Chenq

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 406

Copyright

by

Quan Chen

2008
The Dissertation Committee for Quan Chen Certifies that this is the approved
version of the following dissertation:

Effects of Thermal Loads on Texas Steel Bridges

Committee:

Todd A. Helwig, Supervisor

Karl H. Frank

Michael D. Engelhardt

John L. Tassoulas

Eric B. Becker
Effects of Thermal Loads on Texas Steel Bridges

by

Quan Chen, B.S.; M.S.

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin


August, 2008
Dedication

To my beloved family members


Acknowledgements

I am extremely lucky to be a graduate student under the supervision of Dr. Todd

A. Helwig. He offered me such a wonderful learning experience. As the project director,

he always provided timely and insightful guidance and support. I would also like to thank

my dissertation committee members: Dr. Karl H. Frank, Dr. Michael D. Engelhardt, Dr.

John L. Tassoulas, and Dr. Eric B. Becker. They have always been tremendous sources

of information and ideas. None of this work would be possible without their guidance.

Special thanks are given to Dr. Reagan Herman for her kind advice all the time who used

to work at the University of Houston and is currently working at the John Hopkins

University.

I am grateful to my colleagues including Tanmay Borse at the University of Texas

at Austin, Glenn P. Grisham, Yusef Arikan and Eric Writer at the University of Houston.

Their hard work and sweat made the instrumentation and field monitoring of the bridge

under study possible. I also thank Eric Schell, Dennis Fillip, Blake Stasney, Mike Wason

for their help to the laboratory tests. Special thanks go to Jeff Miller who is the machinist

at the University of Houston for his help to the field monitoring.

Finally, I wish to express my appreciation to my family. They are always with me

and love me at any time. Their support, encouragement, and love made possible the

completion of this dissertation. My wife, Lu Peng, has been supported and encouraged

me strongly and consistently through the entire process by taking extra housework and

enduring some loneliness. Her endless love and support kept me studying.

v
Effects of Thermal Loads on Texas Steel Bridges

Publication No._____________

Quan Chen, Ph.D

The University of Texas at Austin, 2008

Supervisor: Todd A. Helwig

The effects of thermal loads on steel bridges are not well understood. Although

thermal effects are discussed in the AASHTO specifications, the appropriateness of the

recommended thermal gradients is questionable. Thermal effects on the bridges can

impact the design of the steel superstructure, the support bearings, and even the bridge

piers. Previous field monitoring of steel trapezoidal box girder bridges has shown that

thermal stresses on the order of 5 ksi were not uncommon under regular daily thermal

cycles. Stresses induced during annual thermal cycles may be potentially larger than

those during daily thermal cycles. Recent data has shown that the bearings that are to

allow the girders to expand and contract freely due to thermal movements are not

frictionless. Because of the bearing friction, the supporting piers must flex to

accommodate the bridge movements. In curved girder applications, questions have been

raised by designers and contractors regarding the proper orientation of guided bearings.

This research study includes field measurements, laboratory tests and finite

element parametric analyses. The bearings of nine bridges in the Houston area have been

instrumented and monitored for more than a year to measure bearing movements due to

changes in temperature. Instrumentation of the steel girders on one of the Houston


vi
bridges was made utilizing thermocouples and vibrating wire strain gages to measure

temperature distribution and thermal stresses. In addition, strain gages and thermal

couples were applied to the steel girders and concrete bridge deck on a simple twin box

girder bridge located at the Ferguson Structural Engineering Laboratory in Austin, Texas.

The data from the field monitoring and laboratory tests were used to validate a

finite element model. Based on this model, a detailed parametric study was conducted to

investigate the effects of bridge configuration. It is found that under the given weather

conditions, the most critical thermal loads are achieved under the following bridge

configurations: N-S bridge orientation, shorter lengths of the concrete deck overhang,

deeper steel girder webs, thinner concrete decks, and larger spacing between two box

girders. To evaluate the effect of environmental conditions and obtain extreme thermal

loads for design purposes, the most critical configuration of bridge sections was modeled

for thermal analysis with Texas weather data from 1961 to 2005 as the input

environmental conditions. Four cities were considered to bound Texas weather

conditions. Based on the thermal analyses, a 45-year sample data of thermal parameters

were used to describe the temperature field over a section. Extreme value analyses of the

sample data were performed to obtain the relationship between thermal loads and return

periods. The thermal loads with 100-year return period were compared to the ones

suggested by AASHTO.

The thermal loads with 100-year return period were used to investigate structural

response. The effect of bearing orientation and the point of fixity were studied. A rigid

body model was proposed to estimate thermal movements at the ends, which matched

those obtained from field monitoring and finite element analysis. The maximum possible

thermal stresses were also evaluated. Design suggestions are put forward based on the

analysis.
vii
Table of Contents
List of Tables .................................................................................................................... xii

List of Figures .................................................................................................................. xiv

Chapter 1 Introduction .........................................................................................................1


1.1 Overview ...............................................................................................................1
1.2 Heating and Cooling of Bridges ...........................................................................3
1.3 Expansion and Contraction of Bridges .................................................................6
1.3.1 Thermal Strain and Deformation ..............................................................6
1.3.2 Bearing Details..........................................................................................6
1.3.3 Design Assumptions and Problematic Details ........................................12
1.4 Research Objectives and Overview of Study......................................................16

Chapter 2 Background .......................................................................................................18


2.1 Introduction .........................................................................................................18
2.2 Thermal Gradient within a Bridge ......................................................................18
2.3 Thermal Stress ....................................................................................................19
2.3.1 Self-equilibrating Stresses ......................................................................20
2.3.2 Continuity Stresses..................................................................................21
2.4 Importance of Thermal Effects in Bridge Design ...............................................22
2.5 Current Design Specifications ............................................................................23
2.5.1 US Codes ................................................................................................24
2.5.2 Eurocode (CEN, 2003)............................................................................29
2.5.3 Britain (BS5400, 1978) ...........................................................................32
2.6 Literature Review................................................................................................33
2.7 Impact of Background Studies on this Investigation ..........................................44

Chapter 3 Field Monitoring and Laboratory Test ..............................................................46


3.1 Measuring Devices and Data Acquisition System ..............................................46
3.2 Instrumentation ...................................................................................................53
3.2.1 Instrumented Bridges ..............................................................................53
viii
3.2.2 Geometry of Fully Instrumented Bridges ...............................................56
3.2.3 Instrumentation of electronic sensors .....................................................63
3.2.4 Instrumentation of Wax Measuring Devices ..........................................72
3.2.5 Instrumentation of Laser Range Meter ...................................................74

Chapter 4 Finite Element Thermal Analysis ......................................................................76


4.1 Introduction .........................................................................................................76
4.2 Formulation of Thermal Analysis .......................................................................76
4.2.1 Heat Conduction PDE Within Bridges ...................................................76
4.2.2 Boundary Conditions ..............................................................................77
4.2.3 Solar Radiation On A Tilted Sunlit Surface ...........................................81
4.3 Implementation of Thermal Analysis Using ANSYS .........................................85
4.3.1 Three Dimensional Model vs. Two Dimensional Model........................86
4.3.2 Element Types ........................................................................................87
4.3.3 Mesh Density ..........................................................................................89
4.3.4 Time Step ................................................................................................90
4.4 Validation of ANSYS Thermal Model ...............................................................91
4.4.1 Validation with the FSEL Bridge ...........................................................91
4.4.2 Validation with Intercontinental Airport Bridge.....................................96

Chapter 5 Parametric Thermal Analysis ..........................................................................100


5.1 Structural Response to Thermal Loading .........................................................100
5.2 FE Model for Parametric Study ........................................................................106
5.3 Comparison of Two-dimensional and Three-dimensional Models ..................110
5.4 Parametric Study ...............................................................................................114
5.4.1 Bridge Orientation ................................................................................114
5.4.2 Season of the Year ................................................................................120
5.4.3 Overhang Length of Concrete Deck .....................................................126
5.4.4 Thickness of Concrete Deck .................................................................132
5.4.5 Girder Spacing ......................................................................................138
5.4.6 Depth of Steel Web ...............................................................................142
5.5 Vertical Temperature Distribution ....................................................................147
ix
5.5.1 Temperature Distribution During Daily Thermal Cycles .....................148
5.5.2 Proposed Vertical Temperature Gradient .............................................155
5.5.3 Applicability of Proposed Vertical Thermal Gradient ..........................157
5.6 Summary ...........................................................................................................162

Chapter 6 Development of Design Criteria .....................................................................165


6.1 Methodology for Developing Design Criteria ..................................................165
6.2 Extreme Value Analysis ...................................................................................168
6.2.1 Order Statistics and Extreme Value Distribution..................................168
6.2.2 Determination of The Domain of Attraction ........................................170
6.2.3 Parameter Estimation ............................................................................172
6.2.4 Procedure for Extreme Value Analysis.................................................173
6.3 Development of Design Criteria .......................................................................174
6.3.1 Determination of Cumulative Distribution Function ............................176
6.3.2 Calculation of Thermal Parameters ......................................................186
6.4 Effect of Sample Size on Extreme Value Analyses ..........................................188
6.5 Comparison with AASHTO Design Specifications ..........................................189
6.6 Discussion and Summary..................................................................................191

Chapter 7 Finite Element Structural Analysis .................................................................193


7.1 Introduction .......................................................................................................193
7.2 Thermal Loads in Bridge Design ......................................................................195
7.3 Bearing Types ...................................................................................................199
7.4 Finite Element Model of the Bridge .................................................................203
7.4.1 Element Type (ANSYS Manual, 2005) ................................................203
7.4.2 Bearing Models .....................................................................................204
7.4.3 Meshing Density ...................................................................................207
7.5 Validation of ANSYS Model ............................................................................207
7.6 Summary ...........................................................................................................211

Chapter 8 Structural Responses to Thermal Loads ..........................................................213


8.1 Introduction .......................................................................................................213

x
8.2 Basic Bridge Model ..........................................................................................214
8.3 Definition of Quantities Under Consideration ..................................................217
8.4 Fixed Point, Bearing Orientation and Guide Number ......................................218
8.4.1 The Point of Fixity ................................................................................223
8.4.2 Bearing Orientation ...............................................................................226
8.4.3 Number of Guided Bearings .................................................................230
8.5 Estimation of the Stationary Point and Thermal Movements ...........................232
8.5.1 Rigid Body Model.................................................................................233
8.5.2 Simplified Method ................................................................................239
8.5.3 Effect of Pier Flexure ............................................................................240
8.5.4 Effect of Bearing Stiffness ....................................................................247
8.5.5 Effect of Bridge Curvature....................................................................251
8.5.6 Design Shear at Pier Top ......................................................................253
8.6 Thermal Stresses ...............................................................................................256
8.6.1 Effect of Additional Restraints .............................................................257
8.6.2 Thermal Stress With 100-year Return Period .......................................260
8.7 Effect of Errors In Orienting Guides ................................................................266
8.8 Discussion and Summary..................................................................................271

Chapter 9 Summary and Conclusions ..............................................................................274

Appendix A Solar Geometry and Solar Radiation Model ...............................................284

Appendix B Validation of ANSYS Model ......................................................................292

Appendix C Applicability of Proposed Vertical Thermal Gradient ................................318

Appendix D Extreme Value Analysis ..............................................................................338

Reference .........................................................................................................................373

Vita .................................................................................................................................379

xi
List of Tables

Table 2-1 AASHTO LRFD Procedure A Temperature Ranges ........................................26

Table 3-1 List of Bridges Selected.....................................................................................54

Table 4-1 Element Types Used in Thermal Analysis ........................................................87

Table 4-2 Material Properties for Thermal Analysis .........................................................91

Table 6-1 Parameter Estimate for Effective Bridge Temperature ...................................177

Table 6-2 Parameter Estimate for Effective Bridge Temperature ...................................183

Table 6-3 Summary of Parameter Estimates ...................................................................186

Table 6-4 Thermal Parameter Values with Desired Return Period .................................187

Table 6-5 Effect of Sample Size on Extreme Value Analysis .........................................188

Table 6-6 Comparison of Effective Bridge Temperature ................................................189

Table 7-1 Element Types to Be Used ..............................................................................203

Table 7-2 Resistance and Friction Coefficient of Pot Bearings .......................................206

Table 7-3 Material Properties ..........................................................................................208

Table 7-4 Bearing Resistance of Intercontinental Airport Bridge ...................................208

Table 7-5 Comparison between Calculated and Measured Displacements .....................209

Table 7-6 Bearing Shear ..................................................................................................211

Table 8-1 Material Properties for Parametric Structural Analysis...................................215

Table 8-2 Comparison of Maximum Results for the Two Cases.....................................225

Table 8-3 Comparison of Maximum Results for the Three Cases...................................228

Table 8-4 Comparison of Maximum Results for the Two Cases.....................................232

Table 8-5 Summary of Pier Flexure.................................................................................241

Table 8-6 Summary of Bearing Stiffness.........................................................................248

Table 8-7 Summary of Thermal Stresses Under Different Thermal Loads .....................262
xii
Table 8-8 Summary of Finite Element Results ................................................................268

Table D-1 Parameter Estimate for Tv ...............................................................................340


Table D-2 Parameter Estimate for Th ...............................................................................344

Table D-3 Parameter Estimate for Tres,s and Tres,c ............................................................348

Table D-4 Parameter Estimate for T1, T1-T2 and T3-T2 ....................................................353

Table D-5 Parameter Estimate for T1, T2-T1 and T2-T3 ....................................................360

Table D-6 Parameter Estimate for Tres,s and Tres,c ............................................................366

Table D-7 Parameter Estimate for Tweb .........................................................................371

xiii
List of Figures

Figure 1-1 Heating and Cooling of Bridge Structures .........................................................5

Figure 1-2 Multidirectional Pot Bearing ..............................................................................7

Figure 1-3 Fixed Pot Bearing ...............................................................................................8

Figure 1-4 Guided Pot Bearing ............................................................................................9

Figure 1-5 Bearing Layout in five-span curved girder system ..........................................10

Figure 1-6 Thermal behavior under uniform temperature change .....................................10

Figure 1-7 Elastomeric Bearing Pad ..................................................................................11

Figure 1-8 Fabrica Sliding Bearing Pad.............................................................................12

Figure 1-9 Finger Joint.......................................................................................................13

Figure 1-10 Debris in Expansion Joints .............................................................................13

Figure 1-11 Orientation of lateral guides on horizontally curved bridges .........................15

Figure 2-1 TMaxDesign for Steel Girders with Concrete Decks .............................................26

Figure 2-2 TMinDesign for Steel Girders with Concrete Decks .............................................27

Figure 2-3 Positive Vertical Temperature Gradient in Superstructures.............................28

Figure 2-4 Solar Radiation Zones for the United States ....................................................28

Figure 2-5 Vertical temperature difference from Eurocode 1 ...........................................31

Figure 2-6 Vertical temperature distribution for Group 4 section .....................................33

Figure 3-1 Thermocouple ..................................................................................................46

Figure 3-2 CM3 Pyronometer ............................................................................................47

Figure 3-3 Wax Measuring Device ....................................................................................48

Figure 3-4 Example of Grid Drawn on Wax Device .........................................................50

Figure 3-5 Hilti PD32 Laser Range Meter .........................................................................51

Figure 3-6 CR5000 Logger ................................................................................................52

xiv
Figure 3-7 AM416 Multiplexer .........................................................................................53

Figure 3-8 Location of FSEL Bridge .................................................................................55

Figure 3-9 Locations of Bridges Selected ..........................................................................56

Figure 3-10 Cross-Section of Twin Box Girder Bridge.....................................................58

Figure 3-11 Plan View of Instrumented Bridge .................................................................58

Figure 3-12 Placement of Concrete Deck ..........................................................................59

Figure 3-13 Complete Bridge ............................................................................................59

Figure 3-14 Overview of Intercontinental Airport Bridge .................................................60

Figure 3-15 Bottom View of Intercontinental Airport Bridge ...........................................61

Figure 3-16 Plan View of Intercontinental Airport Bridge ................................................61

Figure 3-17 Typical Section of Intercontinental Airport Bridge .......................................62

Figure 3-18 Original Design Plan for Intercontinental Airport Bridge .............................62

Figure 3-19 Bearing Orientation ........................................................................................63

Figure 3-20 Instrumented Section of FSEL Bridge ...........................................................64

Figure 3-21 Foil Strain Gage Layout .................................................................................65

Figure 3-22 Thermocouple Layout ....................................................................................65

Figure 3-23 Instrumented Section for Intercontinental Airport Bridge .............................66

Figure 3-24 Thermocouple Layout for Intercontinental Airport Bridge ...........................67

Figure 3-25 Thermocouple Layout On Exterior Girder of Section 1 ................................68

Figure 3-26 Thermocouple Layout On Central Girder of Section 1 ..................................68

Figure 3-27 Thermocouple Layout On Interior Girder of Section 1 .................................69

Figure 3-28 Thermocouple Layout On Exterior Girder of Section 2 ................................69

Figure 3-29 Thermocouple Layout On Central Girder of Section 2 ..................................70

Figure 3-30 Thermocouple Layout On Interior Girder of Section 2 .................................70

Figure 3-31 Thermocouple Layout On Exterior Girder of Section 3 ................................71


xv
Figure 3-32 Thermocouple Layout On Central Girder of Section 3 ..................................71

Figure 3-33 Thermocouple Layout On Interior Girder of Section 3 .................................72

Figure 3-34 Wax Locations of FSEL Bridge .....................................................................73

Figure 3-35 Wax Locations of Intercontinental Airport Bridge ........................................73

Figure 3-36 Laser Range Meter Assembly ........................................................................74

Figure 3-37 Method for Measuring Pier Flexure ...............................................................75

Figure 4-1 Shadow Due to Rail and Overhang Concrete Deck .........................................84

Figure 4-2 Three-dimensional ANSYS Finite Element Model .........................................86

Figure 4-3 Two-dimensional ANSYS Finite Element Model ...........................................87

Figure 4-4 Hourly Ambient Temperature on Aug. 22, 2006 .............................................92

Figure 4-5 Hourly Solar Radiation on Aug. 22, 2006........................................................93

Figure 4-6 Hourly Wind Speed ..........................................................................................93

Figure 4-7 Comparison at the Location of Thermocouple 01............................................94

Figure 4-8 Comparison at the Location of Thermocouple 02............................................94

Figure 4-9 Sectional Temperature Distribution at 6 AM, 08/22/2006...............................95

Figure 4-10 Sectional Temperature Distribution at 10 AM, 08/22/2006...........................96

Figure 4-11 Hourly Ambient Temperature on July 17, 2006 ............................................97

Figure 4-12 Hourly Solar Radiation on July 17, 2006 .......................................................97

Figure 4-13 Comparison at the Location of Thermocouple 1............................................98

Figure 4-14 Comparison at the Location of Thermocouple 2............................................98

Figure 5-1 Unrestrained Section Subjected to Temperature Field Change .....................100

Figure 5-2 Decomposition of A Temperature Profile ......................................................105

Figure 5-3 2-d Models for Parametric Analyses ..............................................................107

Figure 5-4 Solar Radiation Incident on Horizontal Surfaces ...........................................108

Figure 5-5 Diffuse Solar Radiation Incident on Horizontal Surfaces ..............................108


xvi
Figure 5-6 Ambient Temperature ....................................................................................109

Figure 5-7 Variation of incidence angle of curved bridges .............................................110

Figure 5-8 Longitudinal Variation of Effective Bridge Temperature ..............................111

Figure 5-9 Longitudinal Variation of Vertical Linear Thermal Gradient ........................112

Figure 5-10 Longitudinal Variation of Transverse Linear Thermal Gradient .................112

Figure 5-11 Effect of Orientation on Effective Bridge Temperature ..............................115

Figure 5-12 Effect of Orientation on Vertical Linear Thermal Gradient.........................115

Figure 5-13 Effect of Orientation on Transverse Linear Thermal Gradient ....................117

Figure 5-14 Effect of Orientation on Max Steel Residual Temperature ..........................118

Figure 5-15 Effect of Orientation on Min Steel Residual Temperature ..........................118

Figure 5-16 Effect of Orientation on Max Concrete Residual Temperature ...................119

Figure 5-17 Effect of Orientation on Min Concrete Residual Temperature ....................119

Figure 5-18 Effect of Season on Effective Bridge Temperature .....................................121

Figure 5-19 Effect of Season on Vertical Linear Thermal Gradient ...............................121

Figure 5-20 Effect of Season on Transverse Linear Thermal Gradient ...........................123

Figure 5-21 Effect of Season on Max Steel Residual Temperature.................................124

Figure 5-22 Effect of Season on Min Steel Residual Temperature .................................125

Figure 5-23 Effect of Season on Max Concrete Residual Temperature ..........................125

Figure 5-24 Effect of Season on Min Concrete Residual Temperature ...........................126

Figure 5-25 Effect of Overhang Length on Effective Bridge Temperature.....................127

Figure 5-26 Effect of Overhang Length on Vertical Thermal Gradient ..........................128

Figure 5-27 Effect of Overhang Length on Transverse Thermal Gradient .....................128

Figure 5-28 Effect of Overhang Length on Max Steel Residual Temperature ................130

Figure 5-29 Effect of Overhang Length on Min Steel Residual Temperature ................130

Figure 5-30 Effect of Overhang Length on Max Concrete Residual Temperature .........131
xvii
Figure 5-31 Effect of Overhang Length on Min Concrete Residual Temperature ..........131

Figure 5-32 Effect of Deck Thickness on Effective Bridge Temperature .......................133

Figure 5-33 Effect of Deck Thickness on Vertical Linear Thermal Gradient .................134

Figure 5-34 Effect of Deck Thickness on Transverse Linear Thermal Gradient ............135

Figure 5-35 Effect of Deck Thickness on Max Steel Residual Temperature ..................135

Figure 5-36 Effect of Deck Thickness on Min Steel Residual Temperature ...................136

Figure 5-37 Effect of Deck Thickness on Max Concrete Residual Temperature ............136

Figure 5-38 Effect of Deck Thickness on Min Concrete Residual Temperature ............137

Figure 5-39 Effect of Deck Width on Effective Bridge Temperature .............................138

Figure 5-40 Effect of Deck Width on Vertical Linear Thermal Gradient .......................139

Figure 5-41 Effect of Deck Width on Transverse Linear Thermal Gradient ...................139

Figure 5-42 Effect of Deck Width on Max Steel Residual Temperature ........................140

Figure 5-43 Effect of Deck Width on Min Steel Residual Temperature .........................141

Figure 5-44 Effect of Deck Width on Max Concrete Residual Temperature ..................141

Figure 5-45 Effect of Deck Width on Min Concrete Residual Temperature ...................142

Figure 5-46 Effect on Effective Bridge Temperature ......................................................143

Figure 5-47 Effect on Vertical Linear Temperature Difference ......................................143

Figure 5-48 Effect on Transverse Linear Temperature Difference .................................144

Figure 5-49 Effect on Maximum Steel Residual Temperature ........................................145

Figure 5-50 Effect on Minimum Steel Residual Temperature.........................................146

Figure 5-51 Effect on Maximum Concrete Residual Temperature .................................146

Figure 5-52 Effect on Minimum Concrete Residual Temperature ..................................147

Figure 5-53 Sectional Temperature Distribution at 08:00 ...............................................148

Figure 5-54 Sectional Temperature Distribution at 12:00 ...............................................149

Figure 5-55 Sectional Temperature Distribution at 14:00 ...............................................149


xviii
Figure 5-56 Sectional Temperature Distribution at 18:00 ...............................................150

Figure 5-57 Sectional Temperature Distribution at 22:00 ...............................................150

Figure 5-58 History of Web Average Temperature .........................................................152

Figure 5-59 Average Vertical Temperature Distribution (00:00 ~ 00:06) ......................153

Figure 5-60 Average Vertical Temperature Distribution (00:07 ~ 12:00) ......................153

Figure 5-61 Average Vertical Temperature Distribution (1 pm ~ 6 pm) ........................154

Figure 5-62 Average Vertical Temperature Distribution (7 pm ~ 12 pm) ......................154

Figure 5-63 Proposed Vertical Temperature Distributions ..............................................156

Figure 5-64 Effect of Season on Vertical Thermal Gradient (Spring).............................158

Figure 5-65 Effect of Season on Vertical Thermal Gradient (Summer) ..........................159

Figure 5-66 Effect of Season on Vertical Thermal Gradient (Fall) .................................160

Figure 5-67 Effect of Season on Vertical Thermal Gradient (Winter) ............................161

Figure 6-1 Methodology for Developing Design Thermal Loading ................................167

Figure 6-2 Cities to be Analyzed .....................................................................................175

Figure 6-3 Gumbel Plot of Effective Bridge Temperature for Maxima ..........................177

Figure 6-4 Arithmetic Plot of Effective Bridge Temperature for Maxima ......................178

Figure 6-5 Histogram of Month for Max Effective Bridge Temperature ........................179

Figure 6-6 Histogram of Time for Max Effective Bridge Temperature ..........................179

Figure 6-7 Annual Ambient Temperature........................................................................180

Figure 6-8 Annual Solar Radiation Striking a Horizontal Surface ..................................181

Figure 6-9 Correlation between Maximum Te and maximum Ta.....................................182


Figure 6-10 Gumbel Plot of Minus Te for Maxima .........................................................182

Figure 6-11 Arithmetic Plot of Minus Te for Maxima .....................................................183

Figure 6-12 Histogram of Month for Minimum Te ..........................................................184

Figure 6-13 Histogram of Time for Minimum Effective Bridge Temperature ...............184
xix
Figure 6-14 Correlation between Minimum Te and minimum Ta ....................................185
Figure 6-15 Comparison of Vertical Temperature Gradient for Heating ........................190

Figure 6-16 Proposed Vertical Temperature Gradient for Cooling .................................190

Figure 7-1 Inclusion of Thermal Loads ...........................................................................196

Figure 7-2 Thermal Loads ...............................................................................................199

Figure 7-3 Models of Pot Bearing ...................................................................................201

Figure 7-4 Deformation of Elastomeric Bearing .............................................................202

Figure 7-5 ANSYS Model of Intercontinental Airport Bridge ........................................208

Figure 7-6 Contour of Tangential and Radial Displacement in Deck .............................210

Figure 8-1 Section Sizes for Parametric Structural Analysis...........................................215

Figure 8-2 Chord Oriented Bearing Toward the Fixed Pier ............................................216

Figure 8-3 Vertical Temperature Distribution Over the Depth .......................................217

Figure 8-4 A Simple Curved Girder Subjected to Uniform Temperature .......................220

Figure 8-5 Chord-oriented Bearing Toward Abutment 1 ................................................223

Figure 8-6 Contour Plots when bearings are chord oriented toward bent 2 ....................224

Figure 8-7 Contour Plots when bearings are chord oriented toward abutment 1 ............225

Figure 8-8 Chord-oriented toward stationary point .........................................................227

Figure 8-9 Tangentially Oriented Bearing .......................................................................227

Figure 8-10 Radial Movement when bearings are tangentially oriented .........................228

Figure 8-11 Actual Direction of Thermal Movement at Supports ...................................229

Figure 8-12 Bearing Layout Preferred by AGC ..............................................................231

Figure 8-13 Rigid Body Model ........................................................................................234

Figure 8-14 Orientations of Piers.....................................................................................237

Figure 8-15 Equilibrium Paths when Temperature Drops ...............................................238

Figure 8-16 Stationary Point for Straight Girders............................................................239


xx
Figure 8-17 Parameters for locating Stationary Point......................................................242

Figure 8-18 Effect of Pier Flexure on Stationary Points..................................................245

Figure 8-19 Effect of Pier Flexure on Thermal Movement at Abutment 1 .....................246

Figure 8-20 Effect of Pier Flexure on Thermal Movement at Abutment 4 .....................246

Figure 8-21 Effect of Bearing Stiffness on Stationary Point ...........................................249

Figure 8-22 Effect of Bearing Stiffness on Thermal Movement at Abutment 1 .............249

Figure 8-23 Effect of Bearing Stiffness on Thermal Movement at Abutment 4 .............250

Figure 8-24 Effect of Bridge Curvature on Stationary Point ...........................................251

Figure 8-25 Effect of Bridge Curvature on Thermal Movement at Abutment 1 .............252

Figure 8-26 Effect of Bridge Curvature on Thermal Movement at Abutment 4 .............252

Figure 8-27 Thermal Movements at Abutments 1 and 4 .................................................258

Figure 8-28 Effect of Unintended Restraint on Support Shear ........................................258

Figure 8-29 Effect of Unintended Restraint on Guide Force ...........................................259

Figure 8-30 Effect of Unintended Restraint on Bottom Flange Stress ............................259

Figure 8-31 Vertical Thermal Gradient ...........................................................................262

Figure 8-32 Location of Stationary Points Under Different Thermal Loads ...................265

Figure 8-33 Effect of Different Thermal Loads on Movement At Abutment 1 ..............265

Figure 8-34 Effect of Different Thermal Loads on Movement At Abutment 4 ..............266

Figure 8-35 Example of Misdirected Guides ...................................................................267

Figure 8-36 Definition of Error Direction .......................................................................268

Figure 8-37 Effect of Small Orientation Error on Stationary Point .................................269

Figure 8-38 Effect of Small Orientation Error on Guide Force .......................................269

Figure 8-39 Possible Location of Stationary Points ........................................................270

Figure 9-1 Decomposition of A Temperature Profile ......................................................275

Figure 9-2 Proposed Vertical Temperature Distributions ................................................277


xxi
Figure A-1 Three Basic Solar Angles ..............................................................................286

Figure A-2 Three Derived Solar Angles ..........................................................................286

Figure A-3 All Six Solar Angles ......................................................................................287

Figure A-4 Azimuth Angle of Bridge Surfaces ...............................................................289

Figure A-5 Incident Angle of A Tilted Surface ...............................................................290

Figure B-1 Temperature History at the Location of Thermocouple 3 .............................293

Figure B-2 Temperature History at the Location of Thermocouple 4 .............................293

Figure B-3 Temperature History at the Location of Thermocouple 5 .............................294

Figure B-4 Temperature History at the Location of Thermocouple 6 .............................294

Figure B-5 Temperature History at the Location of Thermocouple 7 .............................295

Figure B-6 Temperature History at the Location of Thermocouple 8 .............................295

Figure B-7 Temperature History at the Location of Thermocouple 9 .............................296

Figure B-8 Temperature History at the Location of Thermocouple 11 ...........................296

Figure B-9 Temperature History at the Location of Thermocouple 12 ...........................297

Figure B-10 Temperature History at the Location of Thermocouple 13 .........................297

Figure B-11 Temperature History at the Location of Thermocouple 14 .........................298

Figure B-12 Temperature History at the Location of Thermocouple 15 .........................298

Figure B-13 Temperature History at the Location of Thermocouple 16 .........................299

Figure B-14 Temperature History at the Location of Thermocouple 17 .........................299

Figure B-15 Temperature History at the Location of Thermocouple 18 .........................300

Figure B-16 Temperature History at the Location of Thermocouple 19 .........................300

Figure B-17 Temperature History at the Location of Thermocouple 23 .........................301

Figure B-18 Temperature History at the Location of Thermocouple 24 .........................301

Figure B-19 Temperature History at the Location of Thermocouple 25 .........................302


xxii
Figure B-20 Temperature History at the Location of Thermocouple 26 .........................302

Figure B-21 Temperature History at the Location of Thermocouple 35 .........................303

Figure B-22 Temperature History at the Location of Thermocouple 36 .........................303

Figure B-23 Temperature History at the Location of Thermocouple 37 .........................304

Figure B-24 Temperature History at the Location of Thermocouple 38 .........................304

Figure B-25 Sectional Temperature Distribution at 0 am ................................................305

Figure B-26 Sectional Temperature Distribution at 2 am ................................................305

Figure B-27 Sectional Temperature Distribution at 4 am ................................................306

Figure B-28 Sectional Temperature Distribution at 8 am ................................................306

Figure B-29 Sectional Temperature Distribution at 12 am ..............................................307

Figure B-30 Sectional Temperature Distribution at 14 am ..............................................307

Figure B-31 Sectional Temperature Distribution at 16 am ..............................................308

Figure B-32 Sectional Temperature Distribution at 18 am ..............................................308

Figure B-33 Sectional Temperature Distribution at 20 am ..............................................309

Figure B-34 Sectional Temperature Distribution at 22 am ..............................................309

Figure B-35 Sectional Temperature Distribution at 24 am ..............................................310

Figure B-36 Temperature History at the Location of Thermocouple 3 ...........................311

Figure B-37 Temperature History at the Location of Thermocouple 4 ...........................311

Figure B-38 Temperature History at the Location of Thermocouple 5 ...........................312

Figure B-39 Temperature History at the Location of Thermocouple 6 ...........................312

Figure B-40 Temperature History at the Location of Thermocouple 7 ...........................313

Figure B-41 Temperature History at the Location of Thermocouple 8 ...........................313

Figure B-42 Temperature History at the Location of Thermocouple 9 ...........................314

Figure B-43 Temperature History at the Location of Thermocouple 11 .........................314

Figure B-44 Temperature History at the Location of Thermocouple 12 .........................315


xxiii
Figure B-45 Temperature History at the Location of Thermocouple 13 .........................315

Figure B-46 Temperature History at the Location of Thermocouple 14 .........................316

Figure B-47 Temperature History at the Location of Thermocouple 15 .........................316

Figure B-48 Temperature History at the Location of Thermocouple 16 .........................317

Figure B-49 Temperature History at the Location of Thermocouple 17 .........................317

Figure B-50 Temperature History at the Location of Thermocouple 18 .........................318

Figure B-51 Temperature History at the Location of Thermocouple 19 .........................318

Figure C-1 Effect of Orientation on Vertical Temperature Distribution (XI=0) .............319

Figure C-2 Effect of Orientation on Vertical Temperature Distribution (XI=30) ...........320

Figure C-3 Effect of Orientation on Vertical Temperature Distribution (XI=60) ...........321

Figure C-4 Effect of Orientation on Vertical Temperature Distribution (XI=90) ...........322

Figure C-5 Effect of Overhang length on Vertical Thermal Gradient (B1=26 in) ...........323
Figure C-6 Effect of Overhang length on Vertical Thermal Gradient (B1=44 in) ...........324

Figure C-7 Effect of Overhang length on Vertical Thermal Gradient (B1=62 in) ...........325

Figure C-8 Effect of Overhang length on Vertical Thermal Gradient (B1=80 in) ...........326

Figure C-9 Effect of The Thickness of Concrete Deck (T1=8 in) ....................................327

Figure C-10 Effect of The Thickness of Concrete Deck (T1=10 in) ................................328
Figure C-11 Effect of The Thickness of Concrete Deck (T1=12 in) ................................329

Figure C-12 Effect of The Thickness of Concrete Deck (T1=14 in) ................................330
Figure C-13 Effect of The Width of Concrete Deck (S=144 in)......................................331

Figure C-14 Effect of The Width of Concrete Deck (S=192 in)......................................332

Figure C-15 Effect of The Width of Concrete Deck (S=240 in)......................................333

Figure C-16 Effect of The Width of Concrete Deck (S=288 in)......................................334

Figure C-17 Effect of The Depth of Steel Girder (D=66 in ) ..........................................335

Figure C-18 Effect of The Depth of Steel Girder (D=88 in ) ..........................................336


xxiv
Figure C-19 Effect of The Depth of Steel Girder (D=110 in ) ........................................337

Figure C-20 Effect of The Depth of Steel Girder (D=132 in ) ........................................338

Figure D-1 Weibull Plot of Tv for Maxima ......................................................................339


Figure D-2 Weibull Plot of Minus Tv for Maxima...........................................................339

Figure D-3 Arithmetic Plot of Tv for Maxima .................................................................340

Figure D-4 Arithmetic Plot of Minus Tv for Maxima ......................................................341

Figure D-5 Histogram of Month for Maximum Tv ..........................................................341

Figure D-6 Histogram of Time for Maximum Tv ............................................................342

Figure D-7 Histogram of Month for Minimum Tv ...........................................................342

Figure D-8 Histogram of Time for Minimum Tv .............................................................343

Figure D-9 Gumbel Plot of Th for Maxima ......................................................................344

Figure D-10 Arithmetic Plot of Th for Maxima ...............................................................345

Figure D-11 Histogram of Month for Maximum Th ........................................................346

Figure D-12 Histogram of Time for Maximum Th ..........................................................346

Figure D-13 Gumbel Plot of Tres,s for Maxima ................................................................347

Figure D-14 Gumbel Plot of Tres,c for Maxima ................................................................347

Figure D-15 Arithmetic Plot of Tres,s for Maxima............................................................348


Figure D-16 Arithmetic Plot of Tres,c for Maxima ...........................................................349

Figure D-17 Histogram of Month for Maximum Tres,s ....................................................350


Figure D-18 Histogram of Time for Maximum Tres,s .......................................................350

Figure D-19 Histogram of Month for Maximum Tres,c ....................................................350

Figure D-20 Histogram of Time for Maximum Tres,c.......................................................351

Figure D-21 Gumbel Plot of T1 for Maxima ....................................................................352

Figure D-22 Gumbel Plot of T1- T2 for Maxima ..............................................................353

Figure D-23 Gumbel Plot of T3- T2 for Maxima ..............................................................353


xxv
Figure D-24 Arithmetic Plot of T1 for Maxima ...............................................................354
Figure D-25 Arithmetic Plot of T1-T2 for Maxima ..........................................................354
Figure D-26 Arithmetic Plot of T3-T2 for Maxima ..........................................................355

Figure D-27 Histogram of Month for Maximum (T1-T2).................................................356

Figure D-28 Histogram of Time for Maximum (T1-T2) ...................................................356

Figure D-29 Histogram of Month for Maximum (T3-T2).................................................357

Figure D-30 Histogram of Time for Maximum (T3-T2) ...................................................357

Figure D-31 Gumbel Plot of T1 for Minima ....................................................................358

Figure D-32 Gumbel Plot of T2- T1 for Maxima ..............................................................359

Figure D-33 Gumbel Plot of T2- T3 for Maxima ..............................................................359

Figure D-34 Weibull Plot of T1 for Minima ....................................................................360

Figure D-35 Arithmetic Plot of T1 for Minima ................................................................361

Figure D-36 Arithmetic Plot of T2-T1 for Maxima ..........................................................361

Figure D-37 Arithmetic Plot of T2-T3 for Maxima ..........................................................362

Figure D-38 Histogram of Month for Maximum (T2-T1).................................................363

Figure D-39 Histogram of Time for Maximum (T2-T1) ...................................................363

Figure D-40 Histogram of Month for Maximum (T2-T3).................................................364


Figure D-41 Histogram of Time for Maximum (T2-T3) ...................................................364

Figure D-42 Gumbel Plot of minus Tres,s for Maxima .....................................................365


Figure D-43 Gumbel Plot of minus Tres,c for Maxima .....................................................366

Figure D-44 Arithmetic Plot of minus Tres,s for Maxima .................................................367

Figure D-45 Arithmetic Plot of minus Tres,c for Maxima .................................................367

Figure D-46 Histogram of Month for Minimum Tres,s .....................................................368

Figure D-47 Histogram of Time for Minimum Tres,s .......................................................368

Figure D-48 Histogram of Month for Minimum Tres,c .....................................................369


xxvi
Figure D-49 Histogram of Time for Minimum Tres,c .......................................................369
Figure D-50 Gumbel Plot of Tweb for Maxima ..............................................................371
Figure D-51 Arithmetic Plot of Tweb for Maxima ..........................................................372

Figure D-52 Histogram of Month for Maximum Tweb ...................................................372

Figure D-53 Histogram of Time for Maximum Tweb .....................................................373

xxvii
Chapter 1 Introduction

1.1 OVERVIEW

Continuous steel girder systems are widely used in highway bridge construction.

