) and γi(t) = generalized coordinates in vertical and torsion directions, respectively.

Using Eqs. (3.2) to (3.7) and considering only the first vertical and first torsion modes for simplicity, Eq. (1) can be rewritten in a matrix form as

[ ] [ ] [ ] bI X C X K X Q+ + = (3.8)

where

(t)

X(t)

ξ = γ (3.9)

bQ

L

b0*h

L

b0*

h(x)L dx

M

(x)M dx

= α

dx

(3.10)

[ ]2 * * 2 * *

h h 1 hh h 2 h h2 * * 2 * *

1 h 2

2 B H G /(2M ) B H G /(2M )C

B A G /(2I ) 2 B A G /(2I )α

α α α α αα α

ζ ω − ρ ω −ρ ω= −ρ ω ζ ω − ρ ω

(3.11)

[ ]2 3 2 * * 3 2 * *h 4 hh h 3 h h

3 2 * * 2 3 2 * *4 h 3

B H G /(2M ) B H G /(2M )K

B A G /(2M ) B A G /(2M )α

α α α αα α

ω − ρ ω −ρ ω= −ρ ω ω − ρ ω

(3.12)

( )L 2

hh 0G h x= ∫

( )L*

0I I xα = α∫

[

dx ( ) ( )L

h h 0G G h x x dα α= = α∫( )2 x dx

; ; ;

and = generalized mass for vertical and torsion modes, respectively;

x ( )L 2

0G xαα = α∫ dx ( ) ( )

L* 2h 0

M m x h x= ∫

]I = unit matrix; and L = bridge length.

The equations of motion contain flutter derivatives that are functions of the reduced

36

Page 44: Dynamic performance of bridges and vehicles under strong wind

frequency K. Therefore, even though bridge vibrations can be decoupled into modal vibrations through the modal superposition method, these flutter derivatives re-couple the equations of motion. Correspondingly, coupled modal vibrations have often been observed in wind tunnel tests, especially in high winds. In the coupled vibration, each mode vibrates about its oscillation frequency, while affected also by the other modes.

3.3.2 Complex Eigenvalue Analysis

The flutter wind speed is commonly determined by eigenvalue method in the frequency domain, i.e., by iteratively searching for a pair of K and ω so that the determinant of the characteristic function of the equations of motion becomes zero (Simiu and Scanlan 1996).

For both buffeting and flutter analyses, one of the most important tasks is to quantify the aerodynamic damping and stiffness matrices, [C] and [K] in Eq. (3.8). These matrices include the contributions from aerodynamic forces, and are functions of the vibration circular frequencies corresponding to different modes. Iterative complex eigenvalue approach has been used to predict modal properties in a wide range of wind speeds, from which the occurrence of flutter can be determined and [C] and [K] can be quantified (Agar 1989; Miyata and Yamada 1999). Once the damping [C] and stiffness [K] matrices in Eq. (3.8) are defined and wind buffeting force bQ is given, buffeting response can also be obtained in the frequency domain.

The flutter critical point can be identified by iteratively solving the complex eigenvalues in the state-space of Eq. (3.8) at each wind velocity. The Eq. (3.8) can be rewritten in the state-space as

FBYYA =+ (3.13)

where [ ][ ]

[ ] [ ][ ] [ ]

b

I 0 0 I 0XY ; A ; B and F

0 I K C QX−

= = = = .

The equation for free vibration becomes

0BYYA =+ (3.14)

which can be transformed to (Cai and Albrecht 2000):

[ ] [ ] [ ] Φ Φ− =1 Y diagλ −1 Y

(3.15)

where [ and

= eigenvector and eigenvalue matrices of ]Φ [ ]λ diag BA 1−− , respectively.

The Eq. (3.15) can be rewritten in a simplified form as

[ ] Z diag= λ Z (3.16)

where ; and [ ] , a matrix consisting of complex conjugate pairs of eigenvalues, has components as

[ ] Z Y= −Φ 1 λ diag

37

Page 45: Dynamic performance of bridges and vehicles under strong wind

2i,i 1 i i i ij 1 , i 1to3+λ = −ζ ω ± − ζ ω = (3.17)

where λ = iith eigenvalue, ω andi ζ i = modal oscillation frequency and damping ratio in ith

mode, and j = unit imaginary number ( 1j −= ).

Once the eigenvalues are iteratively predicted, the oscillation frequency ω and reduced frequency K are known at the given wind speed; and the damping and stiffness matrices [C] and [K] can then be calculated. The flutter wind speed and corresponding flutter frequency can be identified from the eigenvalue solutions at the condition that the total modal damping approaches zero.

3.3.3 Buffeting Analysis

Buffeting is a multi-mode random vibration, where different ω values are used in the calculation of the reduced frequency K for each mode (Cai and Albrecht 2000). The ω value can be either the respective natural circular frequency of that particular mode or the aeroelastically modified natural frequency considering aeroelastic effects (Simiu and Scanlan 1986, 1996), i.e., the oscillation frequency predicted in the complex eigenvalue analysis discussed above. The choice of different ω values for each different mode is justified for buffeting analysis when the wind speed is lower than the flutter wind speed. When the wind speed approaches the flutter wind speed, the buffeting vibration will be dominated by the frequency of the critical mode. The linear buffeting theory has been well documented in the literature and is not repeated here.

3.3.4 Time-series Analysis

It is noted that the aeroelastic forces in Eqs. (3.2) and (3.3) have frequency-dependent terms, K, Hi and Ai, as well as time-dependent terms, h(x, t) and α(x, t). Such a frequency-dependence causes some difficulties to assemble the element matrices in finite element analysis (Cai and Albrecht 2000). In the previous time domain analysis, in order to eliminate the dependence on the frequency, indicial functions or rational functions were adopted to approximately curve fit the aerodynamic forces in Eqs. (3.2) to (3.5) into time-dependent-only functions (Bucher and Lin 1998).

In the present time-series study, it is proposed to first conduct the complex eigenvalue analysis before the time-series analysis is performed. In this way, the frequency-dependent components of flutter derivatives and K will be resolved from [C] and [K] matrices and Eq. (3.8) will be time-dependent only. Consequently, to study the flutter process in time-series, a free vibration response can then be obtained through numerical integration of the homogeneous part of Eq. (3.8) (on the left hand side). Similarly, if buffeting forces (on the right hand side of Eq. (3.8)) are known, then buffeting response can be predicted using either time-series or spectral analysis.

3.4 Numerical Procedure

For flutter in laminar flow, complex eigenvalue analysis will identify the flutter wind speed and frequency. However, we believe that simulation of the free vibration in time-series will disclose the evolution process of flutter that is in nature a self-excited vibration. The free vibration of the coupled system can be solved through a time integration of the equations of motion. The actual frequency characteristics of the bridge vibration can then be obtained

38

Page 46: Dynamic performance of bridges and vehicles under strong wind

through a spectral analysis of the derived time-series response. The buffeting and flutter analyses at the same wind speed will then be compared to investigate their correlations.

This proposed numerical procedure consists of three steps - complex eigenvalue analysis, time-series analysis and spectral analysis. It is noted that neither complex eigenvalue analysis, time-series analysis, nor spectral analysis is new by itself, but their combined use in the present study can clearly demonstrate the flutter evolution process, which helps further understand the transition mechanism from the multi-frequency vibration to single-frequency flutter.

STEP 1 – COMPLEX EIGENVALUE SOLUTION

The first step is to use the complex eigenvalue solution to predict the respective vibration damping ratio and oscillation frequency of each mode under any desired wind speed. These damping ratios and oscillation frequencies include the effect of aerodynamic forces. Though associated with vibration modes, these modal values are different from the traditional natural modal values in which no aerodynamic forces are involved. Therefore, these modal values that include aerodynamic effects are actually (and hereafter called) the effective modal values. These effective modal values are needed to quantify the aerodynamic matrices [C] and [K] in Eq. (3.8). Otherwise, matrices [C] and [K] include unknown variables, i.e., the effective oscillation frequency and damping ratio.

As discussed earlier, the effective oscillation frequency of each mode under each desired wind speed will be obtained with the complex eigenvalue solution of Eq. (3.14). This process is shown in Fig. 3.2 and briefly explained here. The whole process starts at zero wind speed, and the natural frequency and mechanical damping ratio of each mode are chosen as the initial values, ω and 0 0ζ . For the ith step, Ui = Ui-1 + ∆U and iΩ = ω are assumed. With K

i 1−

i = (B /UiΩ i), the corresponding flutter derivatives can be decided from the experimentally measured flutter derivatives versus K curves. Using the complex eigenvalue analysis, modal oscillation frequency iω and damping ratio iζ for each mode are predicted with Eqs. (3.16) and (3.17). The “bars” above i and iω ζ denote that they are temporary values during the solution process. After a convergence is achieved by comparing iω and Ω , the effective modal properties at wind speed U

i

i can be obtained as i iω = ω and i iζ = ζ . If the effective modal damping ratio iζ is greater than zero, the solution process proceeds to the (i+1)th step. Otherwise, the flutter critical point is identified due to the non-positive iζ and the flutter critical wind speed is predicted as Ucr = Ui-1. If no convergence has been achieved, i.e., iω ≠ Ωi , then set i iΩ = ω and repeat the process until it converges.

STEP 2 - TIME DOMAIN ANALYSIS

The second step is to carry out a time-domain analysis using the quantified equations of motion established in Step 1. This will provide a time-series for the spectral analysis of Step 3. Spectral analysis will reveal the frequency characteristics of vibration.

Effective oscillation frequencies from Step 1 are used to determine the respective

reduced frequency )U

BK ω=( for each mode under any wind speed. Accordingly, flutter

derivatives can be decided according to the respective K for each mode. The matrices, [C] in

39

Page 47: Dynamic performance of bridges and vehicles under strong wind

Eq. (3.11) and [K] in Eq. (3.12), can thus be quantified and equations of motion, Eq. (3.8), is ready for a time domain analysis.

In the complex eigenvalue solution, flutter is identified when ζ i changes from positive to negative value. However, a free vibration response in time-series can better demonstrate the occurrence of flutter. As will be shown in the numerical example, time domain analysis can predict convergent, constant, and divergent amplitude vibrations that correspond to stable, critical, and unstable (flutter) conditions of the bridge, respectively. For this reason, a 5th

order Runge-Kutta method is used to solve the equations of motion in the time domain.

STEP 3 - SPECTRAL ANALYSIS ON TIME-SERIES

Once the time-series of vibrations are available from Step 2, the third step is to conduct spectral analysis on the time-series of vibration response. Spectral analyses on these time-series will reveal the vibration characteristics in the frequency domain. This information will help understand the true nature of the coalesced single-frequency vibration of flutter.

3.5 Numerical Example

The Humen Bridge (Fig. 3.3) is a suspension bridge with a main span of 888 m located in Guang-Dong Province, China. The main data and main mode characteristics for the Humen Bridge are given in Tables 3.1 and 3.2, respectively (Lin and Xiang 1995). After examining its modal characteristics performed with finite element analysis, the first symmetric vertical mode (1st mode, hereafter simply called vertical mode) and the first symmetric torsion mode (4th mode, hereafter simply called torsion mode) are identified as two important modes and are thus used for both flutter and buffeting analyses. This numerical example is used to demonstrate the approach discussed above and to explain how the bridge vibration evolves from a multi-frequency vibration to a single-frequency flutter.

3.5.1 Complex Eigenvalue Analysis

The predicted effective oscillation frequencies and damping ratios of vertical and torsion modes versus wind velocity are shown in Fig. 3.4 for the prototype Humen Bridge. It can be seen that with the increase of wind velocity, vertical mode frequency increases gradually while at the same time torsion mode frequency decreases gradually. Correspondingly, the damping ratio of the vertical mode increases, while the damping ratio of the torsion mode remains about constant and decreases at higher wind velocity. Eventually, at the wind velocity of 87 m/s, the damping ratio of the torsion mode becomes zero (which represents the occurrence of flutter) and the damping ratio of vertical mode increases by about 20%.

At the flutter critical velocity (87 m/s), the frequency of the torsion mode is 0.31Hz and vertical mode is 0.185 Hz. Flutter critical velocity predicted with the traditional iterative search method (Simiu and Scanlan 1986) is 91.0 m/s and measurement in wind tunnel test is 89 m/s (Lin and Xiang 1995). These comparisons have generally verified the present prediction.

While complex eigenvalue analysis is able to identify the critical point of flutter, oscillation modal frequencies for the two modes (0.31 vs. 0.185 Hz) do not approach and do not finally become the same value as stated in the literature. Though there is a tendency that the two frequency curves will approach or meet as the increase of the wind velocity, they do

40

Page 48: Dynamic performance of bridges and vehicles under strong wind

not meet at the flutter wind velocity (they may meet beyond flutter wind velocity). When the natural frequencies of the coupled modes are close enough, the modal frequency curves may approach and intersect at the flutter wind velocity (This will be shown later with a simulated example). To further investigate the vibration characteristics and explain the discrepancy between the observations made above and the statement that flutter is a single-frequency vibration, time domain analysis is conducted below.

Ki=BΩi/Ui

Complexeigenvalue analysis

Decide flutter derivative of step i

ωi, ζi

Eqs.(16-17)

ωi ≅ Ωi

No

Ωi=ωi

Ui=Ui-1+∆U

Yes

No

i=1 Ui-1=0ωi-1=ω0,ζi-1=ζ0

ωi=ωi, ζi=ζi

ζi<0

Ucr=Ui-1

Yes

i=i+1

Ωi=ωi-1

Fig.3.2 Flowchart of complex eigenvalue approach

41

Page 49: Dynamic performance of bridges and vehicles under strong wind

Table 3.1. Main Parameters of Humen Suspension Bridge

Main span (m) 888 Lift coefficient at 0o attack angle 0.02

Width of the deck (m) 35.6 Drag coefficient at 0o attack angle 0.84

Clearance above water (m) 60 Torsion coefficient at 0o attack angle 0.019

Equivalent mass per length

(103 *kg/m)

18.34 ( ) o/CL 0=∂∂

αα 0.51

Equivalent inertial moment of mass per length (103 *kg/m)

1743 ( ) o/CM 0=∂∂

αα 0.62

Design wind speed (m/s) 57.2 Structural damping ratio 0.005

Table 3.2 Dynamic Properties of Five Main Modes for Humen Suspension Bridge

Mode number Natural frequency (Hz) Mode type

1 0.172 1st Symmetric Vertical Mode

2 0.225 2nd Symmetric Vertical Mode

3 0.276 1st Asymmetric Torsion Mode

4 0.361 1st Symmetry Torsion Mode

5 0.426 1st Asymmetry Torsion Mode

42

Page 50: Dynamic performance of bridges and vehicles under strong wind

880 m 330 m293 m

(a) Elevation of Humen Bridge

35600 mm

31000 mm

(b) Cross section of Humen Bridge

Fig.3.3 Humen Bridge elevation and cross section

43

Page 51: Dynamic performance of bridges and vehicles under strong wind

3.5.2 Time-Series and their Frequency Content

Once the oscillating frequencies of vertical and torsion modes are predicted as shown in

Fig. 3.4, they are used to calculate the reduced frequency BU

K ω= in order to quantify the

matrices A and B in Eq. (3.13). This state-space equation is then solved in the time domain. The initial displacements of free vibration are assumed to be 0.01 m and 0.01 deg for vertical and torsion modes, respectively. Fig. 3.5 shows the time-series of the vertical and torsion vibrations at U = 30 m/s. It is obvious that they vibrate at two different frequencies, which is also clearly shown in Fig. 3.6, the spectra diagram. The first peak in Fig. 3.6(a) shows that the vertical mode vibrates mainly at its effective frequency that is predicted in the complex eigenvalue analysis. There exists also a small peak that corresponds to the torsion frequency due to the aerodynamic coupling between the two modes. Similarly, the second peak of Fig. 3.6(b) shows that the torsion mode vibrates mainly at its effective frequency with a very small peak at a frequency corresponding to that of vertical mode due to mode coupling.

Figs. 3.7 and 3.8 show the time-series and spectra at wind speed of 60 m/s. At this higher wind speed, the peaks due to modal coupling are more obvious, especially the second peak of vertical vibration shown in Fig. 3.8(a). The two modes still vibrate in different frequencies.

When the wind speed increases up to 87m/s, time-series in Fig. 3.9 show almost constant magnitude vibrations for both vertical and torsion modes, and both modes vibrate about the same frequency. These constant-amplitude vibrations suggest flutter critical condition and the flutter velocity can be identified as 87 m/s. The same vibration frequency of the two time-series at flutter (U = 87 m/s) is clearly shown through their spectral diagrams in Fig. 3.10. The two peaks correspond to the same frequency of 0.31 Hz. For vertical mode, the original dominant peak corresponding to its effective frequency of 0.175 Hz (Fig. 3.6(a)) becomes very trivial as shown in Fig. 3.10.

When the wind speed increases up to 87m/s, time-series in Fig. 3.9 show almost constant magnitude vibrations for both vertical and torsion modes, and both modes vibrate about the same frequency. These constant-amplitude vibrations suggest flutter critical condition and the flutter velocity can be identified as 87 m/s. The same vibration frequency of the two time-series at flutter (U = 87 m/s) is clearly shown through their spectral diagrams in Fig. 3.10. The two peaks correspond to the same frequency of 0.31 Hz. For vertical mode, the original dominant peak corresponding to its effective frequency of 0.175 Hz (Fig. 3.6(a)) becomes very trivial as shown in Fig. 3.10.

The above observations have numerically proven that the vertical and torsion modes exhibit the same vibration frequency at flutter. Such numerical results agree exactly the observation made in the wind-tunnel tests and the traditional belief that the coupled modes should have the same vibration frequency when flutter happens. However, though the vertical mode vibrates at the same frequency as the torsion mode at flutter, the “same frequency” is not due to the change of the effective oscillation frequency in vertical mode. The two effective oscillation frequencies never “meet” in this numerical example. This single frequency vibration is due to the magnification of the coupling effect at flutter so that the vibration corresponding to the torsion-induced coupling is dominant and that corresponding to the effective oscillation frequency of vertical mode is relatively too small to be seen by the observers. Therefore, a single-frequency vibration is observed at flutter.

44

Page 52: Dynamic performance of bridges and vehicles under strong wind

0 20 40 60 800.15

0.20

0.25

0.30

0.35

0.40

M

odal

freq

uenc

y (H

z)

Wind speed (m/s)

Vertical mode Torsion mode

0 20 40 60 800.0

0.1

0.2

(a) Modal frequency

Mod

al d

ampi

ng ra

tio

Wind speed (m/s)

Vertical mode Torsion mode

(b) Modal damping ratio

Fig. 3.4 Modal properties with complex eigenvalue method

45

Page 53: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30-0.01

0.00

0.01

V

ertic

al d

ispla

cem

ent (

m)

Time (s)

(a) Vertical mode

0 5 10 15 20 25 30-0.010

-0.005

0.000

0.005

0.010

Tors

ion

angl

e (ra

d.)

Time (s)

(b) Torsion mode

Fig.3.5 Time history response of free vibration under U = 30 m/s

46

Page 54: Dynamic performance of bridges and vehicles under strong wind

0.0 0.1 0.2 0.3 0.4 0.50.000

0.001

0.002

0.003

Frequency (Hz)

Spec

tral d

ensit

y am

plitu

de

(a) Vertical mode

0.0 0.1 0.2 0.3 0.4 0.50.000

0.002

0.004

0.006

Frequency (Hz)

Spec

tral d

ensit

y am

plitu

de

(b) Torsion mode

Fig.3.6 Power spectrum of free vibration under U = 30 m/s

47

Page 55: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30-0.01

0.00

0.01

V

ertic

al d

ispla

cem

ent(m

)

Time (s)

(a) Vertical mode

0 5 10 15 20 25 30-0.01

0.00

0.01

Tors

ion

angl

e (ra

d.)

Time (s)

(b) Torsion mode

Fig. 3.7 Time history response of free vibration under U = 60 m/s

48

Page 56: Dynamic performance of bridges and vehicles under strong wind

0.0 0.1 0.2 0.3 0.4 0.50.0000

0.0005

0.0010

0.0015

Frequency (Hz)

Spec

tral d

ensit

y am

plitu

de

(a) Vertical mode

0.0 0.1 0.2 0.3 0.4 0.50.000

0.001

0.002

0.003

0.004

Frequency (Hz)

Spec

tral d

ensit

y am

plitu

de

(b) Torsion mode

Fig. 3.8 Power spectrum of free vibration under U = 60 m/s

49

Page 57: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30-0.04

-0.02

0.00

0.02

0.04

Ver

tical

disp

lace

men

t (m

)

Time (s)

(a) Vertical mode

0 5 10 15 20 25 30-0.02

-0.01

0.00

0.01

0.02

Tors

ion

angl

e (ra

d.)

Time (s)

(b) Torsion mode

Fig. 3.9 Time history response of free vibration under U = 87 m/s

50

Page 58: Dynamic performance of bridges and vehicles under strong wind

0.0 0.1 0.2 0.3 0.4 0.50.00

0.01

0.02

0.03

Frequency (Hz)

Spec

tral d

ensit

y am

plitu

de

(a) Vertical mode

0.0 0.1 0.2 0.3 0.4 0.50.000

0.005

0.010

0.015

Frequency (Hz)

Spec

tral d

ensit

y am

plitu

de

(b) Torsion mode

Fig. 3.10 Power spectrum of free vibration under U = 87 m/s

51

Page 59: Dynamic performance of bridges and vehicles under strong wind

3.5.3 Vibration Characteristics of Buffeting

With the same aerodynamic matrices as used for flutter analysis, buffeting response was obtained through a spectral analysis of the equations of motion.

Figs. 3.11 to 3.13 show the buffeting displacement spectra on the edge of the deck for the wind velocities of 30, 60 and 87 m/s, respectively. For comparison, the results of single-mode-based buffeting response are also shown in these figures. When the wind velocity is 30 m/s, vertical and torsion motions vibrate with different dominant frequencies (Fig. 3.11). When the wind velocity increases to 60 m/s, Fig. 3.12 shows that vertical and torsion modes vibrate with stronger coupling effects; in other words, the magnitudes of the second peak of Fig. 3.12(a) and the first peak of the Fig. 3.12(b) increase obviously.

When wind velocity approaches the flutter critical velocity (U = 87 m/s), Fig. 3.13 shows that the vertical motion exhibits the same dominant frequency as the torsion mode. The vibration corresponding to the effective oscillation frequency of vertical mode is greatly damped out (indicated by the flat peak at frequency near 0.175 Hz). The vibration of the vertical mode is mainly from modal coupling indicated by the sharp peak near 0.3 Hz. Physically, the two modes vibrate in the same dominant frequency (about 0.3 Hz), i.e., the divergent buffeting vibration indicates flutter instability – a single frequency vibration.

Comparison between the results of buffeting analyses in Figs. 3.11 to 3.13 and those of flutter analyses in Figs. 3.6, 3.8 and 3.10 suggests that their vibration characteristics are similar, though not exactly the same due to the difference of vibration types (free vibration vs. forced vibration). Overall, the vibration characteristics of flutter and buffeting are well correlated.

In comparison, the results of single-mode-based buffeting analyses are significantly different from that of multi-mode coupled analyses at high wind velocity, especially at flutter critical wind velocity. Fig. 3.11 shows a small difference when wind speed is low (30 m/s) and Fig. 3.13 shows a very significant difference when wind speed is high (U = 87 m/s). It suggests that for future longer and slender bridges whose natural frequencies will even be lower, aerodynamic coupling effect will be more significant.

3.5.4 Discussions of Coupling Effect on Vibration Characteristics

Due to the modal coupling among the coupling-prone modes, small peaks may be observed in the spectrum diagrams for low wind velocity, in addition to the resonant peaks corresponding to the effective modal frequencies. With the increase of wind speed, modal coupling among modes is strengthened by aeroelastic effects so that the original trivial peaks induced by modal coupling may gradually dominate the vibration and become sharp peaks in the spectral diagrams. Correspondingly, the once dominant peaks at the effective frequency may gradually be damped out and become trivial flat peaks (Figs. 3.11 to 3.13). In a two-mode model with vertical and torsion modes, such phenomenon usually happens for vertical mode and finally the dominant frequencies for both modes become the same – close to the effective frequency of the torsion mode.

These observations indicate that the “same frequency” multimode flutter vibration is actually due to the effect of aerodynamic coupling. With the increase of wind velocity, the frequency of vertical mode increases and the frequency of torsion mode decreases correspondingly. However, the effective oscillation frequencies of vertical and torsion modes

52

Page 60: Dynamic performance of bridges and vehicles under strong wind

0.1 0.2 0.3 0.41E-4

1E-3

0.01

0.1

1

N

orm

aliz

ed sp

ectru

m

Frequency (Hz)

Single mode-based Coupled analysis

(a) Vertical mode

0.1 0.2 0.3 0.4

1E-3

0.01

0.1

1

Nor

mal

ized

spec

trum

Frequency (Hz)

Single mode-based Coupled analysis

(b) Torsion mode

Fig. 3.11 Non-dimensional normalized power spectra of buffeting response under U = 30 m/s

53

Page 61: Dynamic performance of bridges and vehicles under strong wind

0.1 0.2 0.3 0.41E-3

0.01

0.1

1

N

orm

aliz

ed sp

ectru

m

Frequency (Hz)

Single mode-based Coupled analysis

(a) Vertical mode

0.1 0.2 0.3 0.41E-3

0.01

0.1

1

Nor

mal

ized

spec

trum

Frequency (Hz)

Single mode-based Coupled analysis

(b) Torsion mode

Fig. 3.12 Non-dimensional normalized power spectra of buffeting response under U = 60 m/s

54

Page 62: Dynamic performance of bridges and vehicles under strong wind

0.1 0.2 0.3 0.4

1E-3

0.01

0.1

1

N

orm

aliz

ed sp

ectru

m

Frequency (Hz)

Single mode-based Coupled analysis

(a) Vertical mode

0.1 0.2 0.3 0.4

1E-3

0.01

0.1

1

Nor

mal

ized

spec

trum

Frequency (Hz)

Single mode-based Coupled analysis

(b) Torsion mode

Fig. 3.13 Non-dimensional normalized power spectra of buffeting response under U = 87 m/s

55

Page 63: Dynamic performance of bridges and vehicles under strong wind

are not necessary to “meet” at flutter critical wind speed as suggested in the literature. Actually, the torsion vibration excites the vertical vibration through aerodynamic coupling at the effective oscillation frequency of torsion mode, and vice versa. When aerodynamic coupling is weak, the dominant vibration is around its effective frequency (Figs. 3.6 and 3.11). With the increase of wind velocity, the resonant vibration at the effective frequency is damped out gradually since the modal damping ratio of vertical mode increases significantly as shown in Fig. 3.4. The dominant vibration frequency thus changes from its own effective oscillation frequency to the effective oscillation frequency of torsion motion. Due to its low damping ratio, the torsion vibration usually keeps its effective oscillation frequency as the dominant one for the entire range of wind speeds.

These observations indicate that the “same frequency” multimode flutter vibration is actually due to the effect of aerodynamic coupling. With the increase of wind velocity, the frequency of vertical mode increases and the frequency of torsion mode decreases correspondingly. However, the effective oscillation frequencies of vertical and torsion modes are not necessary to “meet” at flutter critical wind speed as suggested in the literature. Actually, the torsion vibration excites the vertical vibration through aerodynamic coupling at the effective oscillation frequency of torsion mode, and vice versa. When aerodynamic coupling is weak, the dominant vibration is around its effective frequency (Figs. 3.6 and 3.11). With the increase of wind velocity, the resonant vibration at the effective frequency is damped out gradually since the modal damping ratio of vertical mode increases significantly as shown in Fig. 3.4. The dominant vibration frequency thus changes from its own effective oscillation frequency to the effective oscillation frequency of torsion motion. Due to its low damping ratio, the torsion vibration usually keeps its effective oscillation frequency as the dominant one for the entire range of wind speeds.

As discussed earlier and demonstrated in Fig. 3.4, the vertical and torsion effective frequencies are not necessary to “meet” at flutter critical wind velocity. However, they do “meet” if the modal frequencies are in a specific ratio. To demonstrate this, the torsion natural frequency was reduced numerically to 66% of its original value and the vertical frequency was kept the same. As a result, a new complex eigenvalue analysis shows that the vertical and torsion frequencies meet at the flutter velocity of about 60 m/s as shown in Fig. 3.14. It can be seen from the figure that the two modal frequencies approach gradually together accompanying an increase of the damping ratio. When approaching the flutter wind velocity, the damping ratio of the vertical mode increases while the torsion mode decreases abruptly. When the two modal frequencies eventually “meet” together, the damping ratio of the torsion mode becomes negative, which leads to the flutter instability.

The motion of the example bridge consists of mainly three parts: (1) vertical motion at its effective frequency, (2) vertical motion excited at the frequency of torsion and (3) torsion motion excited at its effective frequency (torsion motion excited by vertical frequency is relatively small and neglected). Stronger coupling could be expected when two modal frequencies are close enough to be able to “meet” when flutter occurs. It also suggests that the natural frequency ratio of the torsion and vertical modes plays an important role in determining the coupling effect and the critical point of flutter.

3.6. Concluding Remarks

With the increase of span length and slenderness, bridge flutter instability and buffeting response become important issues for future bridges. Coupling effect plays an extremely

56

Page 64: Dynamic performance of bridges and vehicles under strong wind

57

important role in both flutter and buffeting of these future bridges. In this paper, a hybrid approach based on complex eigenvalue modal analysis and time domain analysis is used to better understand the natures of evolution process of bridge vibration. With this hybrid approach, the mechanism of the flutter occurrence and the actual transition process from multi-frequency dominated buffeting to the single-frequency dominated flutter are well illustrated through the numerical results of the Humen Bridge. The following conclusions can be drawn from the present study:

(a) The proposed approach provides a convenient way to numerically replicate the transition of vibrations from multi-frequency buffeting or pre-flutter free vibration to a single-frequency flutter. This transition has been observed in wind tunnel tests and is commonly accepted as a fact.

(b) The single-frequency vibration at flutter is not due to the “meeting” of the oscillation frequencies of all modes. It is due to the magnified modal coupling effect at flutter critical wind velocity. This coupling effect is very significant so that it dominates the vibration in uncritical mode. As a result, all modes are observed or felt by the instruments to vibrate in the same frequency as the effective oscillation frequency of the critical mode, though the actual effective frequency of each mode is different.

(c) The present study has shown that flutter and buffeting are well correlated in terms of their vibration characteristics. Flutter can actually be interpreted from divergent buffeting vibration. As shown numerically, by using the same aerodynamic matrices, the buffeting vibration automatically transits into a single-frequency flutter vibration at flutter critical wind velocity.

(d) Comparison of the results from the hybrid method with those from available experiments of the Humen Bridge has validated the proposed approach.

Page 65: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 4. DYNAMIC ANALYSIS OF VEHICLE-BRIDGE-WIND DYNAMIC SYSTEM

4.1 Introduction

In the United States, hurricanes and hurricane-induced strong winds have being greatly threatened the properties and lives of costal cities. In the past twenty years, annual monetary losses due to tropical cyclones and other natural hazards have been increasing at an exponential pace. In addition to the huge lose of properties, live loss is even more stunning. In extreme cases, evacuations are exceptionally necessary to minimize the loss of lives and properties. Smooth and safe transportation is obviously the key factor of a successful evacuation process. Long-span bridges are often the backbones of the transportation lines in coastal areas and are very vulnerable to strong wind loads. Maintaining the highest transportation capacity of these long-span bridges is vital to supporting hurricane evacuations. Hence to reduce the likeness of accidents on the transportation line is extremely important not only for normal operation but also for evacuation purpose when hurricane arrives.

Currently, the setting of driving speed limit and the decision of closing the transportation on bridges and highways when hurricane arrives is mostly based on intuition or subjective experience (Irwin 1999). The driving speed limit could be too high to be safe or too low to be efficient. The closing of the transportation will totally obstruct the evacuation through such transportation line. A rational prediction of the performance of vehicle-bridge system under strong winds is of utmost importance to the maximum evacuation efficiency and safety. Most existent works focus on either wind action on vehicles running on roadway (not on bridges) (Baker 1986, 1999), wind effect on the bridge without considering vehicles (Scanlan and Jones 1990), or vehicle-bridge interaction analysis without considering wind effect (Timoshenko et al. 1974; Yang and Yau 1997; Pan and Li 2002). A comprehensive vehicle-bridge-wind coupled analysis is very rare, if any.

Baker et al. have made extensive studies on the performance of high-sided road vehicles in cross winds (Baker 1986, 1987, 1991a, 1991b, 1999; Coleman and Baker 1990; Baker and Reynolds 1992). Wind effects on vehicles including static wind force and quasi-static turbulent wind force on the vehicles. In his representative works, Baker (Baker 1986) proposed the fundamental equations for wind action on vehicles. The wind speed at which these accident criteria are exceeded (the so-called accident wind speed) was found to be a function of vehicle speed and wind direction (Baker 1986). Through adoption of meteorological information the percentage of the total time for which this wind speed is exceeded can thus be found and some quantification of accident risk was made (Baker 1999). In addition to wind tunnel tests on several vehicle models to identify the wind force on vehicles (Coleman and Baker 1990), some useful statistic information about actual accidents happened in British was also collected and analyzed (Baker and Reynolds, 1992).

Interaction analysis between moving vehicles and continuum structures originated from the middle of 20th century. From an initial moving load simplification (Timishenko et al. 1974), later on a moving-mass model (Blejwas et al. 1979) to full-interaction analysis (Yang and Yau 1997; Pan and Li 2002; Guo and Xu 2001), the interaction analysis of vehicles and continuum structures (e.g. bridges) has been investigated by many people for a long time. In these studies, road roughness was treated as the sole excitation source of the coupled system.

Recently, dynamic response of suspension bridges to high wind and running train was

58

Page 66: Dynamic performance of bridges and vehicles under strong wind

investigated (Xu et al. 2003), while no wind loading on the train was considered since the train was running inside the suspension bridge deck. It was found that the suspension bridge response is dominated by wind force in high wind speed. The bridge motions due to high winds affected the safety of the running train and the comfort of passengers considerably (Xu et al. 2003). The coupled dynamic analysis of vehicle and cable-stayed bridge system under turbulent wind was also conducted recently (Xu and Guo 2003), where only vehicles under low wind speed were explored and the study did not consider many important factors, such as vehicle number, and driving speeds etc.

The present study aims at building a framework for the vehicle-bridge-wind aerodynamic analysis, which will lay a very important foundation for vehicle accident analysis based on dynamic analysis results and facilitate the aerodynamic analysis of bridges considering vehicle-bridge-wind interaction. The framework starts with building a general dynamic-mechanical model of vehicle-bridge-wind coupled system, which includes both the structural part and the loading part. Structural model simulates the bridge and a series of vehicles consisting of any number and also various types of typical vehicles, from two axle four-wheel passenger car to five axle tractor-trailer. The loading part models the external loading, like wind effect and road roughness loading on vehicles and the bridge.

After the framework is established, a series of 2-axle four-wheel high-sided vehicles on long-span bridges under strong winds are chosen as a numerical example to demonstrate the methodology. The external dynamic loading on the bridge consists of wind loading and road roughness and the external dynamic load on the vehicle includes simplified quasi-steady wind force and road roughness. With the mechanical model of the vehicle-bridge system, dynamic performance of vehicles as well as bridges is studied under strong winds.

4.2 Equations of Motion for 3-D Vehicle-Bridge-Wind System

4.2.1 Modeling of Vehicle

In mechanical engineering areas, vehicles are usually modeled in various configurations depending on what is of more concern. For the study of interactions between vehicles and structures, the vehicle model is somewhat simplified to the extent that all the relevant important information will be included. In the present study, a vehicle is modeled as a combination of several rigid bodies connected by several axle mass blocks, springs and damping devices. The bridge deck and the tire of the vehicle are assumed to be point-contact without separation. The suspension system and the elasticity of tires are modeled with springs. The energy dissipation capacities of the suspension as well as tires are modeled with viscous damping. The mass of the suspension system and the tires are assumed to concentrate on idealized mass blocks on each side of the vehicle and the mass of the springs and damping devices are zero (Fig. 4.1-2).

Fig. 4.1-2 shows a complicated tractor-trailer model; however, this model can also be used to represent other simpler vehicles through defining the number and dimensions for each rigid body, mass block, and spring-damping system. The displacements of the ith rigid body of the qth vehicle are expressed as: vertical displacement qi

vrZ , lateral displacement Y , pitching displacement in x-z plane

qivr

qivrθ , and rolling displacement in y-z plane qi

vrφ . The subscript “vr” refers to the rigid body of the vehicle. Vertical displacements of the mass blocks symmetric to the central line of the jth axle are qj

vaLZ and qjvaRZ . Lateral displacements of the mass blocks

symmetric to the central line of the jth axle are Y and Y . In the subscripts, the “L” and “R” qjvaL

qjvaR

59

Page 67: Dynamic performance of bridges and vehicles under strong wind

represent the left and right mass blocks on the jth axle, respectively; and the “a” represents the axle suspension. The superscript “qj” represents the jth axle of the qth vehicle. Same definitions apply hereafter. The longitudinal, lateral, and vertical directions of the bridge are set as x- y- and z-axis, respectively. Assuming a total of nv vehicles are considered in the analysis.

y

Zvrq1

φ vrq1

KvuRq1

CvuRq1

b1b1

Lrq1

bRZ q1

bLZ q1

Yvr

z

YvaRq1

KyuRq1

CyuRq1 KylR

q1

q1

CylRq1

KvlRq1 CvlR

q1

ZvaRq1ZvaL

q1

YvaLq1 CyuL

q1

KyuLq1

CvuLq1

KvuLq1

KylLq1

CylLq1

CvlLq1KvlL

q1

Rrq1

q q

hv

Fig.4.1 General dynamic model for various vehicle: cross section view

KvuL

x

z

z vr z vr z vrθ vr θθ vr

zvaL zvaL zvaL zvaL zvaL

KvuLq4CvuL

q5 CvuLq4

KvuLq3 KvuL

q2 KvuLq1CvuL

q3CvuL

q2 CvuLq1

CvlLq5 KvlL

q4CvlL

q4 KvlLq3

CvlLq3 KvlL

q2CvlL

q2 KvlLq1

CvlLq1

Lq1

rLq5 rL

q4 rLq3 rL

q2 rLq1

zbLq1zbL

q2zbLq3zbL

q4zbLq5

KvlLq5

q3

q3 q2

q2

q4q5 q3 q2 q1

q1

q1

q5

vr

Lq2Lq

3Lq4Lq

5Lq6Lq

7

Lq8Lq

9Lq10Lq

11

Fig.4.2 General dynamic model for various vehicles: elevation view

60

Page 68: Dynamic performance of bridges and vehicles under strong wind

The vertical displacements of upper left (right) vertical spring for the q( )qjvuL R∆ th vehicle are:

1 1 1 11 1

q q q q q qvuL vr vr vr vaL

1qZ L b Zθ φ∆ = − ⋅ + ⋅ − , 1 1 1 11 1

q q q q q qvuR vr vr vr vaR

1qZ L b Zθ φ∆ = − ⋅ − ⋅ − (4.1a, b)

2 1 1 12 1

q q q q q qvuL vr vr vr vaL

2qZ L b Zθ φ∆ = + ⋅ + ⋅ − , 2 1 1 12 1

q q q q q qvuR vr vr vr vaR

2qZ L b Zθ φ∆ = + ⋅ − ⋅ − (4.2a, b)

3 2 2 24 1

q q q q q qvuL vr vr vr vaL

3qZ L b Zθ φ∆ = − ⋅ + ⋅ − , 3 2 2 24 1

q q q q q qvuR vr vr vr vaR

3qZ L b Zθ φ∆ = − ⋅ − ⋅ − (4.3a, b)

4 3 3 36 1

q q q q q qvuL vr vr vr

4qvaLZ L b Zθ φ∆ = − ⋅ + ⋅ − , 4 3 3 3

6 1q q q q q qvuR vr vr vr vaR

4qZ L b Zθ φ∆ = − ⋅ − ⋅ − (4.4a, b)

5 3 3 17 1

q q q q q qvuL vr vr vr vaL

5qZ L b Zθ φ∆ = + ⋅ + ⋅ − , 5 3 3 17 1

q q q q q qvuR vr vr vr vaR

5qZ L b Zθ φ∆ = + ⋅ − ⋅ − (4.5a, b)

The vertical displacements of lower vertical spring ) for the q(qjvlL R∆ th vehicle are:

( )qj qj qj qjvlL vaL bL LZ Z t r∆ = − − , ( )qj qj qj qj

vlR vaR bR RZ Z t r− −∆ = (j =1 to na) (4.6a, b)

where is the vertical displacement of the left (right) mass block for the j( )qjvaL RZ th axle of qth

vehicle; and ( )qj

L Rr is the road surface roughness height at the contacting point between the bridge

deck and the left (right) wheel of the jth axle of qth vehicle; is the semi-width of the q1qb th vehicle

(Fig.4.1); na is the total axle number of the qth vehicle. The displacements on bridge at the points of contacting with left (right) wheel for the jth axle of the qth vehicle ( )( )

qjbL RZ t are expressed as:

( ) ( ) ( ), ,qjbL b qj b qj LqjZ t h x t x t dα= + , ( ) ( ) ( ), ,qj

bR b qj b qj RqjZ t h x t x t dα= + ; (4.7a, b)

where and ( ,bh x t ) ( ),b x tα are vertical and torsion displacement of the bridge, respectively; , are the horizontal distance to the torsion center from the left and right tires of j

Lqjd Rqjdth axle of the qth vehicle, respectively; and xqj is the position along the bridge for the jth axle of

the qth vehicle with j = 1 to na.

As shown in Fig. 4.2, there are two joints connecting the rigid bodies of the tractor-trailer. For vertical displacement continuity at these connecting points, we have:

1 1 28

q q q q qvr vr vr vrZ L Z L 2

9qθ θ+ = − (4.8)

2 2 310 11

q q q q qvr vr vr vrZ L Z L 3qθ θ+ = − (4.9)

The two dependent variables 2qvrθ and 3q

vrθ can be eliminated by solving Eqs. (4.8) and (4.9) as:

(2 2 18

9

1q q qqvr vr vr vrZ Z LL )1q qθ θ= − − (4.10)

( )3 3 2 19 9 10 10 10 8

9 11

1q q q q q q q q qq qvr vr vr vr vrL Z L L Z L Z L LL L

1q qθ θ= − + + + (4.11)

61

Page 69: Dynamic performance of bridges and vehicles under strong wind

The lateral displacements of upper lateral left (right) spring for the q( )qjyuL R∆ th vehicle are:

qj qi qjyuL vr vaLY Y∆ = − , (j = 1 to nqj qi qj

yuR vr vaRY Y∆ = − a) (4.12a, 12b)

The lateral displacements of lower lateral left (right) spring for the q( )qjylL R∆ th vehicle are:

qj qj qjylL vaL bLY Y∆ = − , (j =1 to nqj qj qj

ylR vaR bRY Y∆ = − a) (4.13a, b)

where are the lateral displacements of the bridge at the points of contacting with the left (right) wheels of the j

( )qj

bL RYth axle for the qth vehicle and are defined as follows:

( ) ( ) ( ,qj qjbL bR b qjY t Y t p x t= = ) (4.14)

4.2.2 Modeling of Bridge

The dynamic model of the long-span bridge can be obtained through finite element method using different kinds of finite elements such as beam elements and truss elements. With the obtained mode shapes, the response corresponding to any point along the bridge deck can be evaluated in the time domain.

Uh

p

αTorsional

center

Lq1d

bb

b

Rq1dy

zx

Fig. 4.3 Coupled model of vehicle on bridge

Motions in three directions including vertical, lateral, and torsion of the bridge are expressed with the mode superposition technique as (Fig. 4.3):

h x ,( ) ( ) (1

,n

b ii

t h x tξ=

=∑ )i ( ) ( ) (1

,n

b ii

)ip x t p x tξ=

=∑ , ( ) ( ) (1

,n

b ii

)ix t xα α=

=∑ tξ (4.15a-c)

62

Page 70: Dynamic performance of bridges and vehicles under strong wind

where n is the total number of modes for the bridge under consideration; ( )i xh ( )ip x and ( )i xα are the components in the three directions for the ith mode shape; and is the generalized coordinates of the bridge.

( )i tξ

For bridges immersing in the flow, the wind forces on the bridge in three directions (vertical, lateral and torsion) are denoted as ( ),b

wL x t , ( ),bwD x t and ( ),b

wM x t , respectively:

( ) ( ) ( ), ,bw st ae b ,L x t L L x t L x t= + + (4.16)

( ) ( ) (, ,bw st ae b ),D x t D D x t D x t= + + (4.17)

( ) ( ) ( ), ,bw st ae b ,M x t M M x t M x t= + + (4.18)

where the subscripts “st”, “ae” and “b” refer to static, self-excited, and buffeting force component due to wind, respectively.

The static wind force of unit span length can be expressed as:

20.5st LL U Cρ= B 20.5st DD U Cρ=, , B 20.5st M2M U C Bρ= (4.19-21)

where ρ is the air density; U is the mean wind speed on the elevation of the bridge; B is the bridge deck width; LC , DC and MC are the lift, drag, and moment static wind force coefficients for the bridge that are usually obtained from section model wind tunnel tests of the bridge deck.

The self-excited force can be expressed as:

( ) ( ) ( ) ( ) ( ) ( )2 * * 2 * 2 * * 2 *1 2 3 4 5 6

, , , ,1 ,2

b b b bae b

h x t B x t h x t p x t p x tL U B KH KH K H x t K H KH K H

U U B Uα

ρ α

= + + + + +,b

B

(4.22)

( ) ( ) ( ) ( ) ( ) ( )2 * * 2 * 2 * * 2 *1 2 3 4 5 6

, , , ,1 ,2

b b b bae b

,bp x t B x t p x t h x t h x tD U B KP KP K P x t K P KP K P

U U B Uα

ρ α

= + + + + +B

(4.23)

( ) ( ) ( ) ( ) ( ) ( )2 2 * * 2 * 2 * * 2 *1 2 3 4 5 6

, , , ,1 ,2

b b b b bae b

h x t B x t h x t p x t p x tM U B KA KA K A x t K A KA K A

U U B Uα

ρ α

= + + + + +,

B

(4.24)

where BKUω

= is the reduced frequency; , P and *iH *

i*iA (i = 1 to 6) are the flutter derivatives of

the bridge obtained from the wind tunnel tests of bridge deck; ω is the vibration frequency of the system; and the dot on the cap denotes the derivative with respect to the time.

The buffeting forces for a unit span in vertical, lateral, and torsion directions are:

( ) ( )2 '1 22b L L D

u t w tL U B C C CU U

ρ

= + +

( ) (4.25)

( )21 22b D D

u t w tD U B C CU U

ρ

=

' ( ) + (4.26)

63

Page 71: Dynamic performance of bridges and vehicles under strong wind

( )2 2 '1 22b M M

u t w tM U B C CU U

ρ

=

( ) + (4.27)

where ( )u t and are the horizontal and vertical components of wind turbulent velocity, respectively; and the prime denotes the derivative with respect to the attack angle of wind.

( )w t

4.2.3 Coupled Equations of Vehicle-Bridge Model in Modal Coordinates

By assuming all displacements remain small, virtual works generated by the inertial forces, damping forces, external wind loading, and elastic forces can be obtained for the bridge-vehicle system. The equilibrium condition of bridge under its self-weight without vehicles is chosen as the initial position, which facilitates the direct comparison between the cases of with and without vehicles. The summation of all the virtual works should be zero:

0v b v b v b v binertia inertia damping damping elastic elastic external externalW W W W W W W Wδ δ δ δ δ δ δ δ+ + + + + + + = (4.28)

where

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ), , , , ,binertia b b b b b b b b bW m x h x t h x t I x x t x t m x p x t p x tδ δ α δα= + + ,δ (4.29)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ), , , , ,b h pdamping b b b b b b b b bW C x h x t h x t C x x t x t C x p x t p x tαδ δ α δα= + + ,δ (4.30)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ), , , , ,b h pelastic b b b b b b b b bW K x h x t h x t K x x t x t K x p x t p xαδ δ α δα= + + , tδ

)

δ

δ

δ

t

)

(4.31)

( ) ( ) ( ) ( ) ( ) ( ) (1 1

, , , , , ,v an n

b b b b qj qj qj qjexternal w b w b w b bL GvL bR GvR

q jW L x t h x t D x t p x t M x t x t Z F Z Fδ δ δ δα δ δ

= =

= + + + +∑∑ (4.32)

)(1 1 1 1

v ar qi qi qi qi qi qin n nnvaL vaL vaL vaR vaR vaRv qi qi qi qi qi qi qi qi qi qi qi qi

inertia vr vr vr vr vr vr vr vr vr vr vr vr qi qi qi qi qi qiq i q i vaL vr vr vaR vr vr

M Z Z M Z ZW M Z Z M Y Y I J

M Y Y M Y Y

δδ δ δ θ δθ φ δφ

δ δ= = = =

+= + + + + + + ∑∑ ∑

v

∑ (4.33)

1 1

v aqi qi qi qi qi qi qi qi qi qi qi qin nvuL vuL vuL vuR vuR vuR yuL yuL yuL yuR yuR yuRv

damping qi qi qi qi qi qi qi qi qi qi qi qiq i vlL vlL vlL vlR vlR vlR ylL ylL ylL ylR ylR ylR

C C C CW

C C C C

δ δ δδ

δ δ δ δ= =

∆ ∆ + ∆ ∆ + ∆ ∆ + ∆ ∆ =

+ ∆ ∆ + ∆ ∆ + ∆ ∆ + ∆ ∆ ∑∑ (4.34)

1 1

v aqi qi qi qi qi qi qi qi qi qi qi qin nvuL vuL vuL vuR vuR vuR yuL yuL yuL yuR yuR yuRv

elastic qi qi qi qi qi qi qi qi qi qi qi qiq i vlL vlL vlL vlR vlR vlR ylL ylL ylL ylR ylR ylR

K K K KW

K K K K

δ δ δδ

δ δ δ δ= =

∆ ∆ + ∆ ∆ + ∆ ∆ + ∆ ∆ =

+ ∆ ∆ + ∆ ∆ + ∆ ∆ + ∆ ∆ ∑∑ (4.35)

( ) ( ) ( )1 1

v rn nv qi qi qi qi qi qi

external vr wz vr wy vr wq i

W Z F t Y F t M φδ δ δ δφ= =

= + +∑∑ (4.36)

( ) ( ) (1 1

2a rn n

qj qj qj qivaL vaR vr aGvL R

j i

F M M M g n= =

= − + +

∑ ∑

(4.37)

nr is the total number of rigid bodies for the vehicle, respectively; g is the gravity acceleration; qi

wzF , ( )qiwyF t and qi

wM φ are the wind forces in vertical direction (axis z), lateral direction (axis y), and wind-induced rolling moment along x axis, respectively, on the ith rigid body of the qth vehicle, which will be defined later; F is the gravity force on the bridge at the location of the

( )qj

GvL R

64

Page 72: Dynamic performance of bridges and vehicles under strong wind

left (right) wheel of the jth axle for the qth vehicle due to the weight of the vehicle. It is approximated as the average weight on each wheel of the whole vehicle.

The vertical displacement of the vehicle at the contact point of two tires for the jth axle can also be expressed based on Eqs. (4.7a, b) as:

( ) ( ) ( ) ( )1

nqjbL i qj i qj Lqj i

i

Z t h x x dα=

= +∑ tξ (4.38)

( ) ( ) ( ) ( )1

nqjbR i qj i qj Rqj i

i

Z t h x x dα=

= +∑ tξ

tα ξ δξ

α ξ δξ

α ξ δξ

δξ

δξ

)

(4.39)

Virtual work for the bridge motion is derived by substituting Eq. (4.15) into Eqs. (4.29) to (4.32) as:

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )1 1 1 1

n n n nb

inertia b k i k i b k i k ik i i i

W m x h x h x p x p x I x x x tδ α= = = =

= + + ∑ ∑ ∑ ∑ (4.40)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )1 1 1 1

n n n nb h p

damping b k i b k i b k i k ik i i i

W C x h x h x C x p x p x C x x x t tαδ α= = = =

= + +

∑ ∑ ∑ ∑ (4.41)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )1 1 1 1

n n n nb h p

elastic b k i b k i b k i k ik i i i

W K x h x h x K x p x p x K x x x t tαδ α= = = =

= + +

∑ ∑ ∑ ∑ (4.42)

( ) ( ) ( ) ( ) ( ) ( ) ( )

( ) ( ) ( ) ( ) ( )

1

1 1 1

, , ,

v a

nb b b b

external w i w i w i ii

n n nqj qj

i qj i qj Lqj GvL i qj i qj Rqj GvR iq j i

W F x t h x P x t p x M x t x t

h x x d F h x x d F t

δ α

α α

=

= = =

= + +

+ + + +

∑∑ ∑

(4.43)

The vertical deformations of left and right lower vertical springs are:

( ) ( ) ( ) (1

nqj qjvlL vaL i qj i qj Lqj i L qj

i

Z h x x d t r xα ξ=

∆ = − + − ∑ (4.44)

( ) ( ) ( ) (1

nqj qjvlR vaR i qj i qj Rqj i R qj

i)Z h x x d t r xα ξ

=

∆ = − + − ∑ (4.45)

The deformation velocities of left and right lower vertical springs are:

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )1 1

qjn ni qj i qj L qjqj qj

vlL vaL i qj i qj Lqj i Lqj ii i

h x x r xZ h x x d t d V t t V t

x x xα

α ξ ξ= =

∂ ∂ ∂ ∆ = − + − + − ∂ ∂ ∂

∑ ∑ (4.46)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )1 1

qjn ni qj i qj R qjqj qj

vlR vaR i qj i qj Rqj i Rqj ii i

h x x r xZ h x x d t d V t t V t

x x xα

α ξ ξ= =

∂ ∂ ∂ ∆ = − + − + − ∂ ∂ ∂

∑ ∑ (4.47)

where ( )V t is the driving speed of the qth vehicle and is defined as:

( ) ( )x tV t

t∂

=∂

(4.48)

The virtual displacements are derived from Eqs. (4.44) and (4.45) as: 65

Page 73: Dynamic performance of bridges and vehicles under strong wind

( ) ( ) ( )1

nqj qjvlL vaL i qj i qj Lqj i

i

Z h x x d tδ δ α δξ=

∆ = − +∑ (4.49)

( ) ( ) ( )1

nqj qjvlR vaR i qj i qj Rqj i

i

Z h x x d tδ δ α δξ=

∆ = − +∑ (4.50)

and the virtual works for the vehicle motion are derived from Eqs. (4.33) to (4.36) as:

( )

2 31 1 2 3 110

99 11

32 39 102 2 2 3 2 3 3 3 3

99 11 11

1

q q qq q q q qvr vrvr vr vr vr vrq q q

q q qq qvrq q q q q q q q qvr vr

vr vr vr vr vr vr vr vr vrq q q q

vinertia

q qvr vr

I I LM Z ZL L L

I L LI IM Z Z M Z ZL L L L

W

I

θ θ δ

θ θ δ θ δ

δ

θ

− + +

+ + − + + + =

( )328 101 2 3 1 1 1 1 2 2 2 3 38

9 9 11

q q qq qvrq q q q q q q q q q qvr

vr vr vr vr vr vr vr vr vr vr vr vrq q q

qj qj qj qj qj qj qj qj qj qj qjvaL vaL vaL vaR vaR vaR vaL vr vr vaR vr vr

I L LL I J J JL L L

M Z Z M Z Z M Y Y M Y Y

θ θ δθ φ δφ φ δφ φ δ

δ δ δ δ

− + + + +

+ + +( )

1

1 1

v

a r

n

q

n nqj qi qj qj

vr vr vrj i

M Y Yδ

=

= =

+

∑ ∑

3qφ +

(4.51)

( )( ) ( )

1 1 1 1 2 2 2 2 1

3 3 3 3 2 4 4 4 4 5 5 5 5 3

1 1 1 1

q q q q q q q q qvuL vuL vuR vuR vuL vuL vuR vuR vr

q q q q q q q q q q q q q qvuL vuL vuR vuR vr vuL vuL vuR vuR vuL vuL vuR vuR vr

q q q qvuL vuL vuR vuR v

vdamping

C C C C Z

C C Z C C C C Z

C C C

W

δ

δ δ

δ

∆ + ∆ + ∆ + ∆ +

∆ + ∆ + ∆ + ∆ + ∆ + ∆ +

− ∆ − ∆ +

=

( ) (( )

2 2 2 2 1 4 3 3 3 3 2

4 4 4 4 5 5 5 5 36 6 7 7

1 1 1 1 1 2 2 2

q q q q q q q q q q quL vuL vuR vuR vr vuL vuL vuR vuR vr

q q q q q q q q q q q q qvuL vuL vuR vuR vuL vuL vuR vuR vr

q q q q q q q qvuL vuL vuR vuR vuL vuL vuR vuR

C L C C

L C L C L C L C

b C C C C

δθ δθ

δθ

∆ + ∆ + ∆ + ∆ +

− ∆ − ∆ + ∆ + ∆ +

∆ − ∆ + ∆ − ∆( ) (

( ) (

( ) ( )

2 1 2 3 3 3 3 2

3 4 4 4 4 5 5 5 5 3

1

a

q q q q q q q qvr vuL vuL vuR vuR vr

nq q q q q q q q q q qi qi qi qi qi qi

vuL vuL vuR vuR vuL vuL vuR vuR vr vuL vuL vaL vuR vuR vaRj

qj qj qjvlL vlL vlL i qj i qj Lqj

b C C

b C C C C C Z C Z

C Z h x x d

δφ δφ

δφ δ δ

δ α δ

=

+ ∆ − ∆ +

∆ − ∆ + ∆ − ∆ − ∆ + ∆ +

∆ − +

( )

( ) ( ) ( )

)(

1

1

1

1

a

a

n

jni

nj qj qj qj

vlR vlR vlR i qj i qj Rqj ji

nqi qi qi qi qi qi qi qi qi qi qi qiyuL yuL yuL yuR yuR yuR ylL ylL ylL ylR ylR ylR

i

t

C Z h x x d t

C C C C

ξ

δ α δξ

δ δ δ δ

=

=

=

=

+ + ∆ − +

∆ ∆ + ∆ ∆ + ∆ ∆ + ∆ ∆

∑∑

1

vn

q=

)

)

)

(4.52)

66

Page 74: Dynamic performance of bridges and vehicles under strong wind

( )( ) ( )

1 1 1 1 2 2 2 2 1

3 3 3 3 2 4 4 4 4 5 5 5 5 3

1 1 1 1 2 2

q q q q q q q q qvuL vuL vuR vuR vuL vuL vuR vuR vr

q q q q q q q q q q q q q qvuL vuL vuR vuR vr vuL vuL vuR vuR vuL vuL vuR vuR vr

q q q q q qvuL vuL vuR vuR vuL vuL

vstiffness

K K K K Z

K K Z K K K K Z

K K K

W

δ

δ δ

δ

∆ + ∆ + ∆ + ∆ +

∆ + ∆ + ∆ + ∆ + ∆ + ∆ +

− ∆ − ∆ + ∆

=

( ) (( )

( )

2 2 1 3 3 3 3 24

4 4 4 4 5 5 5 5 36 6 7 7

1 1 1 1 1 2 2 2 2 1 2 3

q q q q q q q q qvuR vuR vr vuL vuL vuR vuR vr

q q q q q q q q q q q q qvuL vuL vuR vuR vuL vuL vuR vuR vr

q q q q q q q q q q q qvuL vuL vuR vuR vuL vuL vuR vuR vr vuL vuL

K L K K

L K L K L K L K

b K K K K b K

δθ δθ

δθ

δφ

+ ∆ + ∆ + ∆ +

− ∆ − ∆ + ∆ + ∆ +

∆ − ∆ + ∆ − ∆ + ∆( )

( ) (

( ) ( ) ( )

3 3 3 2

3 4 4 4 4 5 5 5 5 3

1

1

a

q q q qvuR vuR vr

nq q q q q q q q q q qi qi qi qi qi qi

vuL vuL vuR vuR vuL vuL vuR vuR vr vuL vuL vaL vuR vuR vaRj

nqj qj qjvlL vlL vlL i qj i qj Lqj j

i

qj qjvlR vlR vl

K

b K K K K K Z K Z

K Z h x x d t

K Z

δφ

δφ δ δ

δ α δξ

δ

=

=

− ∆ +

∆ − ∆ + ∆ − ∆ − ∆ + ∆ +

∆ − +

+ ∆

( ) ( ) ( )

)(

1

1

1

1

v

a

a

n

q

n

nj qj

R i qj i qj Rqj ji

nqi qi qi qi qi qi qi qi qi qi qi qiyuL yuL yuL yuR yuR yuR ylL ylL ylL ylR ylR ylR

i

h x x d t

K K K K

α δξ

δ δ δ δ

=

=

=

=

+ − +

∆ ∆ + ∆ ∆ + ∆ ∆ + ∆ ∆

∑∑

)

)

t

(4.53)

( ) ( ) ( )1 1

v rn nv qi qi qi qi qi qi

external vr wz vr wy vr wq i

W Z F t Y F t M φδ δ δ δφ= =

= + +∑∑ (4.54)

The coupled equations can be finally built based on Eq. (4.28) as

G

+ + = +

v vv v v vb v v vb v r w

s v s v b bb b bv b b b bv b b b r w

F + FM 0 γ C C γ K K γ0 M γ C C + C γ K K + K γ F + F F b

(4.55)

where subscripts “b” and “v” represent for bridge and vehicle, respectively; superscripts of “s” and “v” in the stiffness and damping terms for the bridge refer to the terms of bridge structure itself and those contributed by the vehicles, respectively; subscripts “bv” and “vb” refer to the vehicles-bridge coupled terms; “r”, “w” and “G” represent for roughness, wind, and gravity force, respectively; and vγ and bγ are the displacement vectors of the vehicles and the bridge, respectively;

, , , ,= vTn1 q

v v v vγ γ γ γ… (4.56)

1, , , , Tb i nξ ξ ξ=γ (4.57)

( ), , , , ,diag= vn1 2 qv v v vM Μ M M M v

, ( )diag= vn1 2 qv v v vK ,K , ,K , ,KvK ; ( )diag vn1 2 q

v v v v= C ,C , ,C , ,CvC (4.58a-c)

( )diag= vn1 2 qvb vb vb vb vbC C ,C , ,C , ,C ; ( ), , , , ,diag= vn1 2 q

vb vb vb vb vbK K K KK (4.59a-b)

1 1 1 1 1 2 2 2 2 3 3 3 3 4 4 5 5

1 1 1 2 2 2 3 3 3 4 4 5

, , , , , , , , , , , , , , , , ,

, , , , , , , , , , , ,

q q q q q q q q q q q q q q q q qvr vr vr vaL vaR vaL vaR vr vr vaL vaR vr vr vaL vaR vaL vaRq

v q q q q q q q q q q q qvr vaL vaR vaL vaR vr vaL vaR vr vaL vaR vaL va

Z Z Z Z Z Z Z Z Z Z Z Z Z

Y Y Y Y Y Y Y Y Y Y Y Y Y

θ φ φ φ=γ

5

T

qR

(4.60)

Detailed formulations of matrices and force vectors of the qth vehicle are referred to 4.6.

67

Page 75: Dynamic performance of bridges and vehicles under strong wind

4.3 Dynamic Analysis of Vehicle-Bridge System under Strong Wind

Strong wind effects on moving vehicles and bridge are the concerns of this study. As introduced in Section 2, the wind effect on the bridge consists of static wind force, aeroelastic self-excited force and turbulent buffeting force. It is usually assumed that the analysis begins with the static equilibrium position of the bridge under self weight and static wind force action. Only dynamic loads of wind, namely the last two components of Eqs. (4.16-4.18) are considered for the bridge.

4.3.1 Self-excited Forces on the Bridge

The wind dynamic load on the bridge consists of self-excited force and buffeting force. The self-excited force is related to the motion of the bridge and thus frequency dependent. In the time domain analysis, the frequency dependent variables are difficult to be incorporated. Under each wind velocity, the vibration frequency ω at any time should be determined to quantify the self-excited force terms. As an alternative to the rational function approximation (Chen et al. 2000), complex eigenvalue analysis can also predict the vibration frequency iteratively for the dominant motion at any time step under any wind velocity. In the present study, the complex eigenvalue analysis is conducted first to give the vibration frequency corresponding to each time through interactive process. The results are then incorporated into the vehicle-bridge-wind coupled equations to decide the self-excited force terms of the bridge. This approach to deal with aeroleastic terms of the wind force has been adopted by the writers (Chen and Cai 2003a).

4.3.2 Buffeting Force Simulation on the Bridge

To calculate the time-history response of the vehicle-bridge system under the wind action on the bridge, the stochastic wind velocity field should be simulated. The fast spectral representation method proposed by Cao et al. is adopted here (Cao et al. 2000). The time history of wind component u(t), at the jth point along the bridge span can be generated with:

( ) ( ) ( ) ( ) (1 1

2 cfNj

j u mq jm mq mq mqm q

u t S G tω ω ω ω= =

= ∆ +∑∑ )os ψ , j = 1, 2,… Ns (4.61)

where Nf is a sufficiently large number representing the total number of frequency intervals; Ns is the total number of points along the bridge span to simulate; Su is the spectral density of turbulence in along-wind direction; mqψ is a random variable uniformly distributed between 0 and 2π; /up fNω ω∆ = is the frequency increment; upω is the upper cutoff frequency with the

condition that the value of ( )uS ω is less than a preset small number ε when upω ω> ; and

( )2

0 1

1,

1 2

s

j mjm s

j ms

when j m N

G C when m m j

C C when m j N

ω −

≤ < ≤= =

− ≤ ≤

N≤ ≤

(4.62a)

7exp2

CUωπ

− ∆=

(4.62b)

68

Page 76: Dynamic performance of bridges and vehicles under strong wind

where ∆ is the distance between two consecutive simulated points. Similarly, the time history of vertical turbulence component w(t) can be obtained.

In the present study, wind velocity time histories are simulated for various points along the bridge with the four times the element length interval as those used in the finite element dynamic analysis on the bridge (Lin et al. 1998). With the same assumption as in the spectral representation method (Cao et al. 2000), wind velocity time history is assumed the same within each interval. With the wind velocity time history, the buffeting force time history corresponding to each point along the bridge span can be obtained with Eqs. (4.25-27). With the mode shapes obtained from finite element analysis, the wind buffeting forces on the whole bridge can be obtained through integrating all the force time histories along the bridge span.

4.3.3 Time History Simulation of Road Surface Roughness

The road surface roughness is usually assumed to be a zero-mean stationary Gaussian random process and can be expressed on the inverse Fourier transformation on a power spectral density function (Huang and Wang 1992) as:

( ) ( )1

( ) 2 cos 2N

k kk

r x S x kφ φ πφ=

= ∆∑ θ+ (4.63)

where kθ is the phase angle uniformly distributed between 0 to 2π; φ is the wave number (cycle/m); ( )S φ is the Power Spectrum Density (PSD) function for the road surface elevation and Huang and Wang (1992) suggested following PSD function:

( )

2

0rS A φφφ

= (4.64)

where 0φ is the discontinuity frequency of 1/2π (cycle/m) and Ar is the roughness coefficient (m3 /cycle) related to the road condition.

4.3.4 Quasi-static Wind Effect on the Vehicle

Wind action on a running vehicle includes static and dynamic load effects. The quasi static wind forces on vehicles are adopted since a transient type of force model is not available (Baker 1999):

( )212wx a r DF U C Aρ ψ= ; ( )21 ;2wy a r SF U Cρ ψ= A ( )21

2wz a r LF U C Aρ ψ= ; (4.65a-c)

( )212w a r R vM U C Ahφ ρ ψ= ; ( )21

2w a r P vM U C Ahθ ρ ψ= ; ( )212wz a r Y vM U C Ahρ ψ= ; (4.66a-c)

where Fwx, Fwy, Fwz, Mwφ, Mwθ and Mwz are the drag force, side force, lift force, rolling moment, pitching moment and yawing moment acting on the vehicle, respectively. CD, CS, CL, CR, CP and CY are the coefficients of drag force, side force, lift force, rolling moment, pitching moment and yawing moment for the vehicle, respectively. “A” is the frontal area of the vehicle; hv is the distance from the gravity center of the vehicle to the road surface; Ur is the relative wind speed to the vehicle, which is defined as (Fig. 4.3):

( ) ( )( ) ( )( )22 , , cos ,rU x t V U u x t U u x t2

sinβ β = + + + + (4.67)

69

Page 77: Dynamic performance of bridges and vehicles under strong wind

( )( )( )( ), sin

tan, cos

U u x tV U u x t

βψ

β+

=+ +

(4.68)

where V is the driving speed of vehicle; U and u(x, t) are the mean wind speed and turbulent wind speed component on the vehicle, respectively; β is the attack angle of the wind to the vehicle, which is the angle between the wind direction and the vehicle moving direction (Fig. 4.4); ψ is usually between 0 to π.

βU+u(x)

UrV

ψmoving direc

tion of v

ehicle

Fig. 4.4 Velocity diagram on vehicles

4.3.5 Numerical solutions

To simulate the process of vehicles running on the bridge, the position of any vehicle changes with time. Correspondingly, the coefficients of the coupled equations are also time dependent. Therefore, the matrices in Eq. (4.55) should be updated at each time step after the new position of each vehicle is identified. In the present paper, both the Wilson-θ method and the second-order Rouge-Kutta approach are tried and Rouge-Kutta method is finally chosen for the virtues of convenience and accuracy. A computer program based on Matlab is developed to solve the differential equations.

4.4 Numerical Example

4.4.1 Prototype bridge and vehicles

The Yichang Suspension Bridge located in a typhoon zone of south China has a main span of 960 m and two side spans of 245 m each. The height of the bridge deck above water is 50 m. The sketch of the bridge is shown in Fig. 4.5 and its main parameters are shown in Table 2.1 (Lin et al. 1998). Based on a preliminary analysis with modal coupling assessment technique (Chen et al. 2004) and the observation of wind tunnel results, the four important modes considered in the present study are shown in Table 4.1 to consider the three-direction response of the bridge (Lin et al. 1998). Except explicitly specified, the wind attack angle for the bridge is zero.

70

Page 78: Dynamic performance of bridges and vehicles under strong wind

Table 4.1 Dynamic Properties of Four Main Modes for Yichang Suspension Bridge

Mode number Natural frequency (Hz) Mode type

1 0.105 1st Vertical Mode-Asymmetric

2 0.161 1st Vertical Mode-Symmetric

3 0.337 1st Torsion Mode-Symmetric

4 0.423 1st Torsion Mode-Asymmetric

960 m 245 m245 m

(a)

29990 mm

24480 mm

5998 mm 8997 mm 8997 mm 5998 mm

2.5%2.5 %

(b)

Fig. 4.5 Prototype bridge (Lin et al. 1998): (a) elevation view (b) cross-section view

71

Page 79: Dynamic performance of bridges and vehicles under strong wind

Table 4.2 Main parameters of the vehicle used in the numerical example

Parameter Value Unit

Mass of rigid body (Mv1) 4500 kg

Pitching moment inertia (θv1) I v1 5483 kg.m2

Rolling moment inertia (φv1) Jv1 1352 kg.m2

Axle mass MaL1, MaR1 800 Kg

Axle mass MaL2, MaR2 700 Kg

Upper vertical spring stiffness ( ,,1qvuLK 1q

vuRK ,2q

vuLK ,2q

vuRK ) 400 kN/m

Lower vertical spring stiffness ( ,,1qvlLK 1q

vlRK ,2q

vlLK ,2q

vlRK ) 350 kN/m

Upper vertical damping coefficient (C ,,C1qvuL

1qvuR , C 2q

vuL , C 2qvuR ) 20 kN.s/m

Lower vertical damping coefficient ( ,,1qvlLC 1q

vlRC ,2q

vlLC , C 2qvlR ) 1 kN.s/m

Upper lateral spring stiffness ( ,,1qyuLK 1q

yuRK ,2q

yuLK ,2q

yuRK ) 300 kN/m

Lower lateral spring stiffness ( ,,1qylLK 1q

ylRK ,2q

ylLK ,2q

ylRK ) 120 kN/m

Upper lateral damping coefficient (C ,,C1qyuL

1qyuR , C 2q

yuL , C 2qyuR ) 20 kN.s/m

Lower lateral damping coefficient ( ,,1qylLC 1q

ylRC ,2q

ylLC ,2q

ylRC ) 1 kN.s/m

Vehicle frontal area (A) 10 m2

Reference height hv 2 m

L1 2.9 m

L2 5.0 m

b1 1.0 m

Total length of the vehicle 13.4 m

1Lqd , 2Lqd 5 m

72

Page 80: Dynamic performance of bridges and vehicles under strong wind

As shown in Section 4.2, the introduced generalized vehicle model is applicable in simulating various vehicles, from two-axle cars, trucks, buses to five-axle tractor-trailer. To consider the wind effect on one particular vehicle, wind tunnel test should usually be conducted to obtain the wind force coefficients. Unfortunately, available wind tunnel data are limited to very few kinds of vehicles (Baker, 1991a,b). In the present study, one four-wheel vehicle is used in the numerical example. A sketch of the vehicle model is shown in Fig. 4.6 and the main parameters are listed in Table 4.2. The reason to choose this vehicle is mainly because this four-wheel vehicle has quite similar dimensions to that used in Baker’s work (Coleman and Baker 1990), where wind force coefficients from wind tunnel test are available. This vehicle has seven degrees-of-freedom, three for the rigid body of vehicle (vertical displacement Zv1, pitching about y axis θv1 and rolling about x axis φv1) and the other four for the vertical displacements of the four wheels (ZaL1, ZaR1, ZaL2, ZaR2) (Fig.4.1).