The ability to erect the girders with limited or no shoring is a major advantage,

particularly in urban environments. Continuity of the bridge improves the performance

since the design forces and resulting displacements are smaller than would occur in

simply-supported systems. However a difficult aspect of continuous bridge systems is


providing expansion joints and support bearings that can accommodate the displacements

that occur due to daily and annual thermal variations. The complexity dramatically

increases in horizontally curved bridge systems.

The thermal behavior of bridges can be complicated because the bridge exchanges

heat with the surroundings by several means, including radiation, convection and

conduction. No steady-state thermal condition exists within a bridge due to ambient

temperature change and solar radiation [Reynolds and Emanuel (1974), Emerson (1979),

Dilger et al (1981, 1983), Kennedy and Soliman (1986), Fu at el (1990)]. This unsteady-

state condition causes temperature variations within the bridges at any instant of time.

The bridge expands and contracts accordingly since the material expands when heated

and contracts when cooled.

If the bridge is unrestrained at the supports and thermal gradient is linear over the

cross section, only thermal deformations will develop. Although significant efforts are

made to allow bridges to expand and contract freely, thermal deformations are often

restrained at the supports. These support restraints affect the reactions and lead to

internal thermal stresses that develop over the girder cross section. Friction in the

bearings also affects the design forces for substructure elements in the bridge. Nonlinear
1
temperature gradient over the cross-section will produce thermal stresses within any

bridges regardless of whether the bridge is unrestrained at the supports or not.

The thermal response of a bridge is complicated further by the mixture of material

types that are used in bridges such steel-concrete composite girders. Although they are

similar, the coefficients of expansion of steel and a particular concrete mix are not

typically identical. The American Association of State Highway and Transportation

Officials (AASHTO, 2002, 2003, and 2004) stipulates a design coefficient of thermal

expansion of 6.510-6 in/in/F for steel, and a value of 6.010-6 in/in/F for concrete,
which are typical values for these two materials. The fact that the coefficient for

concrete is about eight percent less than that for steel means that there will be differential

expansion in composite girders subjected to uniform temperature changes.

Prior research studies that included field monitoring of steel bridges showed that

thermal stresses on the order of 5 ksi are not uncommon in daily thermal cycles [Helwig

et al (2004), Helwig and Fan (2000), Bobba (2003)]. These daily stress variations are of

the same order as stresses induced by such construction activities as girder erection and

concrete deck placement. Although previous studies have often considered the impact of

daily thermal gradients, very little work has been done to track the behavior during

annual thermal cycles when the deformations can be significant due to large changes in

the ambient temperature. Field measurements presented by Grisham (2005) showed

values of thermal movements at the end supports for a typical three-span continuous steel

bridge as large as 2 inches. Previous investigations have also shown that notable cracks

have occurred in bridges and have seriously affected the serviceability and integrity of

the structure. Leonardt et al. (1965), Zichner (1981), and Massicotte et al. (1994) reported

such cracks in some major bridges and attributed these cracks to the fact that

environmental thermal stresses was inaccurately modeled or completely ignored in the

2
design of these bridges. Results from these previous studies indicate that some of the

thermal stresses may be significant enough for consideration during design.

When thermal expansion and contraction are overestimated, the design

requirements for the bearings and expansion joints will also increase. Accommodating

these movements can significantly increase the cost of bridge since costs increase for

bearings and expansion joints with higher displacement capacities. Underestimating these

movements will result in unconsidered restraints that can produce large internal forces,

which may increase repair costs and reduce the durability of bridge.

Due to the large annual temperature variations, thermal movements must be

considered in bridge design. Longitudinal and lateral movements are accounted for in the

design of bridge bearings. If these movements are inhibited, significant stresses develop

in the girders, concrete deck, and the supporting piers. Bridges are usually designed to

accommodate the movements expected during annual thermal cycles. In practice, uniform

temperature changes are assumed to calculate thermal movements. In actuality, the

temperature experienced by bridges includes vertical, transverse, and longitudinal

thermal gradients; however the impact of these gradients on bridge behavior is normally

not considered in the calculations.

A discussion of the thermal characteristics of a bridge and the environment is

provided in the following section along with an outline of some of the factors that

contribute to poor thermal behavior of steel bridge systems. The chapter concludes with a

section providing an overview of the dissertation.

1.2 HEATING AND COOLING OF BRIDGES

Due to changes in the surrounding environment, bridges continuously lose and

gain heat. The heat transfer occurs through three principal mechanisms: radiation from
3
the sun, convection of heat between the surface and the ambient air, and re-radiation of

the surface to or from the surrounding environment. The intensity of solar radiation

reaching the surface of a bridge is a function of the time of day as well as the time of the

year. The solar radiation reaches the surface of the bridge primarily through direct beam

radiation, diffusive radiation from the sky, and reflected radiation from the surrounding

objects. The incoming radiation reaching the surface may be reflected back or may be

absorbed by the surface and converted to heat. The amount of absorbed energy depends

upon the nature and color of the receiving surface. The surface also loses heat to the

ambient air by convection and re-radiation. Convective heat transfer is a function of the

wind speed and the temperature difference between the surface and the ambient air. The

heat transfer process occurring in a bridge exposed to the environment is illustrated in

Figure 1-1.

During the daytime, bridges undergo significant heating cycles. As the sun rises,

the bridge typically experiences a net gain of heat, which leads to an increase in

temperature throughout the structure. The temperature changes in the bridge can come

from two primary sources: 1) changes in the ambient air temperature and 2) direct

exposure to solar radiation. The heating from changes in the ambient air temperature lead

to uniform changes in the bridge temperature and are generally not that significant during

a given day when compared to the heating from direct solar radiation. Direct solar

radiation tends to heat the bridge very quickly and leads to significant thermal gradients

throughout the bridge. Portions of the bridge that are exposed to direct sunlight undergo

much quicker temperature changes than those that are shaded. In a composite steel girder

bridge, the sun will usually shine directly on parts of the steel and the concrete deck

while others are left shaded. Steel has a relatively high thermal conductivity compared to

concrete. As a result significant temperature gradients will exist within the bridge.

4
Cloud
Diffuse Radiation
Solar Radiation
Convection
Irradiation

Reflected Radiation

Convection

Figure 1-1 Heating and Cooling of Bridge Structures

Once the sun sets, the bridge begins the cooling portion of the daily thermal cycle.

The bridge transfers heat back to the environment. In this situation, the concrete will

tend to hold the heat longer than the steel and again significant thermal gradients may

develop in the bridge cross section. Additional background on heat transfer is provided in

Chapter 2 and detailed formulations are given in Chapter 4.

5
1.3 EXPANSION AND CONTRACTION OF BRIDGES

1.3.1 Thermal Strain and Deformation

The geometrical dimensions of an object are a function of the temperature.

Heating and cooling of the object will result in expansion and contraction, respectively.

The thermal strain, t, that develops as an object is subjected to a change in temperature,


T, can be estimated by Equation 1-1:

t = (T ) (1-1)

where is the coefficient of thermal expansion of material. Thermal strain is

usually designated as positive when it represents expansion and negative when it

represents contraction. Utilizing the definition of engineering strain, the change in length,

t, for a one dimensional uniformly heated free bar can be calculated by Equation 1-2:

t = t L = (T )L (1-2)

where L is the length of the bar. It should be noted that the above equation is

applicable only when the member is able to expand or contract freely in which case no

stresses are produced.

1.3.2 Bearing Details

When calculating the support movements of a bridge for an anticipated change in

temperature, designers will usually employ Equation 1-2 and simply use the length from

a point on the bridge designated as fixed from translation to the point in question. In

using Equation 1-2 for such a calculation, the support conditions on the bridge are
6
idealized as either free or fixed from translation. Translations are usually divided up into

longitudinal and transverse deformations. As the names imply, longitudinal deformations

refer to deformations parallel to the longitudinal axis of the bridge while transverse

deformations are perpendicular to the longitudinal axis. From the standpoint view of

thermal effects, free bearings (translationally free in all directions) are preferred by

designers. However, to resist braking force, acceleration, centripetal force and wind

forces, a bearing guide is usually needed to restrain the displacement in one direction or

in all direction.

Pot
Sole Plate

Bearing Seat
Masonry Plate

Figure 1-2 Multidirectional Pot Bearing

Although many steel bridges in Texas have employed pot bearings, elastomeric

and fabrica sliding bearings have also been successfully employed. Pot bearings typically

consist of an elastomer that allows the girder to rotate on the bearing. The elastomer is

7
contained in the lower part of the bearing that is connected to the pier cap using anchor

bolts to fix the bottom part of the bearing from movement. The upper part of the bearing

consists of a sole plate that is fastened to the bottom flange of the girder. The sole plate

may often require detailed machining to match the super elevation and vertical curve of

the girder at the support in question. A Teflon reaction surface is also often used between

the upper and lower parts of the bearing to minimize friction and allow the girders to

slide on the bearing. Figure 1-2 shows a multidirectional pot bearing in which the bearing

has been designed to permit the bridge girder to translate longitudinally and transversely

in a plane. Key elements of the bearing have been labeled in the figure. The sole plate in

the bearing shown in Figure 1-2 can slide in either the longitudinal or transverse direction

in plane of the bearing surface.

Bridge Girder

Anchorage

Sole Plate

Pot Bearing Seat

Bridge Cap

Figure 1-3 Fixed Pot Bearing (VSL International Ltd, www.vsl.com)

8
Figure 1-3 shows a fixed bearing in which translation between the girder and the

bearing is prevented in both the longitudinal and transverse directions. Although fixed

bearings have translations fully restrained between the upper and lower parts of the

bearings and multidirectional bearings ideally have translations fully free between the

upper and lower parts of the bearing, some bearings must be provided that allow

movement in one direction and restrain the movement in another direction. In these

applications, a unidirectional bearing is necessary as shown Figure 1-4. Unidirectional

bearings are often referred to as guided pots due to the lateral guides that restrict the

movements of the bearing in all but one linear direction.

Underside of Girder

Sole Plate Pot

Guide
Bearing Seat

Figure 1-4 Guided Pot Bearing

In straight bridge systems, the orientation of the lateral guides are obvious since

the guides will be oriented parallel to the longitudinal axis of the bridge. In horizontally

curved bridges, the orientation of the guides is not as clear. Figure 1-5 shows the typical
9
bearing layout that would be used on a bridge consisting of two box girders. The layout

includes fixed bearings, multidirectional bearings, and unidirectional bearings. Common

practice is to orient the guide along a chord of the arc pointing back to the fixed bearing,

which derives from the thermal behavior of the girder system when the bridge is

subjected to a uniform change in temperature.

Fixed Bearing
Multidirectional/Free Bearing
Uni-Directional Bearing
Figure 1-5 Bearing Layout in five-span curved girder system

xx

x
x
(a) (b)

Figure 1-6 Thermal behavior under uniform temperature change

10
This behavior can be understood by considering the thermal behavior of the

circular ring shown Figure 1-6. For a given change in temperature, the diameter of the

ring will either increase (heating) or decrease (cooling). If one point on the ring is fixed

from translation, the other points will simply move along a straight line away or toward

this point. The case shown in Figure 1-6 demonstrates the movement for a positive

temperature change. Therefore, for a uniform change in temperature, the guides should be

oriented along a straight line back towards the fixed bearing.

SteelGirder

SolePlate

ElastomericBearing

BearingSeat

Figure 1-7 Elastomeric Bearing Pad (Bradberry et al, 2005)

In addition to pot bearings, steel bridges may also utilize elastomeric pads such as

the one shown in Figure 1-7. Elastomeric pads provide vertical support while still

allowing transverse movements through shearing of the elastomer as shown by the

movement denoted in the figure.

Another type of pad that is sometimes used on steel bridges is the fabrica pads

shown in Figure 1-8. The girders in the figure have not been plumbed and the picture

11
shows how rigid these pads are as witnessed by the large gap present between the bearing

plate and the pad. Pot bearings and elastomeric pads accommodate girder rotation

through the flexibility in the elastomer. The fabrica are very rigid pads that accommodate

thermal movements by sliding between the bearing surface and a Teflon surface on the

top of the pad.

Fabrica Pad

Bearing Seat

Figure 1-8 Fabrica Sliding Bearing Pad

1.3.3 Design Assumptions and Problematic Details

Several assumptions are made about the thermal boundary conditions during the

design process. These assumptions range from the ability of the bridge to freely expand

and contract to the proper direction for lateral guides to control the thermal movements.

In addition to the bearing pads outlined in the last subsection, expansion joints must be

provided at the ends of continuous girders to isolate the continuous girder system from

adjacent spans or abutments. While these expansion joints isolate adjacent bridge

components, it is important that the detail allows for expansion while not adversely

12
affecting the smooth transition for traffic wheels to roll between adjacent parts of the

structure. One type of joint is the finger joint shown in Figure 1-9. The finger joints allow

expansion and contraction as the fingers of the expansion joint overlap and provide a

longitudinal path that a wheel can roll over.

Figure 1-9 Finger Joint (D.S Brown Company, www.dsbrown.com)

Figure 1-10 Debris in Expansion Joints

13
The expansion joint in Figure 1-10 has a rubber gasket to keep debris from falling

between adjacent bridge segments. However, rocks, dirt, and other debris are often

caught by the rubber gasket and can restrict the joint from closing during hot weather

when the bridge tries to expand. The picture shown in Figure 1-10 was taken less than 6

months after the bridge was put into service.

Another problem that can adversely affect thermal expansion and contraction are

errors in the orientation of the lateral guides on unidirectional pot bearings. Errors in the

design process for one of the bridges instrumented resulted in lateral guides mis-oriented

more than 45 degrees. Such an error can result in the bridge becoming totally constrained

against thermal movements. In addition to gross errors in the guide orientation, proper

tolerances on the orientation of the lateral guides are not clear. In actuality it is unclear

what the proper orientations of the lateral guides are in a curved steel bridge. Although

the chord layout appears to be logical for uniform changes in temperature, significant

restraints can develop in the bearings due to friction that develops between the girders

and the bearings as well as the lateral guides in the upper bearing and the lower bearing.

These restraints cause large shears at the tops of the supporting piers that can cause the

bridges to breath thermally about a stationary point located at somewhere between the

center of mass and the fixed pier. The appropriateness of the assumptions on thermal

boundary conditions in steel bridge design is not clear. In addition, the sensitivity of the

bearings to improper orientation of the lateral guides is not clear and need to be

addressed. With regards to the orientation of the lateral guides, questions have been

raised on the impact of orienting the guides along a tangential layout instead of the chord

layout on horizontally curved bridges. From an economical perspective, the size of the

pier cap for curved girder systems is often dictated by the orientation pot bearing. Figure

14
1-11 demonstrates this for an inverted T-cap with bearings oriented along the chord or a

tangential layout. A comparison of the plan views of the chord layout versus the

tangential layout depicted in Figure 1-11 demonstrates that for a given bearing

dimension, a larger cap would be required for the chord layout.

a) Pier Cap

b) Chord Layout c) Tangential Layout

Figure 1-11 Orientation of lateral guides on horizontally curved bridges

Another area that is not well understood in the design of steel bridges is the

impact of thermal gradients on the stresses and movements in the bridge. The current

thermal design practice typically includes estimates of the amount of thermal expansion

and contraction that occurs throughout an annual thermal cycle. Although thermal

gradients are specified in the AASHTO Specifications (2004), the appropriateness of

these guidelines is not well understood.

15
1.4 RESEARCH OBJECTIVES AND OVERVIEW OF STUDY

This research documents the results of TxDOT (Texas Department of

Transportation) Research Study 0-5040 Effects of Thermal Loads on Texas Steel

Bridges. The purpose of the study was to improve the understanding of the behavior of

steel bridges to daily and annual thermal cycles. The study included field monitoring,

laboratory tests, and parametric finite element investigations. Although this dissertation

provides an overview of the entire study, previous results have been discussed by

Grisham (2005), Arikan (2006), and Writer (2007). The field monitoring included

measurements on a four-span twin box girder bridge as well as eight other steel bridges

located throughout the Houston area. Laboratory tests were conducted on a twin box

girder bridge located at the Ferguson Structural Engineering Laboratory at the University

of Texas at Austin. The field monitoring and laboratory studies included measurements

of thermally induced girder strains, variations in the temperature profile across the bridge,

and movements at the support bearings throughout the year. Results from these

measurements were used to validate the finite element model that was used in the

parametric FEA studies.

The dissertation is divided into nine chapters. Following this introduction,

Chapter 2 provides background and summarizes previous investigations on the thermal

performance of bridges. A description of the geometry of the bridges in the field studies

and laboratory investigations is provided in Chapter 3 along with an overview of the

instrumentation. The finite element thermal model is discussed in Chapter 4 and

comparisons are made with the data from the laboratory studies and field measurements

to validate the modeling techniques. Results from the FEA parametric studies focusing on

the thermal gradients in steel bridges is provided in Chapter 5. Based upon the FEA

investigations, a more accurate model of the thermal gradient is proposed. The thermal
16
loads with desired return periods are developed and compared with those specified in the

AASHTO specification in Chapter 6. In Chapter 7, a finite element structural model is

developed and validated by the data from the laboratory studies and field measurements.

Chapter 8 focuses on the FEA results demonstrating the structural responses of steel

bridge systems as a function of the thermal boundary conditions. Finally the results are

summarized in Chapter 9.

17
Chapter 2 Background

2.1 INTRODUCTION

Chapter 1 provided a general overview of some of the factors that affect the

thermal performance of bridges and demonstrated some of the details that are used to

accommodate thermal movements. An outline of pertinent background information is

provided in this chapter to develop a foundation for understanding the results presented in

later chapters. This overview includes a discussion of previous investigations on


thermal performance in structures as well as an outline of provisions in the AASHTO

design specifications.

2.2 THERMAL GRADIENT WITHIN A BRIDGE

Because of the ever-changing surrounding environment, bridges are almost

continuously in a state of heat transfer. This transfer occurs through three principal

mechanisms: radiation from the sun, convection of heat between the surface and the

ambient air, and re-radiation of the surface to or from the surrounding environment. The

intensity of solar radiation reaching the surface of a bridge varies throughout the day and

also changes as a function of time of year. The solar radiation reaches surfaces of the

bridge primarily through direct beam radiation, diffusive radiation from the sky, and

reflected radiation from the surrounding objects (Duffie and Beckman, 1980). The

incoming radiation reaching the surface may be reflected back or may be absorbed by the

surface and converted to heat. The amount of absorbed energy depends upon the nature

and color of the receiving surface (Hsieh, 1986). Simultaneously, the surface also loses

heat to the ambient air by convection and irradiation. The amount of heat loss depends on

18
the emissivity and convective heat transfer coefficient. The emissivity is a measure of

how effectively a given material radiates absorbed energy. Materials with a high

reflectivity will have a relatively low emissivity while materials that are dark and dull

have an emissivity that approaches unity. The convective heat transfer coefficient is

affected by many factors, such as wind speed, surface roughness, and geometric

configuration of the exposed structure (Emanuel and Hulsey, 1978). Detailed formulation

on heat transfer by each mechanism is given in Chapter 4.

During a daily thermal cycle, the bridge will typically heat up throughout the day

and then cool down throughout the night. The heating of the bridge leads to an increase

in temperature throughout the structure. Depending on the direction of the sun, the top

surface usually becomes warmer than the bottom. Due to the poor thermal conductivity of

concrete, this uneven heating results in temperature gradients throughout the bridge. For

the purposes of a sign convention, a thermal gradient in which the top of the bridge is

warmer than the lower portion of the bridge will be designated as a positive temperature

gradient.

After the sun sets, significant cooling of the bridge occurs. The rate of cooling is

highly dependent on the time of year. With the loss of heat, the bridge temperature

decreases. The top surface becomes cooler than the soffit. Therefore, a negative

temperature gradient is created within the bridge according to the sign convention

described in the preceding paragraph.

2.3 THERMAL STRESS

Depending on the bridge system, thermally-induced stresses can result from either

changes in the uniform temperature or thermal gradients across the section of the bridge.

Changes in the uniform temperature generally only lead to thermally induced stresses in
19
statically-indeterminate structures. However, thermal gradients that are nonlinear over

the cross section will result in the development of thermally-induced stresses in both

statically determinate and indeterminate structures. In statically indeterminate structures,

the deformations due to uniform temperature change are restrained at the supports. This

restraint leads to support reactions and as a result thermal stresses and internal forces

develop in the bridge. Nonlinear thermal gradients induce self-equilibrating stresses.

Thermal stresses that are produced in the longitudinal direction are referred to as

longitudinal stresses. In bridges with closed and wide cross sections such as box girders,

thermal stresses due to temperature change may also be induced in the transverse

direction, and are referred to as transverse stresses. In some cases, the transverse thermal

can be as important as those produced in the longitudinal directions (Imbson et al, 1985).

2.3.1 Self-equilibrating Stresses

When the temperature varies over the cross section of a statically determinate

bridge, the fibers of the cross section attempt to undergo free strain. If the temperature

variation is linear, the free strain will be linear and deformations will occur with no

stresses produced in the cross section. However, in the case of nonlinear variation, the

free thermal strain is no longer linear over the cross section. According to the Bernoulli-

Euler hypothesis that plane sections remain plane after deformation, the self-equilibrating

thermal stresses develop over the cross section to strain the fibers to lie in one plane.

The self-equilibrating stress depends on the shape of the temperature distribution

over the cross section, regardless of the support conditions. In statically determinate

structures, these stresses produce no internal forces at the support since the stress

resultants including both forces and moments are always equal to zero.

20
2.3.2 Continuity Stresses

Sliding pot bearings have typically been used at support locations on steel box

girder bridges in Texas to accommodate the girder movement expected at bearing

locations. The girder movements relative to the abutments/piers at the support locations

will henceforth be referred to as bearing displacements. Discussions of bearing

displacements in this dissertation refer to the displacements in the plane of the bearing.

As discussed in the first chapter, there are three main types of pot bearings: 1)

fixed bearings (where displacements in the plane of the bearing are restrained), 2)

unidirectional or guided bearings (where displacements in a specified direction are

restrained but displacements in an orthogonal direction are unrestricted), and 3)

multidirectional bearings (where displacements are not restrained in any direction).

Figure 1-4 shows a guided bearing and Figure 1-2 shows a multidirectional bearing. The

tops of both pots are covered with a low-friction material.

The actual support conditions are often different than the idealized conditions.

Although efforts are made to allow free movement, friction still develops in the bearings

and provides some restraint to expansion and contraction from thermal changes [Lopez

(1999), Grisham (2005) and Writer (2007)]. In addition, expansion can be further

restrained by debris in the expansion joints. Due to the redundant constraints by the
bearings and expansion joints, a multi-span bridge is generally statically indeterminate.

The change in temperature leads to support reactions and therefore internal forces in its

girders. These forces produce stresses, often referred to as continuity stresses, since they

result from the restraints by the bearings. The magnitude and distribution of the

continuity stresses are dependent on the bridge configuration and bearing conditions.

Although support conditions are often idealized as free and fixed in design, the actual

support conditions are relatively complex.


21
2.4 IMPORTANCE OF THERMAL EFFECTS IN BRIDGE DESIGN

There have been a number of past studies focusing on the performance of bridges

both during construction and in service. A difficult aspect of these investigations has been

isolating stress changes due to construction activities from thermal loads. Three

investigations on steel trapezoidal box girders showed that thermal stresses on the order

of 5 ksi are not uncommon in daily thermal cycles [Helwig et al (2004), Helwig and Fan

(2000), Bobba (2003)], which was of the same order of stresses induced by such
construction activities as girder erection and concrete deck placement. Thermal stresses

on the similar order were also reported in steel-concrete composite bridges by other

researchers [Zuk (1965), Dilger et al (1981, 1983), Soliman and Kennedy (1986), and Fu

et al (1990), Muzumdar (2003)]. These stresses may be even larger over the annual

thermal cycles. Field measurements [Emerson (1976), Glenn (2005), and Writer (2007)]

showed that thermal movement at the end supports might be also significant. Glenn

(2005) and Writer (2007) reported thermal movements as large as 2 inches at the

abutments for a typical three-span continuous steel bridge during annual thermal cycles.

Temperature-related damage to bridges was summarized by Imbsen (1985). Both

thermal stresses and thermal movements can have significant effects on the behavior and

durability of a bridge. The importance of thermal effects on steel-concrete composite

bridge designs was emphasized by many researchers [Zuk (1965), Reynolds and Emanuel

(1974), Emerson (1979), Dilger et al (1981, 1983), Kennedy and Soliman (1986), Fu at el

(1990), Moorty and Roeder (1992), Lopez (1999) and Roeder (2003)].

The majority of steel girder bridges make use of composite action with the

concrete bridge deck. Since movements from thermal changes are a function of the bridge

length, longer spans lead to larger rotations and reactions at the bearings, and larger
22
thermal movements at the ends of the bridge. Expansion joints are often used at the ends

of bridge units to accommodate thermal movements along the length of the bridge.

Overestimating the required thermal movement leads to costly bearings and larger

expansion joints. Erroneous estimates of the required movements can lead to

deterioration of the bridge in these locations. In particular, underestimating the necessary

thermal movements may result in bearing damage that may increase repair costs and

reduce the durability of bridge (Moorty, 1990).

Thermal stresses may be so significant that they have to be added to the stresses

due to dead and live loads in bridge designs. Kennedy and Soliman (1987) pointed out

that thermal stresses might result in magnifying cracking of concrete deck, which can

lead to corrosion of the steel reinforcement, spalling of the concrete, and deterioration of

the concrete by salt-laden water that can seep through the cracks. In addition, bridge

spans are increasing while the weights are becoming less to produce economic designs.

This reduces reserve strength for temperature induced stresses, which makes undesirable

cracking of concrete deck possible since the tensile strength of concrete is very low.

2.5 CURRENT DESIGN SPECIFICATIONS

Provisions from the US specifications are outlined in this section along with a

brief overview of the Eurocode and British Specification for later reference. Additional

information on the specifications in other countries can be found in Imbsen et al (1985)

which includes a review of the specifications from eleven countries.

23
2.5.1 US Codes

The AASHTO specifications contain guidance for consideration of thermal

effects in bridge design. However, some of the code recommendations are difficult to

incorporate in the analysis procedures typically utilized in bridge design. As will be

discussed, both the AASHTO LRFD Bridge Design Specification (2004, 2007) and

AASHTO Guide Specifications for Horizontally Curved Steel Girder Highway Bridges

(2003) provide thermal gradients or temperature differentials through the depth of the

bridge. It is not possible to include such gradients in the line-element analysis packages

typically used by bridge designers. The recommendations for consideration of thermal

effects contained in the Standard Specifications for Highway Bridges (2002), AASHTO

LRFD Bridge Design Specification (2004), AASHTO Guide Specifications for

Horizontally Curved Steel Girder Highway Bridges (2003) are discussed in the following

subsections.

2.5.1.1 AASHTO Standard Specifications for Highway Bridges

The information regarding thermal effects in the AASHTO Standard

Specifications for Highway Bridges (2007) is contained in Section 3.16, which states:

Provision shall be made for stresses or movements resulting from


variations in temperature. The rise and fall in temperature shall be fixed
for the locality in which the structure is to be constructed and shall be
computed from an assumed temperature at the time of erection. Due
consideration shall be given to the lag between air temperature and the
interior temperature of massive concrete members or structures.

For metal structures, the range of bridge temperature should generally be taken as

0 ~ 120 F under moderate climate and -30 ~ 120 F for cold climate (AASHTO, 2007).

24
As indicated above, the standard specification provides a temperature range for

moderate and cold climates to be used in design of metal structures, like steel bridges.

The specification stipulates that provisions should be made for the stresses or movements

generated by the temperature variations in the bridge.

2.5.1.2 AASHTO LRFD Bridge Design Specifications

More detailed information on thermal loading is provided in the AASHTO LRFD

Bridge Design Specification (2004, 2007) than the AASHTO Standard Specifications.

The LRFD Specifications indicate that the design thermal movement associated with a

uniform temperature change can be calculated using either Procedures A or B, which are

discussed in the following subsections.

AASHTO LRFD Design Thermal Movement Procedure A

Procedure A in the LRFD Specification is analogous to the approach

recommended in the AASHTO Standard Specification. The temperature ranges in Table

2-1 address moderate and cold climates, as in the Standard Specification. Section

C3.12.2.1 of the commentary to the LRFD, defines the distinction between moderate and
cold climates based on the number of freezing days per year. If the number of freezing

days is less than 14, the climate is considered to be moderate. Freezing days are defined

as days when the average temperature is less than 32 F.

Using the AASHTO LRFD procedure the change in temperature is defined as the

difference between a lower/upper bound temperature from the table and the temperature

at which the bridge was erected. This temperature range is multiplied by the thermal

coefficient of expansion of the material being designed and by the length of the member

being designed.
25
Table 2-1 AASHTO LRFD Procedure A Temperature Ranges

Climate Steel or Aluminum Concrete Wood

Moderate 0to 120F 10to 80F 10to 75F

Cold -30to 120F 0to 80F 0to 120F

AASHTO LRFD Design Thermal Movement Procedure B

Procedure B utilizes a design temperature range defined through a thermal

contour map of the United States. Different maps are given for concrete decks on

concrete beams and concrete decks on steel beams. The maps provide the maximum

design bride temperature (TMaxDesign) and minimum design bridge temperature (TMinDesign).
The contour maps for steel girder bridges with concrete decks are shown in Figure 2-1

and Figure 2-2.

Figure 2-1 TMaxDesign for Steel Girders with Concrete Decks (AASHTO 2007)

26
Figure 2-2 TMinDesign for Steel Girders with Concrete Decks (AASHTO 2007)

The expressions for the design movements for joints and bearings are based on

the standard Equation 1-2. In the AASHTO LRFD Specification Procedure B, T is

defined based on the difference between TMaxDesign and TMinDesign.

AASHTO LRFD Design Temperature Gradient

The AASHTO LRFD Bridge Design Specification also contains a thermal

gradient, which is shown in Figure 2-3. The gradient varies depending on whether the
beams are concrete or steel. The dashed line represents the thermal gradient for steel

bridges based on the Australian bridge specification (AUSTROADS 1992). In the figure,

dimension A for steel girders is 12 inches and t is the depth of the concrete deck.

Temperature T3 may be taken as zero unless a site specific study of temperatures is

conducted. Temperatures T1 and T2 vary by location, as shown in Figure 2-4.

27
Figure 2-3 Positive Vertical Temperature Gradient in Superstructures (AASHTO 2007)

Figure 2-4 Solar Radiation Zones for the United States (AASHTO 2007)

The solar zones in Figure 2-4 generally run north-south, and are reminiscent of

time zones. The map apparently does not consider latitude to be of importance in

determining the magnitude of the temperature gradient. It is unclear why solar radiation

28
in the United States would vary by longitude and not latitude. A study of thermal contour

map based upon weather data is reported in Chapter 7 of this dissertation.

2.5.1.3 AASHTO Guide Specifications for Horizontally Curved Steel Girder Highway
Bridges

The recommendations for consideration of thermal effects in the AASHTO Guide

Specifications for Horizontally Curved Steel Girder Highway Bridges (2003) follow the

recommendations of the AAHSTO Standard Specification, with the addition of

temperature differential between the deck and girders, which states:

Load effects in the superstructure shall be determined for uniform


temperature changes as specified in AASHTO Article 3.16. A uniform
temperature difference of 25 degrees Fahrenheit between the deck and the
girders shall be considered when the width of the deck is less than one
fifth of the longest span. The load effects due to the temperature
differential shall be added to the effects due to the temperature changes
specified in AASHTO 3.16.

Additional comments regarding orientation of bearing guides are contained in the

commentary of the Guide Specification.

2.5.2 Eurocode (CEN, 2003)

Eurocode 1 requires thermal actions be assessed by the uniform temperature

component and the temperature difference components for a steel-concrete composite

bridge. For special cases where a horizontal temperature difference needs to be

considered a linear temperature difference component may be assumed in the absence of

other information. The maximum value is 9 F if there is no investigation of the actual

site conditions.

29
2.5.2.1 Range of Uniform Temperature Component

The uniform temperature component depends on the minimum and maximum

temperature that the bridge is expected to experience, which is correlated to minimum

and maximum shade air temperature, respectively. Minimum and maximum shade air

temperature for the site represents shade air temperatures for mean sea level in open

country with an annual probability of being exceeded of 0.02, which shall be derived

from national maps of isotherms. When designing bridge piers and expansion joints, the

range of the uniform temperature component should be expanded 20 C upward and

downward. If the bridge temperature during construction is known, the range may be

expanded 10 C.

The setting temperature T0 may be taken as the temperature at the time that the

bridge is restrained. If T0 is not predictable the mean temperature during the construction

period should be taken. If no information is available, T0 may be taken as 10 C.

2.5.2.2 Temperature Difference Component

The vertical temperature difference component should either include the nonlinear

component using the normal procedure or only consider the vertical linear component

using the simplified procedure. The selection of the approach to be used in a given

country may be found in its National Annex.

The simplified method requires that the effect of vertical temperature differences

should be considered by using an equivalent linear temperature difference component

with TM,heat and TM,cool. When the bridge is heating, the top surface is warmer than the

bottom one and TM,heat is taken as 15 C. When the bridge is cooling, the bottom surface

30
is warmer than the top one and TM,cool is taken as 18 C. These values are based on road

and railway bridges with a 50 mm insulation surfacing. For other surface depths of

surfacing these values should be modified accordingly. These values should be applied

between the top and bottom of the bridge deck.

The second approach is labeled as the normal procedure in Figure 2-5 and

requires that the effect of the vertical temperature differences should be considered by

including a nonlinear temperature difference component.

Figure 2-5 Vertical temperature difference from Eurocode 1

2.5.2.3 Others

Eurocode 1 specifies that thermal analysis should be used for important structures.

When uniform temperature components and vertical temperature components are

considered at the same time, several combination cases with different weighted factors
31
for each term are given. The most adverse case should be chosen for design. In structures

where differences in the uniform temperature component between different element types

may cause adverse load effects, a temperature difference of 15 C between main

structural elements should be used to include the effects.

The thermal coefficient of expansion for structural steel is taken as 1210-6

in/in/C (6.6710-6 in/in/F). The thermal coefficient of expansion for concrete is taken as

1010-6 in/in/C (5.5610-6 in/in/F).

2.5.3 Britain (BS5400, 1978)

BS5400 requires that thermal actions should be assessed by the uniform

temperature component and the temperature difference components.

2.5.3.1 Range of Uniform Temperature Component

Maximum and minimum uniform temperatures are determined by maximum and

minimum shade air temperature and the types of bridge cross section using the

correlation expression. Two isotherm maps are used to find the maximum and minimum

shade air temperatures, which have a 120-year return period. The values should be

adjusted according to the sea level and asphalt depth of bridges.

2.5.3.2 Temperature Difference Component

Piecewise linear temperature distributions are assumed to consider the effects of

vertical temperature difference. The shapes of positive and negative temperature

distributions are determined from the cross section type. Figure 2-6 gives the vertical

temperature distribution for steel-concrete composite bridges.

32
Figure 2-6 Vertical temperature distribution for Group 4 section

2.5.3.3 Others

The thermal coefficient of expansion for structural steel is taken as 1210-6

in/in/C (6.6710-6 in/in/F). The thermal coefficient of expansion for concrete is taken as

1010-6 in/in/C (5.5610-6 in/in/F).

2.6 LITERATURE REVIEW

In the past, studies on thermal loading of bridges have been performed. These

investigations include laboratory tests, field investigations and theoretical or numerical

analyses. However, most of these past investigations have focused on the performance of
concrete bridges. Investigations on the thermal performance of steel-concrete composite

bridges are relatively scarce. A thorough literature review on thermal studies on concrete

bridges was given in NCHRP report 276 (Imbsen et al., 1985). The report summarized

bridge design specifications on thermal action from eleven countries. The methods to

quantify thermal gradients and bridge responses to thermal gradients were compared. In

addition, a worked example in the appendix compared moment distributions and extreme

fiber stresses resulting from various thermal gradients. This dissertation focuses primarily

33
on the past investigations related to the thermal action of steel-concrete composite

bridges. Key investigations are reviewed to provide the necessary background for

understanding the importance of results presented later in the dissertation.

Barber (1957)

Barber (1957) proposed a method to calculate the maximum pavement

temperatures from weather reports. This method related the pavement temperatures to

wind speed, precipitation, air temperature and solar radiation. Field measurements were

compared with calculated values and good agreement was achieved.

Zuk (1961, 1965)

Zuk (1961) proposed a method to calculate thermal stresses and deflections in a

statically determinate beam based on rigorous analyses by assuming constant longitudinal

and transverse temperature, uniform temperature through the steel girder and linearly

varying temperature through the concrete slab. Based upon field measurements on four

steel-concrete composite bridges, equations were developed for calculating thermal

stresses.

In 1965, Zuk modified the equation originally proposed by Barber (1957) to

predict the maximum temperature of concrete decks surfaced with a thin topping of

bitumen. To get some insight into the temperature distribution through the depth of the

girder, it was assumed that the temperature of steel girder was uniform and the same as

that of surrounding air temperature. Two equations were presented to determine the

temperature distribution within the concrete deck. One equation was to predict the

maximum temperature differential between the top and bottom concrete deck surfaces of

a normal steel-concrete composite bridge and the other was to predict the temperature
34
distribution through the thickness of concrete deck itself. Field tests were conducted to

verify the predictions. The field measurements matched well with the predicted values

from the proposed equations. The calculated and measured maximum temperatures on the

top of concrete surface were 102 F and 98 F, respectively. The calculated and measured

temperature differentials between the top and bottom concrete deck surfaces were 24 F

and 23 F, respectively. The difference between the calculated and measured thermal

gradients through the thickness of concrete deck was within 10 percent by spot check

comparisons. To obtain thermal stresses and deflections induced by the proposed

nonlinear temperature gradients, Zuk extended his work in 1961 to consider thermal

gradients of any function varying through the depth of the girder.

Berwanger and Symko (1975)

In 1975, Berwanger measured two-dimensional steady-state temperature

distributions of a steel-concrete composite simple span bridge. To achieve a steady-state

thermal condition, the top and bottom surfaces of the bridge were exposed to known

thermal boundary conditions. Temperature and strain variations in the mid-span section

were recorded. The finite element method was used to simulate the bridge conditions to

verify its validity. The calculated temperatures and strains were compared with the

measurements and good agreement was achieved. The calculated temperature could be

predicted with an accuracy of 1 F. The computed strains were slightly higher than the

measured ones.

Hunt and Cooke (1975)

In 1975, Hunt and Cooke proposed a method for calculating temperature

distributions in concrete bridges by modifying the formulation used by Emerson (1973).


35
The two-dimensional sectional problem was reduced to a single dimension by assuming

constant transverse temperature and then the governing partial differential equations were

solved numerically by substituting finite difference approximations for the derivatives. A

significant difference from Emersons formulation is that the Crank-Nicolson implicit

scheme was used, which is unconditionally stable. Another enhancement is that two

layers with different thermal properties were allowed to be used, which made the

consideration of concrete decks with an asphalt topping possible. The calculated

temperature was compared with the measurements taken by Priestly (1972) and good

agreement was achieved. A method was also developed for calculating stress

distributions by solving a two-dimensional plane strain problem in thermoelasticity,

which assumed that the bridge could be treated as an elastic homogeneous isotropic body

and had a relatively large length-thickness ratio. Another assumption behind the

calculation was the uncoupling of longitudinal and transverse stresses.