Wind force coefficients can be expressed as follows (Baker 1987):

For 0 / 2ψ π≤ <

( )0.3821SC a ψ= ; ( )2 1 sin 3LC a ψ= + ; ( )3 1 2sin 3DC a ψ= − + ; (4.69-71)

( )1.774YC a ψ= − ; C a ; C a ; (4.72-74) ( )1.32

5P ψ= ( )0.2946R ψ=

For 0 ψ π< <

( )0.3821SC a π ψ= − ;C a ;( )( )( )2 1 sin 3L π ψ= + − ( )( )( )3 1 2sin 3D π ψ= + − −C a ; (4.75-77)

( )1.774YC a π ψ= − − ; ; C a ; (4.78-80) ( )1.32

5PC a π ψ= − − ( )0.2946R π ψ= −

where 1α to 6α equal to 5.2, 0.94, -0.5, 2.0, -2.0 and 7.3, respectively (Coleman and Baker 1990).

To properly simulate the traffic on the bridge, traffic flow information is required to define the actual vehicle type and distribution. At any different moment, vehicles on the bridge may have quite different numbers and distribution patterns and locate in different lanes randomly. Since each vehicle may have at least several degrees of freedoms, it is still technically difficult to conduct a coupled analysis on all vehicles together based on simulated real traffic flow using current computer techniques. It is especially true when the bridge is long and the traffic is busy. The common practice in analyzing the interaction between vehicles and bridges is to choose only one vehicle or a series of identical vehicles in one line (Yang and Yau 1997; Guo and Xu 2001). In the present study, only one line of up to three vehicles is assumed to be evenly distributed along the side lane with an interval of 5 m and the lateral distance to the torsion center of the bridge dLq1= dLq2= 5m (Fig. 4.3). The position of vehicles and the interval setting are decided based on some considerations to preliminarily simulate the crowded transportation reality during a hurricane evacuation process. The effect of the number of vehicles will also be studied in the following numerical analysis.

73

Page 81: Dynamic performance of bridges and vehicles under strong wind

x

vaL

Lq1Lq2

q2

z

Bridgestructure

z vrθ q1

q1

vr

zKvuL

q2CvuL

q2

KvlLq2

CvlLq2

rLq2

zbLq2

zvaL

KvuLq1 CvuL

q1

KvlLq1

CvlLq1

rLq1

zbLq1

q1

(a)

y

Zvrq1

φ vrq1

KvuRq1

CvuRq1

b1b1

Lrq1

bRZ q1

bLZ q1

Yvr

z

YvaRq1

KyuRq1

CyuRq1 KylR

q1

q1

CylRq1

KvlRq1 CvlR

q1

ZvaRq1ZvaL

q1

YvaLq1 CyuL

q1

KyuLq1

CvuLq1

KvuLq1

KylLq1

CylLq1

CvlLq1KvlL

q1

Rrq1

q q

hv

(b)

Fig. 4.6 Vehicle used in the example: (a) elevation view (b) cross-section view

74

Page 82: Dynamic performance of bridges and vehicles under strong wind

0 50 100 150 200 250 300 350 400

-0.4

-0.2

0.0

0.2

0.4

u(

t)/U

Time (s)

(a)

0 50 100 150 200 250 300 350 400-0.4

-0.2

0.0

0.2

0.4

w(t)

/U

Time (s)

(b) Fig. 4.7 Simulated fluctuating wind velocities at middle point of main span:

horizontal wind velocity u(t) (b) vertical wind velocity w(t)

The turbulent wind velocities are simulated (see Eq. (4.61)) along the bridge span with the simulation interval ∆=25 m, corresponding to the length of four finite elements along the main span. The total number of frequency intervals Nf equals 1024 and the upper cutoff frequency equals to 2π. Figure 4.7 gives the time histories of wind velocity in the middle point of main span when the average wind speed on the deck U= 40 m/s. The roughness displacement is also derived after the adoption of roughness factor of 20×10-6 m3/cycle for good road condition as shown in Fig. 4.8 (Huang and Wang 1992). Second-order Rouge-Kutta approach is adopted to solve the differential equations with an integration time step of 0.015 second and a total steps of 3000.

With zero initial conditions, the time history of response for the bridge only and the

75

Page 83: Dynamic performance of bridges and vehicles under strong wind

vehicles on the road are predicted under the wind action. It is assumed that after 50 seconds, the vehicles enter the main span of the bridge, the integration starts with the initial conditions of the moment after 50 seconds. Only the dynamic responses of the bridge and vehicles during the period of the vehicles running on the bridge are demonstrated hereafter. The assumption of 50 seconds before the vehicles enter the bridge is to make sure the bridge has enough time to be excited by the wind with stable response.

0 200 400 600 800 1000-0.015

-0.010

-0.005

0.000

0.005

0.010

0.015

Rou

ghne

ss (m

)

Distance along the bridge (m)

Fig. 4.8 Simulated vertical road roughness for the bridge

4.4.2 Vehicle Dynamic Response

Among the three vehicles, the first vehicle is chosen to monitor the response in the study. Since the absolute response of the vehicles on the bridge includes the contribution from the bridge motion, the relative response of the vehicles to the bridge gives more valuable information for the vehicle safety. The relative response of vehicles is obtained through deducting the corresponding bridge response from the absolute response of the vehicles. Under wind speed U = 40 m/s, vehicles with speeds of V = 10 and 20 m/s are studied separately. In Fig. 4.9, the relative vertical displacement of the rigid body for the 1st vehicle is displayed. There is an increase of displacement when vehicle driving speed changes from 10 m/s to 20 m/s. Vehicles with different driving speeds have different location on the bridge at the same time, and it is observed that the vertical response of vehicles with 20 m/s driving speeds are increased when they are in the mid-span area (about 24-30 seconds in Fig. 4.9). Same tendency can also be observed for vehicles with 10m/s driving speed in the corresponding mid-span area (about 45 seconds in Fig. 4.9). The phenomena suggest that vertical motion of the vehicles have strong coupling with symmetric modes of the bridge. It is also noted that the mean value of the relative time history corresponds to the static response of the vehicle excited by static wind force

76

Page 84: Dynamic performance of bridges and vehicles under strong wind

component. Since the static wind forces vary with the driving speed that affects the resultant of wind velocity Ur as shown in Eqs. (4.67) and (4.68), the mean values of the curves have slight shift vertically, which contributes partially to the difference of the response between different driving speeds shown in Fig. 4.9. However, the dynamic amplification of the vertical response contributed by the roughness, dynamic component of wind action on the vehicles, and their interaction with the bridge can obviously be observed.

The relative rolling displacement of the rigid body of the 1st vehicle is also shown in Fig. 4.10. Compared with Fig. 4.9, Fig. 4.10 shows a slight increase of rolling response of the vehicles in the location around the quarter span (about 12 seconds for V=20m/s or 24 seconds for V=10 m/s). No obvious difference exists when vehicles pass the middle of the main span. The results suggest that the difference of rolling response of the vehicle under different driving speed is relatively insignificant comparing with differences of vertical response. The moderate dynamic coupling effects exist between vehicle rolling motion and the asymmetric modes of the bridge.

As an important variable to identify the risk of overturning accident, the absolute rolling angular acceleration is also predicted under different wind speeds shown in Fig. 4.11(a) where the absolute angular acceleration changes little for different driving speed. The spectral density amplitude, shown in Fig. 4.11 (b) is obtained from spectral analysis of the time history. It is found that rolling acceleration of the vehicle is mostly dominated by the bridge response, which is indicated by several peaks including one corresponding to the torsional modal frequency of the bridge (marked with a circle in the figure).

The vertical relative response of vehicles on the highway bridge and road are compared with the same road surface conditions, same driving speed (V= 20 m/s) and same wind speed (U= 40 m/s). Results in Fig. 4.12 show that the vehicle dynamic response on the road is steadier, while its response on bridge increases significantly when time is between 15-30 seconds, which corresponds to the time when the vehicles pass the middle of the main span. In other time period, little differences on the vehicle relative response can be observed between the vehicles on the road and on the bridge. Such phenomenon shows again that the vertical relative response of the vehicle interacts with the bridge motion, especially with the symmetric modes.

The relative vertical response of the vehicle on the bridge under different wind speeds is also studied when the driving speed still remains at 20 m/s. Results displayed in Fig. 4.13 clearly suggest that the dynamic response of the vehicles under higher wind speed are much larger than that under low wind speed especially in some critical locations on the bridge (e. g., mid-span). Again, such result further justifies the concern about the safety of the vehicles under strong winds.

In real traffic situations, many vehicles could run on the bridge at the same time. The relative dynamic response of the rigid body of the first vehicle is compared in Fig. 4.14 when there is only one vehicle and totally three vehicles on the bridge. The results show very insignificant difference on the vehicle relative displacement when the number of vehicles changes. However, the absolute response of the vehicle does decrease with the increase of vehicle numbers as shown in Fig. 4.15.

4.4.3 Bridge Dynamic Response

Figs. 4.16 and 4.17 show the time histories of vertical and torsion dynamic response in the middle point of the main span under wind speed U = 40 m/s where two driving speeds V =10m/s

77

Page 85: Dynamic performance of bridges and vehicles under strong wind

and V= 20m/s are studied separately and the results are also compared with the bridge response without any vehicle when wind speed remains at U= 40 m/s. It can be found that the bridge dynamic response decreases slightly with vehicles compared with that without any vehicle. Such decrease is more significant when the vehicles approach the middle area of the main span (t= 20-35 second). Such decrease for the torsion response of the bridge is probably due to the eccentric weight effect of the vehicles locating on the windward side of the bridge, which suppresses the bridge response comparing with the bridge without any vehicle. For vertical modes, strong coupling and the weight of vehicles may all contribute to the decrease of the vertical response. As stated before, for a given time, vehicles with different driving speed are located at different places on the bridge. For example, since the main span of the bridge is 960 m, vehicles with 20 m/s driving speed will roughly pass the middle of the main span at the time of 24 seconds, while vehicles with 10 m/s will pass the quarter point of the span (will need 48 seconds to pass the mid-span). It is found from the figures that only when vehicles are close to the place where the response is monitored (mid-point of the main span, in this case), the bridge response differs with different vehicle driving speeds.

0 5 10 15 20 25 30 35 40 45

-0.1

0.0

0.1

Vehi

cle

rela

tive

verti

cal d

ispl

acem

ent (

m)

Time (s)

V = 10 m/s V = 20 m/s

Fig. 4.9. Vertical relative displacement for the rigid body of the 1st vehicle when U = 40 m/s

78

Page 86: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45-0.1

0.0

0.1

Veh

icle

rela

tive

rolli

ng a

ngle

φ (r

ad.)

Time (s)

V = 10 m/s V = 20 m/s

Fig. 4.10. Vehicle relative rolling displacement for the rigid body of the 1st vehicle when U = 40 m/s

79

Page 87: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40-4

-2

0

2

4

A

bsol

ute φ''

(rad

/s2 )

Time (s)

V = 10 m/s V = 20 m/s

(a)

0 1 2 3 4 5-0.010.000.010.020.030.040.050.060.07

Spe

ctra

l den

sity

am

plitu

de

Frequency (Hz)

V= 10 m/s V= 20 m/s

(b)

1st symmetric torsion frequency

of bridge

Fig. 4.11. Absolute rolling acceleration for the rigid body of the 1st vehicle: (a) Time history (b) Spectral analysis results

80

Page 88: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45

-0.1

0.0

0.1

Veh

icle

rela

tive

verti

cal d

ispl

acem

ent (

m)

Time (s)

On the road On he bridge

Fig. 4.12 Relative vertical displacements for the rigid body of the 1st vehicle on the bridge and on the road

81

Page 89: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45

-0.1

0.0

0.1

Vehi

cle

rela

tive

verti

cal d

ispl

acem

ent (

m)

Time (s)

U = 5 m/s U = 40 m/s

Fig. 4.13 Relative vertical displacements for the rigid body of the 1st vehicle under different wind speed

82

Page 90: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45

-0.1

0.0

0.1

0.2

Veh

icle

rela

tive

verti

cal d

ispl

acem

ent (

m)

Time (s)

1 vehicle 3 vehicles

Fig. 4.14 Relative vertical displacements for the rigid body of the 1st vehicle on the bridge with different number of vehicles when U=40 m/s and V=20 m/s

83

Page 91: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45-1.2

-0.8

-0.4

0.0

0.4

Ve

hicl

e ab

solu

te v

ertic

al d

ispl

acem

ent (

m)

Time (s)

1 vehicle 3 vehicles

Fig. 4.15 Absolute vertical displacements for the rigid body of the 1st vehicle on the bridge with different number of vehicles when U=40 m/s and V=20 m/s

84

Page 92: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45

-4

-3

-2

-1

0

1

2

3

4

Brid

ge v

ertic

al d

ispl

acem

ent (

m)

Time (s)

No vehicle V = 20m/s V = 10 m/s

Fig. 4.16 Vertical displacement of the bridge at wind speed U = 40 m/s with vehicles

85

Page 93: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45

-0.1

0.0

0.1

Brid

ge to

rsio

n di

spla

cem

ent (

rad.

)

Time (s)

No vehicle V = 20m/s V = 10 m/s

Fig. 4.17 Torsion displacement of the bridge at wind speed U = 40 m/s with vehicles

86

Page 94: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45-4

-3

-2

-1

0

1

2

3

Br

idge

ver

tical

dis

plac

emen

t (m

)

Time (s)

1 vehicle 3 vehicles

Fig. 4.18 Vertical displacement of the bridge at wind speed U = 40 m/s

87

Page 95: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20 25 30 35 40 45-0.1

0.0

0.1

B

ridge

tors

ion

disp

lace

men

t (ra

d.)

Time (s)

1 vehicle 3 vehicles

Fig. 4.19 Torsion displacement of the bridge at wind speed U = 40 m/s

Figs. 4.18 and 4.19 show the response of the bridge with different number of vehicles but the same driving speed V=20 m/s. The bridge response has slight change with different vehicles numbers in a more than straightforward manner. It is probably because that there are multiple excitations from vehicles in different locations on the bridge when there is more than one vehicle. Such slight change of the bridge response with different vehicle numbers, however, may become significant when there are many vehicles on the same bridge together. More specific research on the response of the bridge under real transportation loading and wind are deserved, which will be reported later by the authors. The vertical acceleration of the bridge is also shown in Fig. 4.20. The bridge vertical acceleration responses are quite close when the number of vehicles changes except slight suppression effect with the existence of vehicles. Figs. 18-20 suggest that the existence of vehicles on the bridge may have some suppression effect on the bridge response in high winds if they are located in some places on the bridge (e. g. the middle point of main span).

88

Page 96: Dynamic performance of bridges and vehicles under strong wind

25 26 27 28 29 30-3

-2

-1

0

1

2

3

Brid

ge v

ertic

al a

ccel

erat

ion

(m/s

2 )

Time (s)

No vehicle 1 vehicle 3 vehicles

Fig. 4.20 Absolute bridge vertical acceleration when U= 40 m/s and V = 20 m/s

The bridge response with vehicles under low wind speed is also studied. In comparison to high wind speed, the bridge response is affected by the driving speed of vehicles when wind speed is as low as 5 m/s. With the moving of the vehicles towards the middle point of the span, the curves also show a downward tendency, which is an indication of static displacement of the bridge due to the vehicle gravity (Figs. 4.21 and 4.22). It shows that when wind speed is low, wind effect does not dominate the bridge response like the case when wind speed is high. Under low wind speed, alternatively, the roughness and the static gravity effect of vehicles have become dominant on the bridge response. In this case, different driving speeds may cause totally different response of the bridge.

89

Page 97: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40-0.06

-0.05

-0.04

-0.03

-0.02

-0.01

0.00

0.01

0.02

0.03

0.04

Brid

ge v

ertic

al d

ispl

acem

ent (

m)

Time (s)

V= 10 m/s V= 20 m/s

Fig. 4.21 Vertical displacement of the bridge at wind speed U = 5m/s

The bridge response with vehicles without considering any wind effect is studied to further confirm aforementioned conclusions (Figs. 4.23 and 4.24). In the figures, vehicle static gravity effect dominates the vertical displacement of the bridge. For each driving speed, the bridge displacement of monitoring (middle point of main span) reaches the peak when the vehicles pass the monitoring location. For driving speed of 20 m/s, the curve is roughly symmetric about the peak point of 24 seconds when the vehicles pass the mid-span. It can also be found that the two curves reach the peak values at different time. Such differences on the curves, however, are not obvious when there are wind action on the bridge and the vehicles, for example, Figs. 4.16-4.17.

90

Page 98: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40-0.0020

-0.0015

-0.0010

-0.0005

0.0000

0.0005

Brid

ge to

rsio

n di

spla

cem

ent (

rad)

Time (s)

V= 10 m/s V= 20 m/s

Fig. 4.22 Torsion displacement of the bridge at wind speed U = 5m/s

91

Page 99: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40

-0.01

0.00

0.01

B

ridge

ver

tical

dis

plac

emen

t (m

)

Time (s)

V=10 m/s V=20 m/s

Fig. 4.23 Vertical displacement of the bridge with vehicles without considering wind effect

92

Page 100: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40-0.0006

-0.0005

-0.0004

-0.0003

-0.0002

-0.0001

0.0000

0.0001

0.0002

Brid

ge to

rsio

n di

spla

cem

ent (

rad)

Time (s)

V= 10 m/s V =20 m/s

Fig. 4.24 Torsional displacement of the bridge with vehicles without considering wind effect

4.5 Concluding Remarks

A framework for vehicle-bridge-wind aerodynamic analysis is developed to consider different types and arbitrary number of vehicles on the bridge. Each vehicle is modeled with several rigid bodies, mass blocks, and spring-damping systems. The vehicle model is suitable for modeling different types of cars and trucks up to five axles. This framework provides a tool to systematically investigate bridge and vehicle performance in strong winds, such as bridge flutter and buffeting, and vehicle vibration and instability, considering the interaction of bridge, vehicles, and winds. A comprehensive analysis of vehicle accidents on bridges and highways, such as overturning and lateral sliding, are on-going taking advantage of this analytical framework. These results will be reported in the near future.

To demonstrate the capability of the established analytical framework, the present study investigates numerically the dynamic performance of vehicles and the bridge under strong winds and also compared with the results in low wind speed, considering three identical 2-axle vehicles running on a prototype long-span bridge. Effects of different driving speeds and number of vehicles on the dynamic performance of vehicles and bridge are studied. Dynamic wind loading and road roughness loading are applied on the bridge and quasi-steady wind force and road roughness loading are acting on the vehicles. Based on the present numerical analysis, following conclusions can be drawn:

The vehicle driving speed does affect the relative vertical response of vehicles. The rolling relative response is dominated by direct static wind action on the vehicles and only slightly changes with the different driving speeds.

93

Page 101: Dynamic performance of bridges and vehicles under strong wind

The absolute response of vehicles is dominated by the bridge response and thus varies little with driving speed when wind speed is high. Compared with vehicles on the road, vehicles responses are amplified on the bridge especially when the vehicle approaches the middle point of the main span. Vehicles show stronger relative response under high wind speed, which justifies again the importance of study on the vehicle safety under strong winds.

The relative response of the vehicle is hardly affected by the existence of other vehicles, however, the absolute responses of vehicles change with the number of vehicles on the bridge. Multiple-vehicle simulation on the bridge is still necessary to realistically assess the response of vehicles as well as the bridge.

The vehicles on the bridge could reduce the dynamic response of the bridge in some cases under high winds. It is possibly due to the beneficial effects of eccentric gravity of vehicles, which caused some suppression effect of vibration. However, the effects may depend on many facts, such as positions of vehicles on the bridge, vehicle number, wind speed, and driving speed. More studies are needed to draw a general conclusion.

The bridge response is not sensitive to the vehicle with different driving speed when wind speed is high because the wind-induced response dominates the bridge response. However, when wind speed is low, the bridge response is dominated by the static component induced by the gravity of vehicles and dynamic part contributed by road roughness. Different driving speed has more significant effect on the bridge response when wind speed is low.

4.6 Matrix Details of the Coupled System

Please be noted that those terms not listed in the upper triangle of matrices are zero terms. The matrices are symmetric and the lower triangle terms of matrices are thus not given here.

[ ]2 2

1 2 101 2 31,1

9 9 1

1 qq q q

v vr vrq q q

L 3

1

qvrM M I

L L Lχ χ χ

= + +

I ;[ ] ( )

( )

2

8 10 33 21,2

9 11

q qq q

v vq q

L LrM I

L Lχ= −

[ ] ( )( )

210 9 102 3

2 3 21,89 9 11

1q q q

q q qv vrq q q

L L LvrM I I

L L Lχ χ

+ = − −

;[ ]

( )310

3 21,129 11

qq q

v vq q

LrM I

L Lχ= ;

[ ] 12,2

q qv vrM I= ;[ ]

( )( )

( )8 9 102 38

2 32 22,89 9 10 11

q q qqq q q

v vrq q q

L L LLvrq

M I IL L L L

χ χ+

= − − ;[ ]( )

310 822,12

9 11

q qq q

v vq q

L LrM I

L L= ;

[ ] 13,3

q qv vrM J= ;[ ] 1

4,4

q qv vaLM M= ;[ ] 1

5,5

q qv vM M= aR ;[ ] 2

6,6

q qv vaLM M= ;[ ] 2

7,7

q qv vM M= aR ;

[ ] 2r8,8

q qv vM M= ;[ ]

9 103

9 10 11

q qq

v vq q q

L L8,12

qrM I

L L L+

= − ;[ ] 29,9

q qv vrM J= ;[ ] 3

10,10

q qv vaLM M= ;[ ] 3

11,11

q qv vM M= aR ;

[ ] 3qr12,12

qv vM M= ;[ ] 3q

r13,13

qv vM J= ;

[ ] 4qaL14,14

qv vM M= ;[ ] 4q

v vaR15,15

qM M= ;[ ] 5qv vaL16,16

qM M= ;[ ] 5qv vaR17,17

qM M= ;

94

Page 102: Dynamic performance of bridges and vehicles under strong wind

[ ] 118,18

q qv vrM M= ;.[ ] 1

19,19

q qv vM M= aL ;[ ] 1

20,20

q qv vM M= aR ;[ ] 2

21,21

q qv vaLM M= ;[ ] 2

22,22

q qv vM M= aR ;

[ ] 223,23

q qv vrM M= ; [ ] 3

24,24

q qv vaLM M= ;[ ] 3

25,25

q qv vaRM M= ;[ ] 3

26,26

q qv vrM M= ;[ ] 4

27,27

q qv vaLM M= ;

[ ] 428,28

q qv vaRM M= ;[ ] 5

29,29

q qv vaLM M= ;[ ] 5

30,30

q qv vM M= aR ;

[ ]

( ) ( ) ( ) ( ) ( )

1 1 2 21,1

2 22 23 3 4 4 5 5104

2 3 6 79 9 11

q q q q qv vuL vuR vuL vuL

qqq q q q q q q qvuL vuR vuL vuR vuL vuRq q q

C C C C C

LL C C L C C L C CL L L

χ χ

= + + + +

+ + + + +

;

[ ] ( ) ( ) ( )( )

( )

( )( )

( ) ( ) ( ) ( )

2

4 81 1 2 2 3 31 2 2 21,2

9

22 210 8 4 4 5 5

3 6 72

9 11

q qq q q q q q q q q

v vuL vuR vuL vuR vuL vuRq

q qq q q q q q

vuL vuR vuL vuRq q

L LC L C C L C C C C

L

L LL C C L C C

L L

χ

χ

= − + + + + + +

+ + +

[ ] ( ) ( )1 1 2 111,3

q q q q q qv vuL vuR vuL vuRC b C C C C = − + − ;[ ] 1

1,4

q qv vuLC C= − ;[ ] 1

1,5

q qv vC C= − uR ;[ ] 2

1,6

q qv vC C= − uL ;[ ] 2

1,7

q qv vC C= − uR ;

[ ] ( )( )

( ) ( )( )

( ) ( ) ( ) ( )2

2 24 9 4 9 9 103 3 4 4 5 52 3 6 72 21,8

9 9 11

q q q q q qq q q q q q q q q

v vuL vuR vuL vuR vuL vuRq q q

L L L L L LC C C L C C L C

L L Lχ χ

+ +C = − + − + + +

[ ] ( )3 342 21,9

9

qq q q q

v vuq

LC b CL

χ= − −L vuRC ;[ ] 3421,10

9

qq q

v vq

LC CL

χ= uL;[ ] 34

21,119

qq q

v vq

LC CL

χ= uR;

[ ]( )

( )( )( ) ( )( )( )2 24 4 5 5103 6 11 6 7 11 721,12

9 11

qq q q q q q q q q q q

v vuL vuR vuL vuRq q

LC L L L C C L L L CL L

χ = − − + + − + C ;

[ ] ( ) ( )4 4 5 53 103 6 71,13

9 11

q qq q q q

v vuL vuR vuL vuRq q

b LC L C C L CL L

χ = − − − − C ;

[ ] 46 1031,14

9 11

q qq q

v vq q

L LC CL L

χ= uL;

[ ] 46 1031,15

9 11

q qq q

v vq q

L LC CL L

χ= uR ;[ ] 57 1031,16

9 11

q qq q

v vq q

L LC CL L

χ= − uL ;[ ] 57 1031,17

9 11

q qq q

v vq q

L LC CL L

χ= −

( )

uR;

[ ] ( ) ( ) ( ) ( )( )

( )

( ) ( ) ( ) ( )

22 2 4 81 1 2 2 3 3

1 1 2 2 22,29

22 24 4 5 58 10

3 6 79 11

q qq q q q q q q q q

v vuL vuR vuL vuR vuL vuRq

q qq q q q

vuL vuR vuL vuRq q

L LC L C C L C C C C

L

L L L C C L C CL L

χ χ

χ

= + + + + +

+ + +

+;

95

Page 103: Dynamic performance of bridges and vehicles under strong wind

[ ] ( ) (1 1 2 21 1 1 22,3

q q q q q q q q qv vuL vuR vuL vuRC b L C C b L C C= − − + − ) ;[ ] 1

12,4

q q qv vC L C= uL ; [ ] 1

12,5

q q qv vC L C= uR [; ] 2

22,6

q q qv vC L= − uLC ;

[ ] 222,7

q q qv vuRC L C= −

[ ]

;

( )( )

( ) 4 48 10 6

q q qvuL vuRL L L C( ) ( )5 5q

vuRC( )2 2

7q q q

vuLC L3 34 8 9 102 3 22,8

9 9 11

q q q qq q q q

v vuL vuRq q q

L L L LC C CL L L

χ χ + C = − + + + + +

[ ] ( )3 34 82 12,9

9

q qq q q q

v vq

L LC b CL

χ= − −uL vuRC ;[ ] 34 822,10

9

q qq q

v vq

L LC CL

χ= uL;[ ] 34 8

22,11

q qq q

v vqq

L LC CL

χ= uR;

[ ]( )

( )( ) ( )(4 4 5 58 103 6 6 11 7 7 1122,12

9 11

q qq q q q q q q q q q q

v vuL cuR vuL vurq q

L LC L L L C C L L L CL L

χ = − + + + )C+ ;

[ ] ( ) ( )4 4 5 58 103 1 6 72,13

9 11

q qq q q q q q q

v vuL vuR vuL vuRq q

L LC b L C C L C CL L

χ = − − + − q ;[ ] 46 8 10

32,149 11

q q qq q

v vq q

L L LC CL L

χ= uL ;

[ ] 46 8 1032,15

9 11

q q qq q

v vq q

L L LC CL L

χ= uR;[ ] 57 8 10

32,169 11

q q qq q

v vq q

L L LC CL L

χ= − uL;[ ] 57 8 10

32,179 11

q q qq q

v vq q

L L LC CL L

χ= − uR

vuR

[ ] ( )2 1 1 2 213,3

q q q q q qv vuL vuR vuLC b C C C C = + + + ; [ ] 1

13,4

q q qv vC b C= − uL ; [ ] 1

13,5

q q qv vC b C= uR ;[ ] 2

13,6

q q qv vC b C= − uL ;

;[[ ] 213,7

q q qv vuRC b C= ] 1 1

4,4

q q qv vuL vlLC C C= + ;[ ] ;1 1q q

v vuR vlRC+5,5

qC C= [ ] 2 2q quL vlLC+

6,6

qv vC C= ;[ ] 2 2q q

uR vlRC+7,7

qv vC C= ;

[ ] ( ) ( ) ( ) ( ) ( )2

2 23 3 4 4 5 54 9 9 102 3 6 78,8

9 9 11

q q q qq q q q q q q q q

v vuL vuR vuL vuR vuL vuRq q q

L L L LC C C L C C L CL L L

χ χ + + C = + + + + +

[ ] ( )3 32 18,9

q q q qv vuLC b C Cχ= − vuR

328,10

q qv vC Cχ= −;[ ] ;uL [ ] 3

28,11

q qv vC Cχ= − uR ;

[ ] ( )( )

( ) 4 411 6

q q q q qvuL vuRL L= − ( ) ( )2

7q q

vuLC C L+ + ( )5 5qvuRC

9 10

3 628,129 11

q qq

v q q

L LC L

L Lχ

+C+ ;

[ ] ( ) ( ) (1 9 10 4 4 5 43 6 78,13

9 11

q q qq q q q q q q

v vuL vuR vuL vuRq q

b L LC L C C L C

L Lχ

+ = − − )C− ;[ ] ( )6 9 10 4

38,149 11

q q qq q

v vq q

L L LC C

L Lχ

+= − uL

;

[ ] ( )6 9 10 438,15

9 11

q q qq q

v vq q

L L LC C

L Lχ

+= − uR

;[ ] ( )7 9 10 538,16

9 11

q q qq q

v vq q

L L LC C

L Lχ

+= uL

;[ ] ( )7 9 10 538,17

9 11

q q qq q

v vq q

L L LC C

L Lχ

+= uR

C

;

[ ] ( )2 3 319,9q q q

v vuLC b C C = + vuR3

19,10

q q qv vC b= −; [ ] ; uL [ ] 3

19,11

q q qv vC b C= uR ;

[ ] 3 310,10

q q qv vuLC C= + vlLC ;[ ] 3

110,13

q q qv vC b= − uLC [ ] 3 3

11,11

q q qv vuRC C C= + vlR ;[ ] 3

111,13

q q qv vC b C= uR ;

[ ] ( ) ( )2 2

4 4 5 511 6 11 712,12

11 11

q q q qq q q q q

v vuL vuR vuL vuRq q

L L L LC C C CL L

− += + +

C+ ;

[ ] ( )( ) ( )(4 4 5 5111 6 7 1112,13

11

qq q q q q q q q q

v vuL vuR vuL vuRq

bC L L C C L L CL

= − − + + − )C ;

96

Page 104: Dynamic performance of bridges and vehicles under strong wind

[ ] 46 1112,14

11

q qq q

v vq

L LC CL−

= uL ;[ ] 46 1112,15

11

q qq q

v vq

L LC CL−

= uR ; [ ] 57 1112,16

11

q qq q

v vq

L LC CL+

= − uL ;[ ] 57 11412,17

11

q qq q

v vq

L LC CL+

= − uE ;

[ ] ( )2 4 4 5 5113,13

q q q q q qv vuL vuR vuL vuRC b C C C C = + + + ;

[ ] 4 414,14

q q qv vuLC C= + vlLC 4 4

15,15

q q qv vuRC C= +;[ ] ;vlRC [ ] 5 5

16,16

q q qv vuLC C= + vlLC ;[ ] 5 5

17,17

q q qv vuRC C= + vlRC ;

[ ] ( )2

18,181

q qj qjv yuL

j

C C=

= +∑ yuRC 118,19

q qv yC C= −;[ ] ;uL [ ] 1

18,20

q qv yC C= − uR ;[ ] 2

18,21

q qv yC C= − uL ;[ ] 2

18,22

q qv yC C= − uR ;

[ ] 1 119,19

q q qv yuLC C= + ylLC ;[ ] 1 1

20,20

q q qv yuRC C= + ylRC ;[ ] 2 2

21,21

q q qv yuLC C= + ylLC ;[ ] 2 2

22,22

q q qv yuRC C= + ylRC ;

[ ] 3 323,23

q q qv yuLC C C= + yuR

323,24

q qv yC C= −;[ ] ;uL [ ] 3

23,25

q qv yC C= − uR ;[ ] 3 3

24,24

q q qv yuLC C= + ylLC ;[ ] 3 3

25,25

q q qv yuRC C= + ylRC ;

[ ] ( )5

26,264

q qj qjv yuL

j

C C=

= +∑ yuRC 426,27

q qv yC C= −;[ ] ;uL [ ] 4

26,28

q qv yC C= − uR ;[ ] 5

26,29

q qv yC C= − uL ;[ ] 5

26,30

q qv yC C= − uR ;

[ ] 4 427,27

q q qv yuLC C= + ylLC ;[ ] 4 4

28,28

q q qv yuRC C= + ylRC ;[ ] 5 5

29,29

q q qv yuLC C= + ylLC ;[ ] 5 5

30,30

q q qv yuRC C= + ylRC ;

where

1 2 3

1 2 3

1 2 3

1, 0 0 1

1, 1 0 2

1, 1 1 3

r

r

r

and if n

and if n

and if n

χ χ χ

χ χ χ

χ χ χ

= = = =

= = = =

= = = =

vK has the same format as C except the C is replaced by K in the matrices. v

[ ] ( ) ( )1

114,

q

q qvb vlL j j Lqj x x

C C h x x dα=

= − − ;[ ] ( ) ( )1

115,

q

q qvb vlR j j Rqj x x

C C h x x dα=

= − − ;

[ ] ( ) ( )2

226,

q

q qvb vlL j j Lqj x x

C C h x x dα=

= − − ;[ ] ( ) ( )2

227,

q

q qvb vlR j j Rqj x x

C C h x x dα=

= − − ;

[ ] ( ) ( )3

3310,

q

q qvb vlL j j Lqj x x

C C h x x dα=

= − − ;[ ] ( ) ( )3

3311,

q

q qvb vlR j j Rqj x x

C C h x x dα=

= − − ;

[ ] ( ) ( )4

4414,

q

q qvb vlL j j Lqj x x

C C h x x dα=

= − − ;[ ] ( ) ( )4

4415,

q

q qvb vlR j j Rqj x x

C C h x x dα=

= − − ;

[ ] ( ) ( )5

5516,

q

q qvb vlL j j Lqj x x

C C h x x dα=

= − − ;[ ] ( ) ( )5

5517,

q

q qvb vlR j j Rqj x x

C C h x x dα=

= − − ;

[ ] ( )1119,

q qvb ylL j qj

C C p= − x ;[ ] ( )1120,

q qvb ylR j qj

C C p= − x ;[ ] ( )2221,

q qvb ylL j qj

C C p= − x ;

[ ] ( )2222,

q qvb ylR j qj

C C p= − x ;

97

Page 105: Dynamic performance of bridges and vehicles under strong wind

[ ] ( )3324,

q qvb ylL j qj

C C p= − x ;[ ] ( )3325,

q qvb ylR j qj

C C p= − x ;[ ] ( )4427,

q qvb ylL j qj

C C p= − x ;

[ ] ( )4428,

q qvb ylR j qj

C C p= − x ;[ ] ( )5529,

q qvb ylL j qj

C C p= − x ;[ ] ( )5530,

q qvb ylR j qj

C C p= − x .

[ ] ( ) ( ) ( ) ( ) ( )1

1 11 14,

q

q j jq qvb vlL Lq vlL j j Lqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂

[ ] ( ) ( )

;

( ) ( ) ( )1

1 11 15,

q

q j jq qvb vlR Rq vlR j j Rqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂ ;

[ ] ( ) ( ) ( ) ( ) ( )2

2 22 26,

q

q j jq qvb vlL Lq vlL j j Lqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂

[ ] ( ) ( )

;

( ) ( ) ( )2

2 22 27,

q

q j jq qvb vlR Rq vlR j j Rqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂ ;

[ ] ( ) ( ) ( ) ( ) ( )3

3 33 310,

q

q j jq qvb vlL Lq vlL j j Lqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂

[ ] ( ) ( )

;

( ) ( ) ( )3

3 33 311,

q

q j jq qvb vlR Rq vlR j j Rqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂

[ ] ( ) ( ) ( ) ( ) ( )4

4 44 414,

q

q j jq qvb vlL Lq vlL j j Lqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂

[ ] ( ) ( )

;

( ) ( ) ( )4

4 44 415,

q

q j jq qvb vlR Rq vlR j j Rqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂ ;

[ ] ( ) ( ) ( ) ( ) ( )5

5 55 516,

q

q j jq qvb vlL Lq vlL j j Lqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂

[ ] ( ) ( )

;

( ) ( ) ( )5

5 55 517,

q

q j jq qvb vlR Rq vlR j j Rqj

x x

h x xK C V t d K h x x d

x xα

α=

∂ ∂ = − + − + ∂ ∂ ;

[ ] ( ) ( ) ( )11 1119,

q j qq qvb ylL ylL jj

p xK C V t K p

x∂

= − −∂

x ( );[ ] ( ) ( )11 1120,

q j qq qvb ylR ylR j qj

p xK C V t K p

x∂

= − −∂

x ;

[ ] ( ) ( ) ( )22 2221,

q j qq qvb ylL ylL j qj

p xK C V t K p

x∂

= − −∂

x ( );[ ] ( ) ( )22 2222,

q j qq qvb ylR ylR j qj

p xK C V t K p

x∂

= − −∂

x ;

[ ] ( ) ( ) ( )33 1324,

q j qq qvb ylL ylL j qj

p xK C V t K p

x∂

= − −∂

x ( );[ ] ( ) ( )33 3325,

q jq qvb ylR ylR j qj

p xK C V t K p

x∂

= − −∂

x ;

98

Page 106: Dynamic performance of bridges and vehicles under strong wind

[ ] ( ) ( ) ( )44 4427,

q j qq qvb ylL ylL j qj

p xK C V t K p

x∂

= − −∂

x ( );[ ] ( ) ( )44 4428,

q j qq qvb ylR ylR j qj

p xK C V t K p

x∂

= − −∂

x ;

[ ] ( ) ( ) ( )55 5529,

q j qq qvb ylL ylL j qj

p xK C V t K p

x∂

= − −∂

x ( );[ ] ( ) ( )55 4530,

q j qq qvb ylR ylR j qj

p xK C V t K p

x∂

= − −∂

x ;

j=1to n

( ) ( ) ( )1

11 1 1

4q

qvq Lq q q

vlL vlL Lrx x

r xF C V t K r x

x=

∂= + ∂

; ( ) ( ) ( )1

11 1 1

5q

qvq Rq q q

vlR vlR Rrx x

r xF C V t K r x

x=

∂= + ∂

;

( ) ( ) ( )2

22 2 2

6q

qvq Lq q q

vlL vlL Lrx x

r xF C V t K r x

x=

∂= + ∂

; ( ) ( ) ( )2

22 2 2

7q

qvq Rq q q

vlR vlR Rrx x

r xF C V t K r x

x=

∂= + ∂

;

( ) ( ) ( )3

33 3 3

10q

qvq Lq q q

vlL vlL Lrx x

r xF C V t K r x

x=

∂= + ∂

; ( ) ( ) ( )3

33 3 3

11q

qvq Rq q q

vlR vlR Rrx x

r xF C V t K r x

x=

∂= + ∂

;

( ) ( ) ( )4

44 4 4

14q

qvq Lq q q

vlL vlL Lrx x

r xF C V t K r x

x=

∂= + ∂

; ( ) ( ) ( )4

44 4 4

15q

qvq Rq q q

vlR vlR Rrx x

r xF C V t K r x

x=

∂= + ∂

;

( ) ( ) ( )5

55 5 5

16q

qvq Lq q q

vlL vlL Lrx x

r xF C V t K r x

x=

∂= + ∂

; ( ) ( ) ( )5

55 5 5

17q

qvq Lq q q

vlR vlR Lrx x

r xF C V t K r x

x=

∂= + ∂

;

[ ] ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ), 0

l

b b i k b i k b i ki kM m x h x h x m x p x p x I x x x dxα α= + + ∫ (i, k=1 to n)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ), 0

ls pb b i k b i k b i ki k

C C x h x h x C x p x p x C x x x dxα α α = + + ∫ (i, k=1 to n)

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ), 0

ls pb b i k b i k b i ki k

K K x h x h x K x p x p x K x x x dxα α α = + + ∫ (i, k=1 to n)

v vk= +bbK K K vc

b

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

,1 1

qt

v a

qt

qt qt

qtvlL i i Lqt j j Lqt

x xn n

vk qtb vlR i i Rqt j j Rqti j x xq t

qt qtylL i j ylR i jx x x x

K h x x d h x x d

K K h x x d h x x d

K p x p x K p x p x

α α

α α

=

== =

= =

+ + = + + +

∑∑

+

+

99

Page 107: Dynamic performance of bridges and vehicles under strong wind

( ) ( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( ) ( ) ( )

,

qt

qt

qt qt

j jqtvlL i i Lqt Lqt

x x

j jvc qtb vlR i i Rqt Rqti j

x x

j jqt qtylL i ylR i

x x x x

h x xC h x x d d V t

x x

h x xK C h x x d d

x x

p x p xC p x V t C p x V t

x x

αα

αα

=

=

= =

∂ ∂ + + ∂ ∂ ∂ ∂ = + + ∂ ∂

∂ ∂ + ∂ ∂

1 1

v an n

q t= =

+

+

∑∑

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

,1 1

qt

v a

qt

qt qt

qtvlL i i Lqt j j Lqt

x xn n

v qtb vlR i i Rqt j j Rqti j x xq t

qt qtylL i j ylR i jx x x x

C h x x d h x x d

C C h x x d h x x d

C p x p x C p x p x

α α

α α

=

== =

= =

+ + = + + +

∑∑

( ) ( ) ( )

+

+

( ) ( ) ( ) ( ) ( ) ( ) ( )1 1

v a

qt

qt qtn nb L Rqt qt qt qt qt qt

vlL L vlL i Lqt i vlR R vlR i Rqt irq t

x x

r x r xF K r x C V t h x d x K r x C V t h x d x

x xα α

= ==

∂ ∂ = − + + + + + ∂ ∂ ∑∑

2t r an n n= +

where

=bv vbK K

=bv vbC C

100

Page 108: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 5. ACCIDENT ASSESSMENT OF VEHICLES ON LONG-SPAN BRIDGES IN WINDY ENVIRONMENTS

5.1 Introduction

While little statistical information has been collected and published, threats of strong winds on the safety of vehicles have been realized and reported around the world (Scibo-rylski 1975; Coleman and Baker 1990; Baker 1987; Baker and Reynolds 1992). In the United States, the gust winds have also been found to be very important contributors to the accidents of vehicles especially trucks. For example, on March 24, 2001, two trucks collided and one driver was killed on the highway in Madison, Indiana and wind effect was identified to be one potential cause (IDOT, 2001). On January 7, 2003, a tractor-trailer ran off the Interstate 81 in Frederick County, Virginia due to the strong winds (Winchester Star News, 2003). On May 30, 2003, a truck loaded with cattle overturned on I-29 in Iowa when the wind blew it off the road (Iowa channel news, 2003). On February 1, 2002, on the interstate 95/495 near Largo, Maryland, the National Transportation Safety Board believed that the accident was contributed by multiple factors, one of which was also gust wind (NTSB, 2003).

Different from scattered accidents, series concurrent accidents due to the strong winds may turn out to be a catastrophe. On November 10, 1998, south-central and southeast Wisconsin's counties experienced strong winds with sustained wind speed of 30-40 mph gusted to 60-70 mph. In only 17 hours, there were more than 40 semi-trucks that were reported to overturn or side slipping off the highways in that area according to the report (NWS, 1998). Some vehicles that were pushed sideways by gusts also caused multiple vehicle accidents. Transportation was unavoidably delayed and one Interstate highway was reported to be totally closed because of the accidents. Many small cars were pushed into the ditches, one of them was observed as “partially airborne” (NWS, 1998).

In hurricane haunted areas, strong winds can be expected in windy season. In addition to the loss of each individual accident, the more serious issue in hurricane-prone area is that accidents constantly happening on the highways will greatly delay or even obstruct the important transportation line before or upon the landfall of hurricanes. Transportations are usually very busy and more important for hurricane preparations at those moments compared with ordinary days. If accidents happen frequently when evacuations are in progress, the whole evacuation process will be significantly delayed and the safety of those people, who cannot be evacuated in time due to the transportation interruptions, will be inevitably put on stake.

While very little statistical data of accidents on bridges have been collected, vehicles on bridges are more vulnerable to the cross gusts than on roads (Baker and Reynolds, 1992). The reasons include that many vehicles experience a suddenly strengthened crosswind when they just entered the bridge that is usually more open than the road. This is especially true when compared with roads with trees, hills or bushes on both sides. Another critical situation for vehicles on bridges is when vehicles just pass by bridge towers. The vehicle may experience a very short period of no crosswind action and then crosswind action resumes since the bridge towers may block the crosswind on the vehicles temporarily. For long-span bridges, the strong dynamic response of the bridge due to the interaction with wind and vehicles will also contribute to the accidents. To reduce accidents, some safety measures like setting an appropriate driving speed limit and criteria to close the bridge/highway during windy period have been adopted (Baker, 1987). However, in the past, the decision of setting driving speed limit and closing the

101

Page 109: Dynamic performance of bridges and vehicles under strong wind

transportation on bridges and highways arrives is mostly based on intuition or subjective experience (Irwin, 1999). The driving speed limit could be too high to be safe or too low to be efficient. Determination of suitable safety approaches at different situations deserves a more rational research.

Baker et al. have made explorative studies on the performance of high-sided vehicles in crosswinds on roads (Baker 1986, 1991a, 1991b, 1999); In his representative early work, Baker (Baker, 1986) proposed the fundamental equations for wind action on vehicles without considering the dynamic response of the vehicles in most directions. Through the adoption of meteorological information the percentage of the total time for which allowable wind speed is exceeded can thus be found and quantification of accident risk has been made considering driver behavior performance (Baker 1991b, 1999). In addition to wind tunnel tests on several particular vehicle models to identify the wind force on vehicles (Coleman and Baker 1990), some useful statistical information about actual accidents in British was also collected and analyzed (Baker and Reynolds, 1992). All of the above analyses, however, were limited to vehicles on the road. Moreover, the vehicle was modeled with a rigid body with only two degrees of freedom and thus no dynamic vibration excited by road roughness and wind forces were considered. The road roughness effect, the vertical acceleration, pitching and rolling acceleration of the vehicles were assumed to be zero in the existent model (Baker, 1986, 1999). Such simplifications may be reasonable for vehicles on the road. However, the dynamic interaction effects are very important for the vehicles on bridges and should be incorporated into the accident analysis. A more general and more realistic accident analysis model which can be used for vehicles on bridges and on roads is very desirable. To the knowledge of the writers, however, no existing accident model is so far available for such purpose.

In Chapter Four, a three-dimensional coupled bridge-vehicle-wind system was developed and analyzed (Cai and Chen 2004). Each vehicle was modeled as a combination of several rigid bodies, axle mass blocks, springs, and dampers. Dynamic interaction analysis is then conducted on the vehicle-bridge system to predict the “global” bridge and vehicle dynamic responses without considering accident occurrences. The results of the global bridge-vehicle vibrations serve as the basis for the present accident analysis of the “local” vehicle vibrations.

The present study aims at building a framework of general vehicle accident analysis model, which can be used for vehicles on bridges and on roads. Excitations from the road roughness and the wind loading are incorporated. The proposed model is applicable for vehicles on the road with curves, cambers, and grades. After setting up some accident criteria and driving behavior model, the accident risk can be assessed with the derived accident-related responses. To present the methodology, a truck model and a prototype bridge are chosen as the example of applications.

5.2 Dynamic Interaction of Non-Articulated Vehicles on Bridges

This section summarizes the procedures of interaction analysis of bridge-wind-vehicle for the convenience of discussion. Details are referred to Chapter Four.

5.2.1 General Ride Model of the Non-articulated Vehicle

A 2-axle model is widely adopted for the non-articulated vehicles in automobile engineering since non-articulated vehicles with more than two axles usually can be transformed into equivalent 2-axle vehicle models in the dynamic analysis (Wong 1993; Gillespie, 1993). In

102

Page 110: Dynamic performance of bridges and vehicles under strong wind

the present study, a 2-axle four-wheel vehicle is modeled as a combination of several rigid bodies connected by several axle mass blocks with springs and damping devices. The suspension system and the elasticity of tires are modeled with springs. The energy dissipation capacities of the suspension as well as the tires are modeled as damping devices with viscous damping assumed. The mass of the suspension system and the tires are assumed to concentrate on idealized mass blocks on each side of the vehicle and no mass in the spring and damping devices exists (Fig. 4.6). The displacements of the rigid body of the qth vehicle are denoted as: vertical displacement q

vrZ , pitching displacement in x-z plane qvrθ and rolling displacement in y-z

plane qvrφ . In the subscripts, “vr” refers to the rigid body of the vehicle. Vertical displacements of

the mass blocks symmetric to the central line of the jth axle are qjvaLZ and qj

vaRZ . In the subscripts, the “L” and “R” represent the left and right mass blocks on the jth axle, respectively. The lower “a” represents the axle suspension. The superscript “qj” represents the jth axle of the qth vehicle. Same definitions apply hereafter. The longitudinal, lateral, and vertical directions of the bridge are set as x, y, and z axis, respectively.

5.2.2 Coupled Equations of Vehicle-Bridge Model in Modal Coordinates

Assuming all displacements remain small, virtual works generated by the inertial forces, damping forces, and elastic forces acting on each vehicle on the bridge at a given time can be obtained. Assuming there are totally nv vehicles running on the bridge, and the initial conditions are the equilibrium conditions of the bridge under the self-weight of the bridge only without vehicles on it. The coupled equations can be finally built from the principle of virtual work as

G

+ + = +

v vv v v vb v v vb v r w

s v s v b b bb b bv b b b bv b b b r w

F + FM 0 γ C C γ K K γ0 M γ C C + C γ K K + K γ F + F F

(5.1)

where subscripts “b” and “v” represent for the bridge and vehicle, respectively; vγ and bγ are the displacement vectors of the vehicles and the bridge, respectively; superscripts of “s” and “v” in the stiffness and damping terms for the bridge refer to the terms of bridge structure itself and those contributed by the vehicles, respectively; subscripts “bv” and “vb” refer to the vehicles-bridge coupled terms; Matrices M, C and K are the mass, damping and stiffness matrices, respectively. F is the external loading terms. Subscripts “r”, “w” and “G” for the F refer to the loadings due to the road roughness, wind forces, and the gravity of the vehicles, respectively; superscripts of “v” and “b” refer to the forces acting on the vehicles and on the bridge, respectively. Some vectors are shown as follows and details of other terms in Eq. (5.1) can be found in Chapter Four and are omitted here for the sake of brevity.

, , , ,= vTn1 q

v v v vγ γ γ γ… (5.2)

1, , , , Tb i nξ ξ ξ=γ (5.3)

1 1 2 2, , , , , ,Tq q q q q q q q

v vr vr vr vaL vaR vaL vaRZ Z Z Z Zθ φ=γ (5.4)

v v vTv n n n1 1 1 q q q

wz w w wz w w wz w wwF ,M ,M ,0,0,0,0, ,F ,M ,M ,0,0,0,0, , F , M ,M ,0,0,0,0,θ φ θ φ θ φ=F (5.5)

103

Page 111: Dynamic performance of bridges and vehicles under strong wind

Eq. (5.1) can be used to predict the dynamic response of vehicles as well as bridge under wind action. As a special case for vehicles on the bridge model, the dynamic response of vehicles on the road can also be predicted using Eq. (5.1) after removing the bridge-related terms in the equations.

5.2.3 Wind Loading Simulation

For bridges immersing in the flow, the wind forces on the bridge in the three directions (vertical, lateral, and torsion) are denoted as ( ),b

wL x t , ( ),bwD x t and ( ),b

wM x t , respectively:

( ) ( ) ( ), ,bw st ae b ,L x t L L x t L x t= + + (5.6)

( ) ( ) ( ), ,bw st ae b ,D x t D D x t D x t= + + (5.7)

( ) ( ) ( ), ,bw st ae b ,M x t M M x t M x t= + + (5.8)

where the subscripts “st”, “ae” and “b” refer to static wind force component, self-excited wind force component, and buffeting force component, respectively.

With the generated wind velocity time history, the buffeting force time history corresponding to each point along the bridge span can be obtained (Cao et al. 2000). With the mode shapes obtained from finite element analysis, the wind buffeting forces on the whole bridge can be obtained through integrating all the force time histories along the bridge span.

The quasi static wind forces on vehicles are usually adopted (Baker, 1986):

( )212wx a r DF U C Aρ χ= ; ( )21 ;2wy a r SF U Cρ χ= A ( )21

2wz a r LF U C Aρ χ= ; (5.9-11)

( )212w a r R vM U C Ahφ ρ χ= ; ( )21

2w a r P vM U C Ahθ ρ χ= ; ( )212w a r Y vM U C Ahψ ρ χ= ; (5.12-14)

where Fwx, Fwy, Fwz, Mwφ, Mwθ and MwΨ are the drag force, side force, lift force, rolling moment, pitching moment and yawing moment acting on the vehicle, respectively; CD, CS, CL, CR, CP and CY are the coefficients of drag force, side force, lift force, rolling moment, pitching moment and yawing moment for the vehicle, respectively; “A” is the frontal area of the vehicle; hv is the distance from the gravity center of the vehicle to the road surface; Ur is the relative wind speed to the vehicle, which is defined as:

( ) ( )( ) ( )( )22 , , cos ,rU x t V U u x t U u x t2

sinβ β = + + + +

0

(5.15)

where V is the driving speed of vehicle; U and u(x, t) are the mean wind speed and turbulent wind speed component on the vehicle, respectively; β is the attack angle of the wind to the vehicle, which is the angle between the wind direction and the vehicle moving direction (β is 90 degrees if the cross wind is perpendicular to the vehicle driving directions). χ is the angle between the wind direction relative to the moving direction of vehicle (also the positive part of x axis here)

0 0

0 0

,

, 0

if

if

χ χχ

π χ χ

≥= + <

(5.16)

104

Page 112: Dynamic performance of bridges and vehicles under strong wind

where

( )( )( )( )0

, sinatan

, cosU u x t

V U u x tβ

χβ

+= + +

(5.17)

5.2.4 Numerical Approaches

The position of any vehicle running on the bridge changes with time. Correspondingly, the coefficients of the coupled equations are also time dependent. Therefore, the matrices in Eq. (5.1) should be updated at each time step after a new position of each vehicle is identified. Rouge-Kutta method is chosen and a computer program based on Matlab is developed to solve the differential equations.

5.3 Accident Analysis Model for Vehicles on Bridges

In the previous section, the dynamic interaction model of vehicle-bridge-wind system is briefly introduced. Such model is used to consider the dynamic interaction effects between vehicles and the bridge based on the detailed simulation of vertical stiffness and damping effect from the suspension system as well as from the tires. This interaction analysis model, however, is built based on the assumption that each vehicle wheel has full point contact with the bridge surface all the time and there exists no lateral relative movement between the wheels and the bridge surface. Such model predicts the responses of the bridge in all directions and the responses of vehicles only in several directions such as vertical, rolling and pitching directions.

The dynamic responses of vehicles in the vertical, rolling, and pitching directions from global bridge-vehicle analysis will be carried into the local accident analysis. Relative lateral and yaw responses of vehicles, which are not available in the global analysis, however, will be calculated separately with the local accident model which emphasizes on simulating the lateral relative movement and friction effects. The effects from lateral vibrations of the bridge on vehicle dynamics are considered through treating the lateral acceleration of the bridge as the external base excitation source of the vehicles.

5.3.1 General Model of Vehicle for Accident Analysis

In the derivations of section 2, a general case with multiple vehicles is considered and each individual vehicle is generalized as the qth vehicle. In the hereafter accident model derivation, only one typical vehicle is considered. Fig. 5.2 (a) shows the force coordinates for the typical 2-axle vehicle and the four wheels are defined for accident analysis. For most vehicles, the driving wheels are usually the rear ones and the steering wheels are the front ones. So the traction forces T only exists in the two rear wheels and the steering angle δ only exists for the two front wheels.

The wheel rolling resistance forces Fi (moving along the x axis direction) are related to the vertical forces as:

f re fi iF n N= (i=1 to 2) (5.18a, b) r re

iF n N= ri

where nre is a coefficient of rolling friction (with negative value) and can be estimated as a constant or with some simple formulas at elementary level (Gillespie, 1993); superscripts “f” and “r” denote the front wheels and rear wheels; and Ni is the reaction forces of the ith wheel. Aligning moment arising from the tire lateral frictions is omitted here.

105

Page 113: Dynamic performance of bridges and vehicles under strong wind

When the side slipping angle is small, the tire side slipping forces Hi (along the y axis direction) can be related, approximately in a linear way, to the vertical reactions as (Gillespie, 1993):

f la f fi iH m Nα= ( i=1 to 2) (5.19a, b) r la r

iH m Nα= ri

where mla is a cornering stiffness coefficient and fα and rα are the side slipping angles for the front wheels and rear wheels, respectively (Fig. 5.2 (b)).

The side slipping angles for the front wheels and rear wheels can be expressed as (Gillespie, 1993; Chen and Ulsoy, 2001; Shin et al. 2002)

1fi i

LV

ψα γ= − +δ 2ri

LV

ψα γ= + (5.20a, b)

where ψ is the yaw angle of the vehicle at the center of the gravity around axis z; δi is the steering angle of the ith wheel (it is zero for rear wheels and the same for the two front wheels here); L1 and L2 are the horizontal distance between the center of gravity to the front wheels and rear wheels, respectively (Fig. 5.5); γ is the vehicle body side slipping angle at the gravity center and defined by

( )tan / /arc v V v Vγ = − ≈ − (5.21)

where v is the side slipping (lateral) velocity of the vehicle body relative to the road surface; and V(t) is the longitudinal driving speed of vehicle at time t. In Ref. (Baker, 1986, 1991a, b, 1999), γ is chosen to approximate side slipping angles for both front and rear wheels as shown in Eqs. (5.19-20).

For vehicles considering driver behavior as shown in Fig. 5.1, the following force and moment equilibrium equations should be satisfied.

(a) Force equilibrium in the x axis:

( ) ( )2 2

1 1cos sin sinf f r

i i i i i i wx v g vi i

F H F T F M g Mδ δ θ= =

− + + + − =∑ ∑ V (5.22)

where Fwx is the aerodynamic wind force on the vehicle in the x direction as defined in Eq. (5.9); vM is the total mass of the vehicle; the derivative of V enables the acceleration/decelerations to

be considered; gθ is the grade and sinvM g gθ is also called grade resistance; g is the gravity acceleration; Ti is the traction force (be zero for non-drive wheels) or braking force of the ith tire and is believed approximately in proportion to the vertical reaction forces as iN

th riT k N= i (5.23)

where kth is the coefficient varying with the equilibrium conditions and can be identified later.

106

Page 114: Dynamic performance of bridges and vehicles under strong wind

C. G.

N 21 N

N1 N2

2fH

F1f

F2rF1

r

H2f

1fF

H2rH1

r

Fx

Fy

Fz

T1 T2

Mx

My

Mz

b1b1

h1

L1

L2

Mvtotalg

z axis

x axis

y axis

ff

r r

(a)

V

−γ

δαf

δαf

H 1f

H f2

αr

1

H2r

Hr

Ff1

Ff2F

r2

Fr1

C. G.

L2 L1

ψ

-v

(b)

Figure 5.1 The coordinate system and vehicle horizontal layout

107

Page 115: Dynamic performance of bridges and vehicles under strong wind

(b) Force equilibrium in the y axis:

( ) ( )2

1cos sin sinf r f v

i i i i i wy v v bri

H H F F M g M v Y Vδ δ ϕ=

ψ + + + − = + − ⋅ ∑ (5.24)

where Fwy is the aerodynamic wind force on the vehicle in y direction as defined in Eq. (5.10); is the lateral acceleration of the supporting surface under the vehicle, e. g. the bridge,

which is obtained from the previous section; ( )v

brY t

vM Vψ is the centrifugal force and ϕ is the road camber.

(c) Force equilibrium in the z axis:

( ) (2

1cos 0f r

i i wz v vri

N N F M Z g ϕ=

+ + + − =∑ ) (5.25)

where Fwz is the aerodynamic wind force on the vehicle in the z direction as defined in Eq. (5.11); vrZ is the vertical displacement of the vehicle obtained from the previous section in Eq. (5.4).

(d) Moment equilibrium about the x axis.

( ) ( ) ( )2 2

11 1

1 1

1 cos sini f r f r fw i i i i i i i

i i

M N N b H H F hφ vr vrJδ δ+

= =

+ − + − + + =∑ ∑ φ (5.26)

where Mwφ is the aerodynamic wind moment on the vehicle about the x axis (Eq. (5.12)); b is the half width of the vehicle and is the height from the contacting points of wheels with road surface to the center of gravity (c. g.); is the rolling moment inertia about the x axis; and

1

1h

vrJ vrφ is the rolling displacement of the vehicle obtained from the previous section in Eq. (5.4).

(e) Moment equilibrium about the y axis.

( ) ( )2 2

1 2 11 1

cos sinf r f f rw i i i i i i i i

i i

M N L N L F H F T h Iθ vr vrδ δ= =

+ − + + + + =∑ ∑ θ (5.27)

where Mwθ is the aerodynamic pitching moment on the vehicle about y axis (Eq. (5.13)); vrI is the pitching moment inertia about y axis; vrθ is the pitching displacement of the vehicle about y axis obtained from the previous section in Eq. (5.4).

(f) Moment equilibrium about the z axis

( ) ( ) ( )2 2

11 1 2 1

1 1

cos sin 1 cos sinif f r f f rw i i i i i i i i i i i

i i

M H L F L H L F H F T bψ vrδ δ δ δ+

= =

− + + − + − − + + = Θ∑ ∑ ψ (5.28)

where Mwψ is the aerodynamic yawing moment of the vehicle about the z axis (Eq. (5.14); vrΘ is the polar moment inertia about the z axis; ψ is the yaw displacement of the vehicle about the z axis that is to be calculated. Please be noted that aligning moment of the tire is omitted here.

(g) Compatibility conditions 108

Page 116: Dynamic performance of bridges and vehicles under strong wind

Vertical tire displacements for the non-articulated vehicle or any rigid body of the articulated vehicle should remain coplanar. If it is assumed that the tire reaction forces Ni are proportional to the tire displacements, then for conventional 2-axle vehicle, the following equation is satisfied (Baker, 196):

1 2 1f r rN N N N+ = + 2

f (5.29)

It should be noted that for articulated vehicle, compatibility equations exist for each rigid body of the whole vehicle.

Assuming the steering angle is small, approximations of cos 1δ ≈ and sinδ δ≈ can be made. Equations (5.22-23, 25-27 and 29) can be rewritten as following,

( ) ( ) ( )( )2 1 2 1 sin 0re la f f f re th r rwx v v gn m N N n k N N F M V M gα δ θ+ + + + + + − − = (5.30)

( )2 1 2 1 cos 0f f r rwz v vrN N N N F M Z g ϕ+ + + + + − = (5.31)

( ) ( ) ( )( ) ( )

2 1 1 1 1 1 1 1 1 1 2

1 1 1 0

f la f re la f re f la r r

la r rvr vr w

N b h m n h h m b n h N b h m N

h m b N J M φ

α δ α δ α

α φ

+ + + − + + +

− + − =

+ (5.32)

( ) ( ) ( ) ( )1 1 1 2 1 1 2 2 1 0re la f f f th re r rw vr vrL n h m h N N k n h L N N M Iθα δ θ + + + + + − + + − = (5.33)

2 1 1f r fN N N N+ = + 2

r (5.34)

Solving Eqs. (5.30-34) gives

( )( ) ( )0 1 1 2 2 1 1 0 3

3 1 1 0 1

re la fth

n L L m L hk

L hβ β α δ β

β β β

− + − − +=

− − +

β β (5.35)

( )

( )( )( )

3 0 1 11 0 1 2 3 2 1 1

1 2 1 1 11

1 0 2 1 31

1 2 1

2 2 2

14

lala

fre la

h h mh L h mL L b b b V

Nh L n m

hL L b

β ββ β β β β γ ψ

β β βδ

−− −+ − + −

+ = − + + + +

(5.36)

( )

( )( )( )

3 0 1 11 0 1 2 3 2 1 1

1 2 1 1 12

1 0 2 1 31

1 2 1

2 2 2

14

lala

fre la

h h mh L h mL L b b b V

Nh L n m

hL L b

β ββ β β β β γ ψ

β β βδ

−− −− + − +

+ = − + + + +

(5.37)

109

Page 117: Dynamic performance of bridges and vehicles under strong wind

( )

( )( )( )

3 0 1 11 0 1 1 3 2 1 1

1 2 1 1 11

1 0 2 1 31

1 2 1

2 2 2

14

lala

rre la

h h mh L h mL L b b b V

Nh L n m

hL L b

β ββ β β β β γ ψ

β β βδ

−+ −− + − +

+ = − + + + +

(5.38)

( )

( )( )( )

3 0 1 11 0 1 1 3 2 1 1

1 2 1 1 12

1 0 2 1 31

1 2 1

2 2 2

14

lala

rre la

h h mh L h mL L b b b V

Nh L n m

hL L b

β ββ β β β β γ ψ

β β βδ

−+ −− − + −

+ = − + + + +

+

(5.39)

0 sinwx v v gF M V M gβ θ= − − (5.40)

(1 coswz v vrF M Z g )β ϕ= + − (5.41)

2 vr vr wJ M φβ φ= − (5.42)

3 w vrM Iθ vrβ θ= − (5.43)

Eqs. (5.24, 28) give the following results

4 5 6 7dv vdt

β β β ψ β= + + + δ (5.44)

2

8 9 10 112

d vdtψ β β β ψ β= + + + δ

brϕ

(5.45)

where

4 / sin vy vF M g Yβ = − − (5.46)

15

la

v

mVM

ββ = (5.47)

( )3 1 06

la

v

m hV

M Vβ β

β −

=

+

)

(5.48)

( )(( )

1 0 3 2 17

1 2

re la

v

n m h LL L M

β β ββ

+ − −=

+ (5.49)

( )8 0.5 /re thwM n kψβ = − + + Θ 2β (5.50)

110

Page 118: Dynamic performance of bridges and vehicles under strong wind

( ) 1 0 3 1 19

0.5la re thm h n k h

V

β β ββ

− − +=

Θ (5.51)

( ) ( ) 1 1 2 1 2 1 1 0 3

10

0.5la re thm L L L L h n k h

V

β β ββ

− − + + − =Θ

(5.52)

( ) ( ) ( )

( )1 1 1 0 2

111 2

0.5la re re thm n h n k L h L

L L1 3β β β

β + + + − =

+ Θ

− (5.53)

By denoting Y as the lateral displacement, the coupled equations of motion, Eqs. (5.44 and 45), can be written in the state-form as:

ddt

= + ξA B ξ C (5.54)

where

1 0 0 00 1 0 00 0 1 00 0 0 1

=

A (5.55)

5 6

9 10

0 0 1 00 0 0 10 00 0

β ββ β

=

B (5.56)

Y

ψ

=

ξ (5.57)

4 7

8 11

00

β β δβ β δ

= + +

C (5.58)

With the assumed initial conditions about ξ and δ, Eq. (5.54) can be solved at time step t and the reaction forces of four wheels can also be quantified with Eqs. (5.36-39). The new steering angle in time t+∆t can be predicted with the steering angle model (discussed next) based on the obtained response in time t. Then this calculation procedure continues to time t+∆t.

111

Page 119: Dynamic performance of bridges and vehicles under strong wind

Repeat such procedure and the whole time history of ξ, reaction forces Ni and steering angle δ can be derived. The vehicle accident can be identified based on suitable accident criteria.

5.3.2 Driver Behavior Model

The driver behavior is considered as the way that a driver will steer his/her vehicle being blown laterally and rotationally across the road. While each driver’s behavior may be different, it is expected that a driver would set the steering angle in according to the lateral and yaw displacements, velocities and accelerations in order to keep the vehicle in position (Baker 1991b, 1999). In automobile engineering, steering angle is mostly studied about how to negotiate the cornering other than how to correct the driving to avoid accident in a straight route (Wong, 1993; Gillespie, 1992). It was also pointed out in (Chen and Ulsoy, 2001) that the driver behavior model is still a topic with a lot of uncertainties. Hence most existent models cannot be easily used in the current problem. Among very few people working on the driver behavior model serving current problems, Baker (1991b and 1999) once proposed the steering angle model, which is related to the lateral response and the lateral velocity. After some trial-and-error investigations of the driver behavior model, one similar to that proposed by Baker is suggested as follows.

The present driver behavior model is developed based on a simple idea that the steering angle should be adjusted to correct any lateral displacement of the front (steering) wheels. The adoption of the lateral responses of the front wheel other than that of the vehicle body (at C. G.) enables the yawing response can be taken care of as well. As mentioned earlier, each driver may react differently, and therefore, the present model is for demonstration only even though it gives very reasonable results. It is out of the scope of the current work in determining how driver behaviors in actual driving in a windy environment. The model is suggested as:

( ) (1 21 1 2

L L Y L Y LR

δ λ ψ λ+= − + − + )1ψ (5.59)

where R is the radius of turn; λ1 and λ2 are related to the driver behavior and assumed to be constants for the same driver.

5.3.3 Accident Criteria

For vehicles, it is usually believed that three types of typical accidents may happen: overturning accident, rotational (yawing) accident, and side slipping accident (Baker 1991b). All the accident criteria are decided based on some practical considerations. For lateral side slipping and rotational accident, the criteria can be set to avoid the vehicle sliding into other lanes, which causes potential accidents, such as impacting other vehicles, curbs, soft ground or tripping the vehicle into rollover (Gillespie, 1992). Theoretically, a overturning or rollover accident is to happen only when the C. G. of the vehicle is raised up to the rollover point and an experienced driver may be able to stabilize the vehicle when one wheel loses contact with the road surface (Gillespie, 1992). However, for ordinary drivers, setting more conservative criteria still makes sense.

Baker et al. once gave some guidelines for accident identifications (1986, 1991), which will be adopted here as follows: within some distance of the vehicle entering a sharp edged gust, the overturning accident is said to happen when one of the tire reaction forces Nj fell to zero, or side slipping accident is said to happen when the lateral response Y of the vehicle exceeds 0.5 m, or 112

Page 120: Dynamic performance of bridges and vehicles under strong wind

the rotation accident is said to happen if the yawing displacement vψ exceeds 0.2 radius. Such sharp-edged wind field exists when the vehicle just passes the bridge tower which used to block the wind action on the vehicle, or there is a strong gust acting on the vehicles and the bridge suddenly.

In the real world, the strong wind usually may not reign the driving for a very long time. Besides, the strong wind itself will naturally remind the driver to concentrate more on driving. So it is maybe acceptable for ordinary drivers to continually adjusting steering angle for a not too long period. On the other hand, the low wind speed is relatively common in ordinary days and it is not very realistic for drivers to adjust the steering angle in a very high frequency for a long time. Since there is no sufficient available information, it is suggested to set a steering angle adjusting frequency limitation of 1 Hz when wind speed is lower than 15 m/s and 2 Hz when wind speed is 15 m/s above. This additional criterion eliminates some results which are actually impractical in the real world.

5.4 Numerical Example

5.4.1 Prototype Bridge and Vehicles

The Yichang Suspension Bridge located in the south of China has a main span of 960 m and two side spans of 245 m each. The height of the bridge deck above the water is 50 m. The sketch of the bridge is shown in Fig. 4.5 and its main parameters are shown in Table 2.1 (Lin and Chen 1998). Based on a preliminary analysis with modal coupling assessment technique (Chen et al., 2004) and the observation of wind tunnel results, the four important modes considered in the present study are shown in Table 4.1 to consider the three-direction response of the bridge (Lin and Chen, 1998). In this example, the attack angle of the wind to the vehicle β=90o, i.e., the wind is acting perpendicular to the driving direction.

The vehicle model shown in Fig. 4.6 has seven degree-of-freedom, three for the rigid body of the vehicle (vertical displacement Zvr, pitching about the y axis θvr and rolling about the x axis φvr) and the other four for the vertical displacements of the four wheels (ZaL1, ZaR1, ZaL2, ZaR2).

Wind force coefficients used by Coleman and Baker (1990) are adopted here:

For 0 / 2χ π≤ <

( )0.3821SC a χ= ;C a ( )2 1 sin 3L χ= + ; ( )3 1 2sin 3DC a χ= − + ; (5.60a-c)

( )1.774YC a χ= − ; C a ; C a ; (5.61 a-c) ( )1.32

5P χ= ( )0.2946R χ=

For / 2π χ π≤ <

( )0.3821SC a π χ= − ;C a ( )( )( )2 1 sin 3L π χ= + − ; ( )( )( )3 1 2sin 3DC a π χ= + − − ; (5.62 a-c)

( )1.774YC a π χ= − − ; C a ; C a ; (5.63 a-c) ( )1.32

5P π χ= − − ( )0.2946R π χ= −

where a1-a6 equal to 5.2, 0.93, 0.5, 2.0, -2.0 and 7.3 for such vehicle [2], respectively.

In addition to the variables of the vehicle shown in Table 4.2, some additional parameters

113

Page 121: Dynamic performance of bridges and vehicles under strong wind

of the vehicle are shown in Table 5.1.

Table 5.1 Additional parameters of the vehicle used in the numerical example

coefficient of side slipping friction mla 2.5

coefficient of rolling friction nre -0.0075

5.4.2 Vehicle Interaction Analysis with the Bridge

As discussed earlier, with the bridge-vehicle interaction model, the time history of vehicle responses can be predicted for accident analysis. Interaction analysis starts from the first vehicle enters the bridge until the first vehicle leaves the bridge. Figs. 5.2-5.3 show the accelerations of vehicles in the vertical, rolling and pitching directions of the first vehicle with the same driving speed of 22 m/s versus the time under wind speed U=30m/s and U= 5m/s, respectively. The small sketch of the bridge indicates the corresponding location of the first vehicle on the bridge at any time. As will be seen later, vehicle accelerations in these directions will be used in the accident assessment. In Fig. 5.2, relatively larger accelerations can be observed when the vehicle moves to the middle range of the main span with 30 m/s wind speed. Such tendency, however, is not noticeable when the wind speed is as low as 5 m/s (Fig. 5.3). It is maybe because the acceleration contribution from the bridge to the vehicle is relatively small when wind speed is low. Comparing Figs. 5.2 and 5.3 suggests that the vertical and pitching accelerations decrease less significantly than the rolling acceleration does when the wind speed changes from 30 m/s to 5 m/s. It is possibly because that the accelerations of vehicles are contributed more significantly by high frequency vibration than by low frequency vibration. So the road roughness, as the high frequency excitation, may contribute more significantly to the total accelerations than the wind as a wide-band excitation does. Since the varying road roughness is only assumed on the vehicle moving directions (roughness is assumed the same in the bridge width direction), road roughness contributed much less to the rolling accelerations than in other two directions.