Emanuel and Hulsey (1978)

In 1978, Emanuel and Hulsey built a two-dimensional model of a section and

investigated the effects of weather data on the sectional temperature distribution within a

steel-concrete composite girder bridge located in Missouri. Based on the analyses of 20

years of weather data, they proposed two equations to predict the maximum and

minimum ambient air temperature during annual thermal cycles. To relate the

temperature distributions to ambient air temperature, they assumed that the temperature

distributions within a bridge reached extremes when the ambient air temperature

achieved extreme values. By doing so, only two daily cycles needed to be considered: (a)

maximum ambient air temperatures and their corresponding solar flux and wind velocity;

(b) minimum ambient air temperatures and their corresponding solar flux and wind
36
velocity. For each cycle, due to rare film coefficient data, the study was considered

bounded by two extreme values: 5.7 W/m2-C and 28.4 W/m2-C. The finite element
method was used to solve the governing partial differential equations for the above four

cases. The annual extreme distributions were developed.

Rahman and George (1979)

In 1979, Rahman and George developed a computer program to calculate thermal

stresses and displacements of continuous bridges with skewed supports. The bridges were

subjected to nonlinear variations of temperature and material properties through the

thickness using finite element formulation. The transverse and longitudinal gradients

were assumed constant. A two-span bridge with skewed supports was analyzed and the

effect of the skew was investigated. The results showed that the skew angle had a

pronounced impact on thermal stresses and deflections.

Dilger et al. (1981, 1983)

In 1981, Dilger et al. instrumented the Muskwa River Bridge located near Fort

Nelson, British Columbia, Canada, which is a continuous five-span two-lane steel-

concrete composite bridge. The deflection, strain, temperature and crack width were

monitored both before and after the placement of concrete deck. Significant findings

were observed:

A temperature difference as large as 73 F between the concrete deck and the

sunlit steel box developed due to side sunshine in less than 2 hours.

Depending on extreme environmental conditions, the difference could be

magnified.

37
A temperature difference as large as 50 F between the sunlit and shaded steel

webs due to side sunshine was measured. These large temperature differences

might imply the possible development of large thermal stresses.

A box-type composite bridge produced a higher temperature difference

between the sunlit and shaded steel webs due to side sunshine because the

enclosed air inside the cell acts as a heat source and warms the web quickly

due to the high thermal mass of steel after the sunlit side was heated up.

The strain essentially followed a linear distribution through the depth and along

the width of the girder.

In 1983, Dilger et al developed a computer program based on a finite difference

formulation to predict the temperature distributions. The program assumed constant

longitudinal temperature and could handle any steel-concrete composite section of

arbitrary geometry and orientation for a given geographic location and environmental

conditions. The calculated temperatures were compared to the measurements recorded

from the Muskwa River Bridge (1981) and a good comparison was achieved. After

validation by the measurements, the program was also used to conduct parametric

analyses to investigate their effect on the thermal behavior of the bridge. The chosen

parameters included season, bridge orientation, length of overhang concrete slab, depth of

steel web, and the slope of steel web. The following conclusions were reached:

The maximum temperature difference between the steel girder and concrete

deck in the spring is achieved when the bridge is oriented 45 degree from due

south. The east-west and the south-north orientations were the most critical

cases in the winter and in the summer, respectively.

38
A shorter length of concrete slab overhang and a larger depth of steel girder

had similar effects. Both parameters led to a higher temperature difference

between the steel girder and concrete deck.

Sloping of the webs in trapezoidal box girders reduces the incidence angle of

sunshine and thereby results in a smaller temperature difference between the

steel girder and concrete deck. The results from the thermal analysis were

chosen as input for a later stress analysis. A simply supported and a continuous

beam with two equal spans were considered to identify the effects of boundary

conditions. The results showed that the continuous beam experienced larger

thermal-induced stresses. The extreme stresses in the steel girder reached as

high as 14.5 ksi in compression and 6.1 ksi in tension. The maximum tensile

stresses in concrete reached as high as 0.42 ksi.

Berwanger (1983)

In 1983, Berwanger conducted a cooling test on a 0.354-scale model of a three-

span steel-concrete composite bridge. The initial temperature of the bridge was 78.5 F.

During the cooling test, the top concrete surface of the bridge was uniformly covered

with ice in two minutes. The test lasted 77 minutes. The sectional temperature gradients,

strains and deflections were monitored at varying time intervals during the period. A two-

dimensional finite element thermal analysis on the measured section was conducted to

predict the temperature distributions. The calculated temperatures matched well with the

measurements. The accuracy achieved in thermal variations predicted by the FEA model

was within 0.5 F from the measurements. A finite element thermo-elastic analysis was

also used to calculate strains, stresses and deflections. Due to water leaking into the

39
concrete slab, there was some swelling of the concrete slab, which resulted in unsuitable

correlation between the measured and computed strains.

Kennedy and Soliman (1986, 1987)

In 1987, Kennedy and Soliman proposed a vertical temperature distribution for a

steel-concrete composite girder bridge based on a synthesis of several theoretical and

experimental studies on prototype bridges. The temperature through the thickness of the

concrete deck was assumed to be linear while the temperature was uniform through the

depth of the steel girder. The longitudinal and transverse temperatures were assumed to

be constant. To calculate thermal stresses, four combination cases were considered:

Heating in winter

Cooling in winter

Heating in summer

Cooling in summer

The maximum and minimum temperature differentials were also suggested.

In 1986, Soliman and Kennedy simplified the equation originally proposed by

Zuk (1961, 1965) by assuming a temperature distribution based on their synthesis of

several theoretical and experimental studies. Simple formulas were derived for

calculating thermal stresses in simply supported and straight continuous beams. The

proposed method was straightforward and suitable for design purposes. An example was

worked out to illustrate its application. The results indicated that significant thermal

stress could be developed in bridges.

Fu et al. (1990)

40
In 1990, Fu et al. developed a two-dimensional analytical thermal model and

conducted parametric analyses on steel-concrete composite bridges. Three types of

composite girder systems were considered: I-girder section, single box-girder section and

double box-girder section. The thermal model was solved using the computer program

ADINA-T. The effects of wind speed, bridge orientation, the ratio of the slab overhang

length to the girder depth, and ambient air temperature were investigated. They found

that the ratio of the slab overhang length to the girder depth was the most influential

factor on the temperature distribution within a steel-concrete composite bridge. The

entrapped air inside the box girder, ambient air temperature, and wind speed had a

marked impact on the temperature distribution. The influence of initial temperature was

minor. Following thermal analyses, thermo-elastic analyses were also conducted to

compute thermal stresses using SAP-IV. They concluded that thermal stresses were

correlated to thermal gradients. Greater thermal gradient tended to induce higher thermal

stresses.

Moorty and Roeder (1992)

Moorty and Roeder conducted two-dimensional heat conduction analyses by

assuming constant longitudinal temperature to predict temperature distribution within

steel-concrete composite bridges. They also reduced the two-dimensional models to

single dimensional models by further assuming constant transverse temperature. The

models were verified by a field test. The results indicated that a single-dimensional

analysis was generally adequate for calculating temperature distribution in composite

bridge systems. Only when steel webs were exposed to direct sunshine were two-

dimensional analyses needed. The effective bridge temperature range was also examined

41
and it was found that the temperature range suggested by AASHTO were too large for

composite bridges in some locations.

Besides thermal analyses, they performed structural analyses to obtain thermal

movements and the associated thermal stresses. A parametric analysis was conducted.

Bridge geometry, bridge type, temperature distribution and support conditions were

examined. The following important conclusions were reached:

The method recommended by AASHTO was only reasonable for orthogonal

straight bridges;

Transverse movements were significant for sharply skewed bridges;

Multidirectional guides should be used in both curved bridges and sharply

skewed bridges due to marked transverse thermal movements;

Chord-oriented bearing was only applicable when the fixed point is at a rigid

support.

It should be noted that the thermal movements and thermal stresses were purely

analytical and not verified by experimental field data.

Tong et al. (2001)

Based on finite element method, Tong et al. developed a numerical model for the

prediction of the temperature distribution of steel bridges with steel decks. The calculated

temperatures were compared with the measurements taken from scaled models and good

agreement was achieved. A method was proposed to determine the input parameters, such

as film coefficient, absorptivity and emissivity, when they are not available. These

parameters are affected by many factors and are often not clear. By investigating the

effects of these parameters on thermal functions such as effective bridge temperature and

42
thermal moment, values were chosen so that errors between the calculated and measured

thermal functions are minimized.

Roeder (2003)

Roeder improved the design procedure for thermal movements based on the

analysis on more than sixty continuous years of weather data. The methodology was

similar to that used by Emerson (1979). Both of them related effective bridge temperature

to climatic conditions for different bridge types and found the relationship between them.

The extreme effective bridge temperature was calculated by inputting the extreme

temperature of shade air that was based on the analysis on climatic data. Different from

Emersons method was that the Kuppa method was used to estimate the extreme ambient

temperature since this method was proved more accurate according to his comparison

research. Based on his analyses on the climatic data of all 50 states of the US, isothermal

maps were plotted for steel-concrete composite and concrete bridges, respectively. He

also examined the probability distribution of the effective bridge temperatures and found

that the effective bridge temperatures were always skewed toward the higher end of the

temperature range. Based on this finding, the strategies were proposed to determine the

optimized installation temperature and design movements for bearings and expansion

joints from statistical standpoint.

Tindal (2003)

Based on the results of literature review, Tindal (2003) proposed a simple thermal

loading for winter and summer conditions for solar zone 3. This load was used to

examine the structural responses of composite skewed steel highway bridges. Parametric

analyses were conducted to examine the relative influence of each parameter. The
43
parameters that were investigated include span length, section depths, bridge width and

skew angle. Based on the analyses, some design suggestions were given. A design

example was used to demonstrate the applicability of the suggestions.

Ni et al. (2007)

Ni et al. (2007) assessed bridge expansion joints based on the one-year field

monitoring data obtained from the instrumentation of the cable-stayed Ting Kau Bridge

in Hong Kong. This long period field data were used to build a statistical model based on

the theory of extreme value analysis. The temperature-displacement pattern was checked.

The extreme temperature and cumulative movement at the joints were predicted and

verified by the field data. They found that the movements at expansion joints are highly

correlated with the effective temperature by a linear relationship.

2.7 IMPACT OF BACKGROUND STUDIES ON THIS INVESTIGATION

The material presented in this chapter provides a valuable starting point for the

research outlined in this dissertation. This investigation includes field measurements,

laboratory tests and finite element parametric analyses. The bearings of nine bridges in
the Houston area have been instrumented and monitored for more than a year to measure

bearing movements due to changes in temperature. A detailed instrumentation of the

steel girders on one of the Houston bridges was made utilizing thermocouples and

vibrating wire strain gages to measure temperature distribution and thermal stresses. In

addition, strain gages and thermal couples were applied to the steel girders and concrete

bridge deck on a simple twin box girder bridge located at Ferguson Structural

Engineering Laboratory.

44
The data from the field monitoring and laboratory tests were used to validate a

finite element model. Based on this model, a detailed parametric study was conducted to

investigate the effects of bridge configuration. It is found that under the given weather

condition, the most critical thermal loads are achieved under the following bridge

configurations: N-S bridge orientation, shorter lengths of the concrete deck overhang,

deeper steel girder webs, thinner concrete decks, and wider girder spacing. To eliminate

the effect of environmental condition and get extreme thermal loads for design purposes,

the most critical configuration of bridge sections was modeled for thermal analysis with

Texas weather data from 1961 to 2005 as the input environmental conditions. Four cities

were considered to bound Texas weather conditions. Based on the thermal analyses, a

long period of sample data of thermal parameters were used to describe the temperature

field over a section. Extreme value analyses of the sample data were performed to obtain

the relationship between thermal loads and return periods. The thermal loads with 100

year return period were compared to the ones suggested by AASHTO.

The thermal loads with 100 years of return period were used to investigate the

structural responses. The proper bearing orientation and the point of fixity were studied.

A rigid body model, as well as a simplified method, was proposed to estimate thermal

movements at the ends, which match very well with those obtained from field monitoring

and finite element method. The maximum possible thermal stresses were also evaluated.

Design suggestions are put forward based on the analysis.

45
Chapter 3 Field Monitoring and Laboratory Test

3.1 MEASURING DEVICES AND DATA ACQUISITION SYSTEM

The study included both field monitoring and laboratory test of steel girder

bridges with a composite deck. A variety of instruments were used in the study,

including thermocouples, foil strain gages, wax measuring devices, laser range meter and

pyronometers to measure temperatures, solar intensities and thermal movements,

respectively. A brief introduction of these measuring devices and data acquisition


systems are provided at the outset of this chapter.

Thermocouple

Thermocouples are used to measure temperature of air, solids or liquids.

Figure 3-1 Thermocouple

As shown in Figure 3-1, a thermocouple consists of two wires of dissimilar metals

that join at one end. When a temperature change occurs at the joining point of the wires, a

voltage is generated. This voltage is read and interpreted by a data acquisition device and

46
then converted into a temperature, which is the temperature at the junction point. The

thermocouples used in this study consisted of copper and constantan wires. The accuracy

of the thermocouples is within 0.02 F.

Pyronometer

Pyronometers (CM3 type from Kipp & Zonen, www.kippzonen.com) are used to

measure the total solar radiation on a flat surface. A CM3 pyronometer consists of a

thermopile sensor, a housing, a dome, and a cable as shown in Figure 3-2. The thermopile

is coated with a black absorbent coating. The paint absorbs the radiation and converts it

to heat. The resultant temperature difference is converted to a voltage by the copper-

constantan thermopile. This voltage is read and interpreted by a data acquisition device

and then converted into a heat flux, which is the heat flux at that instant in time. The

accuracy is approximately 5%.

Figure 3-2 CM3 Pyronometer


47
Wax Trace Box

An important aspect of this project was to measure the movements at the bearings

of various bridges to gain a measure of the actual boundary conditions at the support for

comparison of the idealized support conditions. Requirements for the device included the

necessity of being resilient in extreme weather conditions since the device would be left

on the bridge for several months. After much consideration the device shown in Figure 3-

3 was developed for measuring the magnitude and direction of the girder movements

relative to the pier/abutment.

Screws C-Channel

Beam Flange
Flat Bar

All-Thread Rod

Acrylic Plastic Sheet

Steel Nut Threaded onto


Rod and Tightened
Against Flat Bar

Plastic Storage Steel Nut Used to


Container with Lid Elevate Acrylic Cover

Figure 3-3 Wax Measuring Device

48
The device includes a stylus (attached to the bottom flange of the bridge girder as

shown in the figure), which etches a trace in the wax contained in the plastic container.

The plastic container containing the wax is approximately 7 in. long by 5 in. wide by

2 in. deep, and approximately 3/8 in. of microcrystalline wax was provided in each

container. The wax that was used was obtained from Vishay Micro Measurements and

is typically used for mechanical and moisture protection of foil strain gages. The

relatively high melting point for the wax provided a stable medium to record bridge

movements during hot weather commonly encountered in Texas summers. A hole was

cut in the lid of each of the plastic containers to allow for the expected stylus movement.

A sheet of acrylic plastic (12 in. by 8 in. by 1/16 in. thick) was attached to the stylus to

provide weather protection to the opening in the container lids. Figure 3-4 shows a closer

view of the top of the wax trace box with the weather protection in place.

A length of in. diameter low carbon zinc-coated all-thread rod was used to form

the holder for the stylus. The length of each rod was selected based on the geometry of

each instrumentation location since the bearing height, flange thicknesses, and other

geometrical dimensions vary from bridge to bridge and girder to girder. The threaded rod

permitted adjustments in the depth of penetration into the wax so that both positive and

negative girder rotations could be accommodated without the stylus lifting out of the wax

or interfering with the bottom of the trace box. The end of the threaded rod was turned

down in a metal lathe to form a stylus in. long by 1/8 in. diameter that left a trace in the

wax. Stainless steel pin was attached to the end of the threaded rod shown in Figure 3-4.

Each threaded rod was attached to a girder using a 6 in. length of in. mild steel

flat bar, with one in. diameter hole drilled and tapped into the ends of the bar as shown

in Figure 3-3. The flat bar holding the stylus rod was welded to 6 in. lengths of 4 in. x 2

in. rectangular structural steel tubing with in. walls cut in half lengthwise to form two

49
C-channels. The resulting bracket was attached with two in. diameter holes drilled

and tapped into the legs of each channel as shown in the figure using two in. diameter

by 1 in. or 2 in. long GR5 hex cap screws with 13 threads per inch.

A grid was placed on the top of each of the wax measuring devices as shown in

Figure 3-4. To produce the grid, slits in. apart were sliced into one piece of plastic film,

and slits running perpendicular to the first were cut into a second piece of film. Each

film was then used as a template to draw grid lines on the wax in each container. The

transverse (X) and longitudinal (Y) directions were labeled on the grid, along with the

origin of the system, as shown in Figure 3-4. The origin of the system marks the location

of the stylus at the time that the device was installed on the bridge.

Max Expansion Max Contraction

Current Position

Figure 3-4 Example of Grid Drawn on Wax Device

50
Laser Range Meter

The Hilti PD 32 laser range meter can measure distances with an accuracy of

1/16 inch by using highly accurate, state of the art technology. The Hilti PD32 meter

sends a laser pulse towards the object in question and this pulse reflects with a phase shift

that is measured by the device. This phase shift is converted to the distance between the

range meter and the object being measured (Hilti, 2007). As shown in Figure 3-5, the

meter has a built-in optical sight viewing window that allows it to measure distances in

direct sunlight up to 200m. This device was used to measure lateral pier deflection during

annual thermal cycle.

Figure 3-5 Hilti PD32 Laser Range Meter

Data Acquisition System

The instrumented bridge had three CR5000 data loggers and sixteen AM 4/16

multiplexers. Each of these devices is manufactured by Campbell Scientific Inc.

The CR5000 data logger as shown in Figure 3-6 is a device that records and stores

the readings obtained from sensors. It records data at a rate ranging between 2000 to 5000

samples per second. In the case of bridge girder, it was used to read and store the

readings from the strain gages and thermocouples. This device has slots to connect the
51
gages. At any given time, 40 thermocouples or 20 strain gages can be read. It can store up

to 900,000 data points, and has slots for memory cards as well. In field applications,

power was provided by a rechargeable battery base or a 12 V external auto-marine

battery source.

Figure 3-6 CR5000 Logger (Campbell Scientific Inc.)

The CR5000 data logger can accommodate 20 strain gages at most. To increase

connections, multiplexers as shown in Figure 3-7 were used. A single multiplexer can

hold up to 32 thermocouples or 16 strain gage connections and it occupies just a single

slot in the data logger. A single data logger can hold 7 such multiplexers. By using

multiplexers, a data logger can hold up to 224 thermocouples or 112 strain gages.

Another advantage in the use of multiplexer is that it can help reduce the amount of

wiring necessary since multiplexers can be placed close to a gage location thereby

reducing the number of wires that need to be extended along the length of the bridge to

the data logger location.

52
Figure 3-7 AM416 Multiplexer (Campbell Scientific Inc.)

The software is used to program the data logger, download data and display real-

time or historical data. All the technical support for this software is provided by

Campbell Scientific Inc.

Vibrating wire strains gages were also instrumented. Laboratory experiments

showed that vibrating wire strain gages are sensitive to direct sunshine and the readings

are not reliable, so the instrumentation are not presented here. Detailed information is

given in Writers thesis (2007).

3.2 INSTRUMENTATION

3.2.1 Instrumented Bridges

The purpose of the field monitoring and laboratory studies was to provide

validation data for the finite element models to ensure proper modeling of the various

bridge components. To improve the accuracy of the modeling, several bridge systems

were modeled. The bridges that were modeled were located around the Houston

Metropolitan area. In addition to the field monitoring, a twin box girder bridge that was

53
constructed at Ferguson Structural Engineering Laboratory (FSEL) in Austin, Texas was

also monitored. This section provides an overview of the location and type of bridge

that was monitored. For practicality many of the bridges were instrumented with only

wax trace boxes. Extensive monitoring of one bridge in Houston and the FSEL bridge

was carried out with the full range of instrumentation outlined in the last section. A total

of nine bridges were modeled throughout the Houston area in addition to the FSEL

bridge. The information about these ten bridges is given in the form of tables, maps, and

photographs which are presented in the remainder of this section.

Table 3-1 List of Bridges Selected

Bridge Name
Description Instrumentation
(Unit No.)
Intercontinental Airport John F. Kennedy Blvd. southbound connection
Full
Bridge (1) to rental car companies
Galleria Wax Trace Box
IH10 underpass @ N. Post Oak Rd.
(2)
Downtown IH10 EB Connection to US59 southbound Wax Trace Box
(3) over-crossing Nance St., upstream unit
Hardy Downstream Hardy Toll Rd. Connection to John F. Kennedy
Wax Trace Box
(4) Blvd. northbound, downstream three-span unit
Hardy Upstream Hardy Toll Rd. Connection to John F. Kennedy
Wax Trace Box
(5) Blvd. northbound, upstream two-span unit
Metro IH45 HOV W Connection to Fuqua Park and
Wax Trace Box
(6) Ride
Greens-point IH45 SB Connection to Beltway 8 westbound,
Wax Trace Box
(7) downstream unit
Woodlands South IH45 underpass @ Lake Woodlands Dr.
Wax Trace Box
(8) eastbound
Woodlands North IH45 underpass @ Lake Woodlands Dr.
Wax Trace Box
(9) westbound
FSEL Laboratory test at the Pickle Research Campus Full
(10) of University of Texas at Austin

54
Table 3-1 lists each bridge selected for the study and contains the bridge name, a

short description of the bridge, and the instrumentation plan for the bridge. Unless

otherwise noted, each bridge is comprised of one unit. As discussed above, only two of

the bridges had thermal, strain, and bearing movements monitored (Full) while the other

bridges had only the bearing movements monitored (Wax Trace Box).

It should be noted that Full Instrumentation indicates that the instrumentation of

the bridge includes strain gauges, thermocouples, and mechanical displacement

measurement devices (wax container). Unit is the length of bridge between expansion

joints. The girders are continuous between expansion joints, and thus are continuous for

the length of the unit.

A map is presented in Figure 3-8 and Figure 3-9 on which are drawn the locations

of the bridges selected for instrumentation. The map is from the Yahoo! Maps website

[Yahoo! Maps (2005)].

FSEL Bridge

Figure 3-8 Location of FSEL Bridge

55
Figure 3-9 Locations of Bridges Selected

3.2.2 Geometry of Fully Instrumented Bridges

This section provides an overview of the geometry and sensor layout on the FSEL

bridge in Austin, Texas and the Intercontinental Airport bridge in Houston, Texas. The

detailed information about these bridges is given in the theses by Grisham (2005), Arikan

(2006), and Espinoza (2007).

The FSEL Bridge is a single span steel twin trapezoidal box girder system used

for the laboratory studies. It is located at latitude 30 17' north and longitude 97 44' west

and the tangent of bridge axis at the south abutment is at a clockwise angle of 26

56
measured from due south. The orientation, latitude and longitude are important

positioning information for the thermal analysis for locating the bridge relative to the sun

during a year. Detailed information about positioning the bridge relative to the sun is

provided in Appendix A. The twin box girders used in this study were originally used

for a High Occupancy Vehicular connector bridge at IH-10 and the 610 West Loop in

Houston, Texas. After the bridge was decommissioned, the girders were moved to

FSEL for a project sponsored by the FHWA and TxDOT investigating the redundancy of

twin box girder bridge systems (Sutton, 2007). The construction of the concrete bridge

deck on these twin box girders provided an opportunity to gain valuable data on the

thermal interaction of the concrete bridge deck and the steel girders. The girders were

erected and assembled, followed by the construction of an 8 inch thick concrete deck

thereby creating a composite bridge system. The pier assemblies, erecting of the girders,

and construction of the concrete deck and rails were performed in coordination with local

contractors and TxDOT officials. The bridge is comprised of two single span units, with

span lengths of 119 and 121 feet respectively. The girders curved slightly with the center

of the radius of curvature located on the west side of the bridge. Therefore, the east

girder had the 121 ft. span. The radius of curvature at the centerline of the bridge was

1365.393 feet. The typical cross section of the bridge is shown in Figure 3-10. Temporary

external solid plate diaphragms were provided at the supports and a top flange lateral

truss system was used along the length of the girders. External cross-frames were

provided at two locations, approximately 12 feet away from mid-span in each direction.

A plan view showing the span lengths, lateral truss orientation, and cross-frame locations

is provided in Figure 3-11.

57
Figure 3-10 Cross-Section of Twin Box Girder Bridge

Figure 3-11 Plan View of Instrumented Bridge

Figure 3-12 and Figure 3-13 show the picture of this bridge with the camera

located on the northwest side of the bridge. Pictures are shown with the deck formwork

in place and after removal of the formwork.

58
Figure 3-12 Placement of Concrete Deck

Figure 3-13 Complete Bridge

The Intercontinental Airport Bridge instrumented in this study is located on the

northeast side of Houston, Texas at the George Bush Intercontinental Airport. The

latitude is 29 45' north and the longitude is 95 23' west. The bridge is indicated on the

map given in Figure 3-8. Pictures of the bridge are shown in Figure 3-14 and Figure 3-15.
59
The bridge is a composite bridge with a 10 inch thick concrete deck on three curved steel

box girders. The bridge connects the car rental facility to the terminals of the airport and

is utilized by buses that transfer incoming and outgoing passengers using rental cars.

The bridge is a four span unit, with span lengths from north to south of 160, 240,

160, and 100 ft, respectively, giving a total length of 660 ft along the centerline of the

bridge. The radius of curvature at the centerline of the bridge was 955.00 feet. A plan

view showing the span lengths and overall orientation of the bridge is provided in Figure

3-16. The pier locations are also labeled in this figure. The cross section of the bridge is

shown in Figure 3-17, using the exterior (E), center (C), and interior (I) girder labels

defined in Figure 3-15. Note the label exterior girder refers to the girder on outside of

the curve and the interior girder refers to the girder on the inside of the curve. The

layout of the bridge scanned from the original design plans for the bridge is shown in

Figure 3-18.

Figure 3-14 Overview of Intercontinental Airport Bridge

60
Figure 3-15 Bottom View of Intercontinental Airport Bridge

N
North Abutment
W
151'

t ier
1s P
er
228

Pi
2 nd
'

In
r

te
Pie

rio
Ce r G
n ird
3 rd

te er
rG
ird (I)
Ex 14 er
te 2' (C
rio )
South Abutment

rG
ird
er
(E
) 99'

Figure 3-16 Plan View of Intercontinental Airport Bridge

61
42' 10"
7' 8.5"
6' 11.5"

10"

5'10" E C I
6'

3'1" 4' 4"

6'9" 14'8" 14'8" 6'9"

Figure 3-17 Typical Section of Intercontinental Airport Bridge

Figure 3-18 Original Design Plan for Intercontinental Airport Bridge

Three different types of pot bearings are used to support the girders of the

instrumented bridge. The first type of pot bearings are fixed bearings where

displacements in the plane of the bearings are restrained. The second type of pot bearings

are multidirectional or free bearings where displacements in the plane of the bearing are

not restrained in any direction. The third type of pot bearings are unidirectional or guided

bearings where displacements in the plane of the bearing in a specified direction are
62
restrained but displacements in an orthogonal direction are unrestricted. On the

instrumented bridge, fixed pot bearings are positioned under each girder at the middle

pier (2nd Pier in Figure 3-16). At other support locations, the girders are supported by

either guided bearings or multidirectional bearings as shown in Figure 3-19. From this

figure, the guides on all unidirectional bearings are oriented on a chord towards the fixed

bearing on that girder line.

Figure 3-19 Bearing Orientation

3.2.3 Instrumentation of electronic sensors

FSEL bridge

Since the bridge is simple and almost straight, no longitudinal thermal gradient is

expected and thereby only one section on both the girders was instrumented as shown in

63
Figure 3-20. The exterior girder had foil strain gages and thermocouples whereas the

interior girder was only instrumented with the foil strain gages and the thermocouples.

A total of 24 foil strain gages and 39 thermocouples installed on the bridge. The

thermocouples were installed on the steel girder as well as inside the concrete deck. To

install the thermocouples inside the concrete, concrete cylinders were prepared first,

having a depth equal to the thickness of the concrete deck. The thermocouples were

placed in the concrete cylinders for higher precision in the placement of the gages relative

to the depth of the slab. The cylinder was placed at the desired location on the concrete

deck and anchored to the reinforcing cage in the slab prior to placement of the concrete

deck. The layout of the foil strain gages and thermocouples are shown in Figure 3-21 and

Figure 3-22, respectively.

North Abutment N
W

Exterior Girder (E)

Instrumented
Section A-A

Interior Girder (I)

South Abutment

Figure 3-20 Instrumented Section of FSEL Bridge


64
Figure 3-21 Foil Strain Gage Layout

Figure 3-22 Thermocouple Layout

It should be noted that five thermocouples hanging in the air were used to

measure the ambient temperatures both inside the boxes and in the open air between the

boxes. To eliminate the effect of plate bending, foil strain gages were placed on both

sides of webs and bottom box girder flanges. This bridge was monitored through in the

period from August to November of 2006.

Intercontinental Airport Bridge

Three sections along the length of the bridge were instrumented as shown in

Figure 3-23. Sensors were installed in all three girders (interior, center, and exterior
65
girders) at each section location. Thermocouples were installed at each section location,

which allowed measurement of the vertical, transverse, and longitudinal thermal

gradients in the bridge. More detailed information on sensor locations, sensor labels,

and wiring details is given by Arikan (2006).

The thermocouple layout on a cross section used at Sections 1, 2, and 3 is shown

in Figure 3-24. There are 30 thermocouples at each of Sections 1, 2, and 3 (90 total).

Beneath the layout of sensors in Figure 3-24 is a plan view marked with the sections that

use the given instrumentation layout.

North Abutment

17' Section 1

N
t ie
r W
1s P on 2
cti
Se

48'
er
Pi
2 nd

Section 3

In
i er

te
rio
r
3 rd P

Ce G
nt ird
e er
Ex rG
te ird
rio er
rG
ird
South Abutment

er

17'

Figure 3-23 Instrumented Section for Intercontinental Airport Bridge

66
E C I
(Multiplexer)

Thermocouple

Section 1 Section 2 Section 3

I
North Abutment

South Abutment
C
E

1st Pier 2nd Pier 3rd Pier

Figure 3-24 Thermocouple Layout for Intercontinental Airport Bridge

The thermocouples hanging in the air were used to measure the ambient

temperatures. This bridge was monitored from Dec., 2005 to Jan., 2007. The numbering

of the thermocouples is given in the following figures.

67
Figure 3-25 Thermocouple Layout On Exterior Girder of Section 1

Figure 3-26 Thermocouple Layout On Central Girder of Section 1

68
Figure 3-27 Thermocouple Layout On Interior Girder of Section 1

Figure 3-28 Thermocouple Layout On Exterior Girder of Section 2

69
Figure 3-29 Thermocouple Layout On Central Girder of Section 2

Figure 3-30 Thermocouple Layout On Interior Girder of Section 2

70
Figure 3-31 Thermocouple Layout On Exterior Girder of Section 3

Figure 3-32 Thermocouple Layout On Central Girder of Section 3

71
Figure 3-33 Thermocouple Layout On Interior Girder of Section 3

3.2.4 Instrumentation of Wax Measuring Devices

Figure 3-34 and Figure 3-35 show the locations of the wax devices on the two

fully instrumented bridges. The drawings also list the beam type as well as the date and

temperature at which each measuring device was placed. There is one schematic for each

bridge. The locations of wax devices on the other bridges are provided by Grisham

(2005). It should be noted that the drawings are not to scale.

72
Wax Device Layout & Placement Temperatures
FSEL Bridge for Laboratory Test
Simple Tub Girders

Notes:
(B) = placed on beam flange
- All abutments and bents are radial.
- Not to scale.
- Tubs placed 8-15-2006.
- Placement Temperature: 85.7o F

Figure 3-34 Wax Locations of FSEL Bridge

Figure 3-35 Wax Locations of Intercontinental Airport Bridge

73
3.2.5 Instrumentation of Laser Range Meter

As shown in Figure 3-36, the laser range meter was placed on a 16-inch long

piece of galvanized 331/8 inch steel angle. Three 8-inch long bolts were threaded

into one leg of the angle through the three holes on the leg to adjust the height and

inclination of the laser range meter as needed. The measurement assembly was placed

on a stool or W-shape section to elevate it to a level that is comfortable for the research

personnel to look into the optical sight on the side of the device to aim toward the target

as shown in Figure 3-36.

Figure 3-36 Laser Range Meter Assembly

The method for measuring the lateral pier deflection is illustrated in Figure 3-37.

To determine whether the pier moved or not due to thermal change, the position of the

top of the cap was measured relative to a reference point that would not move over the

nine month measuring period. As shown in the picture, the fixed reference point was

selected at the base of an adjacent pier. To ensure that the same points on the cap were

74
measured each time the readings were taken, the target points were chosen about 1 ft

beneath bridge girders and marked with indelible pen on the cap.

Mark at base
Target on of pier
pier cap
Measurement
assembly

Figure 3-37 Method for Measuring Pier Flexure

The pier deflection could be obtained by taking the difference between two

measurements of pier position relative to the fixed reference point. It should be noted

that the measured distances that were inclined should be transformed into the horizontal

ones using triangle geometry. Detailed description of the measuring procedure was

given in Writers thesis (2007).

By following the above procedure, six bridges were measured over a full annual

thermal cycle (from October, 2006 to July, 2007): the Intercontinental Airport Bridge,

both upstream and downstream units of the Hardy Bridges, Metro Bridge, Woodlands

South Bridge and Woodlands North Bridges. The descriptions of these bridges were

given briefly in Section 3.2.1 of this dissertation and fully in the theses by Grisham

(2005) and Writer (2007). The measured pier deflections were also given in the Writers

thesis (2007) and will be used in Chapter 7 and Chapter 8 of this dissertation.

75
Chapter 4 Finite Element Thermal Analysis

4.1 INTRODUCTION

A bridge exposed to the environment experiences an exchange of heat energy

between its surfaces and its surroundings, resulting in unsteady thermal conditions within

the bridge. On the boundaries, there are three main mechanisms of heat transfer: (1)

conduction to and from the surroundings through the piers and the supports; (2)

convection with the ambient air; and (3) solar radiation and irradiation from the bridge
surface. Heat transfer on the outer boundaries is mainly due to radiation and convection.

The conduction of heat transfer at the boundaries can be neglected compared with the

amount of heat transfer by convention and radiation (Noda, 2000).

An overview of the finite element model along with validation by comparison

with field data is provided in this chapter. Prior to discussing the finite element model, a

discussion of the governing differential equations required for the thermal analysis is

provided. The FEA model is then discussed along with comparisons of representative

data from the field studies.

4.2 FORMULATION OF THERMAL ANALYSIS

4.2.1 Heat Conduction PDE Within Bridges

The temperature field within a bridge is governed by Fourier heat conduction

partial differential equation (PDE). For a bridge with no internal heat source, the heat

conduction equation can be expressed as follows:

76
T 2T 2T 2T
c = k 2 + 2 + 2 (4-1)
t x y z

where k is the thermal conductivities of the solid; c is the specific heat; and is

the density of the solid; T is the temperature at an arbitrary point (x,y,z); and t is time.

For curved bridges, the temperature varies with time within the bridge cross-

section, as well as from section to section along the bridge length. In such situations, a

three-dimensional analysis is necessary to predict the temperature distribution within the

bridge. For straight bridges with uniform cross-sections, the assumption that the

temperature is constant along the bridge length seems to be reasonable and a two-

dimensional analysis may be sufficient to determine the distribution of temperature

within the bridge. For a two-dimensional analysis, the governing differential equation

reduces to:

T 2T 2T
c = k 2 + 2 (4-2)
t x y

It should be noted that the z-axis is the longitudinal direction of the bridge in

Equations 4-2 and 4-3. A solution to this partial differential equation requires initial

conditions as well as specific boundary conditions.

4.2.2 Boundary Conditions

4.2.2.1 Irradiation

The heat transfer, qr, between bridge surfaces and the ambient air by irradiation
can be calculated by the following expression:
77
q r = F (T 4 T4 ) (4-3)

where F denotes the Stefan-boltzman constant which is equal to to 5.67 w/m2.k4;

is the emissivity of the surface; T denotes the ambient air temperature that depends on

position and time; T is the surface temperature in k.

The emissivity can be measured experimentally using a thermocouple probe and

an infrared radiation thermometer. The surface temperature is first measured using the

surface-type thermocouple probe. The temperature of the same surface is then measured

with the infrared radiation thermometer, adjusting emissivity on the thermometer until

both the thermocouple and infrared radiation thermometer give the same reading. This

emissivity on the infrared radiation thermometer is that of the measured surface (Omega

Engineering Technical Reference, www.omega.com).


The hourly ambient air temperatures are generally available from National

Climatic Data Center. If they are not available, a sinusoidal daily cycle between the

minimum and maximum temperatures, Tmin and Tmax, can be used to interpolate them.

(Kreith and Kreider, 1978).

T (t ) =
(Tmax + Tmin ) + (Tmax Tmin ) sin (t 9) (4-4)
2 2 12

This equation was used by Fu et al. (1990) and Tong et al. (2001).

4.2.2.2 Convection

The heat transfer, qc, between bridge surfaces and the ambient air by convection

can be calculated by the following expression:

78
qc = hc (T T ) (4-5)

where hc denotes the heat transfer coefficient with dimension w/m2.k; and T

denotes the temperature of the ambient air that is dependant on position and time. The

convective heat transfer coefficient hc is influenced by many variables such as wind

speed, surface roughness, and geometric configuration of the exposed structure (Emanuel

and Hulsey, 1978). Experimental tests and empirical formulae are usually used to

determine its value. Ibrahim (1995) used the following empirical formula originally

recommended by Kehlbeck (1975) to determine its value for bridges.

4.67 + 3.83u Top surfaces of bridge deck


2.17 + 3.83u Soffit surfaces

hc = (4-6)
3.67 + 3.83u Outer surfaces of steel webs and concrete slab
3.5 Inside surfaces

where u is wind speed in m/s and hc is in w/m2.k. The hourly wind speed is

generally available from National Climatic Data Center; however in the absence of

hourly measurements, approximations can be made using the average wind speed of the

day.

4.2.2.3 Solar Radiation

Due to solar radiation, heat flows into the bridges through the boundary surfaces.

The heat flux can be expressed as:

q f = I t (4-7)

79
where /n denotes differentiation along the external normal direction n to the

boundary surface. The heat flux on the boundary surface is denoted by qf,. For adiabatic
(impermeable to heat transfer) boundary surfaces, qf is zero. The variable It is the total

solar radiation striking the surface, and is the absorptivity of the surface, which depends

on the nature of surface. The process of calculating the total solar radiation, It, will be

discussed in detail in the following sections.

4.2.2.4 Thermal Contact Between Steel and Concrete

The steel surfaces are assumed to be in perfect thermal contact with the concrete

surfaces [Fu et al. (1990), Moorty (1992), Tong et al. (2001)]. The temperature on the

contact surface and the heat flow through the contact surfaces are the same for both

surfaces as described in the following equations:

Ts T
ks = kc c and Ts = Tc (4-8)
n n

where subscripts s and c refer to steel and concrete surface and n is the common

normal direction on the contact surface, respectively.

4.2.2.5 Initial Conditions

For transient heat conduction problems, the governing partial differential equation

is time-dependent. Therefore, it is necessary to specify an initial condition, which

expresses the temperature distribution within the bridge at the initial time:

T = T (P ) (4-9)

80
where T(P) is the initial temperature distribution and P is a point in the bridge.

Unfortunately, the initial temperature distribution is generally unknown. In the present

study, the initial temperature is assumed uniform and equal to the ambient air temperature

at the initial time. Steps can be taken to reduce the influence of unrealistic initial

conditions. For example, in this study the initial time was typically chosen such that the

temperature distribution is close to uniform, such as between midnight and 5 am when

the bridge temperature is relatively stable and uniform. To further minimize the

influence of the initial time on the behavior, the initial time was typically set 3 days

before the time of interest. The initial temperatures become more realistic over three

days cycle of environmental conditions. For example, if the time period of interest for

the temperature field is on May 5, 2006, the initial conditions were established beginning

around midnight of May 2, 2006.

4.2.3 Solar Radiation On A Tilted Sunlit Surface

The total solar radiation, It, striking a surface consists of three components: beam

radiation Ib, diffuse radiation Id and reflected solar radiation Ir.

It = Ib + I d + I r (4-10)

The beam component of the solar radiation originates from the intense, parallel

suns rays. The diffuse component is derived from the scattering of the beam component

by clouds, dust, fog, smoke, and other particles suspended in the atmosphere. Although

the diffuse component is not as intense as the beam component, its contribution to the

total radiation on a surface is not negligible (Ibrahim, 1995). Solar radiation may also

strike a surface due to reflection from other surrounding surfaces. The reflected
81
component of the solar radiation depends on the reflective properties of the surrounding

surfaces and the total radiation striking the horizontal plane.

4.2.3.1 Beam and Diffuse Components Incident On A Horizontal Surface

The solar radiation incident on a horizontal surface, Ih, can be decomposed into a

beam component and a diffuse component and calculated by the following expression

[Dilger et al (1983), Elbadry and Ghali (1983), Fu et al (1990) and Sandepudi (1991)]:

I h = I bh + I dh (4-11)

where Ibh refers to the beam component striking the horizontal surface; Idh is the
diffuse component striking the horizontal surface; The total solar radiation incident on a

horizontal surface Ih and diffuse component Idh are usually measured by the National

Renewable Energy Laboratory (NREL, www.nrel.gov) or the National Climatic Data

Center (NCDC, www.ncdc.noaa.gov). Once Ih and Idh are available, the beam component

Ibh can be calculated using the above expression. If Idh and Ih are not available, a model

provided in Appendix A can be used.

4.2.3.2 Beam Component On Inclined Surfaces

The beam components striking on an inclined surface can be calculated by:

I bh
I bn = (4-12)
cos
I b = I bn cos (4-13)

82
where Ibn refers to the beam radiation striking a surface that is normal to the suns

rays; is the incidence angle. The method for calculating and according to the sun-

earth geometry is provided in Appendix A.

4.2.3.3 Diffuse Component

The diffuse component Id for a surface at a tilted angle with the horizontal plane
can be calculated using the following equation (Duffie and Beckman, 2006).

1 + cos
I d = I dh (4-14)
2

where the term (1+cos)/2 is the shape factor between the sky and the surface,

which is the fraction of diffuse radiation leaving the sky that strikes the tilted surface.

4.2.3.4 Reflected Component

In calculating the reflected component, it is assumed that:

The reflected component comes from all horizontal surfaces surrounding the

considered surface. The reflected component originates only from the beam

and diffuse component incident on nearby horizontal surfaces, not including

multiple reflections from adjacent surfaces;

All surrounding surfaces have the same reflectivity;

The surrounding surfaces reflect radiation equally in all directions.

Based on the above assumptions, the reflected component on a tilted surface is

given by the following expression (Duffie and Beckman, 2006):

1 cos
I r = rg I h (4-15)
2
83
where rg is the reflectivity of the surrounding surfaces; and Ih is the total radiation

on a horizontal surface adjacent to the considered surface.

4.2.3.5 Shadow Effect

As shown in Figure 4-1, the web of steel fascia girder and the surface of the

concrete deck may be partially or wholly shaded by the concrete deck overhang and by

the rail, which depends on time and the inclined angle of the surface. The shadow has a

significant impact on solar radiation [Dilger et al (1983) and Fu et al (1990)] and it is


therefore important to determine whether a surface is sunlit or not. When the surface is

shaded, only diffuse and reflected components can reach the surface and Ib=0.

Hs

shaded
Sunlit
Hc

shaded

Sunlit
Lc

Figure 4-1 Shadow Due to Rail and Overhang Concrete Deck

The shadow length, Ls, can be calculated as follows (Elbadry and Ghali, 1983).

Lctg
Ls = (4-16)
tg cos + sin cos( )

84
The relationship between the height of rail wall Hw and its shadow length Hs can

be expressed as follows (Hsieh, 1985).

H w cos( )
Hs = (4-17)
tg

where Hc is the length of the overhang concrete deck; is the azimuth angle of

the sun, which is defined as the angle of the suns rays measured in the horizontal plane

from due south, with westward being designated as positive; is the solar altitude angle,

which is referred to as the angle between the suns rays and the horizontal; is the tilted

angle of the surface, which is defined as the angle between the surface and the horizontal

plane. is the surface azimuth angle, which is defined as the angle of the surface normal

measured in the horizontal plane from due south, with westward positive. The methods

for calculating the above angles are provided in Appendix A.

4.3 IMPLEMENTATION OF THERMAL ANALYSIS USING ANSYS

The last section provides a discussion of the partial differential equations

governing heat conduction within a bridge. These differential equations were solved

using the general commercial finite element program ANSYS v10.0 [ANSYS, 2007].

This section provides an overview of the FEA model that was used to conduct the

thermal analysis. Results are presented in this subsection comparing the accuracy of the

FEA model with thermocouple results from the field studies. The field measurements

and laboratory tests were used to validate the accuracy of the model so that detailed

parametric investigations could be used to improve the understanding of the fundamental

thermal behavior of composite steel bridge girders.

85
4.3.1 Three Dimensional Model vs. Two Dimensional Model

To get some insight into the complex response of bridges to environmental

conditions, the entire bridge was modeled and a three-dimensional heat conduction

analysis was conducted. For comparison, a two-dimensional analysis that only modeled

the bridge cross section was also performed to assess the possibility of reducing modeling

requirements for thermal analyses. In both the two- and three-dimensional models, the

temperature was assumed constant through the thicknesses of the steel webs and flanges

since their thickness are very small, which is also justified by the high heat conductivity

of steel (Tong, 2001). Simple finite element analyses of steel plates with various

thicknesses also confirmed this assumption. The respective three- and two-dimensinal

ANSYS models of a typical twin-girder curved bridge are shown in Figure 4-2 and

Figure 4-3, respectively.

Figure 4-2 Three-dimensional ANSYS Finite Element Model

86
Figure 4-3 Two-dimensional ANSYS Finite Element Model

4.3.2 Element Types

Table 4-1 lists all element types used in the thermal analyses.

Table 4-1 Element Types Used in Thermal Analysis

Three-dimensional Two-dimensional
Concrete Deck Solid90 Plane77
Steel Web and Flange Shell132 Plane77
Boundary Heat Transfer Surf152 Surf151

A brief description of these elements is given as follows (ANSYS Manual, 2007).

Solid90

Solid90 is a 20-node Serendipity element with a single degree of freedom,

temperature, at each node. The element has quadratic shape functions and thereby is well

87
suited to model curved boundaries. This element is applicable to a three-dimensional,

steady-state or transient thermal analysis. In this analysis, Solid90 elements were used to

model the concrete deck in the three-dimensional models.

Shell132

Shell1132 is a three-dimensional layered shell that has the capability of modeling

in-plane and thru-thickness thermal conduction. The element has eight nodes. Each node

has up to 32 degrees of freedom, which is applicable to a three-dimensional, steady-state

or transient thermal analysis. In this analysis, Shell132 was used to model the steel

plates that comprise the bridge girder system. Since steel has a high thermal conductivity

and the plate is of small thickness, the thru-thickness thermal conduction is insignificant

and can be ignored. Therefore, only one layer was considered.

Surf152

Surf152 is used to apply various types of loads, which can be overlaid onto an

area face of any three-dimensional thermal element. In this analysis, Surf152 was used to

apply convection, irradiation, and solar radiation boundary conditions on the surfaces of

steel plates and concrete decks in the three-dimensional models.

Plane77

Plane77 is an 8-node Serendipity element with a single degree of freedom,

temperature, at each node. The element has quadratic shape functions and thereby is well

suited to model curved boundaries. This element is applicable to a two-dimensional,

88
steady-state or transient thermal analysis. In this analysis, Plane77 was used to model

steel plates and concrete decks in the two-dimensional models.

Surf151

Like Surf152, Surf151 is the two-dimensional version of Surf152. In this analysis,

Surf151 was used to apply convection, irradiation, and solar radiation boundary

conditions on the surfaces of steel plates and concrete decks in the two-dimensional

models.

4.3.3 Mesh Density

Element sizes can have a significant effect on the accuracy of the results. An

appropriate mesh density is very important to capture the shape of the temperature

gradient within the concrete deck since the temperature is expected to change sharply

through the deck thickness [Emanuel et al (1976), Priestley (1972, 1978), and Fu et al

(1990)]. To determine whether the element size is sufficiently fine, the number of

elements is incrementally increased and comparisons are made between the analyses.

Results obtained from the model with a certain number of elements can be compared to
those obtained from the model with double number of elements. If no significant

difference is observed between them, then the mesh can be deemed adequately fine. It

should be noted that the density can also be reduced by using high-order elements.

Detailed analyses evaluating mesh sensitivity showed that 4 elements along the depth of

the concrete deck are fine enough to capture the nonlinear temperature gradient.

Aspect ratio is another meshing factor that can influence the accuracy of the

results. Therefore, to avoid excessive distortion of the elements, the aspect ratio of

elements must be limited. In practice, the thickness of concrete deck in typically bridge
89
construction ranges from 8 to 14 inches. The element size along the concrete depth is

from 2 to 3.5 inches if four elements are used. For long bridges, a relatively large model

results from using aspect ratios approaching unity. To reduce the number of elements in

the longitudinal direction, a large aspect ratio can be used; however sensitivity analyses

were run to ensure that the larger aspect ratio did not adversely impact the results. Based

upon results from a sensitivity analysis, an aspect ratio in the longitudinal direction

(compared to the width and thickness) was limited to 5.

4.3.4 Time Step

Transient heat conduction is time dependent, so it is also necessary to discretize

time. A good choice of time step is of vital importance since too large of a time step may

miss the peak point of interest while too small of a step leads to poor economy in

computer time. The time step depends on the type of the governing partial differential

equation and the features of the input. The heat conduction PDE is of parabolic type and

expresses a smoothing process (Greenberg, 1998). That is, heat conduction within the

bridge tries to smooth the previous temperature field. According to results from the field

monitoring, the temperature-time curve within a bridge takes on similar shape to that

observed in changes in the ambient temperature. Therefore, it is reasonable to choose an

appropriate time step such that a good approximation can be achieved when we discretize

the ambient temperature data. Numerical tests indicated that an hour interval is small
enough to obtain good results for this analysis.

90
4.4 VALIDATION OF ANSYS THERMAL MODEL

Comparisons were made between the temperature fields generated from the FEA

model with those obtained from the field measurements taken from the bridges

instrumented at FSEL and the bridge located at Intercontinental Airport in Houston. The

detailed information about these two bridges was described in Chapter 3. The thermal

material properties used in the model are given in Table 4-2. The reflectivity of the

ground was taken as 0.2. The convective heat transfer coefficient hc was calculated from
Equation 4-6.

Table 4-2 Material Properties for Thermal Analysis

Material Property Steel Concrete


Thermal Conductivity (w/m.k) 46 1.5
Specific Heat (J/kg.k) 486 960
Absorpivity 0.7 0.5
Emissivity 0.9 0.8
Density (kg/m3) 7850 2400

4.4.1 Validation Comparison with the FSEL Bridge

August 22, 2006 is arbitrarily chosen to demonstrate the comparisons between

measured and predicted results for the purposes of validation. In the interest of clarity

and practicality, select results are presented in the main body of this dissertation. The

values presented are representative of the degree of comparison observed on different

days as well as at various regions throughout the bridge. Additional results are provided

in Appendix B. The environmental data on August 22, 2006 including ambient

temperature, wind speed and solar radiation incident on horizontal surfaces were input

into both the two-dimensional and three-dimensional models. The wind speed was

obtained from National Climate Data Center (www.ncdc.noaa.gov). The ambient

91
temperature and the solar radiation incident on the horizontal surfaces were measured by

thermocouples and CM3 pyronometers in the field. The ambient temperature, solar

radiation incident on the horizontal surfaces and the wind speed are plotted versus time in

the respective Figure 4-4 to 4-6.

Figure 4-4 Hourly Ambient Temperature on Aug. 22, 2006

1000
Total Radiation IH
Diffuse Component IDH

750
Thermal Flux w/m2

500

250

0
0 4 8 12 16 20 24
Time (hr)

Figure 4-5 Hourly Solar Radiation on Aug. 22, 2006


92
Figure 4-6 Hourly Wind Speed

4.4.1.1 Temperature History

Figures 4.6 and 4.7 provide a comparison between the measured temperature

histories at the locations of thermocouples 1 and 2 with those obtained from both the two-

dimensional and three-dimensional ANSYS models on the chosen days. Similar

comparisons at other locations are given in Appendix B. As noted in the figure of the

twin box section, thermocouples 1 and 2 were located on the outside web of the exterior

girder. Additional information about the location and numbering of the instrumentation

was given in Chapter 3. The three-dimensional model generally has slightly better

agreement with the measurements than the two-dimensional model; however both models

follow the general trends very well. Kim (2007) also compared the thermal effects

obtained from two-dimensional and three-dimensional models for modular maglev steel

93
guideways. He found that two-dimensional model is acceptable when the cross section

is uniform along the length of the guideway since longitudinal heat conduction is small.

120

100
Temperature ( F)

80
o

60
23 35 29
30
40 24
25
26
36
37
38
31
32

22 27 28 33
21 1 10 9 11 20 19 34 Measured
20 2 8
39
12
I 18
3 E 7 13 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure 4-7 Comparison at the Location of Thermocouple 01

120

100
Temperature (oF)

80

60

23 35 29
40 24
25
26
36
37
38
30
31
32

22 27 28 33

20 21 1 10 9
39
11 20 19 34 Measured
2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure 4-8 Comparison at the Location of Thermocouple 02

94
4.4.1.2 Sectional Temperature Distribution

The temperature distributions across the width of the bridge were calculated from

the three-dimensional ANSYS models and are also compared with field measurements on

the chosen day. Only the distributions at 6 am and 10 am are shown in Figures 4.8 and

4.9 and similar comparisons at other times on the chosen day are given in Appendix B.

The field measurements are represented by the dots as indicated in the figures, while the

FEA results are graphed as a line along the individual elements of the bridge. At 6 am,

the distribution is close to uniform and approaches ambient temperature after several

hours cooling. At 10 am, it can be seen that the sunlit part of the east web of the exterior

girder is much warmer than the shaded portions. The higher temperatures are also

apparent at the tops of the rails on both sides of the bridge as well as through the

thickness of the slab since the top of the slab would be exposed to direct sunlight. The

three-dimensional FEA model had excellent agreement with the field measurements

across the full width of the bridge.

Figure 4-9 Sectional Temperature Distribution at 6 AM, 08/22/2006

95
Figure 4-10 Sectional Temperature Distribution at 10 AM, 08/22/2006

4.4.2 Validation Comparison with the Intercontinental Airport Bridge

The Intercontinental Airport Bridge also provided valuable validation data for the

ANSYS FEA model. The detailed information about the bridge, the geographic location

and the instrumentation was provided in Chapter 3. July 17, 2006 was arbitrarily chosen

for the comparison; however other days that were compared had similar levels of

agreement between measured and predicted results. Although temperature gradients were

measured at three locations along the length of the bridge, in the interest of space results

are shown only for Section 1 on the bridge. Similar agreement between predicted and

measured results was observed at the other locations as well. The average wind speed of

3.25 m/s on that day is used. The solar radiation incident on the horizontal surfaces is

calculated from the solar model described in Appendix A since the National Climate Data

Center only provides solar radiation measurement between 1991 and 2005. The ambient

temperature was taken from the measurements from the thermocouples hanging inside
96
and outside the box in the field. The hourly environmental data are plotted in Figure 4-11

and Figure 4-12.


120

Am bient Tem perature (o F) 100

80

60

40

20
Inside the Box Girder
Outside the Box Girder
0
0 4 8 12 16 20 24
Time (hr)

Figure 4-11 Hourly Ambient Temperature on July 17, 2006

1250
Total Radiation IH
Diffuse Component IDH

1000
Thermal Flux w/m 2

750

500

250

0
0 4 8 12 16 20 24
Time (hr)

Figure 4-12 Hourly Solar Radiation on July 17, 2006

97
The temperature histories at Thermocouples 1 and 2 calculated from the two-

dimensional ANSYS models are compared with field measurements in Figures 4-13 and

4-14, respectively. Similar comparisons at other locations are given in Appendix B.

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure 4-13 Comparison at the Location of Thermocouple 1

140
Measured
2d ANSYS Model
120

100
Temperature (oF)

80

60
20 24
21 25
22 26
40 23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure 4-14 Comparison at the Location of Thermocouple 2

98
From the above comparisons for both the FSEL bridge and the Intercontinental

Airport bridge, it can be seen that both the three-dimensional and two-dimensional

models give good estimates of the temperature field within the bridge throughout the

daily thermal cycle. Based upon the good agreement observed in the validation

comparisons, the FEA model can be used to provide insight into the thermal behavior of

composite steel bridge systems, which will be the topic of the next chapter.

99
Chapter 5 Parametric Thermal Analysis

5.1 STRUCTURAL RESPONSE TO THERMAL LOADING

The heating and cooling of bridge due to solar radiation, convection and

irradiation lead to complex temperature changes that vary in distribution and intensity

throughout the day and year. The impact of these variations in the thermal environment

can have a significant effect on the structural performance, both in the area of

deformations and the resulting stress distributions. This chapter presents results from a
parametric investigation on the thermal response of composite steel bridge girders. The

first section of the chapter provides an overview of the impact the thermal loads can have

on the distribution of stresses in the cross section. An symmetrical section unrestrained

in the longitudinal direction is subjected to an arbitrary temperature field change T(x,y) as

shown in Figure 5-1.

o x

Figure 5-1 Unrestrained Section Subjected to Temperature Field Change T(x,y)

100
The origin of the coordinate system is chosen at the centroid of the transformed

section, which is found as a function of the material modulus of the steel and concrete. If

both steel and concrete are treated as isotropic, homogenous and elastic materials and the

bond between them is perfect, according to Bernoulli-Euler hypothesis that plane sections

remain plane after deformation, the strain in the z direction can be expressed:

res ( x, y ) y x
z ( x, y ) = + ( x, y )T ( x, y ) = m + + (5-1)
E ( x, y ) y x

where z(x,y) is the strain at Point (x,y); res is the residual self-equilibrating stress;

E(x,y) is the material modulus at Point (x,y); (x,y) is the thermal coefficient of expansion

of the material at Point (x,y); T(x,y) is temperature field change; m is the mean sectional

normal strain; y and x are the strain curvature about y and x, respectively. If no external

forces are applied on the section, the following three equilibrium equations must be

satisfied:


A
res ( x, y )dA = 0 (5-2)


A
res ( x, y ) xdA = 0 (5-3)


A
res ( x, y ) ydA = 0 (5-4)

Solving for res from Equation 5-1 and substituting it into equations 5-2 to 5-4

gives:

101
1 1
m E ( x, y)dA + E ( x, y) ydA + E ( x, y) xdA = ( x, y) E ( x, y)T ( x, y)dA (5-5)
A
y A x A A

1 1
m E ( x, y) ydA + E ( x, y) y dA + E ( x, y) xydA = ( x, y) E ( x, y)T ( x, y) ydA
2
(5-6)
A
y A x A A

1 1
m E ( x, y ) xdA + E ( x, y) xydA + E ( x, y) x dA = ( x, y) E ( x, y)T ( x, y) xdA
2
(5-7)
A
y A x A A

Considering that the origin is chosen at the centroid of the transformed section

and assuming that the section of the bridge is symmetrical about the y axis, produces the

following expressions:

E ( x, y) ydA = 0
A
(5-8)

E ( x, y) xydA = 0
A
(5-9)

E ( x, y) xdA = 0
A
(5-10)

Solving for m, y, x, and res(x,y) from Equation 5-1, and Equations 5-5 to 5-10
gives:

( x, y) E ( x, y)T ( x, y)dA
m = A
(5-11)
E ( x, y)dA
A

102
1 ( x, y) E ( x, y)T ( x, y) ydA
= A
(5-12)
y E ( x, y) y dA
2

1 ( x, y) E ( x, y)T ( x, y) xdA
= A
(5-13)
x E ( x, y) x dA
2

res ( x, y ) y x
res ( x, y ) = = m + + ( x, y )T ( x, y ) (5-14)
E ( x, y ) y x

The above structural responses suggest defining the corresponding thermal


loading parameters to characterize the sectional temperature distribution in the
temperature field. For steel-concrete composite bridges, the coefficients of thermal
expansion, s and c, are similar. The American Association of State Highway and
Transportation Officials (AASHTO, 2002, 2003, 2004, and 2007) stipulates a design
coefficient of thermal expansion of 6.510-6 in/in/F for steel, and a value of 6.010-6
in/in/F for concrete, which are typical values for these two materials. The difference
between these coefficients is about eight percent. If this difference is ignored for
simplicity and the average value is used, the following thermal parameters can be
defined: effective bridge temperature Te, vertical linear temperature difference Tv,
transverse linear temperature difference Th, and residue temperature contributing to self-
equilibrating stresses Tres. For convenient use in the subsequent finite element analyses,
the corresponding discretized forms are also given.

103
n

m
E T ( x, y)dA + E T ( x, y)dA E T A
Ac
c
As
s i i i
Te = = = i =1
(5-15)
Ec dA + Es dA
n

Ac As
E A
i =1
i i

D
EcT ( x, y) ydA + EsT ( x, y) ydA
Ac As
E T A y i i i i
Tv = =D = i =1
(5-16)
y E y 2 dA + E s y 2 dA
n

Ac
c
As
E A y
i =1
i i
2
i

B
E T ( x, y) xdA + E T ( x, y) xdA E T A x
Ac
c
As
s i i i i
Th = =B = i =1
(5-17)
x Ec x dA + Es x dA
2 2 n

Ac As
E A x
i =1
i i
2
i

res ( x, y ) T T
Tres ( x, y ) = = T ( x, y ) Tm v y h x (5-18)
D B

where the subscripts c and s stand for concrete and steel, respectively; A stands
for the section area. B and D are the width and depth of the section; dA is the differential
area of the cross section; the subscript i is the ith element in the finite element model. If
the transverse and vertical linear temperature differences are normalized by the width and
depth of the section, the transverse and vertical linear temperature gradients can be
accordingly defined as follows:

Th
Th = (5-19)
B

Tv
T v = (5-20)
D

104
The two sets of definitions are different only by a constant, so both definitions
will be used interchangeably in the subsequent sections. The decomposition for
composite steel bridge section is illustrated in Figure 5-2. In the figure, the red color
represents the temperature distribution in the concrete deck while the blue color stands
for the distribution in the steel girder. For clarity, a cross section is also plotted in the left
of the picture.

Figure 5-2 Decomposition of A Temperature Profile

If the section is unrestrained, the effective bridge temperature and linear thermal
gradients are associated with the longitudinal displacement and rotation, respectively. If
the section is restrained, the effective bridge temperature and linear thermal gradients are
associated with the axial force and moment, respectively. The residual temperature is the
leftover after subtracting the linear components from the total temperature distribution. It
generally reflects how nonlinear the sectional temperature distribution is and produces
self-equilibrating stresses regardless of whether or not the section is restrained.
105
Noda (2000) derived the structural response of an arbitrary section subjected to an
arbitrary temperature distribution. However, he did not correlate temperature terms in
the thermal field to the corresponding structural responses. Emerson (1979) derived the
effective bridge temperature that governs the longitudinal movement of the deck
neglecting the effects of bending. He obtained the same expressions except that the
mean coefficient of thermal expansion of the composite section was used. However,
no linear thermal gradient and nonlinear thermal gradient in the temperature field
corresponding to rotation and residual stress in the field of structural response were
defined. Emersons expression was used by Moorty et al. (1992) and Tong et al. (2001).
Similar concepts were used for a homogeneous section by Eurocode (2003), which also
suggested splitting the temperature distribution within an individual structural element
into four essential constituent components outlined in the temperature expressions given
in Equations 5-15 to 5-18.

5.2 FE MODEL FOR PARAMETRIC STUDY

The temperature field within steel-concrete composite bridges depends on many


factors including geographic location, environmental conditions, material properties, as
well as the orientation and configuration of the bridge. It is still unclear how these factors
influence temperature gradients within a bridge. In this chapter, results are presented
from using both the two- and three-dimensional models described in Chapter 4 to conduct
analyses to assess the importance of certain parameters on the behavior of composite
bridges. The parameters considered in this chapter include: bridge curvature, bridge
orientation, seasons, overhang length of concrete deck, thickness of concrete deck, the
width of concrete deck and depth of the girder. The parameters are shown as variables
in Figure 5-3. The bridge orientation for the two- and three-dimensional models is
106
defined by the angle, X1, between the tangent of the bridge axis at the end and due south
as shown in Figure 5-3. The geographic locations and the material properties used in the
subsequent analyses are taken as the same as those of the FSEL Bridge discussed in
Chapters 3 and 4. The orientation is chosen as X1=0 unless otherwise specified.
The environmental conditions include solar radiation, ambient temperature, and
wind speed. Four typical seasonally clear days as shown in Figure 5-4 to Figure 5-6 are
preliminarily considered in this chapter.

a) Sectional Notation

b) Orientation of 2-d Model c) Orientation of 3-d Model

Figure 5-3 2-d Models for Parametric Analyses


107
1200
Spring (April 8)
Solar Radiation Incident on Horizontal

Summer (July 20)


1000 Fall (Oct. 5)
Winter (Jan. 12)
Surfaces IH (w/m2)

800

600

400

200

0
0 4 8 12 16 20 24
Time (hr)
Figure 5-4 Solar Radiation Incident on Horizontal Surfaces

300
Spring (April 8)
Summer (July 20)

250 Fall (Oct. 5)


Diffuse Solar Radiation Incident on
Horizontal Surfaces IDH (w/m2)

Winter (Jan. 12)

200

150

100

50

0
0 4 8 12 16 20 24
Time (hr)

Figure 5-5 Diffuse Solar Radiation Incident on Horizontal Surfaces


108
Ambient Temperature (oF) 120

90

60

30
Spring (April 8)
Summer (July 20)
Fall (Oct. 5)
Winter (Jan. 12)
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-6 Ambient Temperature

When investigating other parameters, the summer environmental condition is used


unless specified otherwise. The solar radiations incident on horizontal surfaces are
calculated from the model described in Appendix A. The ambient temperature is
obtained from Equation 4-4 in Chapter 4. The maximum and minimum temperatures
used in the equation were obtained from the National Climatic Data Center (NCDC,
www.ncdc.noaa.gov). Actual environmental conditions over a 45-year period will be
used in an analysis that will be outlined in the next chapter to build a statistical model to
develop design effective bridge temperature range and thermal gradients.
It should be noted that when investigating the influence of a given parameter, the
other parameters were held constant.

109
5.3 COMPARISON OF TWO-DIMENSIONAL AND THREE-DIMENSIONAL MODELS

A complexity in horizontally curved bridges is that the angle of the solar radiation
relative to the bridge section varies along the length of the bridge. As a result, some
sections of the bridge will likely receive more sunlight than other sections since the angle
of inclination of the sunlight varies along the bridge as shown in Figure 5-7. Due to the
difference of the incident angle along the bridges length, a longitudinal thermal gradient
may exist.

Figure 5-7 Variation of incidence angle of curved bridges

As has been described in Chapter 2, the AASHTO Guide Specifications for


Horizontally Curved Steel Girder Highway Bridges (2003) suggests incorporating a
vertical thermal gradient effect by specifying a temperature differential between the
concrete deck and steel girders. In contrast, the AASHTO Standard and LRFD
Specifications (2004) specify a vertical temperature gradient along the depth of the
girder. Although both of these specifications recognize the presence of vertical
temperature gradients, the gradients are quite different. Worthy of mentioning is that
these specifications do not make a distinction between the thermal loading for straight
110
and curved bridges and as a result they do not address the effects of longitudinal
temperature gradients. To assess the importance of bridge curvature, a realistic three-
dimensional finite element model of a sharply curved bridge was developed. The heat
exchange at both the ends is ignored. The sectional sizes of the bridge are the same as
those of the FSEL Bridge and the radius of curvature is taken as 400 ft. The azimuth
angle XI was set to 0.
The summer weather condition described above was used as the input
environment of the finite element model. A total of 21 sections along the bridge length
were selected for evaluating thermal parameters. Each thermal parameter was calculated
according to the discretized forms of Equations 5-15 to 5-18 at an interval of two hours.
Figures 5-8 to 5-10 show the variations of effective bridge temperature, vertical and
transverse linear thermal differences along the bridge axis at the selected times.

120
Effective Bridge Temperature (oF)

100

80

t=8
60 t=10
t=12
t=14
t=16
t=18
40
0 20 40 60 80
Subtended Angle (Deg)

Figure 5-8 Longitudinal Variation of Effective Bridge Temperature


111
25

Vertical Linear Temperature Difference ( oF )


20

15

10

5
t=8
t=10
t=12
0
t=14
t=16
t=18
-5
0 20 40 60 80
Subtended Angle (Deg)

Figure 5-9 Longitudinal Variation of Vertical Linear Temperature Difference

10
Transverse Linear Temperature Difference (oF)

t=8
-5
t=10
t=12
t=14
t=16
t=18
-10
0 20 40 60 80
Subtended Angle (Deg)

Figure 5-10 Longitudinal Variation of Transverse Linear Temperature Difference

112
From the figures, the following observations can be made:
Bridge curvature has a slight effect on the longitudinal effective bridge
temperature as shown in Figure 5-8; however the change in effective bridge
temperature is generally negligible from section to section;
Bridge curvature has a small impact on the vertical linear temperature
difference as shown in Figure 5-9. When the webs of steel girder are open to
sunshine (noticeable in curves for times t=8 and t=18), the girder heats up
rapidly due to the high conductivity of steel. On the contrary, if the webs are
shaded, they will also cool down rapidly. In different times of the day, the part
open to sunshine depends on the sun trajectory path, which results in a small
distinction in the vertical linear temperature difference. The impact is still
within negligible range;
The effect of bridge curvature on transverse linear temperature difference is
significant as shown in Figure 5-10. When the webs are exposed to sunshine,
they heat up rapidly, which significantly changes the linear transverse
temperature difference. For example, at t=10, the webs on the sunlit side heat
up quickly. A positive linear temperature difference occurs. With the increase
of subtended angle, the received sunshine intensity becomes lower due to
decreasing incident angle and thereby the transverse gradient deceases. At
t=18, the webs on the other side are exposed to sunshine and a negative linear
transverse temperature difference is produced. As at t=10, due to the decrease
of incident angle along the bridge length, the transverse temperature difference
also drops.

113
Strictly, the heat conduction within bridges is three dimensional. However, only
the transverse linear temperature differences significantly change with the subtended
angle. The effective bridge temperature and vertical linear temperature differences almost
remain unchanged along the bridge axis. For simplicity, the maximum transverse
temperature difference can be used to bound its effects to eliminate the impact of bridge
curvature. It is also conservative for design purposes. In such circumstances, two-
dimensional FEA models can be used. The subsequent results presented in this
dissertation will focus on the two-dimensional model solutions.

5.4 PARAMETRIC STUDY

5.4.1 Bridge Orientation

The above analysis indicates that temperature gradients within a bridge are
generally influenced by the orientation of the bridge. To examine its effects, the angle XI
was changed at an increment of 30 starting from XI=0. The section size is the same as
the FSEL Bridge. For each orientation, the thermal loading parameters of effective bridge
temperature, vertical and transverse linear temperature differences, maximum and
minimum residual temperature in the steel girder and concrete deck are examined.
Figures 5-11 to 5-17 show the comparisons.
The effect of bridge orientation on the effective bridge temperature is slight. Only
small differences are detected at the time of sunrise or time as shown in Figure 5-11
because the webs are sunlit at these times. This result is consistent with the previous
finding that the effect of bridge curvature on the effective bridge temperature is
negligible.

114
120

Effective Bridge Temperature ( F)


o
100

80

60

40

XI=0
20 XI=30
XI=60
XI=90
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-11 Effect of Bridge Orientation on Effective Bridge Temperature

20
Vertical Linear Temperature Difference (oF)

15

10

5 XI=0
XI=30
XI=60
XI=90
0
0 4 8 12 16 20 24

-5
Time (hr)

Figure 5-12 Effect of Bridge Orientation on Vertical Linear Temperature Difference

115
From Figure 5-12, it can be seen that the vertical linear temperature difference
first decreases and then increases for all four orientations. This is because after the
sunshine in the previous day, the concrete deck holds some heat due to its low
conductivity and the temperature is higher than that of the steel girder from about 20:00
until approximately 08:00 after sunrise when the sun starts to heat up the bridge. When
the concrete deck is hotter than the steel girders the slope of the curves of vertical
gradient versus time are negative. This will be the case during nighttime as the heat
flows from the concrete to steel girder and the surrounding, so the vertical gradient
becomes smaller. When the sun rises and the top surface of the concrete deck becomes
exposed to sunshine, the concrete deck heats up and the vertical gradient becomes larger.
The vertical linear temperature differences begin to change rapidly at the onset of
sunrise (approximately 06:00). The east-facing webs for XI=0 receive more solar
radiation than those for other orientations and therefore heat up quicker. Therefore, the
vertical linear thermal gradient for XI=0 decreases at a higher rate from 06:00 to 11:00.
After 11:00, the whole steel girders are shaded and the distinctions for all four
orientations disappear. Beginning at approximately 15:30, the west-facing webs become
sunlit and those webs for XI=0 receive more solar radiation than the other orientations.
For the same reason, the vertical linear temperature difference is lower.
Another finding worthy of mentioning is that after sunset (beginning at
approximately 19:00), the vertical linear temperature difference becomes higher and
reaches a maximum. This is because the steel girders cool down much more quickly than
concrete deck after sunset. However, the maximum vertical linear thermal gradient the
bridge achieves during a day seems to be independent of the orientation and occurs at
around the time of sunset.

116
6
Transverse Linear Temperature Difference ( F)
o

0
0 4 8 12 16 20 24

-2

-4

XI=0
-6 XI=30
XI=60
XI=90
-8
Time (hr)

Figure 5-13 Effect of Bridge Orientation on Transverse Linear Thermal Gradient

Figure 5.13 demonstrates that the effect of bridge orientation on the transverse
linear temperature difference is very pronounced. For the orientation XI=0, the maximum
positive transverse temperature difference is achieved at approximately 08:00 when the
east-facing webs are sunlit whereas the maximum negative temperature difference occurs
at 17:00 when the west-facing webs are sunlit. The XI=0 orientation is the most critical
since the webs for XI=0 receive more solar radiation than for other orientations during
sunrise and sunset times.

117
30
XI=0

Maximum Steel Residual Temperature (oF)


XI=30
XI=60
XI=90
20

10

0
0 4 8 12 16 20 24

-10
Time (hr)

Figure 5-14 Effect of Bridge Orientation on Maximum Steel Residual Temperature

10
XI=0
XI=30
Minimum Steel Residual Temperature ( F)

XI=60
o

XI=90
5

0
0 4 8 12 16 20 24

-5

-10
Time (hr)

Figure 5-15 Effect of Bridge Orientation on Minimum Steel Residual Temperature

118
25
XI=0

Maximum Concrete Residual Temperature ( F)


o XI=30
XI=60
20 XI=90

15

10

0
0 4 8 12 16 20 24
Time (hr)

Figure 5-16 Effect of Bridge Orientation on Maximum Concrete Residual Temperature

0
0 4 8 12 16 20 24
Minimum Concrete Residual Temperature (oF)

XI=0
XI=30
XI=60
-3 XI=90

-6

-9

-12
Time (hr)

Figure 5-17 Effect of Bridge Orientation on Minimum Concrete Residual Temperature

119
The residual temperature reflects how nonlinear the sectional temperature
distribution is and will produce self-equilibrating stresses that are of interest to bridge
design. The maximum steel residual temperature occurs at the intersecting points of the
web and bottom flange when these points are open to sunshine in the morning or in the
afternoon. The XI=0 case is more critical than the other orientations because the webs for
XI=0 receive more solar radiation than other cases near the times of sunrise and sunset.
The minimum steel residual temperature occurs at the points that are always shaded. The
absolute extreme value can be as large as approximately 21 F. This residual temperature
will induce self-equilibrating stresses of approximately 4 ksi in the steel girders.
The maximum concrete residual temperature occurs at the corner points of the
concrete deck when both the top and side surfaces are open to sunshine in the morning.
The XI=0 case is more critical than the other orientations for the same reason that the
webs for XI=0 received more solar radiation than the other cases at sunrise and sunset
times. The minimum concrete residual temperature occurs at the bottom surface that is
always shaded. The absolute extreme value can be as large as aproximately 22 F. This
residual temperature induces self-equilibrating stresses of approximately 0.5 ksi in the
concrete deck.

5.4.2 Season of the Year

Environmental conditions change from season to season, so the thermal behavior


of the bridges change accordingly. The effects of the season on the thermal performance
of the bridges were considered in the investigation. To outline the impact of the season,
four days are chosen from spring, summer, fall and winter, respectively. The input
weather data for these four days were given previously. The section sizes that were used
are the same as those of the FSEL Bridge. Figures 5-18 to 5-24 depict the variations of
120
effective bridge temperature, vertical and transverse linear temperature differences,
maximum and minimum residual temperature in the steel girder and concrete deck.

120
Effective Bridge Temperature (oF)

90

60

30
Spring (April 8)
Summer (July 20)
Fall (Oct. 5)
Winter (Jan. 12)
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-18 Effect of Season on Effective Bridge Temperature

20
Vertical Linear Temperature Difference (oF)

15

10

0
0 4 8 12 16 20 24
Spring (April 8)
-5 Summer (July 20)
Fall (Oct. 5)
Winter (Jan. 12)
-10
Time (hr)

Figure 5-19 Effect of Season on Vertical Linear Temperature Difference

121
As shown in Figure 5-18, the season has a significant effect on the effective
bridge temperature. As expected, in the summer both the ambient temperature and solar
radiation intensity are higher than the other three seasons, so the effective bridge
temperature is most critical. On the contrary, in the winter, both the ambient temperature
and solar radiation intensity are low, so the effective bridge temperature achieves the
minimum. The spring and fall are similar. The shape of the effective bridge temperature
is similar to that of ambient temperature with the maximum effective bridge temperature
occurring in late afternoon and the minimum occurring in the morning just before sunrise.
As shown in Figure 5-19, the season has a significant effect on the vertical linear
thermal gradient. The maximum vertical linear temperature difference occurs in summer
because solar radiation is the dominating factor of thermal gradients. Actually, the
vertical linear thermal gradient in Eurocode was developed by relating vertical linear
thermal gradient to solar radiation (Sukhov, 1994). The concrete deck holds more heat in
summer than other seasons because the summer daytime is longer and the deck is
therefore exposed to sunshine longer. The steel girder cools down rapidly at sunset
while the deck is still hot, and as a result a larger thermal gradient is induced during
summer. On the contrary, the minimum occurs at sunrise on winter mornings when the
steel girders heat up quickly and the concrete deck is still cool. The maximum difference
can reach up to 16 F in the summer as compared to the minimum of -5 F in winter.
The time of the day at which the vertical linear temperature difference reaches a
maximum varies from season to season. This is because the sunrise and sunset times,
together with the length of the daytime, change throughout the year.