Figs. 5.4-5.5 give the relative responses of the vehicle in the vertical, rolling and pitching directions. Larger relative vertical response can be observed when wind speed U is 30 m/s compared with that when U=5 m/s with the same driving speed. Slightly larger relative responses can be observed when vehicles move in the middle-span region of the bridge compared with vehicles moving elsewhere on the bridge when wind speed U is 30 m/s (Fig. 5.4). Relative responses, however, have no obvious increment in the middle range of the main span when wind speed U is 5 m/s (Fig. 5.5). Such phenomenon shows again that the interaction effect exists between vehicles and the bridge, especially when wind speed is high. To avoid losing interaction information between vehicles and the bridge in the accident assessment process, vehicle dynamic results from interaction analysis seem to be very important.

114

Page 122: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40 50 60-30

-20

-10

0

10

20

30

Rol

ling

accl

erat

ion φ''

(rad

/s2 )

0 10 20 30 40 50 60-30

-20

-10

0

10

20

30

V

ertic

al a

ccle

ratio

n (m

/s2 )

0 10 20 30 40 50 60

-10

0

10

Pitc

hing

acc

lera

tion θ''

(rad

/s2 )

Time (second)

Figure 5.2 Vehicle accelerations when wind speed U=30 m/s and V=22 m/s

115

Page 123: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40 50 60

-20

-10

0

10

20

30

V

ertic

al a

ccle

ratio

n (m

/s2 )

0 10 20 30 40 50 60

-0.4

-0.2

0.0

0.2

0.4

Rol

ling

accl

erat

ion φ''

(rad

/s2 )

0 10 20 30 40 50 60

-10

0

10

Pitc

hing

acc

lera

tion θ''

(rad

/s2 )

Time (second)

Figure 5.3 Vehicle accelerations when wind speed U=5 m/s and V=22 m/s

116

Page 124: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40 50 60-0.2

-0.1

0.0

0.1

0.2

0 10 20 30 40 50 60-0.04-0.03-0.02-0.010.000.010.020.03

0 10 20 30 40 50 60-0.01

0.00

0.01

0.02

0.03

0.04

R

elat

ive

verti

cal d

isp.

(m)

Time (s)

Rel

ativ

e pi

tchi

ng d

isp.

θ (r

ad.)

Rel

ativ

e ro

lling

dis

p. φ

(rad

.)

Figure 5.4 Vehicle relative displacements when wind speed U=30 m/s and V=22 m/s

117

Page 125: Dynamic performance of bridges and vehicles under strong wind

0 10 20 30 40 50 60

-0.05

0.00

0.05

0.10

0 10 20 30 40 50 60

0.0016

0.0018

0.0020

0.0022

0.0024

0 10 20 30 40 50 60-0.02

-0.01

0.00

0.01

0.02

R

elat

ive

verti

cal d

isp.

(m)

Rel

ativ

e ro

lling

dis

p. φ

(rad

.)

Rel

ativ

e pi

tchi

ng d

isp.

θ (r

ad.)

Time (s)

Figure 5.5 Vehicle relative displacements when wind speed U=5 m/s and V=22 m/s

5.4.3 Accident-Related Response of the Vehicle

While the vehicle-bridge interaction analysis gives vehicle responses in several directions including vertical along axis z, rolling around axis x and pitching around axis y, responses in other directions, such as lateral (along axis y) and yawing (around axis z), are to be identified 118

Page 126: Dynamic performance of bridges and vehicles under strong wind

separately. These response are called “accident-related response” since they are critical to the assessment of accidents.

The accident assessment is only given to the first vehicle starting from the moment when the vehicle just enters the main span. It is to simulate the situation that the vehicle just passes by the bridge tower, which may temporarily block the wind loading on the vehicles. Once the truck enters the main span (namely passes one bridge tower), the truck will experience a sudden wind action. In Eq. (5.54), only four unknowns are to be solved: the truck body lateral displacement Y, the truck body yawing angle Ψ and their respective velocities. The initial conditions for the differential equations are assumed to be zero. In any time step, the new steering angle will be decided and vector C in Eq. (5.54) should be updated.

The second-order Rouge-Kutta approach is chosen to solve the differential equations with the time step of 0.01 second. Since the bridge is symmetric, 500 meter driving distance, a little bit longer than one-half of the main span, was considered. The proposed model is first validated with the comparison with the results by Baker’s (1986) without considering driver behaviors of vehicles on the road. To enable a direct comparison, bridge and steering angle related terms in the formulations are omitted and the side slipping angle model in Eqs. (5.19-20) are changed correspondingly. It is found that the same results can be obtained as Ref. (Baker, 1986) with the same vehicle data.

Fig. 5.6 shows the lateral, yawing response and corresponding steering angle under 15 m/s (33 mph) wind speed and vehicle driving speed when no driver behavior is considered, namely with zero driver steering angle. It is found that both lateral and yawing responses keep increasing with the driving distance and quickly exceed the accident criteria. It suggests that a vehicle will quickly be off-lane and “sway the tail” on bridges and roads if no any driver interference is applied on the driving vehicle under moderate wind. While some difference for the lateral side slipping can be observed when the vehicle runs on the bridge or on the road, little difference can be observed for the vehicle yawing response for this moderate wind speed.

Fig. 5.7 shows the accident-related responses when the driver’s steering angle is applied on the driving vehicle. It is found that the lateral side slipping response has been well suppressed around the zero side slipping location (as shown on the top subplot of Fig. 5.7). The yawing response as shown in the middle subplot of Fig. 5.7 is also limited to some value much lower than the accident criteria. The corresponding steering angle is shown in the bottom subplot of Fig. 5.7. For comparison purpose, the results of vehicles on the road with the same wind speed and driving speed are also plotted in the figure. Similar to the case without driver behavior (Fig. 5.6), it is found that lateral side slipping response of vehicles on the bridge is higher than that of vehicles on the road and yawing response only show slight difference between vehicles moving on the bridge and on the road. Correspondingly, more significant change of steering angle is required when vehicles run on the bridge than on the road. The steering angle should be constantly changed during the process of driving (adjusting frequency is about 0.5 Hz from the figure). The coefficients λ1 = 0.1 and λ2 = 0.2 used in this example represent just only one type of driving behavior described in Eq. (5.59).

119

Page 127: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 500

0.0

0.5

1.0

1.5

2.0

2.5

0 100 200 300 400 500-2.5

-2.0

-1.5

-1.0

-0.5

0.0

0 100 200 300 400 500-1.0

-0.5

0.0

0.5

1.0

On the bridge On the road

La

tera

l dis

plac

emen

t Y (m

)

On the bridge On the road

Yaw

ing

angl

e ψ

(rad

.)

Stee

ring

angl

e δ

(rad

.)

Distance (m)

Figure 5.6 Lateral displacement and yawing angle of the truck without steering angle input (U=V=15 m/s)

120

Page 128: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 500-0.1

0.0

0.1

0.2

0 100 200 300 400 500-0.04

-0.02

0.00

0.02

0.04

0 100 200 300 400 500

-0.10

-0.05

0.00

On the bridge On the road

La

tera

l dis

plac

emen

t Y (m

)

On the bridge On the road

Stee

ring

angl

e δ

(rad

.)

Distance (m)

On the bridge On the road

Yaw

ing

angl

e ψ

(rad

.)

Figure 5.7 Response of the truck with driver steering when U=V=15 m/s (λ1=0.2, λ2=0.1)

121

Page 129: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 5000.7

0.8

0.9

1.0

1.10 100 200 300 400 500

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

Windward front wheel

Rea

ctio

n fo

rce

ratio

Distance (m)

On the bridge On the road

Windward rear wheel

Rea

ctio

n fo

rce

ratio

On the bridge On the road

Figure 5.8 Windward wheel reactions of the truck when U=V=15 m/s (λ1=0.2, λ2=0.1)

122

Page 130: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 5000.9

1.0

1.1

1.2

0 100 200 300 400 500

0.9

1.0

1.1

1.2

1.3

1.4

1.5

Leeward front wheel

Rea

ctio

n fo

rce

ratio

Distance (m)

On the bridge On the road

Leeward rear wheel

R

eact

ion

forc

e ra

tio

On the bridge On the road

Figure 5.9 Leeward wheel reactions of the truck when U=V=15 m/s (λ1=0.2, λ2=0.1)

The reaction forces of each wheel are critical to identify the risk of overturning accidents. Fig. 5.8 and Fig. 5.9 show reaction force ratios for the windward and leeward wheels, which are normalized by the static reaction force of the corresponding wheel when the vehicle remains still on the ground. Comparison of results in Figs. 12 and 13 suggests that the reaction force ratios for the windward wheels are smaller than those of the leeward wheels. In addition, the mean value of the reaction force ratio for the windward rear wheel seems to be the smallest among the four wheels (about 0.7), but is still much higher than 0. Similar to the cases of displacements (Figs. 5.6 and 5.7), the reaction force ratios of vehicles on the bridge also have larger variations than on the road. As shown in Figs. 5.7-9, it can be concluded that it is safe for vehicles to run on both

123

Page 131: Dynamic performance of bridges and vehicles under strong wind

the bridge and the road with the given steering angle shown at the bottom of Fig. 5.7 and with the wind speed and driving speed all equal to 15 m/s (33 mph). The absolute value and changing speed of the steering angle are quite low (adjusting frequency is about 0.5 Hz), which means an ordinary driver can achieve the similar maneuver as shown in Figs. 5.7-9.

Vehicles may have different performance under higher wind speed. Fig. 5.10 shows the accident-related responses and the steering angle versus the travel distance when wind speed is 35 m/s (78 mph) and driving speed is 15 m/s (34 mph). As shown on the top of Fig. 5.10, the lateral side slipping displacement is much higher comparing with that under the wind speed of 15 m/s and maintain around some value other than zero. The middle subplot of Fig. 5.10 shows the yawing response of vehicles and it has exceeded the accident criterion. The bottom figure of the steering angle shows very dense curves with large value, which suggests faster and larger change of steering angle is required when wind speed is higher.

The reaction force ratios are shown in Fig. 5.11 for the four wheels. It is found that the bolded curve (windward rear wheel) has turned to negative in some cases, which suggests the possibilities of overturning accidents. It can also be found that the windward wheels all lose some reaction forces compared with still vehicle situation and the windward rear wheel is most likely to lose contact with the road surface. The higher wind speed requires more frequent change of steering angle (adjusting frequency is about 2- 4 Hz) to control the vehicle, which also justify that vehicles are much more difficult to control in strong wind and an ordinary driver may be very prone to have an accident due to more possible mistakes in steering the wheels.

Vehicle performances under low wind speed and high driving speed are also investigated. Fig. 5.12 shows the lateral and yawing displacements and the corresponding steering angle for vehicles with 50 m/s (112 mph) driving speed when wind speed is 5 m/s (11 mph). Lateral and yawing responses are all lower than the accident criteria even the driving speed is quite high. But the steering angle should be changed with a frequency about 0.8 Hz (it takes 500m/(50m/s) = 10s to finish the roughly 8 steering cycles), higher than the case when wind speed and driving speed all equals to 15 m/s as shown in Fig. 5.11 and lower than the case with high wind speed (35m/s) as shown in Fig. 5.10. The phenomenon suggests that in low wind speed, the vehicle can theoretically be driven in a quite high speed. However, more attention on appropriately adjusting the steering angle is necessary to keep the high driving speed safe. Since long time of adjusting the steering angle in a high frequency will quickly accumulate the driver fatigue and then jeopardize the driving safety, high driving speed is practically not safe for drivers. The reaction force ratios as shown in Fig. 5.13 suggest that the two front wheels lose some reaction forces and the vehicle body is not in the same equilibrium condition as the static situation even though the wind speed is quite low.

5.4.4 Accident Driving Speed

In the transportation practice, lowering the driving speed limit or close the bridge or highway is one common option to ensure the vehicle safety under strong winds. A suitable speed limit is of utmost importance to the drivers and the traffic administrators. Accident-related responses of the truck in several typical situations are studied in the previous section for demonstration. For vehicles running on the bridge with wind, it is desirable to know the highest allowable driving speed under any particular wind speed to avoid risks of accidents. Such critical driving speed is called “accident driving speed” in the present study. The three types of typical accidents (overturning accident, rotational accident, and side slipping accident) may happen

124

Page 132: Dynamic performance of bridges and vehicles under strong wind

concurrently or sequentially. The first occurrence is the critical one and the corresponding driving speed is the accident driving speed.

0 100 200 300 400 5000.000.040.080.120.160.200.24

0 100 200 300 400 500

-0.06

-0.04

-0.02

0.00

0 100 200 300 400 500-0.25

-0.20

-0.15

-0.10

-0.05

0.00

Late

ral d

ispl

acem

ent Y

(m)

Yaw

ing

angl

e ψ

(rad

.)

Stee

ring

angl

e δ

(rad

.)

Distance (m)

Figure 5.10 Response of the truck with driver steering when U=35 and V=15 m/s (λ1=λ2=0.3)

125

Page 133: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 500-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Rea

ctio

n fo

rce

ratio

Distance (m)

Windward front Windward rear Leeward front Leeward rear

Figure 5.11 Wheel reaction forces of the truck when U=35 and V=15 m/s (λ1=λ2=0.3)

126

Page 134: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 500

-0.08

-0.06

-0.04

-0.02

0.00

0.02

0 100 200 300 400 500-0.04

-0.02

0.00

0 100 200 300 400 500-0.02

0.00

0.02

0.04

La

tera

l dis

plac

emen

t Y (m

)

Yaw

ing

angl

e ψ

(rad

)

Stee

ring

angl

e δ

(rad

.)

Distance (m)

Figure 5.12 Response of the truck with driver steering when U=5 and V=50 m/s (λ1=0.4 λ2=0.3)

127

Page 135: Dynamic performance of bridges and vehicles under strong wind

0 100 200 300 400 5000.6

0.8

1.0

1.2

1.4

R

eact

ion

forc

e ra

tio

Distance (m)

Windward front Windward rear Leeward front Leeward rear

Figure 5.13 Wheel reaction forces of the truck when U=5 and V=50 m/s (λ1=0.2, λ2=0.3)

In order to predict the accident driving speed for different wind speeds for the truck on the prototype bridge, the accident driving speeds are searched under different wind speeds. Under each wind speed, the driving speed is increased in a step of 1.0 m/s. In each step, the accident-related response and reaction forces are predicted to check if any of the three accidents may occur during the driving process on the bridge. Keep increasing the driving speed to next step if no accident happens during that period. The lowest driving speed under which at least one accident criterion is not satisfied even after different variables (λ1 and λ2 here) of the steering angle model are tried can be identified as the “accident driving speed”. Repeating such process under different wind speeds, a curve for accident driving speeds versus wind speeds can be drawn from the results as shown in Fig. 5.14.

As shown in Fig. 5.18, the accident driving speed generally decreases with the increase of wind speed. When the wind speed increases from 5 m/s up to about 20 m/s, the accident driving

128

Page 136: Dynamic performance of bridges and vehicles under strong wind

speed decreases gradually from 60 m/s to about 30 m/s. When the wind speed keeps increasing, the accident driving speed drops to zero. This phenomenon suggests that when wind is not so strong, lowering the driving speed can maintain the safety of the vehicles on the bridge. However, when wind speed reaches the upper limit, solely lowering driving speed cannot avoid the accident occurrence. In other words, a still truck will also be blown off under some high wind speed, which agrees with the common sense. It is also found in this example that overturning accidents are most likely to happen when wind speed is over 20 m/s, while side slipping accidents are most likely when wind speed is lower than 20 m/s.

For comparison purpose, accident analysis is also conducted for the same truck on the road, and the results are also plotted in Fig. 5.18. It shows that vehicles on the road have higher accident driving speed, and the upper limit wind speed under which the truck cannot keep safe is the same. At this maximum wind speed (about 35 m/s), the driving speeds are approaching zero no matter on the road or on the bridge, which means the truck can not safely move on the bridge or on the road.

5 10 15 20 25 30 35

0

15

30

45

60

Acc

iden

t driv

ing

spee

d V

acci (m

/s)

Wind speed U (m/s)

On the bridge On the road

Figure 5.14 Accident driving speed versus wind speed

5.5 Concluding Remarks

In the present study, an assessment model for vehicle accidents on bridges and on roads under wind action is introduced. The proposed model starts with a full interaction analysis between the bridge and the vehicle, which predicts, in addition to the bridge vibration, the vehicle response in the directions of vertical, rolling and rotation under the wind action and road roughness. Such vehicle and bridge vibration information is carried over to the following

129

Page 137: Dynamic performance of bridges and vehicles under strong wind

130

accident analysis of the vehicle only. With given accident criteria, the accident driving speed can then be predicted under any wind speed.

The following conclusions can be made after the numerical example with a truck model moving on the prototype bridge. It is noted that some of these conclusions are “common sense” qualitatively, but are quantitatively verified through numerical simulation, which provides a base to make scientific decisions for traffic management in windy environments.

1. The proposed accident analysis model can be used to predict the accident-related response. With suggested accident criteria and driving behavior model, the accident risks can be assessed.

2. Lowering driving speed is effective to lower the accident risk only if the wind speed is not extremely high. Setting suitable driving speed limit is important to decrease the likeliness of accident occurrence.

3. When wind speed reaches high to some extent, the vehicle should not be on the bridge no matter what driving speed it has. Rational critical wind speed limit should be set to decide when to close the bridge. In the present study, 32 m/s (71 mph) is the critical wind speed limit based on numerical simulation. Actual limits can be set by considering also other factors.

4. Vehicles on the bridge are more vulnerable to accidents than on the road. Usually lower driving speed limits for vehicles on the bridge than on the road should be rationally defined to avoid the accident when strong wind speed exists.

5. Overturning is most likely to happen on the bridge for high-sided vehicles, like trucks and tractor-trailer. Windward rear wheel is mostly likely to initiate the accident.

6. The present study is to build up the framework for the accident analysis and more insightful studies on the driver behavior model and accident criterion are necessary.

Page 138: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 6. STRONG WIND-INDUCED COUPLED VIBRATION AND CONTROL WITH TUNED MASS DAMPER FOR LONG-SPAN BRIDGES

6.1 Introduction

Long-span bridges undergoing wind excitation exhibit complex dynamic behaviors. Buffeting vibrations induced by wind turbulence happen throughout the full range of wind speed. As the wind speed increases, aerodynamic instabilities such as flutter may occur at high wind speed (Simiu and Scanlan 1996). Much research effort has been made in mitigating excessive buffeting vibrations and improving aerodynamic stabilities for long-span bridges during construction (Conti et al. 1996; Takeda et al. 1998) and at service (Pourzeynali and Datta 2002; Omenzetter et al. 2002; Miyata and Yamada 1999). Among all of the control procedures, dynamic energy absorbers such as tuned mass dampers (TMDs) have been studied in suppressing the excessive dynamic buffeting (Gu et al. 2001) or enhancing the flutter stability of bridges (Gu et al. 1998; Pourzeynali and Datta 2002). As traditional control devices, the dynamic energy absorbers dissipate external energy through providing supplemental damping to the modes of concern (Abe and Igusa 1995; Kareem and Kline 1995).

Jain et al. (1998) analyzed the effects of modal damping on bridge performance of aeroelasticity. It was found that supplemental damping provided through appropriate external dampers could certainly increase the flutter stability and reduce the buffeting response of long-span bridges. In a conventional TMD control design, the TMD frequency is designed or tuned to the modal frequency of the fundamental mode (Fujino and Abe 1993) in order to reduce the so-called resonant vibration and this method is thus called resonant-suppression approach here.

When the modal coupling among the modes is weak, the bridge can be regarded as a simple combination of many single Degree-Of-Freedom (DOF) systems and single mode analysis is usually applicable (Lin and Yang 1983). In such a case, an equivalent damping ratio ζe can be adopted to assess the control efficiency and performance for each individual mode. When white noise excitation was assumed, the root- mean-square (RMS) ratio between the controlled and uncontrolled vibrations was derived as (Fujino and Abe 1993):

( ) 3 20 s e s scon s

3 2uncon s e0 s s s

S / 2 MRatio

S / 2 M

π ζ + ζ ωσ ζ = ≈ =σ ζ π ζ ω + ζ

(6.1)

where S0 = spectral density of white noise excitation; Ms = generalized modal mass; and ωs, ζs and ζe = modal circular frequency, modal damping ratio, and equivalent damping ratio of TMD, respectively.

It is well-known that wind-induced aeroelastic effects result in additional aerodynamic damping and stiffness for long-span bridges (Tanaka et al. 1993). The additional aerodynamic damping and stiffness may vary with wind speed and be different for various bridges. Typically, the aerodynamic damping increases with the increase of wind speed for bending modes but decreases for torsion modes. Consequently, the modal damping ratio ζs for the

131

Page 139: Dynamic performance of bridges and vehicles under strong wind

mode of concern, which consists of mechanical and aerodynamic damping, could increase and be very high under strong wind. As indicated in Eq. (6.1), the control efficiency decreases with the increase of modal damping ratio ζs. It also implies that, for coupled mode vibrations of long-span bridges, the control efficiency of buffeting response of bending mode decreases with the increase of wind velocity since the aerodynamic damping of bending modes usually increases with the wind speed as mentioned above. Since bending modes usually contribute significantly to the overall buffeting response among all of the modes, the decreased control efficiency in bending modes may deteriorate the overall control efficiency of the bridge vibration.

The adoption of slender deck and the increase of bridge span lengths tend to make the frequencies of modes closer, which increases modal coupling effects through aeroelastic effects in high wind velocity (Bucher and Lin 1988; Jain et al. 1996; Katsuchi et al. 1998; Thorbek and Hansen 1998; Cai and Albrecht 2000). Modal coupling effects due to strong wind may result in a significant additional component to the buffeting response of each individual mode, compared with the cases of weak modal coupling. Accordingly, a more efficient control approach than the traditional resonant-suppression method may exist for the coupled buffeting control of bridges in strong wind.

The present study aims at introducing an alternative TMD design approach, which is based on suppression of modal coupling effect among modes under strong wind. Conventional resonant suppression TMD design idea has difficulty to achieve satisfactory control effects because strong modal coupling under strong wind causes high damping ratio of some concerned modes. Different from conventional resonant suppression through supplying additional damping, the proposed design approach is to attain control efficiency under strong wind through suppression modal coupling effects among several coupling-prone modes. With the proposed control approach, a well-designed TMD system can efficiently suppress wind-induced vibrations for the strongly-coupled modes even in high wind speed. Poorer control performance may otherwise be anticipated for TMDs designed based on the conventional resonant-suppression approach.

To better develop and explain the new control approach, approximated closed-form solutions of coupled buffeting response were first derived for a multi-mode coupled bridge system attached with arbitrary number of TMDs. This derivation clearly shows the contributions of all components of the response and indicates how TMDs can be designed to control each part. Examples of a two Degree-Of-Freedom (2DOF) model and a long-span prototype bridge were then used to further demonstrate and validate the efficiency of TMDs on the coupled response control. Finally, the applicability of dual-objective control with passive TMD system designed based on the new control approach was briefly discussed.

6.2 Closed-Form Solution of Bridge-TMD System

To better understand the coupled vibrations and the interaction of the bridge-TMD system, closed-form solutions are derived below. This derivation will give insights and facilitate the discussion in developing a new TMD control approach for coupled vibrations in strong wind.

132

Page 140: Dynamic performance of bridges and vehicles under strong wind

Linda and Donald (1998) once analysed the coupling problem of sound-induced vibrations and proposed an approximated closed-form formation for acoustics. By using a similar procedure, a closed-form formulation is derived below for a coupled wind-induced vibration of a bridge-TMD system. To keep the integrity of the derivation, some well-known formulas are necessarily revisited below.

For a bridge under wind action with a displacement of r(x, t), the buffeting and aeroelastic forces are expressed as functions of the displacement r(x, t) and location ordinate

x as fb(x, t) and f ( , respectively. Assume that a total number of ns x, r, r) 2 TMDs are attached

to the bridge at the location of xp (p = 1 to n2), then the equation of motion is derived as

[ ] [ ]2

p

n

b s p TMDp 1

r(x, t) r(x, t) (x)r(x, t) f (x, t) f (x, r, r) (x x )f (t)L D=

+ + ρ = + + δ −∑ (6.2)

where [ ]⋅L and [ ]⋅D = elastic and viscous damping operators; (x)ρ = mass density; ( )δ ⋅ =

Dirac delta function; and fTMDp = reaction force from the pth TMD on the bridge.

The deflection components of the bridge are represented in terms of the mode shapes and generalized coordinate as (Jain et al. 1996)

1n

B i ii 1

r(x, t) r (x) (t)=

= δ ξ∑ (6.3)

where n1 = total number of natural modes considered; )t(iξ = generalized coordinate; r = h, p

or α; h(x), p(x), and α(x) = vertical, lateral, and torsional mode shape, respectively and

B

1, if r

deck width b, if r h or p

= αδ = =

(6.4)

The multi-mode formulation of the equation of motion for a bridge-TMD system under external excitation can thus be expressed as

'' 'TMDξ + ξ + ξ = +bI A B Q F (6.5)

where = generalized coordinate vector; the superscript prime “`” represents a derivative

with respect to dimensionless time s U

ξ

t / b= ; = identity matrix; Q = excitation force

vector normalized to the generalized mass inertia; F

I b

TMD = reaction force vector of TMD on

133

Page 141: Dynamic performance of bridges and vehicles under strong wind

the bridge normalized to the generalized mass inertia; and A and B = total damping ratio matrix and total stiffness matrix , respectively. The general terms of matrices A and B are

ij i i ij i ijA (K) 2 K J KZ= ζ δ − (6.6)

2ij i ij i ijB (K) K J K T= δ − 2 (6.7)

where iζ = damping ratio for the ith mode; ijδ = Kronecker delta function that is equal to 1 if

i = j and equal to 0 if ; = reduced frequency; ji ≠ K b /= ω U i iK b / U= ω = ith reduced

frequency; b = bridge width; ω = iith circular natural frequency, U = mean velocity of the

oncoming wind; and

4

ii

bJ2Iρ

=l (6.8)

i j i j i j

i j i j i j i j i j i j

* * *ij 1 h h 2 h 5 h p

* * * * * *1 p p 2 p 5 p h 1 h 2 5 p

Z H G H G H G

P G P G P G A G A G A G

α

α α α α

= + +

+ + + + + + α

α

(6.9)

i j i j i j

i j i j i j i j i j i j

* * *ij 4 h h 3 h 6 h p

* * * * * *3 p 4 p p 6 p h 3 4 h 6 p

T H G H G H G

P G P G P G A G A G A G

α

α α α α

= + +

+ + + + + + (6.10)

where = generalized mass inertia for the iiI th mode; ρ = air density; l = bridge length;

= experimentally determined flutter derivatives for the bridge deck; and

the modal integrals ( G ) are computed as

* * *i i iH ,P , ( i 1 6)= −A

jisr

i jr s i j0

dxG r (x)s (x)= ∫l

l (6.11)

where ; and js h . iiii orp,hr α= j j j, p or= α

To solve the equation of motion, Eq. (6.5) is Fourier transformed in reduced frequency K domain as

TMDξ = +bE Q F (6.12)

134

Page 142: Dynamic performance of bridges and vehicles under strong wind

where TMD, ,ξ bQ F = Fourier transformation of b TMD,Q ,and Fξ , respectively; and

1 2

i

n np

TMD ij p jj 1 p 1

F R= =

= µ∑∑ (K)ξ (6.13)

where

pTMD i p j ppij

i

m r (x )r (x )

Iµ = (6.14)

1 2p p pR (K) R (K) i R (K)= + ⋅ (6.15)

( )( )

p p

p p

2 2 2 2TMD TMD TMD1

p 22 2 2 2 2TMD TMD p

K K K K 1 4R (K)

K K 4 K K

p − + ζ =− + ζ

(6.16)

( )( )

p p p

p p

3 2 2TMD TMD TMD2

p 22 2 2 2 2TMD TMD TMD

2 K K 2K KR (K)

K K 4 K

ζ −=

− + ζpK

(6.17)

p

TMDpTMD

BK

= (6.18)

where pTMDω ,

pTMDζ and pTMDm = circular frequency; damping ratio; and mass for the pth

TMD, respectively.

Similar simplification to that by Linda and Donald (1998) is followed below. The ith equation is extracted from Eq. (6.12) and rewritten as

1

i

n2 2i i j ij j

j 1

(K K ) (K) i K K D (K) (K) Q=

− ξ + ⋅ ⋅ ξ =∑ b (6.19)

where

2 2n np 2 p 1ij p ij p

ij p 1 i ij p 1ij

j j j j

R (K) R (K)A J K T

D (K) iK K K K K K

= =

µ µ ⋅ ⋅

= + + −

∑ ∑⋅

(6.20)

For the convenience of discussion, a non-dimensional variable is introduced as

135

Page 143: Dynamic performance of bridges and vehicles under strong wind

i

2 2i i i ii i

i uni b

(K) (K K i K K D ) (K)(K)(K) Q

ξ − + ⋅ ⋅ ξξ = =

ξ (6.21)

where

ibuni 2 2

i i

Q(K)

(K K i K K D )ξ =

− + ⋅ ⋅ ii

(6.22)

The superscript “un” for )K(uniξ stands for the uncoupled single-mode solution that is

obtained from Eq. (6.19) by ignoring all the coupling effect. Eq. (6.21) thus represents the ratio of coupled and uncoupled solutions.

Eq. (6.19) can be rewritten considering Eq. (6.21) as

1nij

i jjj i

jjj

i D(K) (K) 1K K i D

K K≠

⋅ εξ + ξ =

− + ⋅ε∑ (6.23)

where jQijij RDD~ =ε ; j

i

bQ j

b

Q (K)R (K)

Q (K)= ; ε is introduced to order the coupling coefficients;

for weak coupling 1<<ε . The magnitude of ε is chosen to make the normalized coupling

coefficients Dij~ of the order of 1, denoted as (Linda and Donald 1998). [1]O

Since the right hand side term in Eq. (6.23) is O , there exist following results: [1]

jij

j jjj

j

[ ],K Ki DK [1],K KK i DK K

ε ≠⋅ε = ≈− + ⋅ε

OO

(6.24)

According to the definition in Eq. (6.21), ξ and are equal to 1 if the modal

coupling effect is entirely omitted. Otherwise, for weak coupling system, there exists the conditionε . When the i

i (K) j(K)ξ

1<< th mode is under study (defined as current mode thereafter), its reduced frequency Ki is different from Kj. For 1<<ε , following equation can be derived (Linda and Donald 1998).

j (K) 1 ( ) 1ξ = − ε ≈O (6.25)

Consequently, Eq. (6.23) can be further simplified considering Eqs. (6.24) and (6. 25) as 136

Page 144: Dynamic performance of bridges and vehicles under strong wind

1nij

ijj i

jjj

i D(K) 1 1K K i D

K K≠

⋅ εξ + ⋅ =

− + ⋅ε∑ (6.26)

In the above derivation process, coupling effects with the order of O(ε2) are ignored. Physically, this is to say that when the response of the ith mode is determined, the coupling effects between the ith mode and other modes (jth mode) are included in the solution. However, the coupling effects between jth mode and the other modes (except for ith mode) are deemed negligible as indicated in (6.25). Since only the high order small terms are ignored, the accuracy of the solution is not significantly scarified (Linda and Donald 1998).

Converting (6.26) back to the original form (versus dimensionless form) gives

1

i j

nj ij

i b2 2 2 2j ii i ii j j jj

i K K D1(K) Q (K) Q (K)K K i K K D (K K i K K D )≠

b

⋅ ⋅ξ = × −

− − ⋅ ⋅ − − ⋅ ⋅ ∑ (6.27)

If the cross-modal buffeting spectrum is omitted (Simiu and Scanlan 1996), the power

spectral density (PSD) for the generalized displacements of ith mode, , is derived from

(6.27) as

1

i i i i j j

nun un

ijj i

S (K) S (K) (K) S (K)ξ ξ ξ ξ ξ ξ≠

≈ + γ ⋅∑ (6.28)

where

b bi i

i i2 2

Q Qun2n n

2 p 1 p 2ii ii p ii ii p

p 1 p 1

SS

B K R (K) K A R (K)ξ ξ

= =

= − + µ ⋅ + ⋅ + µ ⋅

∑ ∑2

(6.29)

b bj j

j j2 2

Q Qun2n n

2 p 1 p 2jj jj p jj jj p

p 1 p 1

SS

B K R (K) K A R (K)ξ ξ

= =

= − + µ ⋅ + ⋅ + µ ⋅

∑ ∑2

(6.30)

( )

2 2

2 2

2 2n n2 p 1 2 p 2

i ij ij p i ij ij pp 1 p 1

ij 2 2n n2 2 2 p 1 2 p 2i i ii ii p i i i ii ii p

p 1 p 1

J K T R (K) J K Z R (K)

K

K K J K T R (K) 2 K K J K Z R (K)

= =

= =

− + µ ⋅ + − + µ ⋅ γ =

− − + µ ⋅ + ζ ⋅ − + µ ⋅

∑ ∑

∑ ∑

(6.31)

137

Page 145: Dynamic performance of bridges and vehicles under strong wind

Eq.(6.28) indicates that the coupled response of each mode mainly consists of two parts. The first part (the first term) is the uncoupled response of the current ith mode, namely the resonant component of the ith mode buffeting. The second part (the second term) is due to the modal coupling between the ith mode and other modes and is called the coupled component of the ith mode buffeting here. In Eqs. (6.29), (6.30), and (6.31), the term associated with term R represents the contribution of the TMDs to the bridge vibrations. It can be seen from these equations that including TMDs may affect not only the first term of Eq. (6.28) the resonant component, but also the coupling component of the second term. The traditional control approach of resonant-suppression that targets at the first term is hardly able to control the coupling component directly. A new control approach may be naturally inspired to optimize the control efficiency by reducing the total response, not just the resonant vibration. This new control approach will be discussed below with numerical examples.

6.3 Coupled Vibration Control with a Typical 2DOF Model

As discussed above, conventional control strategy is to suppress resonant vibration that is essentially represented by the first term of Eq. (6.28). If the modal coupling among the current ith mode and the other modes is very weak, the second term of Eq. (6.28) will be trivial. In that case, conventional single-mode-based control analysis without considering the effect from the second term of Eq. (6.28) could lead to acceptable results. However, for the modes with strong modal coupling, the contribution of the second term to the total response can be significant. It becomes necessary to consider both the resonant vibration and that from coupling effects to achieve the optimal performance.

To examine this concept and verify the closed-form derivation conducted above, a simple 2DOF system attached with two identical TMDs was considered as shown in Fig. 6.1, where the parameters associated with masses M1, M2 and Mp represent the 1st DOF, the 2nd DOF, and the TMD DOF, respectively. The parameters for this 2DOF model are defined in Table 6.1. Two coupled modes are the most typical and easiest example whose closed-form results can be more conveniently derived. Using two identical TMDs makes it easy to distinguish clearly the control effect on any part of the vibrations. For simplicity but without losing generality, it is assumed that the external excitation is white noise with a power spectral density of S0.

According to Eqs. from (6.28) to (6.31), the solution of the 2DOF model (n1 = 2) may reduce to

i i i i j j

un unijS ( ) S ( ) ( ) S ( )ξ ξ ξ ξ ξ ξω ≈ ω + γ ω ⋅ ω (6.32)

where

( )( ) ( )1 1

un 022 2 1 2

1 11 p 1 1 11 p

SSR ( ) 2 R ( )

ξ ξ ω = ω −ω +µ ⋅ ω + ω⋅ω ⋅ζ +µ ⋅ ω

2 (6.33)

138

Page 146: Dynamic performance of bridges and vehicles under strong wind

( )( ) ( )2 2

un 02 22 2 1 2

2 22 p 2 2 22 p

SSR ( ) 2 R ( )

ξ ξ ω = ω −ω +µ ⋅ ω + ω⋅ω ⋅ζ +µ ⋅ ω

(6.34)

( )( ) ( )

( ) ( )

2 22 1 212 12 p 12 p

12 22 2 1 21 11 p 1 1 11 p

R ( ) R ( )

R ( ) 2 R ( )

ω +µ ⋅ ω + µ ⋅ ω γ ω =2 ω −ω +µ ⋅ ω + ζ ω⋅ω +µ ⋅ ω

(6.35)

It is shown in Eq. (6.32) that the closed-form solution for this simple case (with 2DOF model and two identical TMDs) can be conveniently derived. By assuming that the structural damping ratios of both the 1st DOF (M1) and 2nd DOF (M2) are as low as 0.5% (a typical value for aerodynamic analysis of long-span bridges), the response power spectra were calculated with above formulas and shown in Figs. 6.2 and 6.3. In these figures, the top half is the spectra for the 1st DOF and the bottom half is for the 2nd DOF.