122
It is again noticed that after sunset in the summer, the vertical linear thermal
gradient becomes higher and reaches a maximum because the steel girders cool down
much quicker than the concrete deck.

10
Spring (April 8)
Transverse Linear Temperature Difference ( F)

Summer (July 20)


o

Fall (Oct. 5)
Winter (Jan. 12)
5

0
0 4 8 12 16 20 24

-5

-10
Time (hr)

Figure 5-20 Effect of Season on Transverse Linear Temperature Difference

From Figure 5-20, the effect of seasons on the absolute extreme value of the
transverse linear thermal gradient seems to be insignificant. The maximum positive
transverse gradient is achieved in the morning when the east-facing webs are sunlit
whereas the maximum negative transverse gradient occurs in the afternoon when the
west-facing webs are sunlit. The absolute extremes of transverse linear gradient are
similar in magnitude; however, the time of the day at which the transverse linear thermal
gradient reaches maximum varies from season to season. This is because the sunrise and
sunset times, together with the length of daytime, change during the year.

123
The residual temperatures are affected by many factors. It is hard to see the trend.
From Figure 5-21 to 5-24, it seems that the extreme residual temperature is not sensitive
to season. Generally, the maximum steel residual temperature occurs at the intersecting
points of web and bottom flange when these points are open to sunshine. The minimum
steel residual temperature occurs at the points that are always shaded. The absolute
extreme value can be as large as about 22 F. This residual temperature will induce self-
equilibrating stresses of about 4 ksi in the steel girders.
The maximum concrete residual temperature occurs at the corner points of the
concrete deck when both the top and side surfaces are open to sunshine in the morning.
The minimum concrete residual temperature occurs at the bottom surface that is always
shaded. The absolute extreme value can be as large as 23 F. This residual temperature
will induce self-equilibrating stresses of approximately 0.5 ksi in the concrete deck.

30
Spring (April 8)
Summer (July 20)
Maximum Steel Residual Temperature (oF)

25 Fall (Oct. 5)
Winter (Jan. 12)
20

15

10

0
0 4 8 12 16 20 24
-5

-10
Time (hr)

Figure 5-21 Effect of Season on Maximum Steel Residual Temperature

124
6
Minimum Steel Residual Temperature (oF)
Spring (April 8)
Summer (July 20)
3 Fall (Oct. 5)
Winter (Jan. 12)

0
0 4 8 12 16 20 24

-3

-6

-9

-12
Time (hr)

Figure 5-22 Effect of Season on Minimum Steel Residual Temperature

30
Maximum Concrete Residual Temperature (oF)

Spring (April 8)
Summer (July 20)
25 Fall (Oct. 5)
Winter (Jan. 12)

20

15

10

0
0 4 8 12 16 20 24
Time (hr)

Figure 5-23 Effect of Season on Maximum Concrete Residual Temperature

125
0
0 4 8 12 16 20 24
Minimum Concrete Residual Temperature (oF)

-3

-6

-9

-12 Spring (April 8)


Summer (July 20)
Fall (Oct. 5)
Winter (Jan. 12)
-15
Time (hr)

Figure 5-24 Effect of Season on Minimum Concrete Residual Temperature

5.4.3 Overhang Length of Concrete Deck

The length of the overhang of the concrete deck impacts the area of the steel webs
exposed to direct solar radiation. Therefore, the overhang geometry likely has a
significant influence on the temperature distribution over the section. To investigate the
effect of the overhang geometry on the temperature variation over the section, four
sections with different overhang lengths were considered. The steel section sizes are the
same as those of the FSEL Bridge with the exception of the overhang length B1. The
summer weather condition on July 20 described above was used as the input environment
of the finite element model. Figures 5-25 to 5-31 show the respective variation of
effective bridge temperature, vertical and transverse linear temperature difference, as well
as the maximum and minimum residual temperatures in steel girder and concrete deck.
126
120

Effective Bridge Temperature (oF)

90

60

30 Overhang Length B1 =26 in


Overhang Length B1 =44 in
Overhang Length B1 =62 in
Overhang Length B1 =80 in
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-25 Effect of Overhang Length on Effective Bridge Temperature

The effect of overhang length on the effective bridge temperature is slight.


Throughout the day, the effective bridge temperature does not change significantly as a
function of the overhang lengths. Although a larger portion of the steel web area is
shaded with a longer overhang, the impact on the effective bridge temperature is
negligible. The reason for the small impact is because the effective bridge temperature
is associated with the behavior of the entire bridge section and is calculated using a
weighted average of the temperatures over the section. Although more area of the steel
web is open to direct solar radiation and the temperature distribution over the section
changes, the effect on the weighted average is small and can be ignored. This is also
justified by the fact that the solar radiation is generally the most intense when the sun is
directly overhead. The solar radiation is less intense at the time when the steel webs are
exposed to sunshine (usually at sunrise or sunset time).

127
25
Overhang Length B1 =26 in
Vertical Linear Temperature Difference (oF) Overhang Length B1 =44 in
20 Overhang Length B1 =62 in
Overhang Length B1 =80 in

15

10

0
0 4 8 12 16 20 24

-5
Time (hr)

Figure 5-26 Effect of Overhang Length on Vertical Linear Temperature Difference

10
Overhang Length B1 =26 in
Transverse Linear Temperature Difference ( F)

Overhang Length B1 =44 in


o

Overhang Length B1 =62 in


5 Overhang Length B1 =80 in

0
0 4 8 12 16 20 24

-5

-10

-15
Time (hr)

Figure 5-27 Effect of Overhang Length on Transverse Linear Temperature Difference

128
From Figure 5-26, it can be seen that the vertical linear temperature differences is
the most impacted by the overhang length when the web is exposed to direct sunshine.
The shorter the overhang, the more intense the vertical linear temperature gradient
becomes. The east-facing webs with the shortest overhang length receive more solar
radiation than those with longer lengths and therefore are heated quicker. Therefore, the
vertical linear temperature difference with the shortest overhang concrete deck decreases
at a higher rate from 07:00 to 09:00. After 11:00, the entire steel girder is shaded and
the distinctions in vertical linear temperature difference disappear. Beginning at 15:00,
the west-facing webs become sunlit. Similar to the morning conditions, the webs with
shorter overhangs receive more solar radiation and as a result the temperature in the web
plates increase. Therefore, the vertical linear temperature difference is lower. At
approximately 19:00, the sun reaches a position in the sky that is sufficiently low so that
the impact of the overhang is minimal and all the curves approach one another. After
sunset (at approximately 21:00), the vertical linear temperature difference becomes
higher and reaches a maximum for all cases. The maximum vertical linear temperature
difference the bridge achieves occurs at about sunset time and seems to be independent of
the overhang length of concrete deck. The rate of change in the vertical temperature
difference is relatively gradual after sunset. However, the temperature distribution does
begin to change since the steel girders cool down more quickly than the concrete deck,
which retains heat well into the night.
The results graphed in Figure 5-27 show that the effect of the overhang length of
concrete deck on the transverse linear temperature difference is significant at times
around sunrise and sunset. Shorter overhang lengths result in a smaller shaded area of the
steel web and leads to more solar radiation on the web plates. Therefore, the section
with the shortest overhang length is the most critical. The curves also show that the time

129
of the day at which the transverse linear difference reaches a maximum or a minimum
varies slightly for different overhang lengths.

30
Overhang Length B1 =26 in
Overhang Length B1 =44 in
Maximum Steel Residual Temperature ( F)
o

25 Overhang Length B1 =62 in


Overhang Length B1 =80 in

20

15

10

0
0 4 8 12 16 20 24

-5
Time (hr)

Figure 5-28 Effect of Overhang Length on Maximum Steel Residual Temperature

6
Minimum Steel Residual Temperature ( F)

Overhang Length B1 =26 in


o

Overhang Length B1 =44 in


3 Overhang Length B1 =62 in
Overhang Length B1 =80 in

0
0 4 8 12 16 20 24

-3

-6

-9

-12
Time (hr)

Figure 5-29 Effect of Overhang Length on Minimum Steel Residual Temperature


130
30
Overhang Length B1 =26 in

Maximum Concrete Residual Temperature (oF) Overhang Length B1 =44 in


25 Overhang Length B1 =62 in
Overhang Length B1 =80 in

20

15

10

0
0 4 8 12 16 20 24
Time (hr)
Figure 5-30 Effect of Overhang Length on Maximum Concrete Residual Temperature

3
Minimum Concrete Residual Temperature (oF)

0
0 4 8 12 16 20 24

-3

-6

-9

Overhang Length B1 =26 in


-12
Overhang Length B1 =44 in
Overhang Length B1 =62 in
Overhang Length B1 =80 in
-15
Time (hr)

Figure 5-31 Effect of Overhang Length on Minimum Concrete Residual Temperature

131
Figure 5-28 and Figure 5-29 show that the extreme steel residual temperatures are
affected by the overhang length since the solar radiation received by the surface varies as
a result of the shading. The section with the shortest overhang concrete deck is the most
critical. The maximum steel residual temperature occurs at the intersecting points of web
and bottom flange when these points are open to sunshine. The minimum steel residual
temperature occurs at the points that are always shaded. The absolute extreme value can
was as large as 22 F. This residual temperature will induce a self-equilibrating stress of
approximately 4 ksi in steel girders.
From Figure 5-30 and Figure 5-31, the extreme concrete residual temperature is
not very sensitive to the overhang length of concrete deck since it does not influence the
shaded area of concrete surface. The maximum concrete residual temperature occurs at
the corner points of concrete deck when both the top and side surfaces are open to
sunshine in the morning. The minimum concrete residual temperature occurs at the
bottom surface that is always shaded. The absolute extreme value was approximately 23
F. This residual temperature induces self-equilibrating stresses of approximately 0.5 ksi
in the concrete deck.

5.4.4 Thickness of Concrete Deck

To investigate the effect of slab thickness on the behavior of composite steel


girders on the temperature variation over the section, four thicknesses were considered: 8,
10, 12 and 14 inches. Although 14 inches are rarely used in practice, it is included to see
the trend. Most decks in composite steel girders are in the range of 8 to 10 inches. The
section sizes that were used in the analysis are the same as those of the FSEL Bridge with
the exception of the concrete deck thickness, H1. The summer weather condition on July
20 described above was used as the input environment of the finite element model.
132
Figures 5-32 to 5-38 show the respective variations of the effective bridge temperature,
vertical and transverse linear temperature differences, as well as the maximum and
minimum residual temperature in steel girder and concrete deck.

120
Effective Bridge Temperature ( F)
o

90

60

30
Slab Thickness = 8 in
Slab Thickness = 10 in
Slab Thickness = 12 in
Slab Thickness = 14 in
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-32 Effect of Deck Thickness on Effective Bridge Temperature

As shown in Figure 5-32, the concrete deck thickness does have an impact on the
effective bridge temperature. The concrete deck receives solar radiation mainly through
the top surface. When the concrete deck thickness is increased, the ratio of the area of the
top surface to the volume of the concrete deck becomes smaller and the effective bridge
temperature decreases in the daytime and increases in the evening due to the insulating
properties of the concrete. Therefore, the section with the thinner concrete deck is more
critical.

133
As shown in Figure 5-33, the thickness of concrete deck also has a significant
effect on the vertical linear temperature difference. Increases in the deck thickness,
result in an increase in the positive vertical linear difference and a decrease in the
negative vertical linear difference. The vertical linear difference was actually entirely
positive for the thickest deck that was considered. A positive vertical linear temperature
difference indicates that the concrete deck is warmer than the steel girders.
Results graphed in Figure 5-34 demonstrate that the concrete deck thickness also
has a significant impact on the transverse linear temperature difference. The primary
differences between the curves occur near sunrise and sunset when the steel web is
exposed to direct sunshine. Increases in the concrete deck thickness result in a reduction
in the transverse linear temperature difference.

20
Vertical Linear Temperature Difference (oF)

15

10

0
0 4 8 12 16 20 24
Slab Thickness = 8 in
-5 Slab Thickness = 10 in
Slab Thickness = 12 in
Slab Thickness = 14 in
-10
Time (hr)
Figure 5-33 Effect of Deck Thickness on Vertical Linear Temperature Difference

134
10
Slab Thickness = 8 in

Transverse Linear Temperature Difference (oF)


Slab Thickness = 10 in
Slab Thickness = 12 in
Slab Thickness = 14 in
5

0
0 4 8 12 16 20 24

-5

-10
Time (hr)

Figure 5-34 Effect of Deck Thickness on Transverse Linear Temperature Difference

30
Slab Thickness = 8 in
Slab Thickness = 10 in
Maximum Steel Residual Temperature (oF)

25 Slab Thickness = 12 in
Slab Thickness = 14 in
20

15

10

0
0 4 8 12 16 20 24
-5

-10
Time (hr)

Figure 5-35 Effect of Deck Thickness on Maximum Steel Residual Temperature

135
6

Minimum Steel Residual Temperature ( F)


Slab Thickness = 8 in

o
Slab Thickness = 10 in
Slab Thickness = 12 in
3
Slab Thickness = 14 in

0
0 4 8 12 16 20 24

-3

-6

-9

-12
Time (hr)

Figure 5-36 Effect of Deck Thickness on Minimum Steel Residual Temperature

30
Slab Thickness = 8 in
Maximum Concrete Residual Temperature ( F)
o

Slab Thickness = 10 in
Slab Thickness = 12 in
25
Slab Thickness = 14 in

20

15

10

0
0 4 8 12 16 20 24
Time (hr)

Figure 5-37 Effect of Deck Thickness on Maximum Concrete Residual Temperature

136
0
0 4 8 12 16 20 24

Minimum Concrete Residual Temperature (oF)


-3

-6

-9

-12 Slab Thickness = 8 in


Slab Thickness = 10 in
Slab Thickness = 12 in
Slab Thickness = 14 in
-15
Time (hr)

Figure 5-38 Effect of Deck Thickness on Minimum Concrete Residual Temperature

As shown in Figure 5-35 and Figure 5-36, the thickness of the concrete deck has a
minor effect of extreme steel residual temperature. During the night, there is a
significant difference between the curves for the maximum steel residual temperature;
however the relative magnitudes are small compared to the magnitudes during the day.
During the daytime, there was very little difference between the curves at sunrise with
larger differences occurring in later afternoon and early evening. Increasing the slab
thickness led to larger maximum steel residual temperatures; however the difference was
not too significant. The largest differences on the minimum steel residual temperatures
occurred during the late afternoon and early evening. Thinner slabs resulted in a larger
absolute magnitude of the minimum residual temperature.
The data graphed in Figure 5-37 and Figure 5-38 show that the slab thickness has
a significant effect on the extreme concrete residual temperature. With the increase of the
thickness of concrete deck, the maximum concrete residual temperature becomes larger

137
and the minimum one becomes smaller. Residual temperatures reflect how nonlinear the
sectional temperature distribution is. When the concrete deck thickness is increased, the
temperature distribution over the section becomes more nonlinear.

5.4.5 Girder Spacing

To investigate the effect of the girder spacing on the temperature variation over
the section, four spacings were considered: 144, 192, 240 and 288 inches. A larger
spacing leads to a longer span of the concrete deck between adjacent girders. Most
typical girder spacings for box girders are usually in the range of approximately 192 to
264 inches, which is well within the range of values considered. The section sizes of the
girders are the same as those of the FSEL Bridge with the exception of the girder spacing.
The summer weather condition on July 20 described above was used as the input
environment. Figures 5-39 to 5-45 show the respective variations of effective bridge
temperature, vertical and transverse linear temperature differences, and also the
maximum and minimum residual temperature in the steel girder and concrete deck.

120
Effective Bridge Temperature (oF)

90

60

30 Girder Spacing S = 144 in


Girder Spacing S = 192 in
Girder Spacing S = 240 in
Girder Spacing S = 288 in
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-39 Effect of Deck Width on Effective Bridge Temperature


138
25
Girder Spacing S = 144 in

Vertical Linear Temperature Difference (oF) Girder Spacing S = 192 in


20 Girder Spacing S = 240 in
Girder Spacing S = 288 in

15

10

0
0 4 8 12 16 20 24

-5
Time (hr)

Figure 5-40 Effect of Deck Width on Vertical Linear Temperature Difference

10
Girder Spacing S = 144 in
Transverse Linear Temperature Difference ( F)

Girder Spacing S = 192 in


o

Girder Spacing S = 240 in


Girder Spacing S = 288 in
5

0
0 4 8 12 16 20 24

-5

-10
Time (hr)

Figure 5-41 Effect of Deck Width on Transverse Linear Temperature Difference

139
As shown in Figure 5-39 and Figure 5-40, the effect of girder spacing on the
effective bridge temperature and vertical linear temperature difference is negligible. The
concrete deck receives solar radiation mainly through the top surface. When the girder
spacing is increased resulting in a wider concrete deck, the ratio of the area of the top
surface to the volume of the concrete deck remains unchanged and therefore the effective
bridge temperature and vertical linear temperature difference show very little effect. In
addition, it also indicates that the temperature distribution along the width is relatively
uniform for a given time of day.
From Figure 5-41, the effect of the girder spacing on transverse linear temperature
difference is significant. Similar to the effect of concrete deck thickness, the transverse
linear temperature difference is induced by the temperature increase in steel webs directly
exposed to sunshine. A larger girder spacing results in a wider concrete deck and the
sectional area of steel webs relative to the whole section becomes smaller. Therefore the
transverse linear temperature difference also becomes smaller.

30
Girder Spacing S = 144 in
Maximum Steel Residual Temperature ( F)

Girder Spacing S = 192 in


25
o

Girder Spacing S = 240 in


Girder Spacing S = 288 in
20

15

10

0
0 4 8 12 16 20 24
-5

-10
Time (hr)

Figure 5-42 Effect of Deck Width on Maximum Steel Residual Temperature

140
6

Minimum Steel Residual Temperature (oF)


Girder Spacing S = 144 in
Girder Spacing S = 192 in
3 Girder Spacing S = 240 in
Girder Spacing S = 288 in

0
0 4 8 12 16 20 24

-3

-6

-9

-12
Time (hr)

Figure 5-43 Effect of Deck Width on Minimum Steel Residual Temperature

30
Maximum Concrete Residual Temperature ( F)

Girder Spacing S = 144 in


o

Girder Spacing S = 192 in


25 Girder Spacing S = 240 in
Girder Spacing S = 288 in

20

15

10

0
0 4 8 12 16 20 24
Time (hr)

Figure 5-44 Effect of Deck Width on Maximum Concrete Residual Temperature

141
0
0 4 8 12 16 20 24

Minimum Concrete Residual Temperature (oF)


-3

-6

-9

-12 Girder Spacing S = 144 in


Girder Spacing S = 192 in
Girder Spacing S = 240 in
Girder Spacing S = 288 in
-15
Time (hr)

Figure 5-45 Effect of Deck Width on Minimum Concrete Residual Temperature

As shown in Figure 5-42 to Figure 5-45, the effect of concrete deck width on the
extreme residue temperature is negligible since the sectional temperature distribution
along the width is relatively uniform at a given time of day. This will be further
discussed later in the chapter.

5.4.6 Depth of Steel Web

Similar to the length of concrete deck overhang, the depth of steel webs also
affects the portion of the steel web that is exposed to direct solar radiation. To
investigate the impact of the web depth on the temperature variation over the section,
four depths were considered: 66, 88, 110 and 132 inches. Aside from the girder depth,
the section sizes are the same as those of the FSEL Bridge. The summer weather
condition on July 20 described above was used as the input environment. Figures 5-46
142
to 5-52 show the effect of the web depth variations on the effective bridge temperature,
vertical and transverse linear temperature differences, and also the maximum and
minimum residual temperatures in the steel girder and the concrete deck.

120
Effective Bridge Temperature (oF)

90

60

30 Girder Depth D = 66 in
Girder Depth D = 88 in
Girder Depth D = 110 in
Girder Depth D = 132 in
0
0 4 8 12 16 20 24
Time (hr)

Figure 5-46 Effect on Effective Bridge Temperature

25
Girder Depth D = 66 in
Vertical Linear Temperature Difference (oF)

Girder Depth D = 88 in
20 Girder Depth D = 110 in
Girder Depth D = 132 in

15

10

0
0 4 8 12 16 20 24

-5
Time (hr)

Figure 5-47 Effect on Vertical Linear Temperature Difference

143
10
Girder Depth D = 66 in

Transverse Linear Temperature Difference ( F)


Girder Depth D = 88 in

o
Girder Depth D = 110 in
5 Girder Depth D = 132 in

0
0 4 8 12 16 20 24

-5

-10

-15
Time (hr)

Figure 5-48 Effect on Transverse Linear Temperature Difference

The curves graphed in Figure 5-46 show that the effect of steel web depth on the
effective bridge temperature is relatively minor. The curves of the effective bridge
temperature over time are very similar for all four different depths. The effect of the
girder depth is very similar to that observed earlier for the overhang length. For a given
overhang size, decreasing the web depth results in more shading of the web; however the
change in the effective bridge temperature is negligible since is the calculation is based
upon the weighted average of temperatures over the section. The behavior is also
supported by the fact that the solar radiation is not quite intense at times when the steel
webs are exposed to sunshine (usually at sunrise or sunset time).
From Figure 5-47, it can be seen that the vertical linear temperature differences
become different when part of steel web becomes exposed to direct sunshine.
Considering the behavior at sunrise, deeper east-facing webs receive more solar radiation
compared to shorter girders and as a result the girder heats quicker. Therefore, the
vertical linear temperature difference with deeper steel webs decreases at a higher rate
144
from 07:00 to 10:00. After 12:00, the steel girders are entirely shaded and the distinctions
in vertical linear temperature difference disappear. Beginning around 14:00, the west-
facing webs start to get exposed to the direct sunlight. Deeper webs receive more solar
radiation than others and their temperature increases quicker. Therefore, the vertical
linear temperature difference is lower. After the sun sets (at approximately 20:00), the
vertical linear temperature difference becomes higher and reaches maximum for all cases
because the steel girders cool rapidly and the concrete deck retains heat. The maximum
vertical linear temperature difference the bridge achieves occurs at the time of sunset.
The depth of the steel webs after the sun sets has a negligible effect on the maximum
vertical linear temperature difference.
The curves shown in Figure 5-48 demonstrate that the steel web depth has a
significant impact on the transverse linear temperature difference. Deeper steel webs
result in less shaded area which leads to more solar radiation received. Deeper webs
lead to more critical temperature gradients. The time of the day at which the transverse
linear difference reaches the extreme lightly varies slightly due to different web depths.

30
Girder Depth D = 66 in
Maximum Steel Residual Temperature ( F)

Girder Depth D = 88 in
25
o

Girder Depth D = 110 in


Girder Depth D = 132 in
20

15

10

0
0 4 8 12 16 20 24
-5

-10
Time (hr)

Figure 5-49 Effect on Maximum Steel Residual Temperature

145
6

Minimum Steel Residual Temperature (oF)


Girder Depth D = 66 in
Girder Depth D = 88 in
3 Girder Depth D = 110 in
Girder Depth D = 132 in

0
0 4 8 12 16 20 24

-3

-6

-9

-12
Time (hr)

Figure 5-50 Effect on Minimum Steel Residual Temperature

30
Girder Depth D = 66 in
Maximum Concrete Residual Temperature ( F)
o

Girder Depth D = 88 in
25 Girder Depth D = 110 in
Girder Depth D = 132 in

20

15

10

0
0 4 8 12 16 20 24
Time (hr)

Figure 5-51 Effect on Maximum Concrete Residual Temperature

146
3

Minimum Concrete Residual Temperature (oF) 0


0 4 8 12 16 20 24

-3

-6

-9

Girder Depth D = 66 in
-12
Girder Depth D = 88 in
Girder Depth D = 110 in
Girder Depth D = 132 in
-15
Time (hr)

Figure 5-52 Effect on Minimum Concrete Residual Temperature

Curves demonstrating the impact of the girder depth on the extreme residual
temperatures in the steel and concrete are presented as a function of time in Figure 5-49
to Figure 5-52. With respect to the maximum values, the extreme residual temperature is
relatively insensitive to the depth of steel webs with the exception of the minimum steel
residual temperature.

5.5 VERTICAL TEMPERATURE DISTRIBUTION

The decomposition of an arbitrary temperature distribution over a section


according to structural responses to thermal loading was outlined in Section 5.1. The
decomposition divided the temperature distribution up into four components, including:
1) effective bridge temperature, 2) vertical linear temperature difference, 3) transverse
linear temperature difference, and 4) the residual temperature. Results from a parametric

147
investigation on the thermal behavior as a function of various geometrical properties were
then presented in Sections 5.2 to 5.4. This section focuses on the temperature
distribution over the cross section.

5.5.1 Temperature Distribution During Daily Thermal Cycles

The FSEL bridge was used to investigate the variation in the temperature
distribution across a section of the bridge during a daily thermal cycle. The orientation is
set to XI=0. The weather data was set to that of a typical summer clear day described in
Section 5.2. Figures 5-53 to 5-57 show the sectional temperature distributions at
selected times during the day. The times range from just after sunrise to after sunset.

Figure 5-53 Sectional Temperature Distribution at 08:00

148
Figure 5-54 Sectional Temperature Distribution at 12:00

Figure 5-55 Sectional Temperature Distribution at 14:00

149
Figure 5-56 Sectional Temperature Distribution at 18:00

Figure 5-57 Sectional Temperature Distribution at 22:00

150
From the above figures, it can be seen that:
The top concrete surface is warmest when the bridge is subjected to heating
when the sun is directly overhead as shown in Figure 5-54 and Figure 5-55.
The middle concrete surface is warmest when bridge is subjected to cooling as
shown in Figure 5-57;
The bottom surfaces of the concrete deck on the overhang and between the two
girders tend to be slightly warmer than the regions of the deck over the tops of
the girders due to solar radiation reflected from the ground as shown in Figure
5-55 and Figure 5-56;
When the steel webs are shaded, the temperature distribution in the steel
section looks roughly uniform along the width as shown in Figure 5-55 and
Figure 5-57. The transverse temperature gradient is generally negligible during
these periods and it is reasonable to consider only vertical temperature
differences;
The maximum vertical temperature differences tend to occur at 12:00 while the
maximum transverse gradients tend to occur at 08:00 and 18:00;
When the steel webs are directly exposed to sunshine, it may cause marked
transverse temperature gradient, particularly for the sections with large ratios
of steel web area to sectional area;
To manifest the effect of side sunshine on the steel webs, the histories of the
average temperature of each web during daily thermal cycles are graphed in Figure 5-58.
It can be seen that the difference of the average temperature between the sunlit
and shaded webs can be as high as 25 F during daily thermal cycles. The difference
between the warmest and coldest points on the entire cross-section is even larger.

151
120

Web4
Web Average Temperature ( F) 110
o

100
Web1
Web2 Web3

90

80

Left Web of Left Girder


70
Right Web of Left Girder
Web1 Web2 Web3 Web4 Left Web of Right Girder
Right Web of Right Girder
60
0 4 8 12 16 20 24
Time (hr)

Figure 5-58 History of Web Average Temperature

To evaluate thermal loads, it is necessary to quantify both the transverse and


vertical temperature distributions. For simplicity, the transverse temperature differential
is considered by assuming a linear distribution along the width of the bridge as Eurocode
does (2003). The vertical temperature differential is evaluated as its average along the
width of the section. That is, the temperature at a given elevation on the cross section is
taken as the average temperature of all points along the width at that elevation. Figures 5-
59 to 5-62 show the vertical temperature distribution during daily thermal cycles. The x-
axis in these graphs is located at the interface of the steel girder and the concrete deck.
The thickness of the concrete deck is normalized to unity and the depth of steel girder is
normalized to 5. The temperature gradient through the slab was generally more nonlinear
than in the steel section. The thickness of the deck and the steel section were initially
normalized by the deck thickness; however this made it difficult to clearly see the data
points in the deck due to the smaller thickness relative to the girder depth. To improve
152
the visibility of the data points in the concrete deck, the depth of the deck was normalized
to unity, while the depth of the steel section was normalized to 5. This normalization
procedure does not affect the general trends observed in the behavior.

1
Concrete
0
75 80 85 90 95 100

-1
Relative Height

-2
Steel
-3 t=00
t=01
t=02
t=03
-4
t=04
t=05
t=06
-5
Temperature (oF)

Figure 5-59 Average Vertical Temperature Distribution (00:00 ~ 00:06)

1
Concrete
0
80 90 100 110 120 130

-1
Relative Height

-2
Steel

-3

t=07
t=08
-4 t=09
t=10
t=11
t=12
-5
Temperature (oF)

Figure 5-60 Average Vertical Temperature Distribution (00:07 ~ 12:00)

153
1
Concrete
0
90 100 110 120 130

-1
Relative Height

-2
Steel
-3
t=13
t=14
-4 t=15
t=16
t=17
t=18
-5
Temperature (oF)

Figure 5-61 Average Vertical Temperature Distribution (1 pm ~ 6 pm)

1
Concrete
0
80 90 100 110 120 130

-1
Relative Height

-2
Steel
-3
t=19
t=20
-4 t=21
t=22
t=23
t=24
-5
Temperature (oF)

Figure 5-62 Average Vertical Temperature Distribution (7 pm ~ 12 pm)

154
From the above figures, it can be seen that no steady-state vertical temperature
distribution exists. The distribution always changes with time. The temperature
variation is very close to uniform at 00:70. At 15:00, the positive vertical temperature
gradient reaches maximum. At 01:00, the negative vertical temperature gradient reaches
maximum.

5.5.2 Proposed Vertical Temperature Gradient

To evaluate thermal loads, the vertical distribution of temperature must be


determined. Since the vertical temperature distributions are quite different during heating
and cooling, two vertical temperature distributions are proposed according to
observations of the vertical temperature distribution during daily thermal cycle. The two
proposed temperature distributions for heating and cooling are shown in Figure 5-63.

T1
y
hc

Concrete Deck
T
8

Depth of Superstructure

T2 Steel Girder

4
y
T = T2 + (T1 T2 ) 1
hc + 8
8

T3

(a) Vertical Temperature Distribution for Heating

155
T1

hc/2 hc/2
T2

hc
Concrete Deck

8
Depth of Superstructure
Steel Girder

T3

(b) Vertical Temperature Distribution for Cooling


Figure 5-63 Proposed Vertical Temperature Distributions during heating and cooling

where each parameter stands for:


T1 for heating: the maximum temperature on the top of concrete deck;
T2 for heating: the average temperature of the middle portion of steel web when T1
reaches maximum;
T3 for heating: the average temperature of bottom flanges when T1 reaches maximum.
When bridge is subjected to heating, T3 is larger than T2 due to the reflected solar
radiation from the ground;
T1 for cooling: the minimum temperature on the top of concrete deck;
T2 for cooling: The mid-height temperature of concrete when T1 reaches minimum;
T3 for cooling: The average temperature of the middle portion of steel web when T1
reaches maximum;
hc: the thickness of concrete deck;

156
8 inches: The distance heat penetrates from the portion with higher temperature to
that with lower temperature.

5.5.3 Applicability of Proposed Vertical Temperature Distribution

To ensure that the proposed vertical gradients are suitable for describing the
extreme temperature distribution under various conditions, the impact of several
parameters on the distribution were investigated. The parameters that were examined
include: bridge orientation, seasons, length of concrete overhang, concrete deck
thickness, spacing between adjacent girders, and the depth of the steel girders. The
parametric model is the same as described in Section 5.2 and the ranges of these
individual parameters are the same as those used in Section 5.4. When investigating the
effect of seasons, four typical seasonally clear days as described in Section 5.2 were
considered. When investigating the other parameters, the summer environmental
condition is used unless specified otherwise.
The extreme distributions obtained from the ANSYS models are compared to the
proposed distributions given in Figure 5-63. The x-axis is located at the interface of
steel girder and concrete deck in the following figures.
Only the comparisons for season change are fully given here. These comparisons
are representative of the range of parameters that were considered. Similar comparisons
are listed in Appendix C.

157
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(a) Vertical Temperature Distribution for Heating

10
Concrete
0
-15 -10 -5 0 5 10

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure 5-64 Effect of Season on Vertical Temperature Distribution (Spring)

158
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(a) Vertical Temperature Distribution for Heating

10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure 5-65 Effect of Season on Vertical Temperature Distribution (Summer)

159
10

Concrete
0
-6 -3 0 3 6 9 12 15

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(a) Vertical Temperature Distribution for Heating

10
Concrete
0
-15 -10 -5 0 5 10

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure 5-66 Effect of Season on Vertical Temperature Distribution (Fall)

160
10

Concrete
0
-6 -3 0 3 6 9 12 15

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)

(a) Vertical Temperature Distribution for Heating

10
Concrete
0
-15 -10 -5 0 5 10

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure 5-67 Effect of Season on Vertical Temperature Distribution (Winter)

161
5.6 SUMMARY

This chapter provided an overview of the development of both two- and three-
dimensional finite element models of a composite steel bridge. The results presented
showed that two-dimensional FEA thermal analyses are sufficient for studying the
thermal behavior of these bridge systems. The effective bridge temperature and vertical
linear temperature difference showed little change along the bridge axis. The transverse
linear temperature difference did vary significantly along the bridge axis. To simplify the
problem, the extreme value was used to bound the effect for design purposes.
Based on the two-dimensional model, two methods of parameterizing the
temperature field over a section were examined. The first method was to decompose the
sectional temperatures into four components: 1) effective bridge temperature, 2) vertical
linear temperature difference, 3) transverse linear temperature difference, and 4) residual
temperature. These components were illustrated in Figure 5-2 and the total temperature
distribution can be evaluated by summing the individual components as given in the
following equation.

Th T
T ( x, y ) = Te + x + v y + Tres ( x, y ) (5-21)
B D

The other method relies on the observation that the temperature distribution along
the width is relatively uniform and to combine the vertical linear temperature difference
and the residual temperature into a nonlinear vertical thermal gradient over the depth as
given in the following equation.

Th
T ( x, y ) = Te + x + T ( y) (5-22)
B

162
where T(y) is the nonlinear temperature distribution along the depth. Other
notations are the same as those in Section 5.1.
Following the parameterizations, the effects of bridge geometry and the season of
year on the above thermal parameters were investigated. Based on the parametric
analyses, the following conclusions can be reached:
The extreme effective bridge temperature and vertical linear temperature
differences are generally not sensitive to the bridge orientation, but transverse
linear temperature difference is very sensitive. The N-S oriented bridge
produces a critical transverse linear temperature difference. The impact of
bridge orientation on the residual temperature is also negligible.
The maximum effective bridge temperature occurs on a clear summer day due
to the combination of the high ambient temperature and intense solar radiation.
On the contrary, the minimum effective bridge temperature occurs during a
winter condition when both ambient temperature and solar radiation are low.
The effects of environmental conditions are considered further in the next
chapter.
The length of the concrete deck overhang and the depth of steel girder have
similar effects on all thermal loading parameters. Lower ratios of the
overhang length to girder depth will produce larger transverse linear
temperature difference. The effects of this ratio on other thermal loading
parameters are negligible.
The thickness of the concrete deck has a pronounced effect. Thinner decks tend
to produce higher effective bridge temperatures and transverse linear
temperature gradients. The effects on the other thermal loading parameters are

163
very complex. However, in practice, the variation of the concrete deck
thickness is relatively limited. Most decks are in the range of 8 to 10 inches.
The spacing between adjacent girders has a significant effect on transverse
linear temperature gradient.
Only when the bridge is subjected to side sunshine in the hours shortly after
sunrise or shortly before sunset does the transverse linear temperature
difference needed to be considered. Otherwise, its effect is negligible.
The temperature profile over the section is relatively uniform along the
concrete width, so it is reasonable to represent sectional temperature fields
using effective bridge temperature, vertical thermal gradient and the transverse
linear temperature gradient. By this simplification, a vertical temperature
distribution for heating and another one for cooling are proposed, both of
which have good correlation with results from the FEA studies.
According to the above influences, the most critical thermal loads are achieved
under the following bridge configurations:
N-S bridge orientation;
Shorter lengths of the concrete deck overhang;
Deeper steel girder webs;
Thinner concrete decks;
Wider girder spacing.
It should be noted that only four typical days from different seasons are selected
to consider their effects in this chapter. Actually, environmental factors have complex
and significant effects on thermal loads because of their random features, which will be
further discussed in the next chapter.

164
Chapter 6 Development of Design Criteria

6.1 METHODOLOGY FOR DEVELOPING DESIGN CRITERIA

Results were presented in Chapter 5 from parametric studies conducted to

investigate the thermal behavior of composite steel bridge girders. In those analyses,

weather data was used from four arbitrarily selected clear days from each season

throughout the year. While the results presented in the last chapter demonstrated the

impact of geometrical variations of the bridge section on the thermal behavior of


composite steel girders, the analyses did not consider the effects of environmental factors

and provide information on the specific design requirements for thermal analyses on

bridge systems. To get design thermal loads, rational analyses on the thermal loading

parameters are needed. Since the environmental conditions (solar radiation, wind speed

and ambient temperature) that affect the temperature field within a bridge exhibit random

features [Ho and Liu (1989), Maes (1992), Sukhov (1994, 2000), Silveira (2000), Tong

(2002), Lucas et al (2003)], it is natural to consider thermal loading parameters as random

variables (Maes, 1992). Although provisions in the AASHTO Specifications outlined in

Chapter 2 (2002, 2004, 2007) provide guidelines for thermal analyses, the development

and suitability of these guidelines is not clear for application on Texas Bridges. The

guidelines are a function of the temperature range the bridge will experience during an

annual thermal cycle as well as the thermal gradient on the bridge. Although a vertical

thermal gradient that had good correlation with the FEA solutions was presented in the

last chapter, design temperature magnitudes of the gradient should be established based

upon historic weather data. A statistical analysis on the weather data over several years

can be used to validate the AASHTO provision or develop new guidelines thereby

resulting in reliable temperature data that bridges will realistically be subjected to.

165
Another motivation to use statistical methods is to define thermal loads consistent with

other loads, such as wind and snow loads based on the relationship between thermal loads

and a suitable return period.

Based on statistical analyses on hourly data, two methodologies can be used to

develop design criteria: direct method and indirect method. Sukhov (1994) named them

the exact method and the simplified method, respectively. The words direct and

indirect were used in this dissertation to particularly stress the different data the two

methods use. The direct method calculates the hourly values of thermal loading

parameters of interest for a relatively long period and then uses these values to build a

stochastic process model (Sukhov, 1994). According to the model, the representative

values of the thermal loading parameters can be determined. Maes (1992) used the

direct method to develop design criteria for evaluating the thermal effects in concrete

bridges.

The indirect method builds a stochastic process model based on only

environmental conditions. The thermal loading parameters of interest are related to

environmental conditions. The representative values of the thermal loading parameters

can be obtained by using the representative values of the corresponding environmental

conditions (Sukhov, 1994). The indirect method was widely used to develop thermal

loading for steel or concrete bridges by many researchers [Ho and Liu (1989), Sukhov

(1994, 2000), Silveira (2000), Tong (2002), Lucas et al (2003)]. However, few research

studies were devoted to the development of design criteria for steel-concrete composite

bridges.