Table 6.1. Parameters for the 2-DOF System Attached with Two Identical TMDs

2DOF system TMD system

M1 , M2 (kg) 1 Mp (kg) 0.01

K1 (N/m) ( )22 1π × Kp (N/m) ( )22 0.1π ×

( )22 0.15π ×

K2 (N/m) ( )22 1.5π × ωp = 2π fp (rad.) 2 1π ×

2 1.5π ×

K12, K21 (N/m) ( )22 0.1π × Generalized inertia ratio

µ11, µ22

0.01

ω1=2π f1 (rad.) 2 1π × ζp 0.04

ω2=2π f2 (rad.) 2 1.5π ×

ζ1, ζ2 0.005

S0 1

139

Page 147: Dynamic performance of bridges and vehicles under strong wind

M1

M2

K1

K2

K12/K21

Mp

Mp

ζ1

ζ2Kp

Kp

ζp

ζp

Fig. 6.1 2-DOF Mechanical Model

Two identical TMDs are still considered here. In Fig. 6.2, the two identical TMDs are conventionally designed to suppress the resonant vibration of the 1st DOF. For comparison, both coupled and uncoupled analyses were conducted. It can be seen that when uncoupled vibration analysis is conducted, the vibration power spectrum for each DOF has only one peak from resonant vibration. However, there exist two peaks when coupled analysis is conducted. One peak is induced by resonant vibration corresponding to its modal frequency, while the other is due to the modal coupling effect between the 1st DOF and the 2nd DOF. The modal coupling effects are significant to the dynamic response.

It is shown in Fig. 6.2 that the TMDs designed for the 1st DOF have good control efficiency for the resonant vibration of the 1st DOF (the first peak of Fig. 6.2(a)), and also has some effect on the first peak of Fig. 6.2(b) that is the contribution of the 1st DOF to the 2nd DOF due to modal coupling. However, this design of TMDs doesn’t help reduce the vibrations due to the modal coupling from the 2nd DOF (the second peak of Fig. 6.2(a)) and the resonant vibration of the 2nd DOF (the second peak of Fig. 6.2(b)).

Fig. 6.3 shows the vibration power spectra when the TMDs are designed for the 2nd DOF. Similarly, the TMD helps reduce only the second peak values that are caused by the 2nd DOF, but not the peak values that are caused by the 1st DOF (the first peak of both Fig. 6.3(a) and Fig. 6.3(b)). Figs. 6.2 and 6.3 suggest that the TMDs should be optimally designed to suppress either the resonant vibration (first part in Eq. (6.28)), or the vibration due to modal coupling (second part in Eq. (6.28)) through weakening the modal coupling. When the overall response of the structure other than any single mode is considered, multiple TMDs can be designed to achieve the best control performance under any particular condition.

140

Page 148: Dynamic performance of bridges and vehicles under strong wind

0.75 1.05 1.35 1.65 1.950.1

1

10

100

0.75 1.05 1.35 1.65 1.950.1

1

10

f (Hz)

Uncoupled response w/o control

Coupled response w/o control

Coupled response with control

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(a) First DOF M1

100

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(b) Second DOF M2

Fig. 6.2 Response Spectra of 2-DOF Model with ζ1=0.005 (TMDs optimally designed for M1)

141

Page 149: Dynamic performance of bridges and vehicles under strong wind

0.75 1.05 1.35 1.65 1.950.1

1

10

100

0.75 1.05 1.35 1.65 1.950.1

1

10

100

f (Hz)

Uncoupled response w/o control

Coupled response w/o control

Coupled response with control

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(a) First DOF M1

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(b) Second DOF M2

Fig. 6.3 Response Spectra of 2-DOF Model with ζ1= 0.005 (TMDs optimally designed for M2)

As stated before, wind-induced vibration results in aeroelastic damping so that the total vibrational damping of some modes may be large in strong wind. To simulate such a case that is common for modern long-span bridges, it is arbitrarily assumed that the damping ratio of

142

Page 150: Dynamic performance of bridges and vehicles under strong wind

the 1st DOF is 3%, while that of the 2nd DOF remains to be 0.5%. The corresponding vibration spectral density results are shown in Figs. 6.4 and 6.5.

It can be found that when the 1st DOF vibrates with high damping ratio, the TMDs designed for the 1st DOF (Fig. 6.4) have less control efficiency for its resonant component (the first peak of Fig. 6.4(a)) than that of its counterpart when the TMDs are designed for the 2nd DOF (the second peak of Fig. 6.5(b)). The component of the 2nd DOF due to coupling even increases slightly as observed from the first peak of Fig. 6.4(b). In comparison, it can be seen from Fig. 6.5 that when TMDs are designed for the 2nd DOF with low damping ratio, the control efficiencies of its resonant component (the second peak of Fig. 6.5(b)) and the component of the 1st DOF due to coupling (the second peak of Fig. 6.5(a)) are still high, even though the 1st DOF has very high damping ratio.

For coupled vibrations, these observations have confirmed that the total modal vibration consists of mainly one portion from resonant vibration and another portion caused by coupling effects with other modes. The frequency of conventionally designed TMDs is tuned to that of the targeted mode to control the resonant vibrations and they may not achieve an efficient control especially when the coupling effect is significant. An optimal control strategy should aim at not only the resonant vibration, but also the vibration from modal coupling. Especially for some strongly-coupled modes vibrating in high wind velocity with high damping ratios, there exists a possibility that the vibration can be optimally suppressed even the TMD is not designed around the natural modal frequency of the targeted mode. For example, to control the vibration of the 1st DOF in strongly coupled vibration, the TMD frequency needs to be tuned to the natural frequency of the 2nd DOF rather than that of the 1st DOF. In other words, weakening the coupling effects may sometimes be more efficient than reducing the resonant vibrations when strong modal coupling exists (for maximum efficiency, both resonant and coupling components should be suppressed, but certainly that will be also more costly). To further understand this new control approach of the coupled buffeting response with TMDs, a prototype long-span bridge is studied in the next section.

6.4 Analysis of a Prototype Bridge

The Yichang Suspension Bridge located in the south of China has a main span length of 960 m and two side spans of 245 m each. The height of the bridge deck above water is 50 m. Its main parameters are shown in Table 2.1. The four modes considered in the present study are shown in Table 4.1. Wind tunnel studies have shown that the 1st symmetric bending mode (Mode 2) and the 1st symmetric torsional mode (Mode 3) are the two key modes for buffeting and flutter analyses (Lin et al. 1998). Meanwhile, strong modal coupling between these two modes was observed at high wind velocity due to aeroelastic effects.

Complex eigenvalue approach was used to analyze the modal properties considering modal coupling. Fig. 6.6 shows that the modal damping ratios of the two vertical bending modes increase with the wind velocity. This increase is more significant when the wind speed surpasses 40 m/s. In contrast, at high wind velocity, the modal damping ratio of the symmetric torsional mode decreases with the increase of wind speed and eventually reaches zero. The critical wind flutter velocity is identified as 73 m/s by using the condition of zero total damping. 143

Page 151: Dynamic performance of bridges and vehicles under strong wind

0.75 1.05 1.35 1.65 1.950.1

1

10

100

0.75 1.05 1.35 1.65 1.950.1

1

10

100

)

Uncoupled response w/o control

Coupled response w/o control

Coupled response with control

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(a) First DOF M1

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

f (Hz(b) Second DOF M2

Fig. 6.4 Response Spectra of 2-DOF Model with ζ1= 0.03 (TMDs optimally designed for M1)

144

Page 152: Dynamic performance of bridges and vehicles under strong wind

0.75 1.05 1.35 1.65 1.950.1

1

10

100

0.75 1.05 1.35 1.65 1.950.1

1

10

100

f (Hz)

Uncoupled response w/o control

Coupled response w/o control

Coupled response with control

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(a) First DOF M1

Res

pons

e po

wer

spec

trum

(mm

2 /Hz)

(b)Second DOF M2

Fig. 6.5 Response Spectra of 2-DOF Model when ζ1= 0.03 (TMDs optimally designed for M2)

145

Page 153: Dynamic performance of bridges and vehicles under strong wind

960 m 245 m245 m

Truss element Beam element

(a) Plane view

29990 mm

24480 mm

5998 mm 8997 mm 8997 mm 5998 mm

"Three-beam" model

(b) Cross section view

Fig. 6.6 Yichang Suspension Bridge

As discussed before about Eq. (6.1), the higher vibration damping of the mode may cause the lower control efficiency of a given TMD. If the two vertical bending modes of the Yichang Bridge are deemed as two single modes omitting modal coupling with any other modes, Eq. (6.1) can be applied and implies that the control efficiencies of the two vertical bending modes should decrease at high wind velocity due to their high existent total damping ratios. In other words, it will be more difficult to suppress the vibration of vertical bending modes by using conventionally designed TMDs that essentially add supplemental damping to the concerned modes. Relatively, the control efficiency of torsional modes will be higher due to their low vibration damping ratios.

146

Page 154: Dynamic performance of bridges and vehicles under strong wind

0 20 40 600

2

4

6

8

U (m/s)

U=73 m/s Sym. Torsional

Asym. Torsional

Asym. Bending

Sym. Bending

Fig. 6.7 Modal Damping Ratio Versus Wind Velocity, Yichang Suspension Bridge

147

Page 155: Dynamic performance of bridges and vehicles under strong wind

0 20 40 60

0.0

0.1

0.2

0.3

0.4

0.5

U (m /s)

Asym. Bending

Sym. Bending

Sym. Torsional

Asym. Torsional

fcr =0.272 Hz

Fig. 6.8 Modal Frequency Versus Wind Velocity, Yichang Suspension Bridge

6.4.1 Buffeting Analysis with Conventional TMD Control

For information, Fig. 6.7 shows the change of vibration frequencies with the wind speed. Similarly to the pattern of damping change, the modal frequencies of vertical bending modes increase while the torsional frequencies decrease with the increase of wind velocity, due to the effects of aeroelastic forces.

Fig. 6.8 shows the RMS of displacement response at the mid-span of the main span versus wind speed for the 1st symmetric bending mode and the 1st symmetric torsional mode using single mode analysis and multiple coupled mode analysis, without considering the TMDs. In this figure, the torsional response represents the vertical displacement at the edge of the cross section due to the torsional vibration. Differences between the results of single-mode analysis and coupled analysis are obvious in high wind speed, which also indicates strong modal coupling between these two modes.

To study the performance of TMD-based control, the total generalized mass of all TMDs are assumed to be 1% of that of the 1st symmetric bending mode of the bridge. Totally 12 identical TMDs are distributed evenly on the two sides in the middle area of the main

148

Page 156: Dynamic performance of bridges and vehicles under strong wind

span. To save the installation space of TMDs, lever type of TMDs can be adopted and each of them has the mass of 11 ton for the Yichang Bridge. The basic distribution scheme and structure of the lever-type of TMDs are the same as that used by Gu et al. (2001). Figs. 6.9 and 6.10 show the response power spectral density of the 1st symmetric bending mode and the 1st symmetric torsional mode with and without control, respectively. For the original uncontrolled case, there are two peaks in the response curve of the bending mode. One corresponds to the modal frequency of the 1st symmetric bending mode, and the other corresponds to the modal frequency of the 1st symmetric torsional mode.

0 20 40 60

0

0.4

0.8

1.2

1.6

U (m /s)

1s t s ym . bending m ode

1s t s ym . tors ional m ode

Coupled analys is

S ingle m ode analys is

RM

S σ h

(m)

Fig. 6.9 Root-Mean-Square of Vertical Displacements at Mid-span without Control

149

Page 157: Dynamic performance of bridges and vehicles under strong wind

0.1 0.2 0.3 0.4 0.50.01

0.1

1

D

ispla

cem

ent P

ower

Spe

ctru

m (m

2 /Hz)

f (Hz)

Uncontrolled Controlled

(a) Bending mode

0.1 0.2 0 .3 0 .4 0 .51E-4

1E-3

0.01

0.1

Dis

plac

emen

t Pow

er S

pect

rum

(m2 /H

z)

f (H z)

U ncontrolled C ontrolled

(b) Torsional mode

Fig. 6.10 Displacement Spectra with Bending Mode Based Control (U= 40 m/s)

150

Page 158: Dynamic performance of bridges and vehicles under strong wind

0.1 0.2 0.3 0.4 0.50.01

0.1

1

Disp

lace

men

t Pow

er S

pect

rum

(m2 /H

z)

f (Hz)

Uncontrolled Controlled

(a) Bending mode

0 .1 0 .2 0 .3 0 .4 0 .5

1 E -4

1 E -3

0 .0 1

0 .1

Disp

lace

men

t Pow

er S

pect

rum

(m2 /H

z)

f (H z )

U n c o n tro lled C o n tro lle d

(b) Torsional mode

Fig. 6.11 Displacement Spectra with Torsional Mode Based Control (U= 40 m/s)

151

Page 159: Dynamic performance of bridges and vehicles under strong wind

In Fig. 6.9, the parameters of the TMDs were designed for the 1st symmetric bending mode at U = 40 m/s in the conventional resonant control approach, and modal coupling effects were not considered in the TMD design. Similar to the 2DOF model, it was found that the TMDs designed based on the resonant reduction of the 1st symmetric bending mode are not very efficient in reducing the resonant part of the response for the 1st symmetric bending mode (first peak of Fig. 6.9(a)) due to its already high modal damping ratio. Moreover, they can hardly suppress the response component caused by coupling from the 1st symmetric torsional mode (the second peak of Fig. 6.9(a)). On the other hand, for the response component of the 1st symmetric torsional mode caused by coupling from the 1st symmetric bending mode (the first peak of Fig. 6.9(b)), this TMD design has some insignificant control effect. As expected, resonant component for 1st symmetric torsional mode (the second peak of Fig. 6.9(b)) is not efficiently suppressed.

In contrast, Fig. 6.10 shows that the TMDs designed for the 1st torsional mode are very efficient in not only reducing the resonant component of 1st symmetric torsional mode (the second peak of Fig. 6.10(b)), but also in reducing the second peak of Fig. 6.10(a), which is the response component of the 1st symmetric bending mode caused by coupling from the 1st torsional mode. However, it can also be found that they are not efficient for the response component of the 1st symmetric torsional mode caused by coupling from the 1st symmetric bending mode (the first peak of Fig. 6.10(b)).

6.4.2 Mechanism of Buffeting Control with Strong Coupling Effects

Based on the observations made above, to reduce the vibration in a case of strong modal coupling, TMDs can be designed to suppress peak values in the spectrum. It is obvious in Fig. 6.10 that, to control the response of the 1st symmetric bending mode, TMDs whose frequencies are close to that of the 1st symmetry torsional mode don’t suppress the resonant vibration of the bending mode, but suppress the modal coupling effect between these two modes (the second peak of Fig. 6.10(a)). Suppressing this coupling effect may be significant in reducing the overall vibration.

As has been discussed earlier, Eq (6.1) indicates that the resonant peak of the bending mode is difficult to control due to its high total damping ratio. This also implies that when strong aerodynamic coupling effect exists, the conventionally designed TMDs considering only resonant vibrations may not be efficient for control performance. Optimal variables of TMDs to control any given modes should be searched in a full range (not just near the modal frequencies) in order to optimally control the total response due to resonant vibration and coupled vibration.

For demonstration, two studies were conducted on the Yichang Suspension Bridge under varied wind speed for controlling the vibration of the 1st symmetric bending mode. The first one was to search the optimal TMD frequency considering only the resonant vibration. The optimal TMD frequency was thus only searched around the modal frequency of the 1st symmetric bending mode. In the second one, the optimal TMD frequency was designed considering only modal coupling effect from the 1st symmetric torsional mode. The TMD frequency was thus searched and restricted around the modal frequency of the 1st symmetric torsional mode. 152

Page 160: Dynamic performance of bridges and vehicles under strong wind

0 20 40 601E-3

0.01

0.1

1

10

100

Critical point for the control efficiency

Co

ntro

l effi

cien

cy R

h(%)

Wind velocity U (m/s)

Resonant vibration based control

Modal coupling vibration based control

C A B

Fig. 6.12 Bending Mode Displacement Control Efficiency with Different Control Schemes

For both cases, TMDs were designed under every wind speed (TMD has an optimal frequency under each wind velocity due to the aeroelastic effects) and corresponding control efficiency of the 1st symmetric bending mode is plotted into the two curves that are shown in Fig. 6.11. It is found that there is an intersection of the two control efficiency curves at the wind velocity near 60 m/s. These two curves suggest that, for optimal control efficiency on the 1st symmetric bending mode, the TMDs should be designed to control the resonant vibration (control efficiency is along point A to point B), then should be changed to control the vibration due to modal coupling when wind speed surpasses 60 m/s (control efficiency is along point B to point C). Ideally, if only a single-frequency TMD is designed, the TMD frequency should be adjustable to achieve optimal control efficiency under different wind speed, namely, the optimal control efficiency will be along points A, B, and C. Practically, for TMD unable to adjust the frequency, the optimal design of TMD relies on the wind speed under which the control is expected. For this particular example, if the control objective is to suppress the response under strong wind with velocity over 60 m/s, the TMD (no matter single or multiple) should be designed around the frequency of the 1st symmetric torsional mode since it will result in the best overall control performance. On the other hand, when the target wind speed is lower than 60 m/s for this problem, the conventional TMD design still have the better control performance. In this case, multiple TMDs can be used, some of them are designed for the 1st symmetric bending mode and some for the 1st symmetric torsional 153

Page 161: Dynamic performance of bridges and vehicles under strong wind

mode. The mass distribution among multiple modes can be decided based on the different contribution to the overall bridge response from these modes. Since this is the issue about conventional resonant suppression design, it is not repeated here.

The above results indicate that for modern long-span bridges, there exists an alternative control approach for TMD design other than the conventional one that focuses only on resonant vibration. Optimal parameters for TMD design should be based on two control mechanisms, i.e., resonant and coupled vibrations. With the increase of modal coupling for ultra-long bridges, the proposed new control strategy may play much more important role in the coupled vibration control.

6.4.3 Dual-Objective Control of Coupled Vibration

TMDs have been proven effective in raising the critical flutter wind speed (Pourzeynali and Datta 2002; Gu et al. 1998). With the common adoption of streamlined cross sections for long-span bridges, coupled flutter is the dominant flutter instability. It has been found above that TMDs may also be effective in suppressing coupled buffeting response through reducing modal coupling effects. Similarly, it is also expected to be effective in improving the flutter stability through destroying the pre-existing modal coupling mechanism. For this reason, it is very logical to pursue dual-objective control of the TMDs, namely, suppressing buffeting response and improving flutter stability at the same time. To validate such a statement, the Yichang Suspension Bridge was also studied on the flutter stability with the TMDs designed for suppressing buffeting response considering the modal coupling effects. It was found that when TMDs are designed for buffeting response control at the wind speed of 65 m/s, the critical flutter wind speed can also be improved from 73 to 86 m/s. Such results are very promising for the extra long-span bridges where TMDs can be effectively adopted to enhance the flutter stability and to suppress the excessive wind-induced buffeting vibration.

6.5 Concluding Remarks

Conventional TMD control approach usually focuses on suppressing the resonant vibration by supplying additional damping to the concerned modes. This approach could be inefficient for coupled vibration of long-span bridges in strong wind due to two reasons.

The first is the strong modal coupling effects in strong wind. For slender long-span bridges, the aeroelastic forces from the wind action often cause several vibration modes to couple together. Such coupling effect increases with the increase of wind speed. Coupled buffeting response of each mode usually consists of two major parts: one is resonant component associated with its modal frequency; the other part of response is due to the modal coupling with other modes. For bridges with weak modal coupling effects, the second part is trivial. However, for long-span bridges in high wind speed, modal coupling effects may become quite strong. The latter part of the response is no longer negligible and a control approach focusing on the first part may be inefficient.

The second reason is the increased total modal damping, caused by aeroelastic effects in strong wind, of the concerned modes. Even though damping helps reduce bridge vibration, satisfactory control performance may be extremely difficult to achieve by supplying

154

Page 162: Dynamic performance of bridges and vehicles under strong wind

155

additional damping using the conventionally designed TMDs, since damping of those concerned modes is already high compared with the additional damping provided by the TMDs.

The present study proposes a new control approach that is to attenuate the modal coupling effects, in addition to suppressing the resonant vibration with TMDs. The vibration contributions to the total response from modal coupling (second part of Eq. (6.28)) can be significant at high wind velocity. Weakening the coupling effects with TMDs can significantly reduce the overall responses. The newly introduced control approach also enables a well-designed TMD system to be efficient in controlling buffeting vibration of coupled modes even with high modal damping under high wind velocity. For optimal control efficiency in applications, the TMD frequency needs to be adaptable in order to switch from resonant suppression to coupling suppression, or multiple frequency TMDs are needed in order to control both resonant and coupling effects.

The effects of TMDs on reducing both resonant and coupled vibrations have been demonstrated through the analytically derived closed-form solutions. Numerical analyses on a 2DOF model and an actual long-span bridge have validated that the new control approach may lead to more efficient control performance than the conventional resonant- suppression strategy when the coupling effects are significant and when the damping ratios of those modes of concern are high. Finally, the concept of dual-objective control, i.e., the TMD control system for both flutter stability improvement and buffeting response reduction, is briefly discussed.

Page 163: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 7. OPTIMAL VARIABLES OF TMDS FOR MULTI-MODE BUFFETING CONTROL OF LONG-SPAN BRIDGES

7.1 Introduction

Under wind excitations, long-span bridges exhibit complex aerodynamic behaviors. Buffeting random response induced by the turbulence of airflow happens throughout the full range of wind speeds. As the wind speed increases, aerodynamic instability phenomena such as flutter may occur (Simiu and Scanlan 1996). Much research effort has been made towards mitigating excessive vibrations and improving aerodynamic stabilities for bridges during construction (Conti et al. 1996; Takeda et al. 1998) and at service stages (Gu et al. 1994; Wilde et al. 1999). Among all of the control procedures, dynamic energy absorbers such as tuned mass dampers (TMDs) were studied and adopted in suppressing excessive vibrations or maintaining the flutter stability of bridges (Gu et al. 1998).

In recent years, the importance of aeroelastic modal coupling to the bridge aerodynamic behaviors has been recognized (Tanaka et al. 1993; Bucher and Lin 1988; Lin and Yang 1983; Miyata and Yamada 1999; Cai and Albrecht 2000). It has been concluded that the coupling tendency of two modes depends on their mode shapes and natural frequencies in still air as well as the flutter derivatives of the bridge section (Jain et al. 1996). The adoption of more slender deck and the increase of bridge span length tend to result in closer modal frequencies. As a result, modal coupling effects through aeroelastic forces in high wind speeds increase (Jain et al. 1998; Namini et al. 1992; Katsuchi et al. 1998; Thorbek and Hansen 1998).

The TMD is known to be effective in suppressing single-mode resonant vibrations when its frequency is tuned to the modal frequency of the structure. When the modal frequencies of the bridge are well separated and modal coupling effects are weak, each TMD is mainly designed to control a single-mode vibration while the effects from other modes on the control are omitted (Igusa and Xu 1991; Kareem and Kline 1995). Abe and Igusa (Abe and Igusa 1995) studied the performances of TMDs on a coupled system with closely-spaced natural frequencies. Through the assumption of very close frequencies, some analytical studies were given to the strongly coupled system. Studies on multi-mode wind-induced vibration controls are limited to the cases with very weak coupling effects (Chang et al. 2003) and few works have focused on the vibration controls of bridges with strong aeroelastic modal coupling.

Considering the complexity of bridge conditions under strong winds, an adjustable TMD system is desirable for the control system to be more robust and effective over various circumstances. However, the effects of system properties on the optimal variables of the TMDs have not been sufficiently addressed. Such study is extremely helpful in evaluating the control performance before the real control devices are designed in practice. It also helps in deciding, for the adaptive control system, what parameters of the bridge-flow system are to be monitored in a feed-back control. With such information, the number of variables to be monitored can accordingly be reduced to the least, through which the cost and complexity of the controller can also be minimized.

In this chapter, a comprehensive investigation on the optimal variables of the adjustable TMD system is made. First, a general formulation of the multi-mode buffeting response control with multiple TMDs is developed. Second, a control strategy with “three-row” TMDs is

156

Page 164: Dynamic performance of bridges and vehicles under strong wind

discussed especially to study the coupled vibration controls. Finally, the three most important factors of the bridge-flow system are studied numerically with the Humen Suspension Bridge built in China. This parametric study is conducted to investigate the factors of the bridge-flow system that will affect the optimal variables of TMDs as well as the control efficiency. These analytical results will be very useful in carrying out further studies of adaptive control strategy based on the “three-row” TMD model in order to “smartly” suppress the wind-induced vibrations.

7.2 Formulations of Multi-mode Coupled Vibration Control with TMDs

Consider a general case shown in Fig. 7.1. A bridge has multiple TMDs, displacement r(x, t), and wind forces consisting of buffeting force fb(x, t) and aeroelastic self-excited force . Assuming that a total number of nsf (x, r, r) 1 modes are included in the analysis and a total number of n2 TMDs are attached to the bridge deck at the location of xs (s = 1 to n2), the equation of motion for the bridge-TMD system can be derived as:

1 2 s

s+1 s+2 2s

2s+1 2s+2 n

d

Torsionalcenter

s-1s

2s-1

2n2 -1

Fig. 7.1 Overview of the general placement of “three-row” TMDs

′′ ′+ + =Mη Cη Sη G (7.1)

where

1 21 n 1, ... , , ... n

′= ξ ξ γ γη ; (7.2)

157

Page 165: Dynamic performance of bridges and vehicles under strong wind

1 11

2 1 2 2

1 2n n n n

3n n n n

M M

M I× ×

× ×

2 =

M ; (7.3)

1 1

2 2

1n n

2n n

C 0

0 C×

×

=

C ; (7.4)

1 1

2 2

1n n

2n n

S 0

0 S×

×

=

S ; (7.5)

bQ 0 0=G , ... . (7.6)

ξ= generalized coordinate of the bridge; γ = coordinate of TMDs; a superscript prime “`” represents a derivative with respect to time t; U = mean velocity of the oncoming wind; B = bridge width; n1 = number of modes; n2 = number of TMDs; = unit matrix; and Q = generalized buffeting force. The components of the matrices are:

I b

1ij i i ij ij

ii

1C ( ) 2 B Z2I

ω = ζ ω δ − ρ ω2 (i, j =1...n1) (7.7)

1 2 3ij i ij ij

ii

1S ( ) B T2I

ω = ω δ − ρ ω2 (i, j =1...n1) (7.8)

( )2 2

2 2

n n2 1 1 2 2n n t t t t t tC diag 2 ,2 ,..., 2× = ζ ω ζ ω ζ ω (7.9)

( ) ( ) ( )( )2

2 2

22 2 n2 1 2n n t t tS diag , ,...,× = ω ω ω (7.10)

2ns

ij t i s i s s j s j s s1 s 1ij

ii

I I h (x ) (x )d h (x ) (x )dM

I=

+ +α +α =

∑ (i, j =1...n1) (7.11)

jt i j i j j2

ijii

I h (x ) (x )dM

I

+ α = (i =1...n1; j =1...n2) (7.12)

3ij j i j i iM h (x ) (x )d= +α (i=1...n2; j=1...n1) (7.13)

i j i j i j i j i j

i j i j i j i j

* * * * *ij 1 h h 2 h 5 h p 1 p p 2 p

* * 2 * *5 p h 1 h 2 5 p

Z H G B H G H G P G B P G

P G B A G B A G B A Gα α

α α α α

= + ⋅ + + + ⋅

+ + ⋅ + + ⋅ (7.14)

158

Page 166: Dynamic performance of bridges and vehicles under strong wind

i j i j i j i j i j

i j i j i j i j

* * * * *ij 4 h h 3 h 6 h p 4 p p 3 p

* * * *6 p h 3 h 4 6 p

1 1 1T H G H G H G P G P GB B B

1 P G B A G A G A GB

α α

α α α α

= + + + +

+ + ⋅ + + (7.15)

whereδ = Kronecker delta function that is equal to 1 if i = j and equal to 0 if i ; and ij j≠ iω iζ = circular natural frequency and mechanical damping ratio of ith mode, respectively; ρ = air density; = experimentally determined flutter derivatives;* * *

i i iH , P ,A (i 1 6)= − s st tandζ ω =

damping ratio and circular natural frequency of the sth TMD, respectively; = mstI s (the mass of

the sth TMD) and ds = horizontal distance between the sth TMD and the torsion center of the cross-section (see Fig. 7.1).

The modal integral ( G ) can be expressed as: jisr

i jr s i j0G r (x)s (x)dx

l= ∫ (7.16)

and the mass moment of inertia of the bridge section can be expressed as:

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )ij i j i j i j0I m x h x h x p x p x I x x x dx

l = + + α ∫ α (7.17)

where m(x) = mass per length of the deck for vertical and lateral bending modes; and I(x) = mass moment of inertia per length of the deck for torsion mode; l = bridge length.

The generalized inertia ratio between the sth TMD and the ith mode, isµ , is defined as:

2s

t i s i s sis

ii

I h (x ) (x )dI

+ αµ = (7.18)

Buffeting force due to the turbulence of wind can be expressed as:

( ) ( ) ( )2 'b L L D

u x, t w x, t1L x, t U B C (2 ) (C C )2 U

= ρ + +

U (7.19)

( ) ( ) ( )2 'b D D

u x, t w x, t1D x, t U B C (2 ) C2 U

= ρ +

U (7.20)

( ) ( ) ( )2 2 'b M M

u x, t w x, t1M x, t U B C (2 ) C2 U

= ρ +

U (7.21)

159

Page 167: Dynamic performance of bridges and vehicles under strong wind

where u and w = horizontal and vertical turbulence of wind flow, respectively; and CL, CD and CM = static coefficients of lift, drag and moment of the bridge deck, respectively. A prime over the coefficients represents a derivative with respect to the attack angle.

The generalized buffeting force for the ith mode in Eq. (7.6) can be written as

( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( )i

l L i D i M ib 0

ii L D i D i M i

2 C h x C p x B C x u x, tUBQ2I C C h x C p x B C x w x, t d

+ + ⋅ α + ρ =′ ′ ′+ + + ⋅ α

∫ x (7.22)

Eq. (7.1) can be Fourier transformed into a new format as

η =F G (7.23)

where η and G = Fourier transformation of η and G, respectively. The impedance matrix F has the general form as ( ) ( )2

ij ij ij ijC Si= −ω + ⋅ω ω + ωF M , where subscripts i and j = 1 to (n1+n2)

and 1= −i .

The mean square of displacements in vertical, lateral and torsion directions can be written as follows:

b bi j

2r i j ij Q Q0

i j

ˆ (x) r (x)r (x) F S F dn∞

σ =∑∑ ∫ *ij

*

(7.24)

where ; ; 1ij ijF [ ]−= F * 1

ijijF [ ]−= F n2ω

is the frequency and

( ) ( ) ( ) ( )b bi j

2uu ww uw

Q Q ij uu ij ww ij uwii jj

UB 1S n Y S n Y S n Y S n2 I I

ρ = + + (7.25)

( ) ( ) ( )

( ) ( ) (i j i j i j

i j i j i j i j i j i j

2 2 2uuij L h h D p p M

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

Y 2C H 2C H 2C H

4C C H H 4C C H H 4C C H H

α α

α α α α

= + +

+ + + + + + ) (7.26)

( ) ( ) ( ) ( ) ( )( ) ( ) ( )

i j i j i j i j i j

i j i j i j i j

2 2 2wwij 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 C C H C H C H C C C H H

C C C H H C C H H

α α

α α α α

′ ′ ′ ′ ′= + + + + + +

′ ′ ′ ′+ + + + + (7.27)

( )( ) ( )( ) ( ) ( )( )

i j i j i j

i j i j

i j i j i j i j

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

L D L D D h p p h

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

Y 2 2 C C C H 2C C H 2C C H

C C C C C H H

C C C C C H H C C C C H H

α α

α α α α

′ ′ ′= + + +

′ ′+ + + +

′ ′ ′ ′+ + + + + + +

(7.28)

160

Page 168: Dynamic performance of bridges and vehicles under strong wind

and using

( ) ( ) 1 2

i j

c x x /r s 1 2 1 20

H x s x e dl l l

i j0r − −= ∫ ∫ x dx (7.29)

where ; s h ; iiii orp,hr α= j j j, p or= jαncU

lλ= ( 7 20≤ λ ≤ ).

The power spectra of the wind velocity components in the horizontal and vertical directions u and w in boundary layer can be expressed as (Simiu and Scanlan 1986):

( )( )

2*

uu 5 / 3200zuS n

U 1 50nz / U=

+ (7.30)

( )( )

2*

ww 5/ 3

3.36zuS nU 1 10 nz / U

= +

(7.31)

( )( )

2*

uw 2.414zuS n

U 1 9.6nz / U=

+ (7.32)

whereS and S = wind velocity spectrum in the horizontal and vertical direction, respectively; and S = cross spectrum, respectively.

uu ww

uw

Control efficiency of displacement at the location of x on the bridge span and in the direction of r is defined as

( ) ( )( )

rr

r

ˆR 1 100%

xx

x σ

= − × σ (7.33)

where and are the root-mean-square (RMS) of displacement after and before control at the location of x and in the direction of r, respectively. r=h, p or α representing vertical, lateral and torsion direction, respectively.

( )rˆ xσ ( )r xσ

7.3 Parametrical Studies on “Three-row” TMD Control

7.3.1 “Three-row” TMD model

According to the previous studies on modal coupling, there are only a limited number of modes prone to couple together (Katsuchi et al. 1998). Among all of the coupling cases for streamlined cross sections, the most common modal coupling is between vertical bending mode and torsion mode. Furthermore, in terms of the contribution of individual mode to the total buffeting response as well as the flutter occurrence, the vertical bending and torsion modes

161

Page 169: Dynamic performance of bridges and vehicles under strong wind

usually play the major role. Hence, an appropriate control strategy of TMD system will be developed based on such observed characteristics. It is known that TMDs placed on the center line of the cross section normally have insignificant control effect on the torsion modes. Therefore, we adopt three rows of TMDs: one along the center line of the cross section (named center row hereafter), and other two identical rows along the two sides of the cross section (named side rows hereafter) as shown in Fig. 7.2. This model is to literally separate the TMD control role into vertical bending and torsion modes since they are the main concern of wind-induced vibrations. In other words, the center row TMD is mainly for vertical mode and the two side rows mainly for torsion mode. Such separation of TMD role is very rough and actually only accurate for the situation when modal coupling effect is weak under low wind speed. As will be found later, side rows of TMD will also contribute to the dynamic suppression vertical mode in the high wind speed when strong modal coupling exists.

The difference between the multiple-TMD and single-TMD placements is that the multiple placements are more robust in control since they cover a wider range of frequencies (Kareem and Kline 1995, Abe and Fujino 1994). Since the present study is only to disclose the nature of optimal variables for coupled vibration controls, only one TMD in each row is considered to reduce the complexity while without losing the generalities. Also, since the damping ratio of the TMD is not a very sensitive variable for TMD (Gu et al. 1994), damping ratios of all the TMDs are assumed to be the same as tζ . The frequency and mass of the two identical TMDs on the side rows are assumed to be and m1ω 1, while the frequency and mass of the center TMD are assumed to beω and m2 2 (Figs. 7.2-7.3), respectively. The total generalized mass of TMDs, greatly related to the efficiency and the cost of the control system, is assumed to be 1% of that of the 1st bending mode of the bridge.

In the present study, two cases are considered in the analysis of the optimal variables of the TMDs. In Case 1, the TMD frequencies 1ω and 2ω under a particular wind speed are set to be the optimal values based on single-torsion and single-bending mode vibration controls, respectively. These optimal values were analytically derived by Fujino and Abe (1993). Under the condition of a given total mass of TMDs (1% of the 1st bending mode), the distribution of mass between m1 and m2 is varied and studied. This is to simulate the case when the TMD mass can be adjusted while the control objective of each TMD targets a particular mode (e.g. center row for the bending mode and side rows for the torsion mode). In Case 2, the total mass of the two side row TMDs (2m1) is set to be equal to the center one (m2), maintaining the total TMD mass the same as Case 1. Only the variables , and 1ω 2ω tζ can be adjusted. This is to simulate a case that the mass of each TMD is fixed, while the frequency and damping ratio can be adjusted to obtain the optimal control performance.

7.3.2 Optimal Variables of “Three-row” TMDs

Humen Suspension Bridge with a main span of 888m is chosen as an example here with the basic data shown in Table 7.1 (Lin and Xiang 1995). A coupled vibration analysis has shown that the 1st symmetric vertical bending mode and the 1st symmetric torsion mode are the two modes most prone to couple together. Therefore, only these two modes, with a modal frequency of 0.17 and 0.36 Hz, respectively, are included in the following analysis. Flutter derivatives H*

1 3−

and are shown in Fig. 7.4 (Lin and Xiang 1995). With the increase of wind speed, the *1 3A −

162

Page 170: Dynamic performance of bridges and vehicles under strong wind

modal properties including modal damping and modal frequencies can be obtained using complex eigenvalue approach (Chen and Cai 2003). The results of modal properties are plotted in Fig. 7.5 and the flutter critical wind speed is identified as 87 m/s for Humen Bridge, which is very close to the result from wind tunnel test (Lin and Xiang 1995). Fig. 7.6 gives the buffeting response in vertical and torsional directions in the center of the mid-span section without control. Results from coupled analysis based on the two modes (vertical bending and torsion) and from the single- mode analysis are compared. Strong coupling effects in high wind speed can be observed that the results of coupling analysis differ obviously from that of single-mode analysis when wind speed is high.