The indirect method does not require the calculation of the distribution of thermal

loading parameters of interest, so the representative values obtained from this method are

approximate. Additionally, if the representative values of the thermal loading


166
parameters with different return periods are desired, the representative values of the

corresponding environmental conditions must be input with the corresponding return

periods and the analysis must be rerun. The direct method conducts statistical analyses

on the thermal parameters of interest themselves, so the results accuracy should be

higher. In this dissertation, the direct method will be used to develop design criteria.

The processes of both the direct and indirect methods are illustrated in Figure 6-1.

Texas Hourly Weather Condition Data

Statistical Model
Correlated
Daily Extreme Thermal Daily Extreme
Loading Parameters Weather Conditions

Thermal Loading
Parameters with Correlation Weather Condition with
desired return period Formula desired return period

(a) Indirect Method

Texas Hourly Weather Condition Data

Validated FE Model

Hourly Thermal Loading Parameters

Extreme Value Analysis

Thermal Loading Parameters


with desired return period

(b) Direct Method


Figure 6-1 Methodology for Developing Design Thermal Loading

167
6.2 EXTREME VALUE ANALYSIS

6.2.1 Order Statistics and Extreme Value Distribution

In the statistical approaches that are outlined in this chapter each thermal loading

parameter is treated as a random variable. Let X be the population of a thermal loading

parameter with (X1, X2, , Xn) being a sample of X. If X1, X2, , and Xn are organized
in an increasing order x1 < x2 < < xn, the rth element of this new sequence is the rth

order statistic of the sample of size n (Castillo, 1988). The maxima and the minima of

the sample are given by the following respective expressions for xn and x1:

xn = max( X 1 , X 2 ,L, X n ) (6-1)


x1 = min( X 1 , X 2 ,L, X n ) (6-2)

Obviously, xn and x1 are also random variables. For design purpose, it is of interest
to accurately estimate the thermal loading parameters with desired return period, the key

step of which requires determining the limiting distributions of xn and x1 when n

approaches infinity. Once the limiting cumulative distribution function (CDF) is

known, the thermal loading parameters with a certain return period can be readily

calculated. If the sample of these thermal loading parameters is assumed independent

and identically distributed and comes from a parent population with a known cumulative

distribution function F(x), the CDFs of maxima and minima can be calculated as follows:

CDF for Maxima: Hn(x) = Prob [Xn<x] =Fn(x) (6-3)

CDF for Minima: Ln(x) = Prob [X1<x] =1-[1-F (x)] n (6-4)

168
Only three families of non-degenerated distributions 1) Frechet, 2) Weibull and 3)

Gumbel can occur as the limiting CDFs for maxima and minima of independent and

identically distributed samples if the limiting CDFs exist. This theorem was originally

proved by Fisher and Tippett (1970) and Galambos (1978). For most practical

problems, one of the above three families generally occurs (Castillo, 1988). For any

given CDF F(x), the rules for determining its domain of attraction are also given by

Fisher and Tippett (1928) and Galambos (1978). Two general rules are useful and are

given (Castillo, 1988): 1) A parent distribution with unbounded endpoint in the tail of

interest cannot lie in a Weibull type domain of attraction; and 2) A parent distribution

with finite endpoint in the tail of interest cannot belong to a Frechet type domain of

attraction. The thermal loading parameters are bounded by practical limits, so the

Frechet type of distribution can be excluded. In case of minima, a change of sign of the

data will permit the use of the Gumbel and Weibull CDFs for maxima. Therefore, only

the Gumbel and Weibull types for maxima are introduced here. The Gumbel CDF for

maxima is given by:

x < x <
G ( x ) = exp exp

(6-5)
>0

where and are constants known as the location and scale parameters,
respectively. The Weibull CDF for maxima is given by:

x
exp x
G ( x) = (6-6)

1 otherwise

169
where , , and are constants known as the location, scale and shape
parameters, respectively. Values of and must be positive.

However, the difficulty in determining the limiting distribution is that both the

CDF F(x) of each thermal loading parameter and its domain of attraction are unknown.

Only a sample of hourly data of each thermal parameter is available. In such situations,

the probability paper method can be used as an alternative method, which provides a very

simple graphic method to determine the domain of attraction. In the subsequent section,

this method will be briefly discussed. Detailed information is given in the book by

Castillo (1988).

6.2.2 Determination of The Domain of Attraction

The probability paper method can be used to determine the domain of attraction.

The basic idea is that when the transformed random variable X is plotted against the

transformed CDF F(x) on the probability paper of a given parametric family of

distributions, the plot appears as a straight line if the plotted CDF belongs to that given

parametric family. The reason to transform the CDF and random variable is to make
sure that the plot not belonging to the given family does not look like a straight line

(Castillo, 1988). This means that the plot of any CDF on such probability paper allows

one to decide whether or not it lies in that family. For the problem considered in this

study, the CDFs associated with each thermal loading parameter are unknown and only a

sample can be obtained from the finite element thermal analyses described in Chapter 5.

In such a situation, an empirical CDF can be used to approximate the actual CDF on the

170
probability paper (Resnick, 1987). The following plotting position formulas used by

Maes (1992) can be used.

i
(xi , pi ) pi = (6-7)
n +1

where pi is the empirical cumulative distribution value. That is, pi = Prob (x<xi); xi

is the ith element in the increasing sequence (x1, x2, , xn); and n is the sample size.
As described above, a plot of the empirical CDF against the transformed random

variable can be made to determine the appropriate domain of attraction. In fact, it is not

necessary to fit the whole sample with a straight line because the only portions of the

CDF governing the behavior of the extremes are the tails. The right tail is needed for

the maxima while the left tail is needed for the minima. Resnick (1987) and Castillo

(1988) show a clear connection between tail equivalence and domains of attraction,

which states that: 1) if two distributions are tail equivalent and one belongs to some

domain of attraction, the other must also belong to the same domain of attraction, and 2)

the sequences of normalizing constants coincide. This conclusion implies that a

distribution F(x) can be replaced by a tail equivalent distribution G(x) without altering

either the domain of attraction or the set of admissible constants (Castillo, 1988). If

G(x) belongs to one of the three families of distributions: Gumbel, Weibull and Frechet,

the problem is considerably simplified because each limiting distribution lies in its own

domain of attraction (Leadbetter, et al., 1983). In other words, one of the three

distributions can be fit to the tail of the given distribution of sample and used for an

extreme value analysis since the tails carry adequate information to determine the domain

of attraction of the parent distribution (Castillo, 1988). When a CDF F(x) is plotted on

171
Gumbel probability paper, the concavity in the associated tail tells the type of domain of

attraction that it belongs to: 1) distributions lying in a Weibull type domain of attraction

for maxima show convexity in the right tail; 2) distributions lying in a Frechet type

domain of attraction for maxima show concavity/convexity in the right tail; 3)

distributions lying in a Gumbel type domain of attraction for maxima appear as straight

lines in the right tail (Castillo, 1988).

Subjectivity is the main drawback of the tail equivalence method since no precise

criterion is given to decide what convexity is negligible and what portions of the sample

are defined as tails. Therefore, erroneous rejections of a Gumbel type distribution are

possible. Fortunately, such rejections are insignificant to the accuracy of the results

since a Gumbel type CDF can be approximated as accurately as desired by Weibull

(Castillo, 1988). The wrong rejection of a Gumbel type distribution can be corrected in

the parameter estimation process following this decision.

It should be noted that strictly speaking, the tail-equivalence method can be only

used in the case of stochastically independent random variables. However, Berman

(1971) and Leadbetter (1983) pointed out that it can be extended to the extremes of the

general time series if the correlation is weak. With the increase of the sample size or the

level of extremes, the peaks become more and more independent, which increases the

justification of this method (Maes, 1992).

6.2.3 Parameter Estimation

Once the domain of attraction is determined, the parameters of the limiting

distribution must be estimated. For the Gumbel distribution, the location and scale
are to be determined. For the Weibull distribution, the location , the scale and the

172
shape parameter are to be determined. The least square method will be used. This

method is intended to minimize the error function. The error function for the Gumbel

distribution is given by:

2
n
x
E = i i (6-8)
i =1

The error function for the Weibull distribution is given by:

n
E = [i + (ln ( xi ) ln )]
2
(6-9)
i =1

i = ln( ln pi ) (6-10)

It should be noted that for the Gumbel distribution, the partial derivatives of E
with respect to and can be set equal to zero and the instantaneous equations can be
solved for and . However, for the Weibull distribution, 3 parameters must be solved
for. The corresponding instantaneous equations are transcendental and cannot be solved
by simple algebraic operations. In such a situation, a graphical method can be used to
get the approximate solution. This method reduces the three-parameter estimate to two
by assuming that is known. A plot of the error versus for different values of .
The value of that minimizes the error function can then be found from observation.

6.2.4 Procedure for Extreme Value Analysis

According to the extreme value theory, the procedure for conducting extreme
value analysis is summarized as follows:

173
1) Get the hourly values of a thermal loading parameter for a long period and
identify daily extremes of each thermal loading parameter.
2) Sort the identified daily extremes in an increasing order.
3) Take the first n extremes for extreme value analysis and calculate the empirical
CDF data sets according to Equation 6-7. The number of data points for
extreme analysis is inconsequential. Taking approximately 0.5% of the total
number of data points is generally sufficient (Maes, 1992).
4) Plot the empirical CDF on the Gumbel probability paper: (xi ,i ) and
i = ln( ln pi ) .

5) Determine the domain of attraction of the thermal loading parameter according


to the concavity of the curve in the right tail. If the Gumbel distribution is
rejected, a plot is made of the empirical CDF on Weibull probability paper.
6) Estimate the parameters of the limiting distribution.
7) Calculate the thermal loading parameter with desired return period.

In the subsequent section, results from an extreme value analyses are presented
that were found from following the above procedure. Parameters that were considered
include effective bridge temperature, vertical and transverse linear temperature
difference, as well as the residual temperatures, T1, T1 -T2 and T3-T2 for heating and T1,
T2-T1 and T2 -T3 for cooling.

6.3 DEVELOPMENT OF DESIGN CRITERIA

To obtain a long period of hourly sample data for each thermal loading parameter
of interest, the weather data from 1961 to 2005 were input into the finite element model

174
used in Chapter 5. Four cities were chosen to bound Texas weather conditions since one
city cannot represent the whole state due to the large geographical area. These four
cities are Austin, Wichita Falls, Brownsville and El Paso. A map is shown to indicate
their locations in Figure 6-2 (Yahoo, Map).

Figure 6-2 Cities to be Analyzed

The most critical value for each parameter obtained from these four cities is
considered as the most critical one for the entire state of Texas on that given day. By
conducting the 45-year analyses for each city, a total of 16436 daily extremes were
obtained. The first 150 data points in the right tail were used for extreme value analysis
for maxima. In the case of the minima, the same procedure for maxima was followed
by a change of sign of the data.
The parametric analyses in the previous chapter clearly indicated that the
configuration of cross section and the bridge orientation influences the thermal behavior
175
during daily or annual thermal cycles. To get critical thermal loading parameters for
design, the critical section sizes within realistic ranges and critical orientation were
chosen to get a long period of hourly sample data for a conservative design. The cross
section that was used is the same as the FSEL Bridge except that the thickness of
concrete deck and the depth of steel girder were taken as 10 and 120 inches, respectively.
The normal of bridge section was oriented in the N-S direction. It should be mentioned
that the historic weather data (solar radiation, wind speed, and ambient air temperature)
were provided by the National Renewable Energy Laboratory (NREL, www.nrel.gov)
and the National Climatic Data Center (NCDC, www.ncdc.noaa.gov).

6.3.1 Determination of Cumulative Distribution Function (CDF)

In this section, the cumulative distribution function for each parameter will be
determined. Only the procedures for determining the maximum and minimum effective
bridge temperature are given in this chapter. Similar procedures for determining other
parameters are given in Appendix D. For efficiency in presentation, only the results are
summarized here.

Maximum Effective Bridge Temperature

The plot of the empirical CDFs for the maximum effective bridge temperatures on
the Gumbel probability paper is shown in Figure 6-3.

176
11

9
Reduced Variate, h

3
110 112 114 116 118 120
Effective Bridge Temperature (oF)

Figure 6-3 Gumbel Plot of Effective Bridge Temperature for Maxima

The graph agrees well with a straight line, which indicates that the Gumbel
distribution is applicable. By the least square method described above, the parameters are
estimated as listed in Table 6-1.

Table 6-1 Parameter Estimate for Effective Bridge Temperature

Parameter Domain of Attraction The Location The Scale


Te for Maxima Gumbel 101.72 1.81

To obtain an indication of how good the predictions are, the accepted distribution
is also plotted against the empirical CDF on an arithmetic scale as shown in Figure 6-4.
Trends in the sample follow the least squares prediction very accurately.

177
1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
110 112 114 116 118 120
Effective Bridge Temperature (oF)

Figure 6-4 Arithmetic Plot of Effective Bridge Temperature for Maxima

To clearly determine the environmental conditions under which the maximum


effective bridge temperature is achieved, the histograms of month and time for the
maximum effective bridge temperature are plotted to show their distributions in Figure
6-5 and Figure 6-6, respectively.

178
50%

29%

15%

5%
1%
0% 0% 0% 0% 0% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 6-5 Histogram of Month for Maximum Effective Bridge Temperature

49%

42%

5%
3%
0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 1% 0% 0% 0% 0% 0% 0%

Figure 6-6 Histogram of Time for Maximum Effective Bridge Temperature

As can be seen from the figures, most of the maximum effective bridge
temperature occurs in the hottest season during an annual thermal cycle (June, July and
August). During daily thermal cycles, an absolute majority of the maximum occurs
between 16:00 and 17:00 which is about 2 hours later than the hottest hour of the day.

179
The maximum effective bridge temperature is strongly correlated to the maximum
ambient temperature, the maximum of which also occurs in the same days. The ambient
temperature distribution during a typical year (1980~1981) is plotted in Figure 6-7.

Figure 6-7 Annual Ambient Temperature (Austin, 1980)

This correlation is also justified by the fact that no maximum effective bridge
temperature occurs in April although solar radiation striking a surface is relatively intense
during this month as shown in Figure 6-8.

180
Figure 6-8 Annual Solar Radiation Striking a Horizontal Surface (Austin, 1980)

Based on this observation, Imbsen (1985) and Moorty (1992) related the
maximum effective bridge temperature to the maximum ambient temperature. Based on
the statistical analysis on both effective bridge temperature and ambient temperature, the
maximum effective bridge temperature with a given return period is also related to the
ambient temperature with the same return period in this study. This correlation is plotted
in Figure 6-9. For comparison, the equations that were suggested by Imbsen (1985) and
Moorty (1992) are also plotted in the same figure.

181
120

Maximum Effective Bridge Temperature (oF)


110

100

SuggestedbytheAuthor
SuggestedbyMoorty
SuggestedbyImbson
90
85 90 95 100 105 110
Maximum Ambient Temperature (oF)

Figure 6-9 Correlation between Maximum Te and maximum Ta

Minimum Effective Bridge Temperature

The plot of the empirical CDFs for the minimum effective bridge temperatures on
the Gumbel probability paper is shown in Figure 6-10.

11

9
Reduced Variate, h

3
-18 -14 -10 -6 -2 2 6
Minus Effective Bridge Temperature (oF)

Figure 6-10 Gumbel Plot of Minus Effective Bridge Temperature for Maxima

182
The graph agrees well with a straight line, so the Gumbel distribution is
applicable. By the least square method described above, the parameters were estimated as
listed in Table 6-2.

Table 6-2 Parameter Estimate for Effective Bridge Temperature

Parameter Domain of Attraction The Location The Scale


Te for Minima Gumbel -35.55 4.11

To get an indication of how good the predictions are, the accepted distribution is
also plotted against the empirical CDF on an arithmetic scale as shown in Figure 6-11.
Again as shown previous with the maxima, the least squares prediction follows the trend
in the sample very accurately.

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
-18 -14 -10 -6 -2 2 6
Effective Bridge Temperature (oF)

Figure 6-11 Arithmetic Plot of Minus Effective Bridge Temperature for Maxima

183
To clearly demonstrate the environmental conditions under which the minimum
effective bridge temperature is achieved, the histograms of month and time for it are
plotted to show their distributions in Figure 6-12 and Figure 6-13.

50%

28%

17%

3%
2%
0% 0% 0% 0% 0% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 6-12 Histogram of Month for Minimum Effective Bridge Temperature

39%

18%

11%

3%
2% 2% 1% 1% 3% 1%
0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure 6-13 Histogram of Time for Minimum Effective Bridge Temperature

184
As can be seen from the figures, most of the minimum effective bridge
temperatures occur in Dec., Jan. and Feb., in which both the ambient temperature and the
solar radiation are lowest during seasonal thermal cycles as shown in Figure 6-7 and
Figure 6-8. During daily thermal cycle, most of the minimum temperatures occur at 7 am
and 8 am which is about 2 hours later than the coldest time. Similar to the maximum
effective bridge temperature, the minimum effective bridge temperature is strongly
correlated to the minimum ambient temperature. This correlation is plotted in Figure
6-14. The correlations suggested by Imbson (1985) and Moorty (1992) are also plotted
for comparison.

30
Minimum Effective Bridge Temperature (oF)

20

10

-10
SuggestedbytheAuthor
SuggestedbyMoorty
SuggestedbyImbson
-20
-20 -10 0 10 20 30
Minimum Ambient Temperature (oF)

Figure 6-14 Correlation between Minimum Te and minimum Ta

The extreme value analyses on other parameters are summarized in Table 6-3.
The circumstances in which the extreme values are most probably achieved are also
summarized in the table. Details are given in Appendix D.

185
Table 6-3 Summary of Parameter Estimates

Parameter Domain Location Scale Shape Factor Most Possible Most Possible
of attraction Months Time in a day
Te,max Gumbel 101.72 1.81 N/A June ~ August 4 pm ~ 5 pm
Te,min Gumbel -35.55 4.11 N/A Dec. ~ Feb. 7 am ~ 8 am
Tv,max Weibull 108.00 92.86 66.38 Not obvious 1 am ~ 2 am
Tv,min Weibull 129.00 121.53 65.76 Dec. ~ Feb. 2 pm ~ 3 pm
Th,max Gumbel 0.78 10.71 N/A March ~ April 9 am
Tweb,max Weibull 49.90 31.14 6.31 March ~ April 9 am
Tres,s,max Gumbel 24.05 1.41 N/A March ~ April 9 am ~ 10 am
Tres,s,min Gumbel 6.72 0.58 N/A March ~ May 8 am ~ 9 am
Tres,c,max Gumbel 25.47 1.29 N/A April ~ June 11 am ~ 12 pm
Tres,c,min Gumbel 8.82 1.25 N/A Not obvious 9 pm and 1 am
T1,heating Gumbel 19.07 0.87 N/A May ~ July 1 pm ~ 2 pm
(T1-T2)heating Gumbel 26.06 1.07 N/A April ~ June 1 pm ~ 2 pm
(T3-T2)heating Gumbel 2.83 0.21 N/A May ~ July 1 pm ~ 2 pm
T1,cooling Weibull 11.00 11.25 11.42 Not obvious 9 pm and 1 am
(T2-T1)cooling Gumbel 7.86 1.49 N/A Not obvious 1 am
(T2-T3)cooling Gumbel 14.01 1.70 N/A Not obvious 1 am

Notation (many of these terms are defined in Chapter 5):


Te,max : Maximum effective bridge temperature
Te,min : Minimum effective bridge temperature
Tv,max : Maximum vertical linear temperature difference
Tv,min : Minimum vertical linear temperature difference
Th,max : Maximum transverse linear temperature difference
Tres,s,max : Maximum residual temperature in steel
Tres,c,max : Maximum residual temperature in concrete
Tres,s,min : Minimum residual temperature in steel
Tres,c,min : Minimum residual temperature in concrete

6.3.2 Calculation of Thermal Parameters With Desired Return Periods

Once the limiting distribution for each thermal loading parameter is known, the
values xk with desired return periods can be readily calculated. The probability of
exceedance with return period of k years is 1/365k. Therefore, the cumulative probability
of xk is given by the following expression:

1
pk = p( x < xk ) = 1 (6-11)
365k
186
It should be noted that the units must be consistent since the data are on a daily
basis and the return periods are in years. For the Gumbel and Weibull types of
distributions for maxima, the values xk with desired return periods can be calculated using
the inverse of the CDF functions, which gives:

Gumbel: x k = ln( ln p k ) (6-12)


1
Weibull: x k = exp ln( ln p k ) + ln (6-13)

The above expressions were evaluated with 1 year, 10, 50 and 100 year return
periods for each thermal loading parameter. The results of these analyses are
summarized Table 6-4.

Table 6-4 Thermal Parameter Values with Desired Return Period (F)

Return Period 100 Years 50 Years 10 Years 1 Year


Te,max 120.8 119.5 116.6 112.4
Te,min -7.6 -4.8 1.9 11.3
Tv,max 28.7 27.9 25.9 23.0
Tv,min -25.4 -24.3 -21.7 -17.9
Th,max 18.93 18.39 17.13 15.32
Tweb,max 44.0 43.3 41.4 37.7
Tres,steel,max 38.9 37.9 35.6 32.4
Tres,steel,min -12.8 -12.4 -11.4 -10.1
Tres,con,max 39.0 38.1 36.0 33.1
Tres,con,min -22.0 -21.1 -19.1 -16.2
T1,heating 28.2 27.6 26.2 24.2
(T1-T2)heating 37.3 36.6 34.9 32.4
(T3-T2)heating 5.1 4.9 4.6 4.1
T1,cooling -6.5 -6.2 -5.5 -4.3
(T2-T1)cooling 23.6 22.5 20.1 16.7
(T2-T3)cooling 31.8 30.6 27.9 24.0

187
6.4 EFFECT OF SAMPLE SIZE ON EXTREME VALUE ANALYSES

In the above extreme value analyses, 45 years of weather data were input into the
finite element model to form the sample. The computational effort in processing this
amount of data was significant. In this section, the possibility of using a smaller sample
is examined. Only 10 years of data are to be used here to perform the same analyses as
before. The estimates for the above thermal parameters based on 10-year sample are
compared to those based on 45-year sample. These estimates are listed in Table 6-5. For
comparison convenience, the values based on 45-year sample are also given.

Table 6-5 Effect of Sample Size on Extreme Value Analysis (F)

100-year 50-year 10-year 1 Year


Return Period Return Period Return Period Return Period
Sample Size 45 10 45 10 45 10 45 10
years Years years Years years Years years Years
Te,max 120.8 121.9 119.5 121.0 116.6 118.5 112.4 114.0
Te,min -7.6 -6.7 -4.8 -5.5 1.9 -1.9 11.3 6.4
Tv,max 28.7 28.4 27.9 27.5 25.9 25.5 23.0 22.6
Tv,min -25.4 -26.1 -24.3 -25.0 -21.7 -22.2 -17.9 -18.2
Th,max 18.93 16.90 18.39 16.84 17.13 16.61 15.32 15.75
Tweb,max 44.0 41.4 43.3 41.3 41.4 40.8 37.7 38.7
Tres,steel,max 38.9 35.2 37.9 35.1 35.6 34.6 32.4 33.2
Tres,steel,min -12.8 -12.2 -12.4 -11.9 -11.4 -11.1 -10.1 -10.1
Tres,con,max 39.0 39.7 38.1 38.8 36.0 36.8 33.1 33.9
Tres,con,min -22.0 -21.2 -21.1 -20.3 -19.1 -18.2 -16.2 -15.2
T1,heating 28.2 29.8 27.6 29.0 26.2 27.3 24.2 24.9
(T1-T2)heating 37.3 38.4 36.6 37.6 34.9 35.7 32.4 33.1
(T3-T2)heating 5.1 5.3 4.9 5.1 4.6 4.8 4.1 4.2
T1,cooling -6.5 -6.3 -6.2 -5.9 -5.5 -5.1 -4.3 -3.9
(T2-T1)cooling 23.6 22.5 22.5 21.5 20.1 19.1 16.7 15.7
(T2-T3)cooling 31.8 30.4 30.6 29.2 27.9 26.6 24.0 22.8

The estimates based on a 10-year sample are relatively close to those obtained
from the 45-year sample while the computational effort is significantly reduced. All the
discrepancies between them are within acceptable ranges. Therefore, a smaller size of

188
sample is sufficient for extreme value analysis without significantly affecting the
accuracy.

6.5 COMPARISON WITH AASHTO DESIGN SPECIFICATIONS

The results obtained in this study were compared with those provided in the
AASHTO LRFD Bridge Design Specification (2004). The comparison of the range of the
effective bridge temperature for the state of Texas is given in Table 6-6.

Table 6-6 Comparison of Effective Bridge Temperature

Return Period 100 Years 65 Years 16 Years 1 Year AASHTO (2004)


Maximum Te 120.8 120.0 116.6 112.4 120
Minimum Te -7.6 5.84 0.08 11.3 0

The upper temperature limit is equivalent with the value with return period of 65
years and also very close to the value with return period of 100 years. However, the lower
temperature limit given by AASHTO LRFD (2004) has only a return period of only 16
years, which is much higher than the value with return period of 100 years.
The positive vertical temperature gradient for Texas provided in The AASHTO
LRFD Bridge Design Specification (2004) was also compared with the gradient for return
period of 100 years obtained in this study as shown in Figure 6-15. The gradient
suggested by AASHTO is conservative in concrete deck. One reason is that the gradient
is based on a one-dimensional conduction analysis, which ignores the solar radiation
reflected by the ground. The reflected component reduces the positive vertical
temperature gradient by heating the bottom of the concrete deck.

189
37.3 37.3
(38.0) Concrete Deck (39.0) Concrete Deck
Proposed Proposed
AASHTO AASHTO

Steel Girder
Steel Girder

5.1
5.1
(0.0)
(0.0)

(a) Deck Thickness = 8 inch (b) Deck Thickness = 10 inch

Figure 6-15 Comparison of Vertical Temperature Gradient for Heating (F)

It should be noted that neither transverse thermal gradient nor vertical thermal
gradient for cooling is suggested in the AASHTO specification, so only the proposed
gradients are given in Figure 6-16.

-23.6
0.5h

Concrete
h
12"

Steel
-31.8

Figure 6-16 Proposed Vertical Temperature Gradient for Cooling (F)

190
The transverse linear temperature gradient during cooling with a return period of
100 years is approximately 18.9 F, which is on the same order of the vertical gradient.

The average temperature difference between the sunlit and shaded webs is as large as 44
oF for a return period of 100 years. Therefore, the impact of the transverse gradient on
bridge design also needs further assessment.

6.6 DISCUSSION AND SUMMARY

A description for the procedure for development of design thermal loads was
presented in this chapter based on the direct method suggested by Mae (1992). To get
extreme thermal loads for design purposes, the most critical configuration of bridge
sections was modeled for thermal analysis with Texas weather data from 1961 to 2005 as
the input environmental conditions. Four cities were considered to bound Texas weather
conditions. Based on the thermal analyses, a long period of sample data of thermal
parameters were used to describe the temperature field over a section. Extreme value
analyses of the sample data were performed to obtain the relationship between thermal
loads and return periods. The results are summarized in Table 6-3. The thermal loads
with 100 year return period were compared to the ones suggested by AASHTO. Based
upon this comparison the following conclusions can be reached.
The upper limit of effective bridge temperature is very close to that with return
period of 100 years. However, the lower limit given by the AASHTO LRFD
(2004) is equivalent with that with return period of 16 years. The current
AASHTO limit of 0 F should be lowered to -8 F to provide a 100 year return
period.

191
The positive vertical temperature gradient during heating provided in The
AASHTO LRFD Bridge Design Specification (2004) is conservative because
the effects of reflected solar radiation from the ground are ignored.
Transverse linear temperature gradient is on the same order of the vertical
gradient. The impact on bridge design needs further assessment.
The analyses in this chapter addressed only the unsurfaced bridge and did not
consider the effect of black-top such as asphalt wearing surfaces. Priestley (1976) studied
the effect of black-top thickness and pointed out that the black-top would induce a
smaller temperature gradient due to the insulation of the black-top. Based on his study, a
design gradient that varied with the black-top thickness was suggested.
In this chapter, only the thermal loads were developed based on the historic
weather data of Texas. The structural responses to thermal loading have not been
addressed. As discussed in Chapter 1, thermal loads will produce both thermal
deformations and thermal stresses in bridges. These effects are examined in the next
two chapters.

192
Chapter 7 Finite Element Structural Analysis

7.1 INTRODUCTION

Structural materials such as steel and concrete expand or contract with increasing

or decreasing temperatures, respectively. If a structural member subjected to a

temperature change is allowed to expand and contract freely, then no thermal stresses will

develop in the member. However, if thermal deformations are inhibited, large thermal

stresses can develop in various elements of the structural system. The largest thermal
stresses will be generated in a system that is fully restrained against thermal movements,

but significant stresses can also develop in a system that is partially restrained.

For structural systems with mixed materials such as a steel-concrete composite

girder, the response under thermal loads is further complicated. The coefficients of

expansion of a specific steel grade and a particular concrete mix are not typically

identical, however they are usually similar. The American Association of State Highway

and Transportation Officials (AASHTO, 2002, 2003, and 2004, 2007) stipulates a design

coefficient of thermal expansion of 6.510-6 in/in/F for steel, and a value of 6.010-6
in/in/F for concrete, which are typical values for these two materials. The fact that the

coefficient expansion for concrete is approximately eight percent less than the value for

steel means that there will be differential expansion in composite girders subjected to

uniform temperature changes. This difference will generate internal restraints to

thermal displacements that will lead to thermal stresses. Although this average coefficient

of thermal expansion of the two materials was used for simplicity when deriving the

decomposition of temperature field over a cross section, this difference will be

considered in the following finite element structural analyses. Based upon the results

presented in Chapter 5, significant thermal gradients are generated during daily thermal
193
cycles that may also produce significant thermal stresses within a steel-concrete

composite bridge. In addition to differences in the thermal properties, other factors such

as the geometry of a particular bridge as well as flexibility in the bearings and piers can

affect the bridges responses to thermal loads.

The magnitudes of thermal stresses can be large enough in some cases to justify

adding the stresses to those caused by dead and live loads during the design phase. Prior

field monitoring of three steel trapezoidal box girders showed that thermal stresses on the

order of 5 ksi are not uncommon in daily thermal cycles (Helwig et al. 2004, Helwig and

Fan 2000, Bobba 2003), which can be of the same order of stresses induced by such

construction activities as girder erection and concrete deck placement. Zuk (1965),

Dilger et al (1981, 1983), Soliman and Kennedy (1986), and Fu et al (1990), Muzumdar

(2003) reported thermal stresses of similar order in steel-concrete composite bridges.

Kennedy and Soliman (1987) pointed out that thermal stresses may result in magnifying

cracking of the concrete deck. Such cracking can significantly reduce the durability of the

bridge deck since it leads to corrosion of the steel reinforcement, spalling of the concrete

and further deterioration in bridge decks that may be subjected to salt-laden water that

can seep through the cracks. Although thermal stresses have not been found to play a

significant role in past structural collapses, the likelihood of problematic issues are

increasing since bridges are becoming longer to fit practical needs and lighter to get an

economic design. This reduces reserve strength for temperature induced stresses, which

makes undesirable cracking of concrete deck possible since the tensile strength of

concrete is relatively low.

In addition to thermal stresses, the bridge movements necessary to accommodate

the changes in the thermal environment are important factors to consider in bridge design.

Emerson (1976) reported significant thermal movement at the end supports. Glenn (2005)

194
and Writer (2007) measured longitudinal movements from changes in the thermal

environment as large as about 2 inches developed at the abutments for a typical three-

span continuous steel bridge during annual thermal cycles. Proper considerations of

thermal movements are of vital importance because they affect not only bearing and

expansion joint design, but also pier design as well. Expansion joints are often used at the

ends of bridge units to accommodate thermal movements along the length of the bridge.

Overestimating these movements often leads to costly bearings and larger expansion

joints. Underestimating these movements results in restraints that produce internal

forces, which may reduce the durability of the bridge and increase repair costs (Moorty,

1990).

Thermal stresses and thermal movements both can have significant effects on the

behavior and durability of a bridge. The importance of thermal effects on steel-concrete

composite bridge designs was emphasized by many researchers [Zuk (1965), Reynolds

and Emanuel (1974), Emerson (1979), Dilger et al (1981, 1983), Kennedy and Soliman

(1986), Fu at el (1990), Moorty and Roeder (1993) and Roeder (2003)]. However, the

effects of thermal loads on bridge behavior are not clear. Further investigation is needed.

In this chapter, a structural finite element model is developed and validated by the field

measurements taken from the Intercontinental Airport Bridge in Houston. In the next

chapter, the validated model is used for parametric analyses to assess the effects of bridge

geometry, bearing orientation and stiffness, pier flexibility, and locations of the fixed

points on the bridge superstructure.

7.2 THERMAL LOADS IN BRIDGE DESIGN

In Chapter 5, two methods were used to parameterize the temperature field over a

section. These methods of thermal analysis also suggest two methods for including
195
thermal actions in bridge design. The first method is to apply the effective bridge

temperature, vertical linear temperature gradient, and transverse linear temperature

gradient. These components induce continuity stresses within bridges. The magnitudes

are dependent upon the bridge geometry and the constraints imposed on the bridge. The

residual temperatures induce self-equilibrating stresses, which have been evaluated based

on extreme value analyses in Chapter 6. This method is illustrated in Figure 7-1 (a).

The other method is to apply effective bridge temperature, nonlinear vertical

temperature gradient, and transverse linear temperature gradient and then evaluate their

effects, which is illustrated in Figure 7-1 (b).

Thermal Loads Structural Responses


Vertical Linear Gradient Thermal Movements
Transverse Linear Gradient Continuity Stress
Effective Temperature Self-equilibrating Stress

Residual Temperature
(a) Method 1: Linear Components

(b) Method 2: Nonlinear Components

Figure 7-1 Inclusion of Thermal Loads

It should be noted that both the above methods assume a transverse linear gradient

that is not actual at all. Transverse temperature gradient is mainly contributed by the

outside web directly exposed to side sunshine as pointed out in Chapter 5, which is

196
significantly dependent upon the ratio of overhang length of concrete slab to steel girder

depth, the bridge position relative the sun, and bridge geometry. The transverse gradient

is also highly nonlinear along the width like the vertical gradient. A linear gradient is

only used to approximate the thermal effects for simplicity.

The first method only requires the application of the linear temperature

components, the effects of which can be included by any structural analysis programs

with beam elements. However, it should be noted that only the component of continuity

stress is calculated by conducting such a structural analysis with beam elements. To

calculate the total thermal stress, the self-equilibrating component should be added as

well. For example, if the total thermal stress at an arbitrary point (x,y) is desired, the self-

equilibrating stress has to be calculated from the same point (x,y) and added to the

continuity stress obtained from the structural analysis with beam elements. Therefore, to

calculate the stress field, the whole temperature field has to be known. In Chapter 6,

only the extreme residue temperature was evaluated. If the extreme self-equilibrating

stress is at the same location as the extreme continuity stress, it is impossible to evaluate

the magnitude of the total thermal stresses. The second method requiring the inclusion

of nonlinear thermal gradients is much more difficult. The typical beam element

generally has no capacity of including them. Special techniques are needed or a more

sophisticated structural analysis programs. However, the total thermal stress that is a

concern in bridge design is evaluated. Therefore, the second method will be used in this

dissertation.

As discussed in Chapter 6, the maximum and minimum effective bridge

temperatures occur during the summer and winter months, respectively. However, the

maximum and minimum vertical temperature gradients are essentially daily behavior and

may occur in almost in any given month. Therefore, a combination of effective bridge

197
temperature and vertical thermal gradients produces four critical temperature gradients

along the depth as shown in Figure 7-2. The first two cases represent the gradients during

heating and cooling when the maximum effective bridge temperature is achieved in

summer conditions. The last two cases represent the gradients during heating and cooling

when the minimum effective bridge temperature is achieved in winter conditions.

However, it should be noted that applying the full combination is too conservative since

results from the analysis summarized in Chapter 6 showed that the extreme effective

bridge temperature and the extreme vertical thermal gradients do not occur

simultaneously. Therefore, appropriate combination factors may be needed to get more

realistic thermal loads.

The transverse temperature load is of the same order in magnitude as the vertical

temperature load for steel-concrete composite bridges, so transverse temperature loads

also need to be considered. Transverse temperature loads have two cases: warmer interior

girder and warmer exterior girder. It should be noted that vertical thermal gradients do

not occur at the same time as transverse ones. When the maximum vertical gradients are

achieved, the transverse gradients are generally negligible. Therefore, the specific

combination of vertical and transverse temperature loads need not be considered. To sum

up, only the following cases need to be considered.

Heating in summer

Cooling in summer

Heating in winter

Cooling in winter

Transverse temperature gradients with warmer interior girder

Transverse temperature gradients with warmer exterior girder

198
Te,max Te,max
Tset Tset

(a) Heating in Summer (b) Cooling in Summer

Tset Tset
Te,min Te,min

(c) Heating in Winter (d) Cooling in Winter

Figure 7-2 Thermal Loads

7.3 BEARING TYPES

Sliding pot bearings are typically used at support locations on steel box girder

bridges in Texas to accommodate the girder movement expected at bearing locations,

which will be referenced as the bearing displacement. In this dissertation, the

displacements are in the plane of the bearing unless specified otherwise. In recent years,
199
elastomeric bearings are preferred in place of mechanical pot bearing in Texas due to

economics. Although a discussion of the types of bearings was provided in Chapter 1, a

brief overview on the bearings is provided in this section.

Pot Bearing

There are three main types of pot bearings, including fixed bearings (where

displacements in the plane of the bearing are restrained), unidirectional or guided

bearings (where displacements in a specified direction are restrained but displacements in

an orthogonal direction are unrestricted), and multidirectional bearings (where

displacements are not restrained in any direction). Figure 1-4 and Figure 1-2 show a

guided bearing and a multidirectional bearing.

The tops of both pots consist of a material such as Teflon so that sliding friction is

minimized when the girders move relative to the piers. Even with a low friction surface,

some friction does develop that inhibits the girder/bearing movement. Once the friction

is overcome, the bridge slides over the pot to accommodate the longitudinal bearing

displacements. Before the friction has been overcome, the friction force builds up, which

results in a longitudinal shear on the top of the pier. A bilinear model given in Figure 7-

3 (a) can be used to approximately characterize the behavior of pot bearings.

Since there is no relative slip between the pot and sole plate before sliding, the

bearing stiffness K1 is the combined stiffness of the bearing seat, masonry plate, and the

pot and sole plates. All these components have big section areas and small thicknesses, so

their shear stiffness is very large compared to that of the pier. The total stiffness of the

pier-bearing system prior to slip consists of the combination of the bearing and pier

stiffness in series.

200
1 1 1
= + (7-1)
K total K1 K pier

V V

Friction
K2 Friction

Shear
Shear

K1


Displacement Displacement
(a) Bilinear Model (b) Idealized Model

Figure 7-3 Models of Pot Bearing

K1 is generally much greater than kpier in a bridge, so it will not change the

stiffness of the pier-bearing system significantly. Therefore, prior to slip at the bearings

it is reasonable to assume infinite stiffness of bearing assembly. Once the friction is

overcome, the bearing maintains the shear resistance and its stiffness approaches zero

since the girder slips on the pot. Thus, K2 is assumed to be zero for simplicity.

Simplifying the bilinear model gives a simple rigid-plastic model as shown in Figure 7-3.