Cross section ofthe bridge

d s

ω2,m2,ζt ω1,m1,ζt ω1,m1,ζt

Fig. 7.2 “Three-row” TMD model

Mid-span

ω2,m2,

ω1,m1,

ω1,m1,

Fig. 7.3 Overview of the “three-row” TMDs placement for case study of Humen Suspension Bridge

In following control studies, the chosen location of interests is at the edge of the mid-span section where the largest vertical displacements contributed by both the symmetric bending and torsion modes are expected. The vertical displacement in that location can combine the 163

Page 171: Dynamic performance of bridges and vehicles under strong wind

contributions from the vertical bending mode as well as the torsion mode. The contribution to the vertical displacement at the edge of the mid-span section by the torsion mode is calculated with the torsion displacement multiple the half of bridge width. The TMDs are placed as shown in Fig. 7.3. The total generalized mass of TMDs, chosen as 1% of that corresponding to the 1st bending mode, is 80,000 kg. For Case 2, the center row and two side rows have the same mass, i.e., 2m1 = m2 = 40,000 kg. The horizontal distance from the side row TMD to the center of torsion ds equals to 14 m for Humen Bridge.

If only a single-mode-based vibration is considered, the optimal frequency of TMDs at wind velocity 30 m/s can be obtained with formulas by Fujino and Abe (1993) as:

2 0.99h

ωω

= (7.34)

1 0.97α

ω=

ω (7.35)

where, ωh and ωα = natural circular frequencies of vertical and torsion modes, respectively. As discussed earlier, these single-mode-based optimal frequencies are chosen for the TMDs in Case 1.

There exist many factors affecting the coupling effects among modes. From the existing knowledge of the modal coupling, the frequency ratio between the coupling-prone modes and the wind speed are the main possible factors that may affect the optimal variables of TMDs (Katsuchi et al. 1998). The wind speed greatly affects the aeroelastic modal coupling effects and affects the buffeting contribution from coupling-prone modes. These factors are discussed below.

7.3.3 Effect of Wind Speed

Aeroelastic coupling is of a great concern in wind-induced vibration of long-span bridges especially when wind speed is quite high. For most streamlined cross sections, the aeroelastic coupling is directly related to the wind speeds. When the total mass of TMDs is fixed as in Case 1, the predicted optimal mass distributions among the TMDs vary significantly for different wind speeds as shown in Fig. 7.7. At the wind speed of about 60 m/s, the masses of center and side TMDs are about equal (2m1/mtotal ≅ m2/mtotal ≅ 0.5). At low wind speed (say 40 m/s), much more mass needs to be assigned to m2 (m2/mtotal ≅ 0.85) with the bending-single-mode-based optimal frequency. However, at higher wind speed (say 80 m/s), much more mass should be assigned to m1 with the torsion-single-mode-based (2m1/mtotal ≅ 0.80). There are probably two reasons for such phenomena: one is in high wind speed, the contribution to the total vertical response at the edge of the mid-span section from the torsion mode increases; the other is that the resonant part of bending mode response with large modal damping (Fig. 7.6) is difficult to be suppressed. Another part of bending response in high wind speed is due to the modal coupling effects between bending and torsion modes (Chen and Cai 2003), and it can be suppressed with TMD with torsion mode frequency. As stated earlier, the frequencies of TMD are set to be equal to the optimal frequency considering single-mode-based vibrations in Case 1. In terms of vertical response control at the edge of the mid-span section, it is noted that the dominant vibration mode changes from vertical to torsion mode at the wind speed of about 60 m/s.

164

Page 172: Dynamic performance of bridges and vehicles under strong wind

For Case 2, since the mass of each TMD is fixed, namely 2m1 = m2, the optimal frequency changes with the wind speed especially when the wind speed is high as shown in Fig. 7.8, where ωh and ωα are the modal frequencies of bending and torsion modes considered in this study, namely, 0.17 and 0.36 Hz, respectively. Fig. 7.8 also indicates that with the increase of the wind speed, the optimal frequency of TMD in the center row may change dramatically from around the modal frequency of the bending mode (indicated by ω2/ωh =1.0) to around the torsion mode frequency (indicated by ω2/ωh = 2.0 since ωα is about twice of the bending frequency ωh). This drastic change occurs at the wind speed of 60 m/s where dominant vibration mode for the vertical response control at the edge of the mid-span section changes from bending mode to torsion mode.

Fig. 7.9 shows the respective optimal damping ratio of TMDs for Case 1 and Case 2. Compared to other variables such as mass distribution and frequency, the damping ratio of the TMDs seems to be less sensitive to the change of wind speed. Therefore, it has relatively the least necessity to be adjusted in an adaptive control.

With the change of wind speed, the optimal control efficiency of the vertical displacement on the edge at the mid-span section. Rh also varies accordingly as shown in Fig. 7.10. It is observed that the control efficiency decreases with the increase of wind speed until around 60 m/s, where the dominant vibration control mode changes from bending mode to torsion mode. Then, the control efficiency increases with the increase of wind speed. Such phenomenon cannot be observed in a single-mode-based control analysis. Hence, this observation is extremely important for the design of a special controller that, for example, controls bridge vibrations under hurricane-induced strong winds. An ideal TMD system in this case should target the bending vibration when wind speed is less than 60 m/s and then is adjusted to target the torsion mode vibration.

Table 7.1 Main Parameters of Humen Bridge

Main span (m) 888 Lift coefficient at 0o attack angle

0.02

Width of the deck (m) 35.6 Drag coefficient at 0o attack angle

0.84

Clearance above water (m)

60 Pitching coefficient at 0o attack angle

0.019

Equivalent mass per length (103 *kg/m)

18.34 ( ) o/CL 0=∂∂

αα 0.51

Equivalent inertial moment of mass per length

(103 *kg/m)

1743 ( ) o/CM 0=∂∂

αα 0.62

Structural damping ratio 0.005 ds (m) 14 Natural frequency of 1st

symmetric vertical bending mode (Hz)

0.17 Natural frequency of 1st symmetric torsion

(Hz)

0.36

Natural frequency of 1st asymmetric vertical mode (Hz)

0.28 Natural frequency of 1st asymmetric torsion (Hz)

0.43

Design wind speed (m/s) 57

165

Page 173: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20

-4

0

4

8

12

Flut

ter d

eriv

ativ

es

U/(fB)

H1*

H2*

H3*

(a)

0 5 10 15 20-0.5

0.0

0.5

1.0

1.5

2.0

2.5

Flut

ter d

eriv

ativ

es

U/(fB)

A1*

A2*

A3*

(b)

Fig. 7.4 Flutter derivatives of Humen Suspension Bridge

166

Page 174: Dynamic performance of bridges and vehicles under strong wind

0 20 40 60 80 100

0.00

0.04

0.08

0.12

0.16

Ucr=87 m/s

Vertical mode Torsion mode

Wind velocity U (m/s)

Mod

al d

ampi

ng ra

tio o

f ver

tical

mod

e

-0.03

-0.02

-0.01

0.00

0.01

Modal dam

ping ratio of torsion mode

(a)

0 20 40 60 80 1000.15

0.20

0.25

0.30

0.35

0.40

Ucr=87 m/s

Mod

al fr

eque

ncy

(Hz)

Wind velocity U (m/s)

Vertical mode Torsion mode

(b)

Fig. 7.5 Variation of modal properties for Humen Suspension Bridge with wind speed

167

Page 175: Dynamic performance of bridges and vehicles under strong wind

20 40 60 800.0

0.2

0.4

0.6

0.8

1.0

Ucr=87m/s

RMS

of v

ertic

al d

ispl

acem

ent σ

h(m)

Wind speed U (m/s)

Dual-mode coupled buffeting Single-mode buffeting

20 40 60 800.000

0.025

0.050

0.075

Ucr=87m/s

RMS

of to

rsio

n di

spla

cem

ent σ

α(rad)

Wind speed U (m/s)

Dual-mode coupled buffeting Single-mode buffeting

(a)

(b)

Fig. 7.6 Buffeting response at the centre of mid-span for Humen suspension bridge without control

168

Page 176: Dynamic performance of bridges and vehicles under strong wind

0 20 40 60 800.0

0.2

0.4

0.6

0.8

1.0

2m1/mtotal

Wind speed U (m/s)

2m1/m

tota

l

0.0

0.2

0.4

0.6

0.8

1.0

m2/mtotal

m2 /m

total

Fig. 7.7 Optimal mass distribution of TMDs versus wind speed (Case 1)

0 20 40 60 800.94

0.95

0.96

0.97

0.98 ω1/ωα

Wind speed U (m/s)

ω1/ω

α

0.8

1.0

1.2

1.4

1.6

1.8

2.0

ω2/ωh

ω2 /ω

h

Fig. 7.8 Optimal frequencies of TMDs versus wind speed (Case 2)

169

Page 177: Dynamic performance of bridges and vehicles under strong wind

0 20 40 60 800.05

0.06

0.07

0.08

ζ t

Wind speed U (m/s)

Case 1 Case 2

Fig. 7.9 Optimal damping ratio of TMDs versus wind speed

0 20 40 60 80 100

20

30

40

Cont

rol e

ffici

ency

Rh (%

)

Wind speed U (m/s)

Case 1 Case 2

Fig. 7.10 Optimal control efficiency of “three-row” TMDs versus wind speed

170

Page 178: Dynamic performance of bridges and vehicles under strong wind

1.0 1.2 1.4 1.6 1.80.0

0.2

0.4

0.6

0.8

1.0

2m1/mtotal

ωα/ωh

2m1/m

tota

l

0.0

0.2

0.4

0.6

0.8

1.0

m2/mtotal

m2 /m

total

Fig. 7.11 Optimal mass distribution of TMDs versus frequency ratio of coupled modes (Case 1)

7.3.4 Effect of Frequency Ratio

With an optimal searching of the TMD mass under the condition of a fixed total mass, the optimal distribution of the total mass among the center and side TMDs has been obtained above. By numerically varying the torsion/bending modal frequencies ratio of the bridge, the effects of the bridge frequency ratio on the TMD optimal variables are studied below considering the wind speed of 30 m/s. It should be noted that for Case 1, the ratios between the TMD and the bridge modal frequencies are fixed as shown in Eqs. (7.34) and (7.35), i.e., fixed to the optimal frequency for the single-mode-based vibration control. Therefore, the numerical values of the TMD frequencies ω1 and ω2 vary with the change of the natural frequencies ωh and ωα accordingly.

Fig. 7.11 shows that the optimal mass of each TMD depends on the frequency ratio between the torsion and bending modes ωα/ωh. When the value of ωα/ωh is around 1.15, 2m1/mtotal is approximately equal to m2/mtotal. When ωα/ωh is around 1.76, about 95% of the total mass should be allocated to m2 (that targets the bending vibration) for the most efficient control.

171

Page 179: Dynamic performance of bridges and vehicles under strong wind

1.2 1.4 1.6 1.80.94

0.95

0.96

0.97

ω1/ωα

ωα/ωh

ω1/ω

α

0.96

0.97

0.98

0.99

ω2/ωh

ω2 /ω

h

Fig. 7.12 Optimal frequencies of TMDs versus frequency ratio of coupled modes (Case 2)

Fig. 7.12 shows the optimal frequencies of the TMDs versus the ωα/ωh ratio when wind speed is 30m/s, considering Case 2 where 2m1 = m2. It can be found from the figure that the ωα/ωh ratio affects the optimal frequency of the TMDs for coupled buffeting control. When the frequencies of bending and torsion modes are well separated (with high ωα/ωh values, say 1.6), the optimal frequencies of TMDs shown in Fig. 7.12 are quite close to those of single-mode-based cases shown in Eqs. (7.34) and (7.35). This finding justifies the common assumptions that the control strategy for weakly coupled vibration can be simplified as that of single-mode-based control.

Fig. 7.13 shows that the optimal damping ratio of the TMD is also affected by the ratio of ωα/ωh for both Cases 1 and 2. With the increase of ωα/ωh, the damping ratio of the TMDs approaches to that of single-mode-based case (Fujino and Abe 1993). The control efficiency of the vertical buffeting vibration (the total vertical displacement from both the vertical vibration of the bending mode and the rotation of the torsion mode) at the edge of the mid-span section is shown in Fig. 7.14. With the increase of ωα/ωh, the control efficiency decreases and approach to that of a single-mode-based control.

The results discussed above show that the “three-row” TMD control system has the collaborative control effect for the vibration when the frequencies of coupling-prone modes are close, i.e., when the ratio of ωα/ωh is low. It has been found that when the ratio of ωα/ωh is low,

172

Page 180: Dynamic performance of bridges and vehicles under strong wind

the difference of the optimal variables for the multi-mode-based and those for single-mode-based controls is significant. When the modal frequencies are well separated (with high ωα/ωh values), both the optimal variables and control efficiency approach to those of single-mode-based controls. In this case, the TMDs can be designed based on the single-mode control without significantly scarifying the accuracy compared to that of multi-mode based control.

1.0 1.2 1.4 1.6 1.8

0.04

0.05

0.06

0.07

ζ t

ωα/ωh

Case 1 Case 2

Fig. 7.13 Optimal damping ratio of “three-row” TMDs versus frequency ratio of coupled modes

7.3.5 Effect of Modal Contributions

Different buffeting response contributions among modes actually indicate the energy distributions of modes and can be varied through changing the static force coefficients in the buffeting force terms, Eqs. (7.19) to (7.21). To simulate the relative contributions among modes, the static force coefficient Cm that is related to the contribution of torsion mode is increased numerically with an amplification factor β. Through changing the quantity of buffeting force for torsion mode, the relative contribution to buffeting response among modes can be adjusted. Again, the wind speed is fixed at 30 m/s in the following discussions.

Fig. 7.15 shows the optimal mass distributions of the TMDs versus the amplification factor β. With the increase of the torsion mode contribution due to the increase of β, the optimal mass m1 that is responsible for the control of torsion mode response increases. The mass of TMD for torsion mode (2m1) is the same as that for bending mode (m2) when β is about 7. It can also be

173

Page 181: Dynamic performance of bridges and vehicles under strong wind

found in Fig. 7.15 that when β approaches 20, essentially all the mass should be assigned to side rows in order to control the vibrations from the torsion mode. When β is less than 3, majority of the mass should be assigned to center row in order to control the vibration from the bending mode at the given wind velocity of 30 m/s.

1.0 1.2 1.4 1.6 1.8

28

32

36

40

44

Co

ntro

l effi

cien

cy R

h (%)

ωα/ωh

Case 1 Case 2

Fig. 7.14 Optimal control efficiency of “three-row” TMDs versus frequency ratio of coupled modes

Fig. 7.16 shows the dependency of the optimal frequencies ratio 1 / αω ω and on the factor β. With the increase of the β, ω

2 /ω ωh

2/ωh increases from about 0.982 up to about 0.986 and keeps stable, while ω1/ωα decreases from about 0.969 to 0.966. These variations are essentially insignificant.

Fig. 7.17 shows the optimal damping ratio of the TMDs versus β. Since the damping ratio of the TMDs has relatively insignificant effect on the control efficiency of the whole control system, such a variation range of the optimal damping ratio is not that critical. Fig. 7.18 shows that with the optimal variables of TMDs, the control efficiency increases quickly with the increase of buffeting contribution of the torsion mode, i.e., the increase of β. This implies, for this particular case, that the buffeting response induced by the torsion mode is easier to control compared with that induced by bending mode.

174

Page 182: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 20

0.0

0.2

0.4

0.6

0.8

1.0

2m1/mtotal

Amplification factor β

2m1/m

tota

l

0.0

0.2

0.4

0.6

0.8

1.0

m2/mtotal

m2 /m

total

Fig. 7.15 Optimal mass distribution of TMDs versus amplification factor β (Case 1)

0 5 10 15 20

0.966

0.967

0.968

0.969

ω1/ωα

Amplification factor β

ω1/ω

α

0.982

0.983

0.984

0.985

0.986

ω2/ωh

ω2 /ω

h

Fig. 7.16 Optimal frequencies ratio of TMDs versus amplification factor β (Case 2)

175

Page 183: Dynamic performance of bridges and vehicles under strong wind

0 5 10 15 200.03

0.04

0.05

0.06

0.07

ζ t

Amplification factor β

Case 1 Case 2

Fig. 7.17 Optimal damping ratio of “three-row” TMDs versus amplification factor β

0 5 10 15 2020

30

40

50

60

70

80

Cont

rol e

ffici

ency

Rh (%

)

Amplification factor β

Case 1 Case 2

Fig. 7.18 Optimal control efficiency of “three-row” TMDs versus amplification factor β

176

Page 184: Dynamic performance of bridges and vehicles under strong wind

177

In summary, different modal contributions of the buffeting vibrations has less significant effect on the optimal frequency and damping ratio if the mass of each TMD is fixed. On the other hand, the optimal mass for each row varies with the change of the contribution factor β, meaning that mass allocation among rows depends on the contribution factor β to some extent. Different bridges have different contributions of buffeting response from each individual mode. Since other variables depend less significantly on the contribution factor β, perhaps only the mass allocation should be made specifically for each bridge. Such feature is helpful to build a general control scheme with robust control performance for different bridges.

7.4 Concluding Remarks

The mathematical formulation of the bridge-TMD system is developed and a “three-row” TMD strategy is discussed. In this strategy, conceptually, the center TMDs are mainly to control the vibration from bending modes and the side TMDs are for the vibration from torsion modes. However, in high wind speed when strong modal coupling effect exists, the side TMDs will also suppress the coupling part of the bending mode response. The optimal variables of the TMDs are predicted based on multi-mode coupled vibrations, instead of single-mode-based mode-by-mode analysis. The following conclusions can be drawn through the present study:

1. Wind speed has significant effect on the optimal variables of TMDs, especially when wind speed is high. To efficiently control buffeting vibration over a wide range of wind speeds, an adaptive semi-active TMD control system that can adjust the optimal variables is necessary.

2. The modal frequency ratio between the torsion and bending modes has large effect on the optimal frequencies of the TMDs as well as the mass distribution when the total mass of the TMDs is fixed. When the frequencies of the coupling-prone modes are close, the optimal variables of the TMDs based on multi-mode coupled vibration control are significantly different from those of single-mode-based control. In these cases, a specific design for coupled vibration control should be considered. When the frequencies of the coupling-prone modes are well separated (weakly coupled vibrations), the optimal variables of the TMDs are close to those of single-mode-based control. In this case, a control strategy based on the single-mode vibration can be used in practice.

3. The change of buffeting response contribution from the torsion and bending modes has relatively less significant effect on the optimal frequency and damping ratio of the TMDs, while it has significant effect on the mass distribution among the “three-row” TMDs.

4. The present finding verifies the common assumption that single-mode-based control strategy can be used for bridges with well-separated modal frequencies. However, for coupling-prone bridges with low frequency ratio, the control strategy should be based on the analysis of coupled vibrations. Many modern long-span bridges may fall in this category.

Page 185: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 8. WIND VIBRATION MITIGATION OF LONG-SPAN BRIDGES IN HURRICANES

8.1 Introduction

Despite the massive population growth in the south and southeast along the hurricane coast of the United States, the transportation infrastructure has not increased its capacity accordingly. Long-span bridges are usually the backbones of transportation lines along the coastal areas. When hurricane is approaching, these long-span bridges sometimes have to be closed in order to ensure the safety of the bridge as well as the transportation on them due to excessive wind-induced vibrations, which however greatly reduces the capability of hurricane evacuation through the bridges.

To date, bridge vibration controls in high wind speeds have not been adequately addressed. Most previous control work dealt with the bridge buffeting under moderate wind speeds (Gu et al. 2001), along with some cases of flutter controls in high wind speeds (Phongkumsing et al. 2001). While active control devices may provide satisfactory multi-objective control performance in a full range of wind speeds (Gu et al. 2002), their dependence on external energy supply has hindered their applications to the disaster evacuations. Recently, some aerodynamic controls using flaps were proposed to control flutter instability (Wilde et al. 1999). However, their applicability to buffeting control has not been reported and established.

Spring-Damper–Subsystems (SDSs) is a mechanical model, which includes the spring, damper and mass blocks. Tuned Mass Damper is a typical example of SDSs and many other passive controllers, such as Tuned Liquid Dampers, Tuned Liquid Column Dampers, can also be simply modeled as equivalent SDS. Even vehicles on bridges can also be very roughly treated as a sort of Spring-Damper–Subsystems (SDSs) to the bridge (Guo and Xu 2001; Park et al. 2001; Zaman et al. 1996). The SDS is used here as a general terminology to differentiate with Tuned Mass Dampers (TMDs). The objective of the present study is to investigate the effects of different SDSs (with different vibration frequencies) on the bridge performance during hurricane evacuations and develop a truck-type of movable passive SDS. The passive nature makes the control approach more reliable than the active one, considering the reality that power may not be available during the hurricane disasters. The temporary/movable SDS can be conveniently driven on the existing bridges when necessary, and be removed when it is not needed.

It has been reported that the gust wind speed during hurricanes could be up to 60-80 m/s or more in the United States and other areas (Schroeder et al. 1998; Sharma and Richards 1999). Though the duration may be short, the consequence of such strong winds may be catastrophic for both the safety of the bridge and the safety of traffic on the bridge.

Flutter instability problem is the most critical wind-related issue for long-span bridges. It has been known that hurricane-induced strong winds have much higher turbulence intensity than that of moderate winds. The effects of turbulence on the flutter stability are still controversial (Cai et al. 1999 a, b) and nonlinear aerodynamic analysis by Chen et al. (2000) confirmed that turbulence might destabilize the bridge. Before fully understanding the turbulence effect, raising the flutter instability limit of the bridge to a conservative level, as proposed in the present study, seems to be an appropriate way to maintain the safety of the bridge.

178

Page 186: Dynamic performance of bridges and vehicles under strong wind

For service performance, the main threat to the bridge and vehicles is the excessive acceleration response in the vertical and lateral directions when the bridge is subjected to strong winds. Lateral large acceleration may cause the overturning or loss of control of the vehicles (Baker 1994; Sigbjornsson and Snabjornsson 1998). The excessive acceleration in vertical direction may cause the discomfort problems of drivers and passengers (Irwin 1991). Therefore, effectively reducing the acceleration response of the bridge may maximize the transportation capacity and possibly save lives and properties in hurricane-prone areas.

Irwin (1991) suggested the following control guidelines:

When wind speed U ≤ 13 m/s, peak vertical acceleration should be no more than 0.05g (g = gravity acceleration)

When wind speed U >13 m/s, peak vertical acceleration should be no more than 0.1g

8.2 Equations of Motion of Bridge-SDS System

For the model shown in Fig. 8.1-8.2, assuming that a bridge has a displacement of r(x, t) consisting of h(x) in vertical direction, p(x) in lateral direction, and α(x) in torsion direction, and wind forces consisting of buffeting force fb(x, t) and the aeroelastic self-excited force f ( . Assuming also that a total number of n

s x, r, r)1 modes are included in the analysis and a total number of n2

SDSs are attached to the bridge at location xs (s =1 to n2). For each mode, vibrations in three directions h, p and α are considered. The equations of motion for the bridge-SDS system can be derived as

′′ ′+ + =Mη Cη Sη G (8.1)

where

1

T

1 n 1 n, ... , , ...= ξ ξ γ γη2

2

(8.2)

1 11

2 1 2 2

1 2n n n n

3 unitn n n n

M M

M I× ×

× ×

=

M ; (8.3)

1 1

2 2

1n n

2n n

C 0

0 C×

×

=

C ; (8.4)

1 1

2 2

1n n

2n n

S 0

0 S×

×

=

S ; (8.5)

bQ 0 0=G , ... . (8.6)

179

and where η = vector of the generalized coordinate of the bridge-SDS system; ξ= generalized coordinate of the bridge; γ = coordinate of the vertical motion of SDSs; a superscript “T” represents

Page 187: Dynamic performance of bridges and vehicles under strong wind

transpose of vectors and a superscript prime “’” represents a derivative with respect to time t; U = wind velocity; M, C, and S = mass, damping, and stiffness matrices, respectively; n1 = number of modes of the bridge; n2 = number of SDSs; = unit matrix; G = external force vector from wind buffeting; and = generalized buffeting force. All the components of the matrices can be found in Chapter 7 and are not repeated here.

unitIbQ

SDS

d S

SDS SDS

K

Bridge system

K CS S

1 2 2

C K KC C1 1 2 2 n2 2n

n

Fig. 8.1 Outline of bridge-SDS system (Elevation)

SDS

Cross sectionof the bridge

CK

S

S S

Fig. 8.2 Cross section of the bridge with multiple SDS

180

Page 188: Dynamic performance of bridges and vehicles under strong wind

8.3 Solution of Flutter and Buffeting Response

A carefully designed SDS system can increase the flutter critical wind speed for the combined bridge–SDS system. The homogenous part of Eq. (8.1) is expressed in the state-space format as:

+ =AY BY 0 (8.7)

and the complex conjugate pairs of eigenvalues are obtained as:

2j, j 1 j j j j1i+λ = −ζ ω ± ⋅ − ζ ω j =1 to (n1+n2) (8.8)

where λj and jω jζ = eigenvalue, modal frequency and damping ratio of the jth mode,

respectively; and i is the unit imaginary number( 1i = − ). If the modal damping ratio jζ changes from positive to negative, the corresponding wind speed is identified as the flutter critical wind speed Ucr. Details for the complex eigenvalue analysis procedure can be referred at Chapter 3.

To evaluate the control performance, the control efficiency of flutter is defined with R1 as:

cr1

cr

UR 1 100%U

= − ×

(8.9)

where U a = flutter critical wind speed with and without SDS control, respectively. cr crˆ nd U

To solve the buffeting response, Eq. (8.1) can then be Fourier transformed into a new system as

η =F G (8.10)

where η and G = Fourier transformation of η and G, respectively; the impedance matrix F has the general form of ( ) ( )2

ij ij ij ijF M C Si= −ω + ⋅ω ω + ω , where subscripts i, j = 1 to (n1+n2) and

1i = − .

The mean square of displacements in vertical, lateral and torsion directions is related to the buffeting force spectra

b bi jQ QS as follows:

b bi j

2r i j ij Q Q0

i j

ˆ (x) r (x)r (x) F S F dn∞

σ =∑∑ ∫ *ij

*

(8.11)

where F [ ; . 1ij ij]−= F * 1

ijijF [ ]−= F

The control efficiency of buffeting displacement is defined with R2 as

181

Page 189: Dynamic performance of bridges and vehicles under strong wind

r2

r

ˆR 1 100%

σ= − × σ

(8.12)

where and are the root-mean-square (RMS) of displacement with and without SDS control, respectively.

rσ rσ

8.4 Numerical Example: Humen Bridge-SDS system

To better demonstrate the applicability of the developed procedure, the Humen Suspension Bridge is analyzed. This bridge with a main span of 888m is located in the south of China, where hurricane (typhoon) is a serious threat. The basic data of the bridge are shown in Table 7.1. Four modes (two symmetric and two asymmetric) are listed in the table (Lin and Xiang 1995). Existent analysis has shown that the midpoint of the main span has the largest vibration response and the 1st symmetric vertical bending mode and the 1st symmetric torsion mode are the two modes most prone to couple together. Therefore, these two symmetric modes are the most important modes for buffeting response (asymmetric modes contribute nearly nothing at the mid-span point) and flutter instability. For simplicity and for demonstration purpose, only these two symmetric modes are considered in the evaluation of the SDS control efficiency for bridge flutter instability and buffeting response. The mass of the SDS, expressed as the percentage of the bridge mass, ranges from 1.0 to 1.5% in the present study.

0.0 0.2 0.4 0.6 0.8 1.084

88

92

96

100

104 Gen. Mass Ratio =1.0% =1.25% =1.5%

Ucr = 87 m/s without SDS

Ucr(m

/s)

fSDS(Hz)

Fig. 8.3 Flutter critical wind velocity versus SDS frequency

182

Page 190: Dynamic performance of bridges and vehicles under strong wind

Using the complex eigenvalue modal analysis approach introduced earlier, the flutter instability of the bridge-SDS system was analyzed with varied frequencies of the SDS. It is found in Fig. 8.3 that when the frequency of the SDS is very low (close to zero), the SDS is actually acting as a static mass block on the bridge. Since the SDS is placed on one side of the cross-section at the mid-point of the span, it acts essentially as a static eccentric load when its frequency approaches to zero. In the case of 1% mass ratio, the flutter critical wind velocity can be improved from about 87 m/s (without SDS) to 93 m/s (R1= 7%). Such a result also agrees with the conclusion that eccentric mass can increase the flutter stability of the bridge (Phongkumsing et al. 2001). However, when the SDS frequency becomes quite high (larger than 0.5 Hz), the SDS has no effect on the flutter stability at all as reflected by the flat horizontal line that corresponds to the case without SDS (Ucr = 87 m/s). The same tendency can also be observed in the cases of larger mass ratios (1.25% and 1.5%). These results suggest that the vibration of traditional vehicles may have insignificant effect on the flutter instability of bridges under wind action since the frequency of the vehicles (normally over 1.0 Hz) is relatively too high to affect the bridge stability. Fig .8.3 also indicates that the optimal SDS frequency is about 0.25 Hz. This frequency corresponds to the torsion modal oscillation frequency (modified from the natural frequency by aerodynamic forces).

Fig. 8.4 shows the acceleration peak response under different wind speeds in cases with and without SDS control for the case of 1.25% mass ratio with optimal SDS frequency (0.25 Hz, determined from flutter analysis). As expected, the service criteria (indicated by the horizontal dash lines) can be satisfied in a higher wind speed when SDS is used. More specifically, for the same control criteria, the service wind speed limit is raised from originally 50 m/s (without SDSs) to 61 m/s (with SDSs).

Fig. 8.5 shows the acceleration reduction ratio versus the different SDS frequencies under a wind speed of 60 m/s. It is found that the maximum reduction ratio can be about 28%, 32% and 35% for the SDS mass ratio of 1.0%, 1.25%, and 1.5%, respectively. However, the SDS with high or low frequency has small reduction effect on the buffeting (acceleration) response. This means that, similarly to the flutter instability, a well-designed SDS with its frequency tuned to the optimal one will have an efficient control effect on the vertical accelerations. In this particular example, the SDS should have a frequency around 0.2 Hz for an optimal buffeting control for wind velocity of 60 m/s. In that case, the SDS with a pre-tuned frequency becomes a kind of TMD.

The numerical results discussed above have shown that a well-designed SDS system can raise flutter velocity and reduce buffeting vibrations. Movable SDS control devices are thus specifically designed for extreme situations when a hurricane forecast is made and the control need is identified. The preliminary design concept of a vehicle-type of control approach is introduced as follows.

To avoid the problem of vehicle instability or lateral overturning, the proposed vehicle-type of SDS (Fig. 8.6) has almost the same height as a typical truck. Lever-type design (Gu et al. 2002), which can greatly reduce the required vertical clearance of a typical vertical control device, is adopted here. The mass block uses iron to minimize the required space. Each vehicle-type of SDS has a gross mass of 10,000 kg. The dimensions of a vehicle-type of SDS can be seen in Fig. 8.6 (a). The SDSs can be spread along the bridge as shown in Fig. 8.6(b), which is similar to the case of traditional multiple tuned mass dampers (MTMDs).

183

Page 191: Dynamic performance of bridges and vehicles under strong wind

0 20 40 60 80

0.01

0.1

1 Uncontrolled Controlled

61 m/s50 m/s

Peak

acc

eler

atio

n re

spon

se(g

ravi

ty g

)

Wind speed U (m/s)

Fig. 8.4 Peak acceleration response versus wind velocity

184

Page 192: Dynamic performance of bridges and vehicles under strong wind

0.0 0.2 0.4 0.6 0.8 1.0

0

10

20

30

40 Gen. Mass Ratio =1.0% =1.25% =1.5%

Acc

eler

atio

n re

duct

ion

effic

ienc

y (%

)

fSDS(Hz)

Fig. 8.5 Acceleration reduction ratio versus SDS frequency at U=60 m/s

185

Page 193: Dynamic performance of bridges and vehicles under strong wind

Mass blockm=10,000kg

1.5 m 0.5 m

3 m

0.7 m

1 m

1 m

0.3 m

(a) Vehicle-type SDS controller

Vehicle type SDScontrollers

Wind comingdirection

Traffic flow

Center of span

Central line of bridge

(b) Placement of SDS on bridges

Fig. 8.6 Outline of movable SDSs

As in the case of designing MTMDs, the optimal frequency bandwidth ratio (Bf), the ratio of the central frequency of MTMD series (fav) to the modal frequency of the concerned mode, and the

186

Page 194: Dynamic performance of bridges and vehicles under strong wind

damping ratio are the three main design variables (Gu et al. 2001). Optimal variables can usually be obtained through numerical searching. There are also some approximate design formulas to help attain the optimal frequency bandwidth ratio and central frequency of MTMDs (Kareem and Kline, 1994). The procedure of searching optimal variables is not repeated here and the reader can be referred to (Gu et al. 2001; Kareem and Kline 1994).

The number of SDSs is directly related to the control effect and the investment on control facilities. Using the Humen Bridge as example, Table 8.1 compares the control efficiencies corresponding to one-lane and two-lane SDS placements and various numbers of SDSs. All of the optimal design variables with respect to different numbers of SDSs are listed in the table as well. In the table, Bf highest lowest

av

f ff−

= is the bandwidth ratio of the even frequency distribution of SDSs (fhighest

and flowest are the highest and lowest frequencies among all SDSs, respectively), fav is the average frequency of multiple SDSs and ζt is the damping ratio for all of the SDSs. The control performances are quantified using two variables: the new Usev and the new Ucr. The former is the new service wind velocity limit under which service criteria shown in Fig. 8.4 can be satisfied and the latter is the new critical flutter wind velocity with SDSs. It has been shown in Table 8.1 that the control performance basically changes positively with the increase of the numbers of SDSs and for most cases, significant control effects can be observed.

It is noted that this exploratory research is to investigate an alternative to improve the bridge performance in extreme events for long-span bridges during hurricanes. Placement of movable SDS system on bridge during evacuation will certainly block traffic and is not a perfect solution. However, it is better than otherwise to completely close bridge in evacuation or see the bridge being damaged or collapse. In extreme case, to protect the bridge otherwise from being damaged or failure, the movable SDSs can also be placed on the bridge when the traffic is completely closed. Certainly, some issues as how to fix the movable SDSs on the bridge under strong wind need to be addressed before actual implementation.

8.5 Concluding Remarks

To maximize the service capability and maintain the safety of the bridge itself, a movable/temporary passive control approach has been proposed based on a general formulation of the bridge-SDS system. The effect of vehicles on the dynamic performance of long-span bridges subjected to wind action is then investigated with the Humen Suspension Bridge. The following conclusions can be drawn:

- A well-designed movable vehicle-type of control facility can effectively and conveniently increase the maximum wind velocity limit for bridge service in hurricane evacuations. For the case of 1.0% mass ratio, it reduces the peak acceleration by around 28% and simultaneously increases the flutter critical wind speed by about 14%.

- It is noted that this exploratory research is to investigate an alternative to improve the bridge performance in extreme events for long-span bridges during hurricanes. Placement of movable SDS system on bridge during evacuation will certainly block traffic and is not a perfect solution. However, it is better than otherwise to completely close bridge in evacuation.

187

Page 195: Dynamic performance of bridges and vehicles under strong wind

188

- In extreme case, to protect the bridge otherwise from being damaged or failure, the movable SDSs can also be placed on the bridge when the traffic is completely closed. Some issues for practical implementation certainly need to be further addressed.

Table 8.1 Optimal variables of SDS for Humen Suspension Bridge

(a) One lane placement

Number of SDS (n2)

Gen. Mass

Ratio (%)

fav/fα Bf

ζt New Usev ∗

(m/s)

New Ucr ∗∗

(m/s)

10 0.8 0.94 0.15 0.06 55 92

16 1.25 0.93 0.16 0.05 61 99

20 1.6 0.93 0.16 0.05 68 110

26 2.1 0.93 0.18 0.04 74 115

(b) Two lanes placement

10 0.8 0.95 0.15 0.06 60 95

16 1.25 0.95 0.15 0.06 67 103

20 1.6 0.95 0.16 0.05 73 112

26 2.1 0.94 0.16 0.05 78 117

Note: ∗ Usev without SDS equals to 50 m/s; ∗∗ Ucr without SDS equals to 87 m/s.

Page 196: Dynamic performance of bridges and vehicles under strong wind

CHAPTER 9. CONCLUSIONS AND FURTHER CONSIDERATIONS

9.1 Summary and Conclusions

The dissertation can be roughly classified as three interrelated parts: (a) background knowledge of the problem for a deeper insight of long-span bridge aerodynamics; (b) investigation on the interaction of vehicle-bridge-wind systems; and (c) mitigation of excessive responses of bridges.