This model has only one parameter: bearing friction or resistance. In the subsequent finite

element analysis, this simple model is used to model the behavior of pot bearings. The

method of determining the bearing resistance is discussed separately later.

Elastomeric Bearing

201
Elastomeric bearings accommodate the longitudinal bearing displacements by

shearing deformation in the bearing. This is quite different from pot bearings, which

accommodate the displacements by sliding. Figure 7-4 shows typical longitudinal and

transverse shear deformations of a steel-concrete bridge in service. These two pictures

were taken at the direct connector interchange between US290 and IH35 in Austin, Texas

by Bradberry et al (2005).

SteelGirder Underside
ofgirder SolePlate


SolePlate

ElastomericBearing
BearingSeat

BearingSeat

(a) Longitudinal Deformation (b) Transverse Deformation

Figure 7-4 Deformation of Elastomeric Bearing (Bradberry et al, 2005)

Yura el at (2001) conducted extensive analyses and tests on the behavior of

elastomeric bearings. The shear stiffness tests showed that elastomeric bearings exhibit

a slightly nonlinear response for lower shear values (Yura et al, 2001). The values of the

shear stiffness depend on its definition. Fortunately, the load-displacement curve is not

highly nonlinear and largely linear instead. For simplicity, elastic behavior of elastomeric

bearings is assumed. The stiffness can be calculated by the familiar formula:

GA
k bearing = (7-2)
d

202
where A is the sectional area of bearing; d is the bearing thickness; and G is the

50-percent secant shear modulus as defined in the report (Yura, 2001). This modulus is

defined such that it gives the correct value of the maximum shear force when the strain

reaches the maximum design level (Yura, 2001). Obviously, this definition somewhat

underestimates the shear stiffness of bearings since in practice the working strain is lower

than the maximum design strain.

7.4 FINITE ELEMENT MODEL OF THE BRIDGE

The structural responses to thermal loads are governed by a set of partial

differential equations, the solutions of which are dependent on the boundary conditions

for the structural system. The equations were solved using a three-dimensional model

developed in the general finite element commercial program ANSYS v10.0. The

modeling details are discussed in the following sub-section.

7.4.1 Element Type (ANSYS Manual, 2005)

The element types to be used are listed in Table 7-1. A brief introduction of each

type is given as follows.

Table 7-1 Element Types to Be Used

Element Type Notes


Concrete Deck Solid95 20-node quadratic solid element
Steel Web and Flange Shell93 8-node shell element
Bracing Link8 3-d truss element
Pot Bearing Combin39 Nonlinear spring element
Elastomeric Bearing Beam4 3-d beam element
Pier Beam4 3-d beam element

203
Solid95

Solid95 is a 20-node Serendipity element with three degrees of freedom at each

node: translations in the x, y, and z directions. The element has quadratic shape functions

and therefore is well suited to model curved boundaries.

Shell93

Shell93 is a three-dimensional shell element. The element has six degrees of

freedom at each node: translations in the x, y, and z directions and rotations about the x,

y, and z-axes. The element has quadratic shape functions in both in-plane directions and

therefore is well suited to model curved boundaries.

Combin39, Beam4 and Link8

Combin39 is a unidirectional nonlinear spring element, which is defined by a

generalized force-deflection curve and can be used in any analysis. Beam4 is a three-

dimensional beam element that has six degrees of freedom at each node: translations in

the x, y, and z directions and rotations about the x, y, and z axes. Link8 is a space truss

element that has three degrees of freedom at each node: translations in the x, y, and z

directions.

7.4.2 Bearing Models

The simplified models described in Section 7.3 were used in the finite element

analyses. Although sliding bearings are often idealized as frictionless in the design

process, in reality friction does develop that produces a lateral shear at the top of the pier.

As a result thermal deformations are accommodated by a combination of sliding at the

204
bearing and lateral movement of the piers. The absolute pier deformation was

monitored in the field to obtain realistic measurements of the distribution of movement

between the bearing and the pier. The amount of movement in the pier is a function of

the resistance that develops in the bearing since this leads to a lateral shear applied at the

top of the pier. Data from the field measurements were used to determine the bearing

resistance and the elastic bearing stiffness for both pot and elastomeric bearings,

respectively. A discussion of the method used to determine these two parameters is

provided in this section.

Resistance for Pot Bearings

The resisting force that develops in the pot bearing, R, can be calculated as

follows:
R = K pier pier / n (7-3)

where Kpier is the pier stiffness; pier is the absolute movement of the top of the
pier; and n is the number of bearings on the pier. In multi-column bents, Kpier is the sum
of the column stiffness per bent and n is the total number of bearings on the bent cap.
The pier can be reasonably viewed as a cantilever beam with a stiffness of 3EI/H3, where
E is the modulus of elasticity of the material in the piers, I is the moment of inertia about
the axis of bending for the pier, and H is the height of the pier measured from the ground
surface to the location of the bearings. For the purposes of validating the finite element
model of the interaction between the super- and substructure elements, measurements
were taken in the field of the absolute movement at the tops of the piers. The elastic
deformation in the bearing before sliding was since it is generally insignificant compared
to pier flexure. Once the resistance is known, the friction coefficient in the sliding
bearing can be calculated as follows:
205
nR
= (7-4)
N
where N is the axial force on the pier. The friction coefficient is assumed an
intrinsic feature of bearing systems. Five piers in four bridges were measured. The
calculated resistances and friction coefficients according to the above method are listed in
Table 7-2. Since the accuracy of the laser device that was used to measure pier
deflections is 1/16 inch, the maximum errors of the resistances and friction coefficients
may be as high as 25%.

Table 7-2 Resistance and Friction Coefficient of Pot Bearings

Bridge Name Bent No. nR (k)


Intercontinetal Airport 2 175.1 0.132
Bridge. 4 109.9 0.128
Woodlands (North) 2 130.3 0.159
Woodlands (South) 2 113.5 0.139
Hardy Toll Road 2 111.9 0.093

Stiffness for Elastomeric Bearings

As discussed before, the stiffness of elastomeric bearings can be calculated by


Equation 7-2. According to Yuras report (2001), the shear modulus typically ranges
between 100 to 200 psi. Bradberry (2005) reported the dimensions of elastomeric
bearing that are typically used in Texas. According to his summary, the minimum
sectional size of elastomeric bearings is 13" 20" while the maximum is 23" 26". The
thickness of the bearings typically used in box girders usually range from 3 to 10 inches.
Therefore, the stiffness of each elastomeric bearings has a range of 3 ~ 40 k/in. In the
next chapter, results are presented from parametric investigations utilizing this range of
bearing stiffness to assess the impact on the bridge behavior.

206
7.4.3 Meshing Density

As outlined in Section 4.3.3 of this dissertation that focused on the thermal


analysis, the element density has a significant effect on the accuracy of the results.
Numerical tests were used on this study to determine the necessary mesh density to
produce good results. To determine whether the element size is sufficiently fine, the
number of elements is incrementally increased and comparisons are made between the
individual analyses. The results obtained from the model with a certain number of
elements are compared to those obtained from the model with double elements. If no
significant difference is detected between them, the mesh can be deemed fine enough. It
should be noted that high-order elements are used to save elements. The element aspect
ratio is another meshing factor to influence the accuracy of the results. To avoid
excessive distortion of elements, the aspect ratio of elements was limited to 5 in this
investigation.

7.5 VALIDATION OF ANSYS MODEL

To validate the ANSYS structural model, the calculated pier movements and the
bearing movements relative to the pier were compared to those obtained from the field
measurements taken from the fully instrumented bridge located at Intercontinental
Airport. Detailed geometrical information about this bridge is given in Chapter 3. The
bearing orientations are assumed to be perfectly chord-guided which is the orientation
often used as outlined in Chapter 1. The sizes of the piers were obtained from the
original construction drawings. The ANSYS model is shown in Figure 7-5.
The material properties used in the model are listed in Table 7-3. The bearing
resistances are calculated using the above method and given in Table 7-4.

207
Table 7-3 Material Properties

Steel Concrete
Youngs Modulus (ksi) 29000 3420
Poissons Ratio 0.3 0.2
Coefficient of Thermal Expansion (10-6/F) 6.5 5.5
Density (lb/ft3) 490 150

Table 7-4 Bearing Resistance of Intercontinental Airport Bridge

Abutment 1 Bent 2 Bent 3 Bent 4 Abutment 5


Resistance (k) 64.8 175.1 Fixed 109.9 46.8

Abutment 5

Bent 4

Bent 3

Bent 2

Subtended Angle

Abutment 1

Figure 7-5 ANSYS Model of Intercontinental Airport Bridge

The temperature change was taken as a 60 F increase to match the temperature


change between the two sets of field measurements. The comparison between the field
measurements and the FEA solutions are given inTable 7-5.

208
Table 7-5 Comparison between Calculated and Measured Displacements

Pier Movement (in) Relative Bearing Movement (in)


FEA Measurement FEA Measurement
Abutment 1 N/A N/A 1.66 1.64
Bent 2 0.24 0.26 0.70 0.78
Bent 3 0.12 0.15 0 0
Bent 4 0.54 0.49 0.29 0.44
Abutment 5 N/A N/A 1.28 1.15

It can be seen from the table that the FEA model gives reasonable estimates of
both pier movements and bearing movements relative to the piers in comparison with the
measurements. It should be noted that the wax measuring devices used to measure the
relative bearing movements have an accuracy of 1/8 inch and the accuracy of the laser
devices used to measure pier deflections is 1/16 inch.
The vibrating wire strain gages that were used on the Intercontinental Airport
bridge are significantly affected by sunshine and the measured stresses are not reliable in
instances where significant temperature variations occur in the gages (Writer, 2007). Due
to the larger variations in temperature in the field, no correlation between the measured
stress and the calculated values can be achieved. Therefore only calculated stresses are
presented here. Besides thermal movements and thermal stresses, the pier shear and the
pattern of thermal movement are also calculated. These results are also presented here.

Stationary Point

When orienting the bearings, it is often assumed that one of the interior piers is
fixed and all bearings are oriented along the chord line back towards the fixed pier.
However, due to pier flexibility and the friction that develops at the sliding bearings, the
fixed pier is not actually stationary. Both field measurements and the finite element

209
analysis show that the true stationary point is not at the assumed fixed pier. Certainly,
there exists such a stationary point since the tangential movement U varies continuously
from a negative value at one end to a positive value at the other end as shown in Figure
7-6, which shows graphs of the contour of tangential displacements in the concrete deck.

(a) Tangential Displacements (b) Radial Displacements


Figure 7-6 Contour of Tangential and Radial Displacement in Concrete Deck

It can be seen from the contour of tangential displacements that the actual
stationary point is along the zero tangential displacement line as indicated by the dash red

210
line. To determine the stationary point, the contour of radial displacement Ur is also
needed, which is shown in Figure 7-6. The stationary point should be also in the zero
radial displacement line as indicated by the dash red line. The intersection of both zero
lines is the stationary point, which is marked in Figure 7-6. This stationary point is
located 28 ft away from the assumed fixed bent 3 towards the center of mass of the
bridge in the longitudinal direction. In the radial direction, this stationary is about 5 ft
offset the bridge axis towards the center of the circle. However, it should be noted that
zero radial displacement may lie outside the girder depending up the bridge geometry,
pier flexure and bearing stiffness. In such situation, it is difficult to locate the stationary
point according to finite element results. Detailed discussions are given in Chapter 8.
The maximum tangential stress component in the bottom flange of steel girders is
1.89 ksi and the maximum principal tensile stress in concrete deck is 0.25 ksi. The shears
carried by the bearings are listed in Table 7-6.

Table 7-6 Bearing Shear (k)

Abutment 1 Bent 2 Bent 3 Bent 4 Abutment 5


74.6 166.0 62.0 114.0 54.4

7.6 SUMMARY

A description of the ANSYS finite element model for structural analysis that was
conducted on a composite bridge under the action of thermal loads was presented in this
chapter. The calculated pier movements and the bearing movements relative to the pier
were compared to those measured on the fully instrumented Intercontinental Airport
bridge. The comparison indicates that the model produced good estimates of the thermal
movements. The stationary point and the pattern of thermal movement are discussed in
detail. The results show the true stationary point is not at the assumed fixed pier but
211
instead lies somewhere between the assumed fixed pier and the center of mass of the
bridge.

212
Chapter 8 Structural Responses to Thermal Loads

8.1 INTRODUCTION

Thermal loads produce thermal movements and thermal stresses in bridges.

Thermal stresses impact the design of the steel components of these bridges while

thermal movements at supports are the basis of bearing and expansion joint design.

Thermal stresses can be divided into self-equilibrating stresses and continuity stresses.

The latter depends on bridge type, bearing type, support conditions, as well as the thermal
loads. The most frequently used cross sections for composite steel girders consist of box-

girders and I-girders. Two types of bearings are commonly used in Texas: pot bearing

and elastomeric bearing. Although the bearings are often idealized as directionally free

or fixed in design, in reality friction develops that restrains the girders. The amount of

friction in the bearings is dependent on several factors including unintentional

misalignment. Even when the bearing alignment follows exactly the direction specified

in the plans, the results in Chapter 7 showed that in curved bridges the fixed pier is

actually not the stationary point along the bridge length. This means that the directional

guides should actually be oriented towards a point between the fixed pier and the center

of mass of the bridge. The actual stationary point is a function of the amount of friction

in the bearing as well as the flexibility of the piers. In addition to friction at the

bearings, unintended restraints may also exist at the ends of the bridge from debris in the

expansion joints that is not readily compressed. The impact of these restraints on the

bridge behavior and the magnitude of the continuity stresses are not clear.

In addition to the direction of the lateral guides on the bearings, there are a variety

of other questions that have been raised about bearing details. The recent detailing

practice for bearings in twin box girder systems is to have one girder guided while the
213
second girder is supported on multi-directional bearings. The Association of General

Contractors (AGC) of Houston has expressed a preference for using larger numbers of

guided bearings to increase stability during construction. The layout and orientation of

the lateral guides is an area that is not well understood. The current practice of directing

the guides along a chord to the fixed layout is based upon the deformations that occur

when the bridge is subjected to a uniform temperature change. As shown in Chapter 7,

the girders actually experience significant temperature change and the combination of

bearing friction and pier flexibility often result in the stationary point positioned between

the fixed bearing and the center of mass of the bridge.

Results from a detailed parametric investigation on impact of bearing details on

the thermal deformations are presented in this chapter. The impact of errors in the guide

placement, as well as the behavior of utilizing a tangent layout for the bearings, is

investigated. The use of a tangent layout could reduce the required size of the pier caps

as shown in Figure 1-11. A method that can be used to estimate the location of the

stationary point on the bridge is also developed.

8.2 BASIC BRIDGE MODEL

The general finite element model outlined in Chapter 7 was used to conduct a

detailed parametrical analysis on the impact of thermal loads on the structural behavior of

composite steel bridges. The basic model was based on the twin trapezoidal box girder

bridge located near downtown Houston that was instrumented. The three-span

continuous bridge has dimensions as shown in Figure 8-1. The three spans had arc

lengths of 180, 300 and 180 ft along the girder length (measured along the bridge

centerline). The reinforced concrete piers in the analysis had cross-sections of 6 ft 4 ft

with a height of 20 ft. The piers were aligned in the radial direction. The cap is of the
214
same section size as the piers. Although the actual radius of horizontal curvature of the

bridge was approximately 1200 ft, the value used in the analysis was 600 ft. The material

properties are listed in Table 8-1. The basic twin girder system with three spans was

selected since this basic geometry has been widely used on highway interchanges in

Texas over the past 15 years.

Figure 8-1 Section Sizes for Parametric Structural Analysis

Table 8-1 Material Properties for Parametric Structural Analysis

Steel Concrete
Youngs Modulus (ksi) 29000 3625
Poissons Ratio 0.3 0.2
Coefficient of Thermal Expansion (10-6/F) 6.5 5.5
Density (lb/ft3) 490 150

Bradberry (2005) pointed out that the Texas Department of Transportation has

moved toward the use of elastomeric bearings for steel trapezoidal box girders in place of

pot bearings. As discussed in Chapter 7, pot bearings behave quite differently than

elastomeric bearings as given in Figure 7-3 and Equation 7-2. Therefore, they have to

be treated separately. Only elastomeric bearings were focused on in this chapter. For

elastomeric bearings, the elastic stiffness Kbearing has to be input. The value of it for each
215
bearing was taken as 20.7 k/in, which is the average of the lower and upper stiffness

ranges of the elastomeric bearings that are typically used in Texas as discussed in Section

7.4.2. In the layout of the bearings, one of the interior supports is assumed as a fixed

bearing that restrains both the longitudinal and lateral movements of the girders. Multi-

directional bearings that do not restrain the lateral or longitudinal movement were

modeled at the other interior supports. At the exterior supports a guided bearing was

provided under one of the boxes while a multi-directional bearing was placed under the

other girder. The orientations of the guided bearings are directed along a chord line back

to the fixed support as shown in Figure 8-2.

Fixed
Multi-direction Guide
Single Direction Guide
Figure 8-2 Chord Oriented Bearing Toward the Fixed Pier

The thermal loads were assumed to be those for heating in summer with the 100-
year return period obtained from the thermal analysis presented in Chapter 6. The upper

limit of temperature range is 120 F. The setting temperature at girder erection was

assumed to be 60 F. Therefore, the uniform temperature loads are a 60 F rise. The

vertical temperature gradient during heating through the depth was taken using the curve

with 100-year return period proposed in Chapter 5 and repeated in Figure 8-3. The

thermal load described here was used in the subsequent analyses unless specially stated

otherwise.

216
38 F
Concrete Deck
4
y
T = 381
16

Steel Girder

5 F

Figure 8-3 Vertical Temperature Distribution Over the Depth

The model described above is viewed as the reference model. The values of all

the parameters given in this model will be used in the subsequent analyses unless

specially specified otherwise. It should be also noted that when investigating the

influence of a given parameter, the other parameters were held constant. The results

obtained from the parametric analysis in the subsequent sections will be compared to the

corresponding ones obtained from this reference model.

8.3 DEFINITION OF QUANTITIES UNDER CONSIDERATION

Several terms are used in the subsequent presentation of the results from

parametric analyses and therefore a clear definition of these terms is warranted. These

terms include the following:

Movement: Magnitude of resultant displacement vector at the specified point

of interest.

Tangential Movement: Tangential component of displacement vector at the

specified point.

217
Radial Movement: Radial component of displacement vector at the specified

point

Bridge Movement: Magnitude of resultant displacement at the bottom flange

location of the girders. This movement must be accommodated by the

bearing and pier system. It should be noted that the displacements in the

bottom flanges are different from those at the top of concrete deck because the

girders are fairly deep and the rotations are not negligible. In other words, the

displacement varies with the depth of girder.

Support Shear: Resultant shear force carried by the bearings at the abutments

or resultant shear force carried by interior piers.

Guide Force: Total force that the guides exert on the girder when the bridge

displacements at the bearings are not in the direction parallel to the guide.

Bottom Flange Stress: Tangential component of stress tensor at a point on the

bottom flange. Tensile stresses are taken as positive while compressive

stresses are negative.

Concrete Stress: The first principal stress at a point in the concrete deck.

Tensile stresses are taken as positive while compressive stresses are negative.

8.4 FIXED POINT, BEARING ORIENTATION AND GUIDE NUMBER

While the design and characteristics of guides are addressed in both Section

14.7.9 of the American Association of State Highway and Transportation Officials

(AASHTO) Load and Resistance Factor Design (LRFD) Bridge Design Specifications

(2004) and in Section 14.6.9 of the AASHTO Standard Specifications for Highway

Bridges (2002), neither specification discusses the requirements for proper orientation of

the guides. However, the orientation of bearing guides is addressed in the AASHTO
218
Guide Specifications for Horizontally Curved Steel Girder Highway Bridges (2003) in

Commentary Section C3.6 which states:

Orienting bearing guides toward a fixed point and allowing the bridge
to move freely along rays emanating from the fixed point causes thermal
forces to be zero if the structure changes temperature uniformly.
However, there are other load conditions that may affect decisions
regarding bearing orientation.

Although the specification does not give a location for the fixed point, the point

is usually assumed to be a fixed bearing in the line of the continuous beam.

Similar instructions on guide orientation are given by the National Highway

Cooperative Research Program (NCHRP) Report 424, Improved Design Specifications

for Horizontally Curved Steel Girder Highway Bridges, in Section 8.3 which states:

Thermal forces due to a uniform temperature change will not be


generated if the bearings are oriented to permit free translation along
rays emanating from a single point. Any other orientation will induce
thermal forces in both the superstructure and substructure. Practical
considerations, however, often do not allow the bearings to be oriented in
such a manner.

The report does not specify the location of the origin of the rays.

A variety of bearing layouts, in terms of the specific numbers of fixed,

multidirectional, and unidirectional bearings have been historically been employed in

Texas bridges. A system like the one shown in Figure 8-2 has been recently used in the

construction of a twin box girder bridge in Houston (Helwig et al., 2004). As shown in

the figure, at the location of one interior pier, fixed bearings were positioned under both

girders. At the other supporting locations a multidirectional bearing was located under

the interior girder and a unidirectional bearing was used under the exterior girder. The

219
guides on the unidirectional bearings were oriented on a chord line to the fixed support to

permit free expansion under uniform thermal loads. As indicated by the AASHTO Guide

Specifications for Horizontally Curved Steel Girder Highway Bridges, the thermal forces

will be zero only if the temperature change is uniform. This rule is true for a simply

supported curved girder made of one material under uniform temperature changes, which

is illustrated graphically in Figure 1-6. Alternatively, the movements can be

mathematically calculated as follows.

y x
B
y
s B (x2,y2)

Fixed A (x1,y1)
x

Figure 8-4 Displacement of A Simple Curved Girder Subjected to Uniform Temperature

As shown in Figure 8-4, a simple curved girder AB with the end A (x1,y1) fixed

and the other end B (x2,y2) free is subjected to uniform temperature change T. Under

the action of this thermal load, the end B moves to B. The thermal movement in the x

direction can be calculated as:

x
dx 2

x = T cos ds = T ds = Tdx =T ( x2 x1 ) (8-1)


s s
ds x1

Similarly,

220
y = T ( y2 y1 ) (8-2)

where s is the arc length along the axis of curved bridges. It can be seen from
above that the direction of the movement is along the chord. Unfortunately, the

assumptions implied in these equations are far from accurate representations of the actual

problem. First, the fixed bearing is assumed to be the point which the bridge expands

and contracts about due to the thermal loads. The piers supporting the bridge are

usually relatively tall and thereby very flexible. Although the sliding bearings are often

idealized as frictionless, significant forces develop in the bearings as a result of friction

between the sliding surfaces. The friction is often intensified due to improper

orientation of the guides as is subsequently explained below. Due to the lateral forces at

the bearings, the piers tend to deflect laterally and the fixed pier is actually not the

stationary point on the bridge. Secondly, by nature composite bridges are made of two

materials with very different properties. Steel is a good thermal conductor while concrete

is a good thermal insulator. This difference causes significant temperature gradients

within a girder due to solar radiation. Therefore, the thermal loads acting on the girder are

not uniform. If the temperature of a girder is not uniform, points on the girder do not

tend to move in a straight path towards the fixed bearing. Consequently, if the lateral

guides dictate movement along the chord towards the fixed bearing, thermal stresses

develop in the bridge. Since the guides are not pointed towards the actual stationary

point on the bridge, friction in the guides is increased further shifting the stationary point

away from the pier with the fixed bearing. Lastly, most curved bridges do not behave

like an individual curved beam. The actual system is often continuous over several

supports and consists of multiple girders joined by the concrete deck as well as cross-

221
frames and/or plate diaphragms. Such a complex system likely behaves differently from a

simply supported curved girder.

Because of the uncertainty in guide orientation on the bearings, some bridge

designers have raised questions on the possibility of using bearing guides with an

orientation tangent to the girder curvature at the supporting piers. Such a practice could

reduce the necessary size of the pier caps on the bridge as illustrated in Figure 1-11.

The width of the pier cap is often controlled by the footprint of the bearing with the

required guide orientation. Bearings with tangentially oriented guides would be easier to

install and use of tangential guides would permit use of smaller pier caps on some

bridges. It should be noted that the AASHTO Guide Specifications for Horizontally

Curved Steel Girder Highway Bridges (2003) cautions against pointing the guides

tangentially. Section C3.6 of this specification states:

The orientation of bearing guides and the freedom of bearing movements


is important to thermal forces. Sharp curvature and sharply skewed
supports can cause excessive lateral thermal forces at supports if only
tangential movement is permitted.

The proper orientation of the bearing guides has not been clearly established due

to the many variables that affect the problem. The impact of misalignment of the

bearings, bearing friction, pier flexibility, and expansion joint restraint due to debris has

not been studied. Establishing the proper orientation to minimize thermal restraints

based upon a parametric structural analysis is necessary to properly understand the

problem of thermal expansion and contraction. Once the proper orientation is

established, the impact of slight misalignments of guided bearings as well as the

tangential orientation can be properly assessed. Such a study will provide designers

with a clear understanding of the problem as well as various alternatives to improve the
222
economics of curved girder designs. The structural FEA model validated in the last

chapter was used to complete this parametrical study.

8.4.1 The Point of Fixity

To establish the proper point for fixity, two respective cases were examined in

which the fixed bearing was located at an interior pier and then an abutment at the end of

the bridge. In the first case, the bearings were chord-oriented toward interior bent 2 as

illustrated in Figure 8-2. In the second case, the bearings were chord-oriented toward

Abutment 1 as illustrated in Figure 8-5. Abutments at the ends of bridges are extremely

stiff and are assumed rigid in the analysis. As outlined earlier in this chapter, the piers

had a cross section of 6 ft x 4 ft and a height of 20 ft. Therefore the intermediate piers

introduce some flexibility into support system when the fixed bearing is located on the

pier. In the case shown in Figure 8-5, the stationary point must be at abutment 1 since

the support point is rigid. In the case shown previously in Figure 8-2, although the

relative movement between the bridge and the top of the pier is zero with the fixed

bearing at this location, the bridge system can still displace due to the pier flexibility.

Fixed
Multi-direction Guide
Single Direction Guide

Figure 8-5 Chord-oriented Bearing Toward Abutment 1

Contour plots of the radial and the tangential movements of the top surface of the

concrete deck are shown in Figures 8.5 and 8.6. The locations of the bents and

223
abutments are indicated in the figures. In both cases the movements in the tangential

and radial directions are indicated on the respective figures.

The results shown in Figure 8-7 demonstrate that when Abutment 1 is fixed and

the bearings are chord-oriented toward this abutment, the zero-tangential displacement is

at Abutment 1, while the maximum tangential movement is at the other abutment. When

bearings are chord-oriented towards Bent 2, as shown in Figure 8-6 the zero tangential

displacement occurs between Bent 2 and the middle (center of mass) of the bridge. The

farther the location is away from this zero displacement point, the larger the tangential

displacement is. The maximum tangential movement for the second case is much larger

than the first case. For both cases, the maximum radial movements are at the corner of

abutment 4, but the magnitude is much larger when the bearings are oriented toward

Abutment 1.

(a) Tangential Movement (b) Radial Movement


Figure 8-6 Contour Plots when bearings are chord oriented toward bent 2

224
(a) Tangential Movement (b) Radial Movement
Figure 8-7 Contour Plots when bearings are chord oriented toward abutment 1

Table 8-2 compares the maximum guide force, bridge movement, support shear

and bottom flange stresses for both cases.

Table 8-2 Comparison of Maximum Results for the Two Cases

Toward Bent 2 Toward Abutment 1


Guide Force (kip) 39.7 60.8
Bridge Movement (inch) 1.77 2.53
Support Shear (kip) 78.2 103.6
Bottom Flange Stress (ksi) 1.59 2.89

The comparison shows that when bearings are oriented toward the fixed

abutment, the maximum guide force is 53% higher, the maximum bridge movement 43%

larger, the maximum support shear 33% higher and the maximum bottom flange stress

225
82% higher. When elastomeric bearings are used in bridge design, the thickness of

bearings is usually determined according to the shear deformation demands. At

abutments, the bridge movements must be accommodated by the shear deformation in the

elastomeric pads. Smaller deformation demands are preferred. For pot bearings,

increases in the guide forces, support shears and bottom flange stresses are also not

desirable. Since these demands are increased when bearings are oriented toward

Abutment 1 compared to the Bent 2 orientation, the bearing orientation towards a fixed

abutment can lead to poor behavior. In addition, the demand on the expansion joint at

abutment 2 goes up significantly when the bearing is fixed at one of the abutments.

Therefore, the common practice of positioning the fixed bearing at one of the flexible

interior bents is the better choice.

8.4.2 Bearing Orientation

Results presented in the previous section demonstrated that positioning fixed

bearings at interior bent is more desirable than fixed abutment. However, the proper

orientation of bearings still must be assessed, which is the topic of this section. Three

bearing orientation schemes were investigated:

Case 1: chord oriented toward the fixed interior pier as illustrated in Figure 8-2
Case 2: chord oriented toward the stationary point as illustrated in Figure 8-8
(required an iterative analysis).

Case 3: tangentially oriented as illustrated in Figure 8-9

226
As noted above, a trial and error procedure was necessary for Case 2 since the

stationary point is unknown. Estimating the location of this stationary point is revisited

in detail later in this chapter.

Fixed
Multi-direction Guide
Single Direction Guide

Figure 8-8 Chord-oriented toward stationary point

Fixed
Multi-direction Guide
Single Direction Guide

Figure 8-9 Tangentially Oriented Bearing

The patterns of tangential movements of the top surface of the concrete deck are

similar for all the three cases as shown in Figure 8-6. The patterns of radial movements of

the top surface of the concrete deck when bearings are chord-oriented toward stationary

point are also similar to the case when bearings are chord-oriented toward Bent 2. It

should be noted that the word similar pattern used here and in the subsequent sections

means that the displacement distributions are similar although the magnitudes may be

quite different. A contour plot of the radial movements of the top surface of the

concrete deck when bearings are tangentially oriented is shown in Figure 8-10. Unlike

the other two cases, the maximum radial movement is not at the corner of Abutment 4 as

shown in Figure 8-6, but is instead at the center of the bridge. The magnitude is also
227
much smaller than for the other two cases. Obviously, part of the radial displacement is

inhibited by the rigid guides while the other part is accommodated by the bearing/pier

system.

Figure 8-10 Contour Plots of Radial Movement when bearings are tangentially oriented

Table 8-3 compares the maximum guide force, bridge movement, support shear

and bottom flange stresses for all the three cases.

Table 8-3 Comparison of Maximum Results for the Three Cases

Toward Toward Tangentially


Bent 2 Stationary Point oriented
Guide Force (kip) 39.7 23.8 34.9
Bridge Movement (inch) 1.77 1.77 1.83
Support Shear (kip) 78.2 72.2 74.3
Bottom Flange Stress (ksi) 1.59 1.56 1.69

The comparison shows that the maximum bridge movement, support shear and

bottom flange stresses are roughly similar for all three cases considered. The support

228
shear and the bottom flange stress are slightly smaller when the bearings are chord-

oriented toward the stationary point. However, the maximum guide force when the

bearings are chord-oriented toward the stationary point is 67% smaller than the case with

the bearings chord-oriented toward Bent 2 and 47% smaller than the case with the

tangential orientation.

When the bearings are tangentially oriented, the bridge displacements are not

along the tangent as can be seen from Figure 8-10. Radial movements at the locations

of bearings are not zero, which is accommodated by the pier movements. When the

bearings are chord-oriented toward Bent 2, the resultant bridge movement at the bearings

is not along the chord from the fixed bent 2, but is instead a few degrees off the chord

line from bent 2 due to the pier deflections. When the bearings are chord-oriented toward

the stationary point, the resultant bridge displacements are nearly along the chord from

the stationary point as shown Figure 8-11. In the figure, the red lines are the actual

directions of girder movements at the supports. The assumed directions as indicated by

the blue lines are also plotted for comparison.

Figure 8-11 Actual Direction of Thermal Movement at Supports (To scale)


229
When determining the stationary point, the values of the bent stiffness and the

bearings stiffness are considered to make the guides parallel to the direction of the

resultant bridge movement. Therefore, the actual displacements at the bearings are closer

to the desired direction and consequently the guide force is also smaller than the other

two cases.

Based upon the comparison of the results with the three different guide

orientations, the forces and stresses that occurred with the tangential bearing orientation

were not significantly different than those occurring using the current methods of

orienting the guides towards the fixed bearing. However, with the tangential guide

orientation the radial bridge movement was primarily accommodated by pier deflection.

When the bridge is very wide and sharply curved, the radial and tangential movements

will be comparable. If such a large movement is accommodated by a stiff pier, a large

internal force will be accordingly generated and such an orientation may lead to poor

behavior. The best behavior was achieved by orienting the bearings towards the true

stationary point since the bridge movement experiences displacements in the expected

directions and smaller guide forces are achieved. However, in this analysis the

stationary point was not known and was determined by trial and error. During the

design process, such a method is not desirable and instead a rational procedure is needed

to locate this stationary point. This issue is addressed in the following section.

8.4.3 Number of Guided Bearings

To assist in maintaining stability during construction, the Association of General

Contractors (AGC) of Houston has raised questions about the impact of utilizing larger

numbers of guided bearings. The guided bearings increase the redundancy of the
230
support system and may create a safer system during erection and construction. The

example shown previously in Figure 8-2 utilized two fixed bearings, three

multidirectional bearings, and three unidirectional bearings. The example depicted in

Figure 8-12 shows the case with replacing the three multidirectional (free) bearings with

three unidirectional (guided) bearings preferred by AGC.

Fixed
Multi-direction Guide
Single Direction Guide
Figure 8-12 Bearing Layout Preferred by AGC

Although the use of more guided bearings may assist in maintaining stability

during intermediate girder construction stages and allow the contractor more flexibility in

construction procedures, the recent trend in laying out guides to optimize the structural

behavior under thermal loads is to use fewer guided bearings, so that there is less restraint

against thermal movements in the bridge. In Section C3.6 of the AASHTO Guide

Specifications for Horizontally Curved Steel Girder Highway Bridges (2003) it is noted

that wide bridges are particularly prone to large lateral thermal forces because the
bridge expands radially as well as longitudinally. This comment would seem to support
using fewer, rather than more, lateral bearing guides. The number of guided bearings

utilized on steel bridges is also an issue that was investigated in this study. The above

bearing layout proposed by the AGC was examined using the above basic model. Two

cases were considered, the first of which consisted of guiding only the exterior girder as

shown in Figure 8-2. In the second case, both girders are guided as shown in Figure 8-12.

231
The patterns of the radial and tangential movements of the top surface of the

concrete deck are similar for both cases as shown in Figure 8-6. Table 8-4 lists the

maximum guide force, bridge movement, support shear and bottom flange stresses for

both cases.

Table 8-4 Comparison of Maximum Results for the Two Cases

Only Exterior Girder Guided Both Girders Guided


Guide Force (kip) 39.7 271.5
Bridge Movement (inch) 1.77 1.76
Support Shear (kip) 78.2 82.4
Bottom Flange Stress (ksi) 1.59 1.71

The comparison shows that the maximum bridge movements are similar for both

cases. The bottom flange stress and the maximum support shear are slightly larger when

both girders are guided. As stated in Section C3.6 of the AASHTO Guide Specifications

for Horizontally Curved Steel Girder Highway Bridges (2003), the bridge expands in

both the tangential and radial directions. If both girders are guided, the radial

movements must be accommodated by the rigid guides and larger internal forces are

generated in the guides. When both girders are guided, the guide forces dramatically

increase from 39.7 kips to 271.5 kips. The parametric analyses show this situation
becomes even worse when the bridge becomes wider and the degree of curvature

increases. Therefore, the current practice of only guiding one girder results in much better

performance and guiding both girders is not recommended.

8.5 ESTIMATION OF THE STATIONARY POINT AND THERMAL MOVEMENTS

When designing a continuous bridge system, an important aspect is providing

expansion joints and support bearings that can accommodate the displacements that occur
232
due to daily and annual thermal variations. When thermal expansion and contraction are

overestimated, the design requirements for the bearings and expansion joints also

increase. Accommodating these movements can significantly reduce the bridge

economy since the cost of the bearings and expansion joints increase with higher

displacement capacities. Underestimating these movements leads to unaccounted

restraints that may produce large internal forces, which may increase repair costs and

reduce the durability of bridge (Moorty, 1992). Therefore, accurate estimates of thermal

movements at abutments are of primary importance in the design of expansion joints and

bearings.

The current method is to assume thermal movements are along the chord line

between the bearing and the fixed pier as illustrated previously in Figure 8-2. The typical

design methodology assumes the piers are rigid and evaluates the thermal movement as

the product of the thermal expansion coefficient, the design temperature change and the

chord length using Equation 1-2. When the pier is relatively flexible, this methodology

yields inaccurate calculations of the bearing movements and can also affect the demand

on the expansion joints. Recorded bearing displacements by Grisham (2005) indicated

that the bearings supporting the bridge girders on the piers often do not work as desired.

Equation 1-2 often overestimates the movement at one end and underestimates the

movement at the other end. The stationary point about which thermal expansions and

contractions take place is not located at the fixed bearings as assumed.

8.5.1 Rigid Body Model

Results presented in the previous sections of this chapter have shown that the

actual thermal movements are along the line between the bearing and a stationary point.

To reduce the guide force, the bearings should be chord-oriented toward this stationary
233
point. However, determination of this point required a trial and error procedure that is not

practical for design. In this section, a simple model, referred to as a rigid body model, is

developed to estimate the location of the stationary point. The following assumptions

were made in the development of the model:

Axial deformations due to thermally induced stresses are neglected (girders


assumed rigid). The girders experience only shortening or elongation only

due to thermal changes following Equation 1-2.


The reaction in the plane of the fixed bearings passes through stationary point.
The bearings are oriented towards a stationary point that is to be determined.
The orientation is illustrated in Figure 8.12. The guides are oriented parallel to

the line connecting the stationary point and the bearing so that no force is

generated in the guide.

The magnitudes of the thermal movements are along rays emanating from the
stationary point and are proportional to the distance between the stationary

point and the bearing under consideration.

Figure 8-13 Rigid Body Model

234
To locate the bearing, a Cartesian coordinate system is created. The origin is set

at the center of circle and the y-axis is aligned with the symmetry line of the bridge. From
geometry, the coordinates can be determined as follows:

A: (R1cosA, R1sinA) E: (R2cosE, R2sinE)


B: (R1cosB, R1sinB) F: (R2cosF, R2sinF)
C: (R1cosC, R1sinC) G: (R2cosG, R2sinG)
D: (R1cosD, R1sinD) H: (R2cosH, R2sinH)
S: (rcos, rsin)

The Ri and i are the coordinates in the corresponding polar coordinate system.

The value (r,) is the coordinate of stationary point in the polar coordinate system that is

to be determined.