(a) Deeper insight of long-span bridge aerodynamics (Chapters 2 and 3)

The major objective of the dissertation is to investigate the dynamic performance of the bridge and transportation under the wind loading. To achieve this goal, some research related to bridge aerodynamics has been conducted first. Since modal coupling is very common in modern bridges under wind loading, knowing the coupling characteristics among modes is extremely important, in order to select only those appropriate modes for coupled aerodynamic analysis and to better understand aerodynamic behavior. With incorporating only those modes which are actually needed in the aerodynamic analysis, the whole process can be greatly simplified. Such simplification makes a big difference in calculation efforts when many vehicles are modeled together with the bridge in the second part of the dissertation. On the other hand, to have knowledge about modal coupling of the bridge under wind action is also desirable. However, there is no quantitative method to assess this coupling effect so far.

In this study, a general modal coupling quantification method is introduced through analytical derivations of coupled multimode buffeting analysis. As a result, a Modal Coupling Factor is proposed to quantitatively assess the modal coupling effect and to select key modes. An approximate method for predicting the coupled multimode buffeting response is also proposed. The main conclusions of this study are:

With the proposed Modal Coupling Factor, modal coupling effect between any two modes can be quantitatively assessed, which will help to better understand the modal coupling behavior of long-span bridges under wind action.

Such an assessment procedure will also help to provide a quantitative guideline in selecting key modes that need to be included in coupled buffeting and flutter analyses.

The proposed approximate method for predicting the coupled multimode buffeting response, derived through a closed-form formula, is with acceptable accuracy compared with the “accurate” approach .

After modal coupling assessment and key mode selection techniques have been developed, they are applied to the hybrid aerodynamic analyses of buffeting and flutter. Most existent works deal with these two analyses separately, with different approaches in the frequency domain. To consider the vehicle moving and interaction with the bridge, analyses in the time domain are necessary. Most existent approaches in the time domain are based on direct finite element modeling of the structures, and are used for buffeting analysis only. A hybrid analysis approach

189

Page 197: Dynamic performance of bridges and vehicles under strong wind

is introduced in the present study to facilitate the following analysis on vehicle-bridge-wind system. In the meantime, such an approach is also helpful to understand the progressive natures of the two phenomena, buffeting and flutter, in a more consistent way and the main conclusions are as follows:

• A hybrid approach based on complex eigenvalue modal analysis and time domain analysis provides a more convenient tool for unified buffeting and flutter analyses in a time domain.

• The proposed approach provides a convenient way to numerically replicate the transition of vibrations from multi-frequency buffeting or pre-flutter free vibration to a single-frequency flutter. This transition has been observed in wind tunnel tests and is commonly accepted as a fact.

• The mechanism of the flutter occurrence and the actual transition process from multi-frequency dominated buffeting to the single-frequency dominated flutter are well illustrated through the numerical results of the Humen Bridge.

(b) Investigation of vehicle-bridge-wind system (Chapters 4 and 5)

With the modal coupling techniques and hybrid aerodynamic analysis approaches developed in the first part, the vehicle-bridge-wind system is systematically studied in the dissertation. This part covers two consecutive topics: interaction analysis of the coupled system and accident assessment of vehicles. A rational prediction of the performance of the vehicle-bridge system under strong winds is of utmost importance to the maximum evacuation efficiency and the safety of vehicles and bridges. Most existent works focus on either wind action on vehicles running on a roadway (not on bridges), wind effect on the bridge without considering vehicles on the bridge, or vehicle-bridge interaction analysis without considering wind effect. A comprehensive vehicle-bridge-wind coupled analysis is very rare.

The present study aims at building a framework for the vehicle-bridge-wind aerodynamic analysis, which will lay a very important foundation for vehicle accident analysis, based on dynamic analysis results, and facilitate the aerodynamic analysis of bridges, considering vehicle-bridge-wind interaction. The framework starts with building a general dynamic-mechanical model of a vehicle-bridge-wind coupled system. After the framework is established, a series of 2-axle four-wheel high-sided vehicles on long-span bridges under strong winds are chosen as a numerical example to demonstrate the methodology. With the mechanical model of the vehicle-bridge system, dynamic performance of vehicles as well as bridges is studied under strong winds.

Based on the dynamic interaction analysis results, an assessment model for vehicle accidents on bridges and on roads under wind action is introduced. All the existent accident analysis models are only for vehicles on roadways, and dynamic vibrations of the vehicles are not considered. The proposed model starts with a full interaction analysis between the bridge and the vehicle, which predicts, in addition to the bridge vibration, the vehicle response in the directions of vertical, rolling and rotation under the wind action and road roughness. Such vehicle and bridge vibration information is carried over to the following accident analysis of the vehicle only. With given accident criteria, the accident driving speed can then be predicted under any wind speed. The main conclusions include:

190

Page 198: Dynamic performance of bridges and vehicles under strong wind

• The proposed accident analysis model can be used to predict the accident-related response. With suggested accident criteria and driving behavior model, the accident risks can be assessed.

• Lowering driving speed is effective in lowering the accident risk only if the wind speed is not extremely high. Setting suitable driving speed limits is important to decrease the likeliness of accident occurrence.

• When wind speed reaches a certain high level, the vehicle should not be on the bridge, no matter what its driving speed. Rational critical wind speed limits should be set to decide when to close the bridge.

• Vehicles on the bridge are more vulnerable to accidents than those on the road. Lower driving speed limits should be set for vehicles on bridges than in the road to avoid accidents when strong wind speed exists.

(c) Mitigation of excessive responses of bridges (Chapters 6-8)

Excessive responses of the bridge usually exist when wind speed is high. Under strong winds, modal coupling effects usually become stronger for long-span bridges. It may result in a significant additional component to the buffeting response of each individual mode, compared with cases of weak modal coupling. In addition to the traditional resonant suppression control mechanism of Tuned Mass Dampers, a more efficient control approach may exist for the coupled buffeting control of bridges in strong wind. Approximated closed-form solutions of coupled buffeting response are first derived for a multi-mode coupled bridge system attached with an arbitrary number of TMDs. This derivation clearly shows the contributions of all components of the response and indicates how TMDs can be designed to control each part. Finally, the applicability of dual-objective control with a passive TMD system, based on the new control approach, is briefly discussed.

After the new features of TMD, control on vibrations are studied. Optimal variables of TMDs in the coupled vibration control are numerically studied. The finding in this part verifies the common assumption that single-mode-based control strategy can be used for bridges with well-separated modal frequencies. However, for coupling-prone bridges with low frequency ratio, the control strategy should be based on the analysis of coupled vibrations. Many modern long-span bridges may fall into this category.

As an exploratory application of vibration control for long-span bridges in a hurricane-prone area, a movable vehicle-type of control facility is designed and its efficiency is studied. Placement of movable control system on bridge during evacuation will certainly block traffic and is not a perfect solution. However, it is better than otherwise to completely close bridge in evacuation. In extreme cases, to protect the bridge from being damaged, the movable control system can also be placed on the bridge when the traffic is completely closed. Some issues for practical implementation certainly need to be further addressed.

9.2 Future Work

The writer believes that the following issues deserve further research:

191

Page 199: Dynamic performance of bridges and vehicles under strong wind

192

• In the dissertation, determinant results are given to assess the risks of vehicle accidents. However, the driving process of vehicles on the road or on the bridge is actually affected by many uncertain factors, so the reliability-based accident analysis is needed and can be carried out in the future work.

• Wind- and vehicle-induced fatigue is a very important topic. All existent works treat vehicle-induced fatigue and wind-induced fatigue separately. Based on the interaction analysis of the vehicle-bridge-wind system, the combined fatigue problem can be rationally predicted, which is a very interesting direction for future research.

• Vehicle performance on roadways under wind loading deserves more study. In some typical highway locations with various curvatures, slopes, and road surface conditions, vehicles may exhibit different accident-related performance. To envisage the transportation scenario of vehicles from several typical locations of the highway to even the whole highway system in one specific area will give the transportation authorities much valuable information. This is another very promising future direction based on the present study.

• The driver-behavior model, which can be adopted in the accident analysis model, should be improved by adopting statistic approaches, since every driver behaves differently. As a very challenging task, it deserves further study.

• In the dissertation, some temporary approaches with Tuned Mass Dampers are proposed to mitigate the vibration of long-span bridges under strong wind. Some more adaptive and economical ways can be studied, based on more advanced materials and mechanical engineering techniques.

Page 200: Dynamic performance of bridges and vehicles under strong wind

REFERENCES Abe, M. and Fujino, Y. (1994). Dynamic characterization of multiple tuned mass dampers and

some design formulas. Earthquake Engineering & Structural Dynamics, 23 (8), 813-835. Abe, M. and Igusa, T. (1995). Tuned Mass Dampers for structures with closely spaced natural

frequencies. Earthquake Engineering & Structural Dynamics, 24 (2), 247-266. Abdel-Rohman, M. and Leipholz, H. H. (1978). Active Control of Flexible Structures, J. of

Structural Division, ASCE, Vol. 104, No. 8 1251-1266. Adikari, R., and Yamaguchi, H. (1994). Adaptive control of nonstationary wind-induced

vibration of tall building, Proc. First World Conf. Structural Control, TA4, 43-52. Agar, T. (1989). Aerodynamic flutter analysis of suspension bridges by a modal technique,

Engineering Structures, 11(2), 75-82.

Amann, O. H., von Karman, T., and Woodruff, G. B. (1941) The failure of Tacoma Narrows Bridge, Federal Works Agency.

Anderson, E.Y. and Pederson, L.(1994). Structural monitoring of the Great Belt East bridges,

Proc. of 3rd symposium on strait crossing, ed. Krokeborg, J., Alesund, Norway, 12-15 June, 207-216.

Baker, C. J. (1986). A simplified analysis of various types of wind-induced road vehicle

accidents, J. of Wind Engineering and Industrial Aerodynamics, 22, 69-85. Baker, C. J. (1987). Measures to control vehicle movement at exposed sites during windy

periods, J. of Wind Engineering and Industrial Aerodynamics, 25, 151-161. Baker, C. J. (1991a). Ground vehicles in high cross winds .1. Steady aerodynamic forces, J. of

Fluids and Structures, 5, 69-90. Baker, C. J. (1991b). Ground vehicles in high cross winds .2. Unsteady aerodynamic forces, J. of

Fluids and Structures, 5, 91-111. Baker, C. J. (1991c). Ground vehicles in high cross winds .3. The interaction of aerodynamic

forces and the vehicle system, J. of Fluids and Structures, 5, 221-241. Baker, C. J. (1994). The quantification of accident risk for road vehicles in cross winds, J. of

Wind Engineering and Industrial Aerodynamics, 52, 93-107. Baker, C. J. (1999). The quantification of accident risk for road vehicles in cross winds, J. of

Wind Engineering and Industrial Aerodynamics, 52, 93-107. Baker, C. J. and Reynolds, S. (1992). Wind-induced accidents of road vehicles, Accident

Analysis & Prevention, 24, 6, 559-575.

193

Page 201: Dynamic performance of bridges and vehicles under strong wind

Blejwas, T. E., Feng, C. C. and Ayre, R. S. (1979). Dynamic interaction of moving vehicles and structures, J. of Sound and Vibration, 67, 513-521.

Boonyapinyo, V., Miyata, T. and Yamada, H. (1999). Advanced aerodynamic analysis of

suspension bridges by state-space approach, J. of Structural Engineering, ASCE, 125 (12), 1357-1366.

Brancaleoni, F. and Diana, G. (1993). The Aerodynamic Design Of The Messina-Straits Bridge. J. of Wind Engineering and Industrial Aerodynamics, 48(2-3), 395-409.

Bucher, C. G. and Lin, Y. K. (1988). Stochastic stability of bridges considering coupled modes,

J. Engineering Mechanics, ASCE, 114(12), 2055-2071. Bucher, C. G. and Lin, Y. K. (1989). Stochastic stability of bridges considering coupled modes:

II., J. of Engineering Mechanics, ASCE 115(2), 384-400. Cai, C. S. (1993). Prediction of long-span response to turbulent wind, Ph. D dissertation, The

University of Maryland. Cai, C. S., and Albrecht, P. (2000). Flutter Derivatives Based Random Parametric Excitation

Aerodynamic Analysis, Computers & Structures, Elsevier Science, Ltd, UK, 75(5), 463-477. Cai, C. S., Albrecht, P., and Bosch, H. R., (1999a). Flutter and Buffeting Analysis: Finite

Element and RPE Solution, J. of Bridge Engineering, ASCE, 4(3), 174-180. Cai, C. S., Albrecht, P., and Bosch, H. R., (1999b). Flutter and Buffeting Analysis: Luling and

Dear Isle Bridges, J. of Bridge Engineering, ASCE, 4(3), 181-188. Cai, C. S. and Chen, S. R. (2004a). Framework of vehicle-bridge-wind dynamic analysis, J. of

Wind Engineering and Industrial Aerodynamics (accepted for publication). Cai, C. S. and Chen, S. R. (2004b). Wind Hazard Mitigation of Long-span Bridges in

Hurricanes, J. of Sound and Vibration (in press). Cao, Y. H., Xiang, H. F. and Zhou, Y. (2000). Simulation of stochastic wind velocity field on

long-span bridges, J. of Engineering Mechanics, ASCE, 1261(1),1-6. Carlson, J. D., and Spencer, B. F., Jr. (1996). Magneto-rheological fluid dampers for semi-active

seismic control, Proc., 3rd Int. Conf. on Motion and Vibration Control, III, 35–40. Chang, C. C., Gu, M. and Tang, K. H. (2003). "Tuned mass dampers for control of suspension

bridges in dual-mode buffeting vibration." J. of Bridge Engineering, ASCE 8(4), 237-240. Chen, L. K. and Ulsoy, A. G. (2001). Identification of a driver steering model, and model

uncertainty, from driving simulator data, J. of Dynamic systems, Measurement, and Control, ASME, 123 (12), 623-629.

194

Page 202: Dynamic performance of bridges and vehicles under strong wind

Chen, S. R. and Cai, C. S. (2001). Modal coupling prediction and application on wind-induced vibration of long-span bridges, Proc. of 5th National workshop on bridge research in progress, Oct. 8-10, 2001, Minneapolis, sponsored by NSF, 189-192.

Chen, S. R. and Cai, C. S. (2002). New control mechanism of Tuned Mass Damper for strong

wind-induced coupled vibration of long-span bridges in hurricane prone area, 15th ASCE Engineering Mechanics Division Conference, June 2-5, Columbia University, New York, CD Proceeding.

Chen, S. R. and Cai, C. S. (2003a). Evolution of wind-induced vibration for long-span bridge-

numerical simulation and discussion, Computer & Structures, 81(21), 2055-2066. Chen, S. R., Cai, C. S., Chang, C. C. and Gu, M. (2004). Modal coupling assessment and

approximated prediction of coupled multimode wind vibration of long-span bridges, J. of Wind Engineering and Industrial Aerodynamics (in press).

Chen, S. R., Cai, C. S., Gu, M. and Chang, C. C. (2003). Optimal Variables of TMDs for Multi-

mode Buffeting Control of Long-span Bridges, Wind & Structures, 6 (5), 387-402. Chen, S. R. and Cai, C. S. (2004a). Accident assessment of vehicles on long-span bridges in

windy environments, J. of Wind Engineering and Industrial Aerodynamics (submitted for review).

Chen, S. R. and Cai, C. S. (2004b). Strong wind-induced coupled vibration and control with

Tuned Mass Damper for long-span bridges, J. of Sound and Vibration (in press). Chen, X., Matsumoto, M. and Kareem, A. (2000). Aerodynamic coupling effects on flutter and

buffeting of bridges, J. of Engineering Mechanics, ASCE, 126(1), 17-26. Chen, X. Z. and Kareem, A. (2001). On the veering of eigenvalue loci and the physics of coupled

flutter of bridges, Proc. 5th Asia-Pacific Conference on Wind Engineering (Kyoto, Japan), 461-464.

Chen, X., Kareem, A. and Haan, F. L. (2000). Nonlinear Aerodynamic Analysis of Bridges under Turbulent Winds: the New Frontier in Bridge Aerodynamics, Proceedings of the International Conference on Advances in Structural Dynamics (ASD 2000), Hong Kong, China, December 2000 475-482. Cheung, Y. K., Au, F. T. K. Zheng, D. Y. and Cheng, Y. S. (1999). Vibration of multi-span non-

uniform bridges under moving vehicles and trains by using modified beam vibration functions, J. of Sound and Vibration, 228, 611-628.

Coleman, S. A. and Baker, C. J. (1990). High sided road vehicles in cross winds, J. of Wind

Engineering and Industrial Aerodynamics, 36, 1383-1392. Conti, E., Grillaud, G., Jacob, J. and Cohen, N. (1996). Wind effects on Normandie cable-stayed

bridge: Comparison between full aeroelastic model tests and quasi-steady analytical approach, J. of Wind Engineering and Industrial Aerodynamics, 65, 189-201.

195

Page 203: Dynamic performance of bridges and vehicles under strong wind

D'Asdia, P. and Sepe, V. (1998). Aeroelastic instability of long-span suspended bridges: a multi-mode approach, J. of Wind Engineering and Industrial Aerodynamics, 74-6, 849-857.

Ding, Q., Lee, P. and Lo, S. H. (2000). Time domain buffeting analysis of suspension bridges

subjected to turbulent wind with effective attack angle, J. of Sound and Vibration, 233 (2), 311-327.

Fischetti, Mark (2001). Drowning New Orleans, Scientific American, October, pp. 76-85 Frandsen, J. B. (2004). Numerical bridge deck studies using finite elements. Part I: flutter,

Journal of Fluids and Structures, 19(2). Fujino, Y. and Abe, M. (1993). “Design formulas for tuned mass dampers based on a

perturbation technique.” Earthquake Engineering. & Structural Dynamics, 22, 833-854. Green, M. F., Cebon, D. and Cole, D. J. (1995). Effects of vehicle suspension design on

dynamics of highway bridges, J. Structural Engineering, ASCE, 121, 272-282. Gu, M., Chang, C. C., Wu, W. and Xiang, H. F. (1998). Increase of critical flutter wind speed of

long-span bridges using tuned mass dampers, J. of Wind Engineering and Industrial Aerodynamics, 73(2), 113-123.

Gu, M., Chen, S. R. and Chang, C. C. (1999). Buffeting control of the Yangpu Bridge using

multiple tuned mass dampers, Proc. 10th International Conf. on Wind Engineering, Larsen, Larose and Livesey (eds), Balkema, Rotterdam, 893-898.

Gu, M., Chen, S. R. and Chang, C. C. (2001). Parametric study on multiple tuned mass dampers

for buffeting control of Yangpu Bridge, J. of Wind Engineering and Industrial Aerodynamics, 89, 11-12, 987-1000.

Gu, M., Chen, S. R. and Chang, C. C. (2002a). Control of wind-induced vibrations of long-span

bridges by semi-active lever-type TMD, J. of Wind Engineering and Industrial Aerodynamics, 90, 2, 111-126.

Gu, M., Chen, S. R. and Chang, C. C. (2002b). Background component of buffeting response of

cable-stayed bridges”, J. of Wind Engineering and Industrial Aerodynamics, 90, 2045-2055. Gu, M., Xiang, H. F. (1992). Optimization of TMD for suppressing buffeting response of long-

span bridges, J. of Wind Engineering and Industrial Aerodynamics, 41-44, 1383-1392. Gu, M., Xiang, H. F. and Chen, A. R. (1994). A practical method of passive TMD for

suppressing wind-induced vertical buffeting of long-span cable-stayed bridges and its application, J. of Wind Engineering and Industrial Aerodynamics, 51, 203-213.

Guo, W. H. and Xu, Y. L. (2001). Fully computerized approach to study cable-stayed bridge-

vehicle interaction, J. of Sound and Vibration, 248(4), 745-761. Holmes, J. D. (2001). Wind loading of structures, Spon Press.

196

Page 204: Dynamic performance of bridges and vehicles under strong wind

Huang, D. Z. and Wang, T. L.(1992). Impact analysis of cable-stayed bridges, Computers &

Structures, 31, 175-183. Hwang, E. S. and Nowak, A. S. (1991). Simulation of dynamics of bridges, J. Structural

Engineering, ASCE, 117, 1413-1434. IDOT, Indiana Department of Transportation, Programs/Events- In Memoriam, (2001),

http://www.state.in.us/dot/programs/inmemoriam/61.html Igusa, T. and Xu, K. (1991). Vibration reduction characteristics of distributed tuned mass

dampers, Proc. 4th International Conf. on Recent Advances in Structural Dynamics, Southampton, USA, 596-605.

Irwin, P. A. (1999), Motion criteria, Technical notes of the Rowan Williams Davies & Irwin Inc.

(RWDI), Ontario, Canada. Jain, A., Jones, N. P. and Scanlan, R. H. (1996). Coupled flutter and buffeting analysis of long-

span bridges, J. Structural Engineering, ASCE, 122(7), 716-725. Jain, A., Jones, N. P., and Scanlan, R. H. (1998), Effect of modal damping on bridge

aeroelasticity, J. of Wind Engineering and Industrial Aerodynamics, 77-78, 421-430. Kareem, A. and Kline, S. (1995). Performance of multiple mass dampers under random loading,

J. Structural Engineering, ASCE, 121(SE2), 349-361. Katsuchi, H., Jones, N. P., Scanlan, R. H. and Akiyama H. (1998a). Multi-mode flutter and

buffeting analysis of the Akashi-Kaikyo Bridge, J. of Wind Engineering and Industrial Aerodynamics, 77&78, 431-441.

Katsuchi, H., Jones, N. P., Scanlan, R. H. and Akiyama, H. (1998b). A study of mode coupling

in flutter and buffeting of the Akashi-Kaikyo Bridge, J. of Structural Mechanics /Earthquake Engineering, JSCE, 598/I-44, I75s-I90s.

Katsuchi, H., Jones, N. P., Scanlan, R.H. and Akiyama, H. (1999). Multimode coupled flutter

and buffeting analysis of the Akashi-Kaikyo Bridge, J. Bridge Engineering, ASCE 125,1, 60-70.

Lin, Y. K. and Yang, J. N. (1983). Multimode bridge response to wind excitation, J. Engineering

Mechanics, ASCE, 109(2), 586-603. Lin, Z. X., Chen, A. R. and Xiang, H. F. (1998). Technical report on the wind tunnel test of

Yichang Suspension Bridge, State key wind tunnel laboratory of China (in Chinese). Lin, Z. X. and Xiang, H. F. (1995). Technical report of state key laboratory of wind tunnel in

China (in Chinese), Full-scale wind tunnel experiment of Humen Bridge.

197

Page 205: Dynamic performance of bridges and vehicles under strong wind

Linda, P. F. and Donald, B. B. (1998). A discussion of modal uncoupling and an approximated closed-form solution for weakly coupled systems with application to acoustics, J. of Acoustics Society of America, Acoustical Society of America, 103(4), 1923-1932.

Mileti, D. (1999). Disasters by Design: A Reassessment of Natural Hazards in the United States,

Joseph Henry Press, Washington D.C. Murakami, S. (2002). Setting the scene: CFD and symposium overview, Wind and Structures, 5

(2-4): 83-88. Miyata, T. and Yamada, H. (1999). New idea on the aero-elastic coupled behavior control of the

super long span bridges. Proc. 2th World Conf. on Structural Control (Kyoto, Japan 1999), 2, 843-850.

Namini, A., Albrecht, P., and Bosch, H. (1992). Finite element-based flutter analysis of cable-

suspended bridges, J. Structural Engineering, ASCE, 118(6), 1509-1526. National Transportation Safety Board, safety recommendation H-03-08, 2003,

http://www.ntsb.gov/Recs/letters/2003/H03_08_09.pdf National Weather Service, Storm data and unusual weather phenomena, 1998,

http://www.crh.noaa.gov/mkx/sd/1998/1198.pdf Olsson M. (1985). Finite element modal co-ordinate analysis of structures subjected to moving

loads, J. of Sound and Vibration, 99, 1-12. Pan, T. C. and Li, J. (2002). Dynamic vehicle element method for transient response of coupled

vehicle-structure systems, J. of Structural Engineering, ASCE, 128, 2, 214-223. Park, K., Heo, S. and Baek, I. (2001). Controller design for improving lateral vehicle dynamic

stability, JSAE review, 22, 481-486. Pesterev, A. V. and Bergman, L. A. (1997). Response of elastic continuum carrying moving

linear oscillator, J. of Engineering Mechanics, ASCE, 123, 8, 878-884. Phongkumsing, S., Wilde K. and Fujino, Y. (2001). Analytical study on flutter suppression by

eccentric mass method on FEM model of long-span suspension bridge, J. of Wind Engineering and Industrial Aerodynamics, 89, 515-534.

Rakha, H., Lucic I., Demarchi, S. H. Setti, J. R. and Aerde, M. V. (2001). Vehicle dynamics

model for predicting maximum truck acceleration levels, J. of Transportation Engineering, ASCE, 127, 5, 418-425.

Scanlan, R. H. (1978). The action of flexible bridges under wind. Part I: Flutter theory, J. Sound

and Vibration, 60(2), 187-199. Scanlan, R. H. (1987). Interpreting aeroelastic models of cable-stayed bridges, J. of Engineering

Mechanics, 113(4), 555-575.

198

Page 206: Dynamic performance of bridges and vehicles under strong wind

Scanlan, R. H. and Jones, N. P. (1990). Aeroelastic analysis of cable-stayed bridges, J. Structural

Engineering, ASCE 116(2), 279-297. Scibo-rylski, A. J. Road vehicle aerodynamics, Pentech Press, London, 1975. Schroeder, J. L., Smith, D. A. and Peterson, R. E. (1998). Variation of turbulence intensities and

integral scales during the passage of a hurricane, J. of Wind Engineering and Industrial Aerodynamics, 77&78, 65-72.

Sharma, R. N. and Richards, P. J. (1999). A re-examination of the characteristics of tropical

cyclone winds, J. of Wind Engineering and Industrial Aerodynamics, 83, 21-33. Shin, M., Bae, S., Lee, J. M., Lee, J., Heo, S. J. and Tak, T. O.(2002). New vehicle dynamics

model of yaw rate estimation, Vehicle System Dynamics Supplement, 37, 96-106. Sigbjornsson, R. and Snabjornsson, J. T. (1998). Probabilistic assessment of wind related

accidents of road vehicles: A reliable approach, J. of Wind Engineering and Industrial Aerodynamics, 74-76, 1079-1090.

Simiu, E. and Scanlan, R. H. (1986). Wind effects on structures – An Introduction to Wind

Engineering, second edition, John Wiley & Sons Publication. Simiu, E. and Scanlan. R. H. (1996). Wind effects on structures – Fundamentals and

Applications to Design, John Wiley & Sons Publication. Soong, T. T. and Grigoriu, M. (1993). Random vibration of mechanical and structural systems.

Prentice-Halll, Inc., Englewood Cliffs, N. J. Takeda, T., Niihara, Y., Ohshio, M., Nakano, R., Kozuma, F., and Ogawa, A. (1998). Vertical

gust response control of long span cable-stayed bridge under cantilever construction by active mass damper, Proc. of the 2nd World Conference on Structural Control, 835-842.

Tanaka, H., Yamamura, N. and Dung, N. N. (1994). Multi-mode flutter analysis and two and

three dimensional model test on bridges with non-analogous modal shapes, J. Structural Mechanics and Earthquake Engineering, Tokyo, Japan, 10(2), 35-46.

Tanida, K., Koike, Y., Mutaguchi, K. and Uni, N. (1991). Development of hybrid active-passive

damper, Active and Passive Damping, 2 PVP-Vol. 211, 21-26. The Iowa Channel news, May 30, 2003,

http://www.theiowachannel.com/news/2239713/detail.html The Winchester Star news, Jan 8, 2003, Winchester, Virginia,

http://www.winchesterstar.com/TheWinchesterStar/030108/Area_accident.asp Timoshenko, S. Young, D. H. and Weaver, W. (1974). Vibration problems in engineering, New

York: Wiley, fourth edition.

199

Page 207: Dynamic performance of bridges and vehicles under strong wind

Thorbek, L. T. and Hansen, S. O. (1998). Coupled buffeting response of suspension bridges, J. of

Wind Engineering and Industrial Aerodynamics, 74-76, 839-847. Wang, T. L., Huang, D. Z. and Shahawy, M. (1996). Dynamic behavior of continuous and

cantilever thin-walled box girder bridges, J. Bridge Engineering, ASCE, 1, 67-75. Wilde, K., Fujino, Y. and Kawakami, T. (1999). Analytical and experimental study on passive

aerodynamic control of flutter of a bridge deck, J. of Wind Engineering and Industrial Aerodynamics, 80, 105-119.

Wolshon, B., Urbina, E. and Levitan, M. (2001). National Review of Hurricane Evacuation

Plans and Policies, Hurricane Center, Louisiana State University. Xu, Y. L., Xia, H. and Yan, Q. S. (2003). Dynamic response of suspension bridge to high wind

and running train, J. of Bridge Engineering, 8 (1), 46-55. Xu, Y. L. and Guo, W. H. (2003). Dynamic analysis of coupled road vehicle and cable-stayed

bridge system under turbulent wind, Engineering Structures, 25, 473-486. Yang, J. N., Wu, J. C. and Hsu, S. Y. (1994). Parametric control of seismic-excited structures,

Proc. First World Conf. Struct. Control, WA 1.13-22. Yang, Y. B. and Yau, J. D. (1997). Vehicle-bridge interaction element for dynamic analysis, J. of

Structural Engineering, ASCE, 123, 1512-1518. Yang, Y. B. and Lin, H. B. (1995). Vehicle-bridge interaction analysis by dynamic condensation

method, J. of Structural Engineering, ASCE, 121, 1636-1643. Yang, Y. B. and Yau, J. D. (1997). Vehicle-bridge interaction element for dynamic analysis, J.

Structural Engineering, ASCE, 123, 1512-1518. Yao, J. T. P. (1972). Concept of structural control, J. of Structural Division, ASCE, 98, 1567-

1574. Zaman, M., Taheri, M. R. and Khanna, A. (1996). Dynamic response of cable-stayed bridges to

moving vehicles using the structural impedance method, Appl. Math. Modeling, 20, 877-889.

200

Page 208: Dynamic performance of bridges and vehicles under strong wind

VITA

Mr. Suren Chen was born in 1973 in Jiangsu Province, China. Before pursuing a doctoral degree at Louisiana State University, starting in August 2001, he was in the doctoral program at Kansas State University. Mr. Chen got his Master of Science and Bachelor of Science degrees from the Department of Bridge Engineering, School of Civil Engineering, Tongji University, China, in 1997 and 1994, respectively.

Mr. Chen has worked as a Graduate Research Assistant at Louisiana State University and Kansas State University since January 2001. Prior to this, he worked as a visiting scholar at Hong Kong University of Science & Technology from January 1998 to July 1998 and January 1999 to January 2000. From September 1994 to March 1997, Mr. Chen worked as a Graduate Research Assistant at Tongji University, China.

Mr. Chen has been involved in researches in several areas, such as wind engineering, structural dynamics, structural control, and vehicle dynamics. He has nearly 30 publications including the ones listed below:

1. S. R. Chen, C. S. Cai, C. C. Chang and M. Gu (2004). "Modal Coupling Assessment and Approximated Prediction of Coupled Multimode Wind Vibration of Long-span Bridges", Journal of Wind Engineering and Industrial Aerodynamics (in press).

2. S. R. Chen, C. S. Cai (2004). "Strong Wind-Induced Coupled Vibration And Control With Tuned Mass Damper For Long-Span Bridges", Journal of Sound and Vibration (in press).

3. S. R. Chen and C. S. Cai (2003). "Evolution of wind-induced vibration for Long-span Bridge- numerical simulation and discussion", Computer and Structures, 81(21), 2055-2066.

4. S. R. Chen, C. S. Cai, M. Gu and C. C. Chang (2003). "Optimal Variables of TMDs for Multi-mode Buffeting Control of Long-span Bridges", Wind & Structures An International Journal, 6(5), 387-402.

5. S. R. Chen, C. S. Cai (2003). “Accident assessment of vehicles on long-span bridges in windy environments”, Journal of Wind Engineering and Industrial Aerodynamics (under review)

6. C. S. Cai, S. R. Chen (2004)."Wind Hazard Mitigation of Long-span Bridges in Hurricanes", Journal of Sound and Vibration (in press).

7. C. S. Cai, S. R. Chen (2004). “Framework of vehicle-bridge-wind dynamic analysis”, Journal of Wind Engineering and Industrial Aerodynamics (accepted).

8. M. Gu, S. R. Chen and C. C. Chang (2002). "Background component of buffeting response of cable-stayed bridges", Journal of Wind Engineering and Industrial Aerodynamics, 90(12-15), 2045-2055.

9. M. Gu, S. R. Chen and C. C. Chang (2002). "Control of wind-induced vibrations of long-span bridges by semi-active lever-type TMD", Journal of Wind Engineering and Industrial Aerodynamics, 90(2), 111-126.

10. M. Gu, S. R. Chen and C. C. Chang (2001). "Parametric study on multiple tuned mass dampers for buffeting control of Yangpu Bridge", Journal of Wind Engineering and Industrial Aerodynamics, 89(11-12), 987-1000.

11. M. Gu, S. R. Chen and H.F. Xiang (1998). "MTMD control of the buffeting for YangPu Bridge", Journal of Vibration Engineering, 11(1), 1-8. (in Chinese)

12. M. Gu, S. R. Chen and H. F. Xiang (1997). "Study on the characteristics of MTMD on controlling buffeting of bridges", Vibration and Shock, 16(1), 1-7. (in Chinese)

13. M. Gu, S. R. Chen and H. F. Xiang (1997). "To include the background response in the buffeting calculation of suspension bridges", Journal of Civil Engineering (CSCE), 30(6), 18-24. (in Chinese)

201

Page 209: Dynamic performance of bridges and vehicles under strong wind

202

14. M. Gu, S. R. Chen and H.F. Xiang (1997). "A simplified method in including the background part of buffeting for cable stayed bridges", Journal of Tongji University, 25(5), 497-501. (in Chinese)

15. M. Gu, S. R. Chen and H.F. Xiang (1997). "Study on the background response of buffeting for cable stayed bridges", Journal of Tongji University, 25(1), 1-6. (in Chinese)

16. S. R. Chen, C. S. Cai (2003). " Multi-objective wind hazard mitigation for long-span bridges in hurricane-prone area", Proc. of 16th ASCE Engineering Mechanics Conference, Seattle, University of Washington, July 16-18, 2003(CD proceeding)

17. S. R. Chen, C. S. Cai (2003). “Dynamic performance of car on the bridge under hurricane-induced strong wind”, Proc. of 11th International Conference of Wind Engineering (presented by Suren), Texas, US, June 2-5, 2003.

18. S. R. Chen, C. S. Cai (2003). “Control of Wind-Induced Coupled Vibration of Long-span Bridges with Tuned Mass Dampers”, Proc. of 11th International Conference of Wind Engineering, Texas, US, June 2-5, 2003.

19. C. S. Cai and S. R. Chen (2003). “Wind hazard mitigation of long-span bridges in Hurricanes”, Proc. of 11th International Conference of Wind Engineering (presented by Suren), Texas, US, June 2-5, 2003.

20. S. R. Chen, C. S. Cai (2002). "New control mechanism of TMD on strong wind-induced vibration control of long span bridges in hurricane-prone area", Proc. of 15th ASCE Engineering Mechanics Conference, New York, Columbia University Jun 2-5, 2002(CD proceeding)

21. S. R. Chen, C. S. Cai (2001). "Modal coupling prediction and application on wind-induced vibration of long-span bridges", Proc. of 5th National workshop on bridge research in progress, Oct. 8-10, 2001, Minneapolis, sponsored by NSF, 189-192.

22. M. Gu, S. R. Chen and C. C. Chang (2001). " Background Component of Buffeting Response of Cable-stayed bridges", J. of Wind Engineering (APCWE-5), Kyoto, Japan, Oct., 89: 273-276, 2001

23. M. Gu, S. R. Chen and C. C. Chang (1999). "Buffeting control of the Yangpu Bridge using multiple tuned mass dampers", in: Proc. 10th International Conf. on Wind Engineering, Larsen, Larose and Livesey (eds), Balkema, Rotterdam, 1999, pp.893-898, 1999.

24. C. S. Cai and S. R. Chen (2004). “Accident assessment of vehicles on long-span bridges under strong wind”, Proc. of 17th ASCE Engineering Mechanics Conference (accepted)

25. C. S. Cai, W. J. Wu and S. R. Chen (2004). “Reduction of cable vibration using MR damper”, Proc. of 17th ASCE Engineering Mechanics Conference (accepted)

26. C. S. Cai, Wenjie Wu, Suren Chen, and George Voyiadjis (2003). “Applications of Smart Materials in Structural Engineering” LTRC Project No. 02-4TIRE, Louisiana Department of Transportation and Development.

27. C. S. Cai and Suren Chen (2002). "Structural Performance Evaluation of An Experimental Low Cost Building System" A Final Report Submitted to David Baird Studio, 2241 Christian St., Baton Rouge, LA 70808

28. C. C. Chang, S. R. Chen and M. Gu (2001). “Rectangular Tuned Liquid Dampers under Off-Axial Rotational Motion”, unpublished report, Hong Kong University of Science and Technology.