Since it is assumed that thermal movement is proportional to the chord length, the

equilibrium equation can be developed as follows:


K A SA+ K B SB + K C SC + K D SD + K E SE + K F SF + K G SG + K H SH = 0 (8-1)

The values of K in the equation are the combined stiffness of the bearing/pier
system. Writing this vector equation in terms of its components produces the following

expressions:

D H

K iX (R1 cos i r cos ) + K iX (R2 cos i r cos ) = 0


i= A i=E
D H
(8-2)
K (R sin r sin ) + K (R sin r sin ) = 0

i= A
iY 1 i
i=E
iY 2 i

235
From these two equations, the two unknowns r and can be solved. Assuming the

stiffness of the two bearings in the x direction at each support is equal results in the
following expressions:

K1X
K AX = K EX = (8-3)
2
K
K BX = K FX = 2X (8-4)
2
K
K CX = K GX = 3X (8-5)
2
K4 X
K DX = K HX = (8-6)
2

The K1X, K2X, K3X and K4X are the total stiffness of the bearing system at each
support. Similar equations can be obtained in the y direction. The two equations can be
further simplified as:

4
K iX cos i
x = r cos = i =1 R = K1 X cos1 + K 2 X cos 2 + K 3 X cos 3 + K 4 X cos 4 R
4
K1 X + K 2 X + K 3 X + K 4 X
K iX
i =1 (8-7)

4

K iY sin i
i =1 R = K1Y sin 1 + K 2Y sin 2 + K 3Y sin 3 + K 4Y sin 4 R
y = r sin =
4
K1Y + K 2Y + K 3Y + K 4Y
K iY
i =1

The 1, 2, 3 and 4 are the coordinates of each support in the polar coordinate
system, and R is the radius of the bridge. For example, for a twin girder bridge, the
bridge radius would be given by:

236
R1 + R2
R= (8-8)
2

In practice, the pier lines are usually oriented in the radial direction as illustrated
in Figure 8.14. From the figure, the global coordinate system does not coincide with the
local system of the piers. For example, the angle between the local and global coordinate
systems is as shown in Figure 8.14. When calculating the stiffness of the bearing-

pier system, the orientations of the bearings and piers must be considered. However,
when the radius becomes larger and the piers become stiffer, the difference due to their
orientations is negligible. Therefore, in practice, for simplicity, the geometrical properties
in the local coordinate system can be used as given in the following equations.

Figure 8-14 Orientations of Piers

237
4
K ix cos i
x = r cos = i=1 R = K1x cos 1 + K 2 x cos 2 + K 3 x cos 3 + K 4 x cos 4 R
4
K1x + K 2 x + K 3 x + K 4 x
K ix
i=1 (8-9)

4

K iy sin i
i=1 R = K1 y sin 1 + K 2 y sin 2 + K 3 y sin 3 + K 4 y sin 4 R
y = r sin =
4 K1 y + K 2 y + K 3 y + K 4 y
K iy
i=1

Once the stationary point is determined, the thermal movements at each bearing
can be calculated using Equation 1-2.
The assumption was mad that the bridge is axially rigid and the guides exert zero
force. In reality, the bridge does deform and the guides exert forces to the bridge. These
forces slightly change the assumed equilibrium path. Consequently, the stationary point is
not exact. In addition, another equilibrium path is possible. For example, for a reduction
in temperature, the equilibrium path shown in Figure 8.15 is possible.

Figure 8-15 Equilibrium Paths when Temperature Drops

238
To verify the applicability of the proposed model, a parametric finite element
study was conducted and the results are presented in the following sections.

8.5.2 Simplified Method

For straight bridge systems, the location of the stationary point is illustrated in
Figure 8.15.

D S A
C B

Abutment 4 Bent 3) Bent 2 (fixed) Abutment 1

o x
Figure 8-16 Stationary Point for Straight Girders

To locate the bearing, a Cartesian coordinate system is created. The origin is set at
the center of bridge with the x-axis is aligned with the longitudinal axis of the bridge. In
the graph, Point S stands for stationary point to be located. The bearing coordinates are:

A: (LA,0) B: (LB,0)
C: (LC,0) D: (LD,0)
S: (x,0)

The thermal movement is assumed proportional to the distance between the


bearing and the stationary point. The equilibrium equation can be built as follows:


K A SA+ K B SB + K C SC + K D SD = 0 (8-10)

239
As in the last section, the values of K represent the combined stiffness of the
bearing/pier system. Writing this vector equation in terms of its components gives:

K (L x ) = 0
i= A
i i (8-11)

Solving it gives:

K A LA + K B LB + K C LC + K D LD
x= (8-12)
K A + K B + KC + K D

Curved bridges with a relatively large radius behave similar to a straight girder.
For simplicity, a similar equation for curved bridges can be developed by treating the
curved girder with arc length, S, as a straight girder by using the above expression with
the arc length as follows:

K A S A + K B S B + KC SC + K D S D
S= (8-13)
K A + K B + KC + K D

Si is the bearing coordinates in the curved coordinate system along the bridge
axis. The stationary point is assumed to be on the bridge axis. The applicability of this
equation is also verified by the following parametric finite element study.

8.5.3 Effect of Pier Flexure

As outlined earlier, under the action of thermal load, the piers deflect laterally.
The length of the piers has a significant impact on the flexural stiffness since the cube of

240
the length is in the denominator. This section focuses on the effect of pier flexure on
thermal movements. Four cases were considered by varying the pier height. The pier
section size was not changed. The stiffness of the bearings and piers is summarized in
Table 8-5. Ktotal in the table denotes the combined stiffness of bearings and piers, which
reflects the net constraints imposed on the bridge by the supports. The total system
stiffness is determined from Equation 7-1 for springs in series.
Ktotal is less than the smallest value of Kpier and Kbearing. For pier heights less
than about 20 feet, the stiffness is dominated by the bearing stiffness. For larger pier
heights, the system stiffness is much more sensitive to the pier length.

Table 8-5 Summary of Pier Flexure

Pier Height (ft) Kpier (k/in) Kbearing (k/in) Kpier/Kbearing Ktotal (k/in)
Case 1 10 4176.0 41.4 101.0 41.0
Case 2 20 522.0 41.4 12.6 38.4
Case 3 35 97.4 41.4 2.4 29.1
Case 4 50 33.4 41.4 0.8 18.5
Case 5 100 4.2 41.4 0.1 3.8

When deriving stationary point in the previous section, the stationary point was
located by the coordinates in the selected coordinate system. It is more convenient to
locate the stationary point by the respective distances along the arc and radius, S and
R, as illustrated in Figure 8.16.
S is the arc length from bent 2 measured along the bridge axis while R is the
distance off the bridge axis measured in radial direction. This locating method is used in
the subsequent analysis.

241
Figure 8-17 Parameters for locating Stationary Point

The following part demonstrates how to use both methods to locate stationary
points of curved bridges with elastomeric bearings and then calculate thermal
movements. As an example, the thermal movements of the exterior girder at Abutment 4
for Case 2 (the piers are 20 ft high) is calculated.

Rigid Body Model

The subtended angle at the location of each support is:


180 + 300 / 2
1 = = 1.021 rad
2 600
300 / 2
2 = = 1.321 rad
2 600
300 / 2
3 = + = 1.821 rad
2 600
180 + 300 / 2
4 = + = 2.121 rad
2 600
The stiffness of the bearing-pier system at the location of each support is:

242
K1 = K bearing = 41.4 kip / in
K 2 = K pier = 522.0 kip / in

K 3 = K total = 38.4 kip / in


K 4 = K bearing = 41.4 kip / in

The coordinates of the stationary point are:


K1x cos 1 + K 2 x cos 2 + K 3 x cos 3 + K 4 x cos 4
x = r cos = K1x + K 2 x + K 3 x + K 4 x
R = 111.7 ft


y = r sin = K1 y sin 1 + K 2 y sin 2 + K 3 y sin 3 + K 4 y sin 4 R = 576.9 ft
K1 y + K 2 y + K 3 y + K 4 y

Solving the above equation for r and gives:


r = 587.6 ft
= 1.379 rad

S and R can be calculated as:


R = 600 587.6 = 12.4 ft
S = 600 (1.379 1.321) = 35.2 ft

The coordinates of the exterior bearing at Abutment 4 considered are:


x = R2 cos 4 = 319.4 ft

y = R2 sin 4 = 520.9 ft

The distance between the bearing considered and the stationary point is:
L= ( 319.4 111.7 )2 + (520.9 576.9)2 = 434.7 ft

The resultant movement of the exterior girder at Abutment 4 is:


= mix TL = 5.63 106 60 434.7 12 = 1.76 inches

The coefficient of thermal expansion in these calculations is the average value of


steel and concrete weighted by their area that can be calculated as follows:

243
Asteelsteel + Aconcreteconcrete 594 6.5 106 + 4032 5.5 106
mix = = = 5.63106 in / in / F
Asteel + Aconcrete 594 + 4032

Simplified Method

The arc length between the origin and each support is:
S A = 180 + 300 / 2 = 330 ft

S B = 300 / 2 = 150 ft
SC = 300 / 2 = 150 ft

S D = (180 + 300 / 2 ) = 330 ft

The stiffness of the bearing-pier system at the location of each support is:
K A = K bearing = 41.4 kip / in
K B = K pier = 522.0 kip / in

K C = K total = 38.4 kip / in


K D = K bearing = 41.4 kip / in

The coordinates of the stationary point are:


K S + K B S B + K C SC + K D S D
S= A A = 112.9 ft
K A + K B + KC + K D

S can be calculated as:


S = 150 112.9 = 37.1 ft

R is zero according to the assumption. The chord length between the bearing
considered and the stationary point is:
S S 330 + 112.9
L = 2 R sin 4 = 2 600 sin = 432.9 ft
2R 2 600
The resultant movement of the exterior girder at Abutment 4 is:
= mix TL = 5.63 106 60 432.9 12 = 1.75 inches

The coefficient of thermal expansion used here is the same as before.

244
By following the above procedures, the movements at other locations can be
calculated. Similar calculations can be done for the other four cases. Figure 8.17 shows a
comparison of the stationary points estimated from both the rigid body model and the
simplified method to those obtained from the finite element models. The value of R
usually lies outside the girder, so it is difficult to determine in the finite element model
and only S is compared to the estimated values.
The respective thermal movements at Abutments 1 and 4 are compared in Figures
8.18 and 8.19. For comparison, the movements based on the current design method that
assumes a fixed bent 2 are also plotted in the respective figures.

150
ArcDistanceFromBent2(S) (ft)

120

90

60

30 FiniteElementModel
RigidBodyModel
SimplifiedMethod
0
0 20 40 60 80 100
PierHeight(ft)

Figure 8-18 Effect of Pier Flexure on Stationary Points

245
1.6

BridgeMovementatAbutment1(inch)
1.2

0.8

0.4 FiniteElementModel
RigidBodyModel
SimplifiedMethod
CurrentDesignMethod
0
0 20 40 60 80 100
PierHeight(ft)

Figure 8-19 Effect of Pier Flexure on Thermal Movement at Abutment 1

2
BridgeMovementatAbutment4(inch)

1.6

1.2

0.8

FiniteElementModel
0.4 RigidBodyModel
SimplifiedMethod
CurrentDesignMethod
0
0 20 40 60 80 100
PierHeight(ft)

Figure 8-20 Effect of Pier Flexure on Thermal Movement at Abutment 4

246
It can be seen from the finite element analyses that the stationary point is close to
the fixed bent 2 (11 ft away from bent 2 to the center of the bridge) when the pier is very
rigid (H=10 ft). The offset of stationary point from the center of bridge leads to a much
larger tangential displacement at one end (abutment 4) than that at the other end
(abutment 1). With the increase of the pier flexibility, the stationary point shifts towards
the center of mass. When the pier is 50 ft high, the stationary point is almost at the center
of the bridge (1 ft away from the center of the bridge to the fixed bent 2). Since the
stationary point is in the middle, the tangential displacements at both ends are similar. It
is clear that it is important to estimate the location of stationary point if a good estimate
of the maximum thermal displacement is desired. In contrast, the current design method
that assumes a fixed bent 2 is only applicable when the pier is relatively rigid.
From the above comparisons, it can be seen that both the proposed rigid model
and the simplified method give similar estimates. The difference between them is
negligible. The estimates of stationary points by both methods are reasonable compared
to finite element method. When the piers are 50 ft high, the error in S is about 10 ft
and reaches maximum. The estimates of thermal movements at the ends by both methods
are relatively accurate compared to the finite element method. When the piers are 100 ft
high, the error is about 0.1 inches (about 8%) and reaches maximum.

8.5.4 Effect of Bearing Stiffness

As the girders move under the action of thermal loads, the bearings transfer some
loading into the piers and thereby carry some shear forces. Under the action of these
shear forces, the bearings will deform accordingly. The effect of bearing stiffness on
thermal movements is similar to pier flexibility as described above. In this study, the
bearing stiffness was taken as a fraction of the pier stiffness. Four ratios were considered:
247
0.001 (very flexible), 0.04, 0.08, and 0.15. The stiffness of the bearings and piers is
summarized in Table 8-6. The value given in the fourth column is the stiffness of one
bearing and there are two bearings at each location of support. These values were
selected according to the stiffness ranges (3 ~ 40 k/in) of elastomeric bearings that are
typically used in box girders as discussed in Section 7.4.2. The fixed bearing was
located at Bent 2 while unidirectional or multidirectional bearings were simulated at the
other piers and abutments. The bearing stiffness given in Table 8-6 represents the
bearings located at Abutment 1, Bent 3 and Abutment 4. A very flexible bearing (Case
1) will develop very little force while a stiff bearing (Case 4) can develop large
restraining forces at the tops of the piers/abutments.

Table 8-6 Summary of Bearing Stiffness

Pier Height (ft) Kpier (k/in) Kbearing (k/in) Kbearing/Kpier Ktotal (k/in)
Case 1 20 522.0 0.26 0.001 0.52
Case 2 20 522.0 10.2 0.04 19.7
Case 3 20 522.0 20.7 0.08 38.3
Case 4 20 522.0 40.3 0.15 69.8

Figure 8.20 provides a comparison of the estimates of the stationary point location
obtained from both the rigid body model and the simplified method to those obtained
from finite element models.

248
80

ArcDistanceFromBent2(S)(ft)
60

40

20
FiniteElement Model
RigidBodyModel
SimplifiedMethod
0
0 0.03 0.06 0.09 0.12 0.15
Kbearing /Kpier

Figure 8-21 Effect of Bearing Stiffness on Stationary Point

The comparisons of the respective thermal movements at Abutments 1 and 4 are


graphed in Figures 8.21 and 8.22. For comparison, the movements based on the current
design method that assumes a fixed bent 2 are also plotted in the respective figures.

1.2
BridgeMovementatAbutment1(inch)

0.8

0.4
FiniteElementModel
RigidBodyModel
SimplifiedMethod
CurrentDesignMethod
0
0 0.05 0.1 0.15
Kbearing /Kpier

Figure 8-22 Effect of Bearing Stiffness on Thermal Movement at Abutment 1

249
2

BridgeMovementatAbutment4(inch)
1.5

0.5 FiniteElementModel
RigidBodyModel
SimplifiedMethod
CurrentDesignMethod
0
0 0.05 0.1 0.15
Kbearing /Kpier

Figure 8-23 Effect of Bearing Stiffness on Thermal Movement at Abutment 4

From the finite element analyses, the stationary point is almost at the fixed bent 2
(0.3 ft offset from bent 2 to the stationary point) when the bearing is very flexible
(Kbearing/Kpier=0.001). Similar to the change of pier heights, the offset of stationary point
from the center of bridge leads to a much larger tangential displacement at one end
(abutment 4) than that at the other end (abutment 1). When the bearing becomes stiffer,
the stationary point shifts towards the center of mass. When Kbearing/Kpier=0.15, the
stationary point is about 60 ft from bent 2.
From the above comparisons, it can be seen that both the proposed rigid model
and the simplified method give similar estimates. The difference between them is minor.
The estimates of stationary points by both methods are accurate compared to finite
element method. The error in S is within 2 ft for all the four cases considered. The
estimates of thermal movements at the ends by both methods are also very accurate
compared to the finite element method. The error is within 0.05 inches (4%) for all the

250
four cases considered. The current design method that assumes a fixed bent 2 is only
applicable when the bearings are very flexible compared to the piers.

8.5.5 Effect of Bridge Curvature

Curved bridges respond to thermal loads in a more complicated manner than


straight bridges. Unlike straight girders, changes in temperature cause lateral
displacements for curved girders. The effect of the girder curvature on thermal
movements of bridges is presented in this section. The curvature was changed by varying
the radius of the bridge from 400 ft to 2000 ft. Four radii were considered: 400, 600, 800
and 2000 ft.
Figure 8.23 shows a comparison of the estimates of the stationary point obtained
from both the rigid body model and the simplified method to those obtained from finite
element models.

50

40
ArcDistanceFromBent2(S)(ft)

30

20

10 FiniteElementModel
RigidBodyModel
SimplifiedMethod
0
400 800 1200 1600 2000
RadiusofCurvature(ft)

Figure 8-24 Effect of Bridge Curvature on Stationary Point

251
The comparisons of the respective thermal movements at Abutments 1 and 4 are
plotted in Figures 8.24 and 8.25.

1.2
BridgeMovementatAbutment1(inch)

0.8

0.4

FiniteElementModel
RigidBodyModel
SimplifiedMethod
0
400 800 1200 1600 2000
RadiusofCurvature(ft)

Figure 8-25 Effect of Bridge Curvature on Thermal Movement at Abutment 1

2
BridgeMovementatAbutment4(inch)

1.5

0.5
FiniteElementModel
RigidBodyModel
SimplifiedMethod
0
400 800 1200 1600 2000
RadiusofCurvature(ft)

Figure 8-26 Effect of Bridge Curvature on Thermal Movement at Abutment 4

252
From the finite element analyses, it can be seen that the stationary point does not
change significantly with the curvature. When the radius changes from 400 ft to 2000 ft,
S only changes from 33 ft to 40 ft. Due to this insensitivity of stationary point to the
bridge curvature, the simplified method can predict both stationary points and thermal
movement at the ends as accurately as the rigid body model.
The above comparisons indicate that both the proposed rigid body model and the
simplified method give estimates of both stationary points and thermal movement at the
ends with acceptable accuracy. The difference between these two methods is minor. The
estimates of stationary points by both methods are reasonable compared to finite element
method. The error in S is within 4 ft for all the four cases considered. The estimates of
thermal movements at the ends by both methods are also very accurate compared to finite
element method. The error is within 0.05 inches (4%) for all the four cases considered.

8.5.6 Design Shear at Pier Top

Under the action of thermal loads, the elastomeric pad will transfer some shear
force to the tops of the supporting piers. It is important to determine the most adverse
shear for design. Based on the above rigid body model and the simplified method, the
extreme values can be obtained by assuming all the bearings at the pier tops are
completely locked. The following part demonstrates how to use both methods to
determine the extreme shear force at the pier tops due to thermal loads. As an example,
the shear force at Bent 3 for Case 2 (the piers are 20 ft high) is calculated.

Rigid Body Model

The subtended angle at the location of each support is:

253

180 + 300 / 2
1 = = 1.021 rad
2 600
300 / 2
2 = = 1.321 rad
2 600
300 / 2
3 = + = 1.821 rad
2 600
180 + 300 / 2
4 = + = 2.121 rad
2 600
The stiffness of the bearing-pier system at the location of each support is:
K1 = K bearing = 41.4 kip / in
K 2 = K pier = 522.0 kip / in
K 3 = K pier = 522.0 kip / in
K 4 = K bearing = 41.4 kip / in

The coordinates of the stationary point are:


K 1x cos 1 + K 2 x cos 2 + K 3 x cos 3 + K 4 x cos 4
x = r cos = K + K + K + K
R = 0 ft
1 x 2 x 3 x 4 x

y = r sin = K 1 y sin 1 + K 2 y sin 2 + K 3 y sin 3 + K 4 y sin 4 R = 576.2 ft


K1y + K 2 y + K 3 y + K 4 y

The coordinates of Bent 3 considered are:


x = R cos 3 = 148.6 ft
y = R sin 3 = 581.3 ft

The distance between the bearing considered and the stationary point is:
L= (0 + 148.6)2 + (576.2 581.3)2 = 148.6 ft

The resultant movement at Bent 3 is:


= mix TL = 5.63 10 6 60 148.6 12 = 0.603 inches

The shear force on the top of bent 3 is:


Vbent 3 = K 3 = 522.0 0.603 = 314.5 kips

254
Simplified Method

The arc length between the origin and each support is:
S A = 180 + 300 / 2 = 330 ft

S B = 300 / 2 = 150 ft
SC = 300 / 2 = 150 ft

S D = (180 + 300 / 2 ) = 330 ft

The stiffness of the bearing-pier system at the location of each support is:
K A = Kbearing = 41.4 kip / in
K B = K pier = 522.0 kip / in
K C = K pier = 522.0 kip / in
K D = K bearing = 41.4 kip / in

The coordinates of the stationary point are:


K S + K B S B + KC SC + K D S D
S= A A = 0 ft
K A + K B + KC + K D

S can be calculated as:


S = 150 0 = 150 ft

R is zero according to the assumption. The chord length between the pier
considered and the stationary point is:
S SC 0 + 150
L = 2 R sin = 2 600 sin = 149.6 ft
2R 2 600
The resultant movement at Bent 3 is:
= mix TL = 5.63 10 6 60 149.6 12 = 0.606 inches

The shear force on the top of bent 3 is:


Vbent 3 = K 3 = 522.0 0.606 = 316.6 kips

255
8.6 THERMAL STRESSES

Bridges continuously exchange heat with the surrounding environment, so the


impact of thermal effects is a concern in bridge design. The adverse influences of
thermal effects are due to either thermal movements or thermal stresses. Past field
studies have also shown that changes in the thermal environment can lead to significant
stress changes. Prior field monitoring of three steel trapezoidal box girders (Helwig et
al 2004, Helwig and Fan 2000, Bobba 2003) show that thermal stresses on the order of

5 ksi are not uncommon in daily thermal cycles, which is of the same order of stresses
induced by such construction activities as girder erection and concrete deck placement. In
annual thermal cycles, thermal stresses may be even larger. Although most typical
bridges have not experienced difficulties because of thermal stresses in the past, bridge
designs have become more complex in recent years. To get an economic design and fit
the practical needs, designers have increased the span and minimized superstructure
weights. These changes have reduced the reserve strength for temperature induced
stresses, and can lead to undesirable cracking of the concrete deck slab, which can
expedite degradation of the structure. Past surveys on problematic bridges have shown
that notable cracks have occurred in bridge structures and have seriously affected the
serviceability and integrity of the structure. Leonardt et al. (1965), Zichner (1981), and
Massicotte et al. (1994) reported such cracks in some major bridges and attributed these
cracks to the fact that environmental thermal stresses was inaccurately modeled or
completely ignored in the design of the bridges. Thermal movements were discussed in
detail in the previous sections and this section will mainly focus on thermal stresses.

256
8.6.1 Effect of Additional Restraints

Expansion joints are often used to accommodate thermal movement along the
length of the bridge; however debris often accumulates in these joints as shown in Figure
1-10. The debris often consists of rocks and concrete pieces, which are not easily
compressible resulting in restraints that develop during hot weather periods when the
bridge tries to expand. These restraints can lead to additional continuity stresses that
develop due to this unexpected restraint. The effects of these restraints have not been
significantly studied. The actual restraint conditions due to the debris are relatively
complex. For simplicity, the effect due to the debris was modeled using a smeared linear
spring. The stiffness of the spring per ten feet was taken as a fraction of the pier stiffness.
Five ratios were considered: 0 (no unintended restraint), 0.1, 1, 2 and 5.
Figure 8.27 shows the thermal movements at Abutments 1 and 4. For comparison
purposes, the sum of the movements at both abutments is also graphed. It should be noted
that in the figure no sign convention was defined for thermal movement and only the
magnitudes of them were considered.
The bridge expands with increasing temperature. The unintended debris at the
expansion joints restrain the bridge expansion and exert an axial force to the girder. This
force changes the equilibrium path assumed before. The stationary point shifts toward
Abutment 4 accordingly. From the figure, it can be seen that the movement at
Abutment 4 decreases while the movement at Abutment 1 increases. It is also noted that
the sum of thermal movements reduces with increasing unintended restraints, which
indicates that the axial shortening is not negligible. The effect of unintended restraints on
the support shears and guide force are plotted in Figures 8.28 and 8.29.

257
2.8

2.4

2
BridgeMovement(inch)

1.6

1.2

0.8
Abutment1
0.4 Abutment4
SumofMovement atAbutments1and4
0
0 1 2 3 4 5
Kunintended/Kpier

Figure 8-27 Thermal Movements at Abutments 1 and 4

500

400
SupportShear(kip)

300

200

100

0
0 1 2 3 4 5
Kunintended/Kpier

Figure 8-28 Effect of Unintended Restraint on Support Shear

258
2000

1500
GuideForce (kip)

1000

500

0
0 1 2 3 4 5
Kunintended/Kpier

Figure 8-29 Effect of Unintended Restraint on Guide Force

The effect of unintended restraints on the maximum bottom flange stress is


plotted in Figures 8.30.

8
MaximumBottomFlangeStress(ksi)

0
CompressiveLongitudinalStress

TensileLongitudinalStress
4

8
0 1 2 3 4 5
Kunintended/Kpier

Figure 8-30 Effect of Unintended Restraint on Bottom Flange Stress


259
From the above figures, it can be seen that with the increase of the stiffness of
unintended restraints, the maximum support shear, guide force, and bottom flange
stresses dramatically increase. The stress in the concrete deck also slightly increases.
When the stiffness ratio is 5, the respective values of the guide force and the
support shear may reach as high as 1657 kips and 431 kips. Such a large stress may
damage the guides. Although it is rare that the debris accumulated at the expansion joints
is 5 times as stiff as the piers, clearing of the debris at the expansion joints during routine
maintenance is important to avoid unintended internal forces as a result of thermal
movements.

8.6.2 Thermal Stress During Thermal Cycles With 100-year Return Period

This section is intended to focus on the maximum thermal stresses induced by the
thermal loads with 100-year return period. The basic model described previously was
used to study the impact of these annual cycles. The pier was assumed to be 10 ft high
and the bearing stiffness was assumed to be 20 kip/in. Both piers and bearings are taken
as feasibly stiff as possible because larger thermal stresses are expected when stiffer piers
and bearings are used. As pointed out in Chapter 7, two methods can be used to
consider thermal effects: the linear component method and the nonlinear component
method. The latter method was used in this analysis.
During annual thermal cycles, a bridge experiences daily heating and cooling,
daily solar radiation from the sides, and season changes that lead to fluctuations in the
ambient temperature. As discussed in Chapter 7, these situations can be conservatively
considered using the following eight types of thermal loads.

260
Set I: In summer
Case 1: Heating
Case 2: Cooling
Case 3: Direct side solar radiation on interior girder
Case 4: Direct side solar radiation on exterior girder
Set II: In Winter
Case 1: Heating
Case 2: Cooling
Case 3: Direct side solar radiation on interior girder
Case 4: Direct side solar radiation on exterior girder
The thermal load with 100-year return period obtained from Chapter 6 was used.
The temperature range considered based upon the Chapter 6 conclusions ranged from -8
to 120 F. The temperature when the girders were set was assumed to be 60 F.
Therefore, the uniform temperature loads consisted of a 60 F increase and a 68 F
decrease.
The transverse temperature linear gradient was used to consider the effects of
direct sunshine on the webs. This difference Th was evaluated based on the extreme value
analyses in Chapter 6. The value of 19 F with 100 years of return period was used.
The vertical temperature gradient through the depth was taken from the gradients
proposed in Chapter 5 as shown in Figure 8.31.

261
24 F
38 F

4"
Concrete Deck Concrete

8"
4
y
T = 381

12"
16

Steel Girder
Steel

32 F
5 F
(a) Heating (b) Cooling

Figure 8-31 Vertical Thermal Gradient

Table 8-7 Summary of Thermal Stresses Under Different Thermal Loads (ksi)

Case Tensile Bottom Compressive Bottom 1st Principal Stress in


Number Flange Stress Flange Stress Concrete Deck

Set I: 1 1.06 -1.60 0.46

Set I: 2 2.91 -0.74 0.61

Set I: 3 0.80 -4.31 0.52

Set I: 4 0.45 -3.10 0.50

Set II: 1 3.24 -1.53 0.34

Set II: 2 6.13 -0.65 0.60

Set II: 3 2.30 -1.41 0.18

Set II: 4 2.39 -1.84 0.17

262
The maximum thermal stresses are summarized in Table 8-7. It can be seen
from the table that the maximum tensile thermal stress in bottom flange occurs when the
bridge is subjected to cooling in winter condition. Under this condition, the concrete
deck is warmer than the steel girder and the whole bridge contracts due to temperature
drop, which means that both uniform temperature loads and temperature gradient over the
depth produce tensile stresses in bottom flange. On the contrary, for Case 2 of Set I, the
temperature gradient produces tensile stresses in bottom flange while the uniform
temperature change produces compressive stresses. These two stresses partially cancel
each other and the resultant tensile stresses in the bottom flange become much smaller.
The maximum compressive bottom flange stresses are achieved at the outside
edge of the exterior girder when the interior web is directly exposed to sunshine in
summer condition. Under uniform temperature change, the curved bridge tends to be
flattened, which produces compressive stress at the outside fiber of the bottom flange of
the exterior girder. If the interior web of the interior girder is further directly exposed to
sunshine, the tendency of being flattened of the bridge will be aggravated and larger
compressive stress is thereby generated at the most critical fiber.
The maximum first principal stresses in concrete deck occur when the bridge is
cooling. It is also noted that the magnitudes of it are similar for both summer and winter
conditions. This seems to imply that the concrete thermal stresses are primarily
contributed by vertical thermal gradients rather than uniform temperature change. For
normal weight concrete of compressive strength of 4 ksi that is commonly used in bridge
design, the cracking stress is nearly 0.4 ksi that is smaller than the maximum 1st principal
stress (0.6 ksi). Therefore, the possibility of concrete cracking due to thermal loads
cannot be ruled out. However, it should be kept in mind that the piers and bearings used
in the analysis are very stiff and the thermal loads that are of 100-year return period were

263
obtained using the most critical section configuration in practice and the most critical
orientation. For most of bridges, the thermal stresses should be much smaller than this
magnitude. In addition, the thermal stresses measured in the field monitoring may be
even smaller from the standpoint of statistics.
The stationary point was also calculated for each thermal load condition as shown
in Figure 8-32. In the figure, the small red hollow circles stand for the stationary points
calculated from the finite element analysis. The black solid dot is the stationary point that
was used in the analysis and predicted from the Rigid Body Model proposed in Section
8.5.1. The dashed line plotted the guide orientation of unidirectional bearings. It can be
seen that the stationary point moves slightly in the radial direction under the action of
different thermal loads, but it does not move in the tangential direction as much as in the
radial direction. This finding is very important and practical in that it implies that only
effective bridge temperature (or uniform temperature change) needs to be considered
when calculating thermal movements at the abutment. In other words, although thermal
gradients have some effects on thermal movements, the effect is generally so slight that it
can be neglected in practical design.
The girder movement at Abutments 1 and 4 were calculated from both the Rigid
Body Model and the simplified method proposed in Section 8.5.1 and compared to the
finite element results as shown in Figures 8-33 and 8-34, respectively. In the figure, the
horizontal axis is the case number. The dots above the x-axis are the movements during
the summer condition while those below x-axis are the movements during the winter
condition. It should be noted that the expansion is assumed to be positive displacement
in the figure.

264
Figure 8-32 Location of Stationary Points Under Different Thermal Loads ( To Scale)

1.5
SummerWeatherCondition
ThermalMovementatAbutment4(inch)

0.5 FiniteElement Analysis


RigidBodyModel
SimplifiedMethod
0
1 2 CaseNo. 3 4
0.5

WinterWeatherCondition
1.5

Figure 8-33 Effect of Different Thermal Loads on Girder Movement At Abutment 1

265
3
SummerWeatherCondition

ThermalMovementatAbutment4(inch)
2

FiniteElement Analysis
1
RigidBodyModel
SimplifiedMethod
0
1 2 CaseNo. 3 4
1

WinterWeatherCondition
3

Figure 8-34 Effect of Different Thermal Loads on Girder Movement At Abutment 4

It can be seen from the above comparisons that both the Rigid Body Model and
the simplified method can predict the movements at the abutments with good accuracy.
The errors are within 10%, so both of the proposed methods have good applicability
under different thermal loads the bridge may experience during the service life.

8.7 EFFECT OF ERRORS IN ORIENTING GUIDES

In the field the lateral guide orientation is often established using the naked eye.
It is therefore difficult to orient the guides exactly towards the desired direction without
significant error. The impact of guide misalignment on the bridge movements and girder
stresses has not been previously studied. The Greenspoint bridge (IH45 Southbound
connection to Beltway 8 Westbound in northwest Houston, Texas) that had the bearings
instrumented, had guides oriented approximately 45 away from the fixed bearing, as
shown in Figure 8.32. Clearly gross errors resulted in this guide placement and such
errors are not likely to commonly happen. However, such an example does indicate the

266
potential for errors in bearing alignment ranging from a few degrees offset to gross
misalignments.

Figure 8-35 Example of Misdirected Guides - Greenspoint Bridge

The basic model described previously was used to examine the effect of bearing
misalignment. Instead of chord-orienting bearings toward bent 2, the chord orientation
toward the stationary point was used. Five cases were considered.

Case 1: the bearing at abutment 4 was misaligned by 10 clockwise error


Case 2: the bearing at abutment 4 is misaligned by 10 counterclockwise error
Case 3: all bearings are oriented toward stationary point (reference case)
Case 4: all guided bearings are misaligned by 5 clockwise error
Case 5: all guided bearings are misaligned by 5 counterclockwise error
The clockwise direction is defined in Figure 8-36.
267
Bent 3
Bent 2

Fixed
Abut 4 Multi-direction Guide
Single Direction Guide Abut 1

Figure 8-36 Definition of Error Direction

Table 8-8 summarizes the finite element results for the five cases considered. It
can be seen from the table that small errors in orienting bearings does not significantly
affect bridge movements, support shears, or thermal stresses. This observation also
indicates that some errors in predicting the location of stationary points using the
proposed rigid methods are acceptable.

Table 8-8 Summary of Finite Element Results

Case 1 Case 2 Case 3 Case 4 Case 5


Movement at Abutment 1 (inch) 0.89 0.92 0.90 0.90 0.90
Movement at Abutment 4 (inch) 1.73 1.84 1.77 1.78 1.77
Max. Support Shear (kip) 92.0 88.7 85.9 86.1 86.5
Max. Tensile BF Stress (ksi) 0.91 1.14 0.96 0.95 0.98
Max. Compressive BF Stress (ksi) -1.62 -1.57 -1.56 -1.57 -1.55
Max. Concrete Stress (ksi) 0.32 0.32 0.32 0.32 0.32

The stationary point only moves slightly due to small errors in orienting bearings
as shown in Figure 8.33. Among them, the stationary point for Case 2 moves most among
the five cases. It is interesting to find that the guide force for Case 2 is also the largest as
268
shown in Figure 8.34. This is not a coincidence. As discussed in Section 8.5.1, the guide
force changes the assumed equilibrium path. Therefore, it seems that the largest guide
force makes the stationary point move the most.

41.8
37.3 37.2 36.1 38.2

Case1 Case2 Case3 Case4 Case5

Figure 8-37 Effect of Small Orientation Error on Stationary Point (ft)

57.9

31.4
23.8 25.3
21.1

Case1 Case2 Case3 Case4 Case5

Figure 8-38 Effect of Small Orientation Error on Guide Force (kip)

Only small errors in bearing alignment were investigated. For very large errors,
the rigid guide may inhibit the movement at the abutments similar to the unintended
restraint from the debris in the expansion joint discussed earlier. In this situation, large
internal forces may be produced in the guides. Therefore, although small errors in
bearing orientation are acceptable, larger errors may result in complete locking of the
guided bearings and should be avoided.

269
From the rigid body model, the stationary points are only possibly located in the
shade area as shown in Figure 8-39. The exact location depends on two parameters: the
span to radius ratio, and the ratio of pier to bearing stiffness.

Figure 8-39 Possible Location of Stationary Points

According to the results presented in Sections 8.5.3 and 8.5.4, when the ratio of
pier stiffness to bearing stiffness becomes larger, the stationary point will become closer
to Bent 2. By examining the two above parameters using the rigid body model, it was
found that the errors between chord-alignment toward Bent 2 and chord-alignment
toward the stationary point are within 10 degree for most of bridges. When the radius of
bridge curvature is larger than 800 ft and the span length is less than 600 ft, all stationary
points calculated from rigid body model lie in the area within 10 degree error for the
bridge with the ratio of pier to bearing stiffness larger than 3. For most bridges without
a very tall pier, this ratio is greater than 3. This implies that it is appropriate to guide the
bearing along the chord toward Bent 2 for most bridges with the radius larger than 800 ft
since 10 degree error can be tolerated and does not make much difference from the
analyses in this section.

270
8.8 DISCUSSION AND SUMMARY

This chapter presented FEA solutions evaluating the effects of thermal loads with
100-year return periods developed in the previous chapters on curved bridges with
elastomeric bearings. Based upon the results presented, the following conclusions can be
reached.
The positioning of fixed bearings on straight and curved bridges was
considered. Two schemes were examined: placing the fixed bearing at a rigid
abutment and at a flexible interior bent. The results indicate that the demands
of guide force, bridge movements at the ends, support shears and bottom flange
stresses when bearings are oriented toward a rigid abutment are much larger
than those when bearings are oriented toward a flexible interior bent. None of
the above demands is desirable in bridge design, so the positioning of fixed
bearings should be placed at interior bents.
The proper bearing orientation was evaluated. It was found that neither
tangential orientation nor chorded orientation toward a fixed bearing at an
interior bent is appropriate. The bridges actually experience displacements
towards a stationary point located away from the fixed bearing. When orienting
bearings towards this stationary point, smaller guide forces are achieved
compared to the traditional orientation towards the fixed bearing. A
tangential orientation was also considered. In general the tangential
orientation did not lead to a significant difference than orienting the bearing
towards the fixed bearing; however larger effects may result with wider bridges
or girders with a smaller radius of curvature. In comparing the different
orientations, the best behavior is achieved when the bearing is oriented towards
the actual stationary point on the bridge.
271
The possibility of guiding both girders as proposed by the AGC Contractors
was examined. The results showed that the expansion in the radial direction
must be accommodated by the rigid guides and large undesirable internal
forces are generated in the guides. Therefore, it is not appropriate to guide both
girders of curved bridges.
A rigid body model was proposed to estimate the location of the stationary
point and maximum thermal movements at the bridge ends. Using the arc
length approach analogous to straight bridges, a simplified method was also
evaluated. Both methods were compared to finite element solutions that were
verified in the previous chapter. It was shown that the proposed methods could
accurately predict both the stationary point and thermal movements at the ends
under different pier flexure, bearing stiffness and bridge curvature. If rigid
body model is used, the chord-orientation toward the calculated stationary
point is recommended since best behavior is achieved. If the simplified method
is preferred, it is recommended the chord-orientation toward the fixed bent is
adopted because the simplified method cannot estimate the offset of the
location of stationary point in the radial direction for the following reasons: 1)
if the pier is rigid, the stationary point is actually close to the fixed bent; 2) if
the pier is flexible, most of thermal movements are accommodated by pier
deflection and the bearing orientation becomes less important.
By assuming that all bearings at pier tops are completely locked, the procedure
for calculating the most adverse shear force at pier tops using the both methods
is illustrated by an example. Both methods give similar estimates.
The maximum magnitudes of thermal stresses were also evaluated under
different possible load conditions that may occur during the service life of

272
bridges. The results showed that the maximum tensile thermal stress in the
bottom flange may be as high as 6.1 ksi while the maximum compressive stress
is 4.3 ksi under the thermal load with 100-year return period. The 1st principal
stress in concrete deck is 0.61 ksi, which may crack concrete. The effect of
the unintended restraints due to the debris built up at the expansion joints was
also considered, which showed that large guide forces, support shears and
thermal stresses may be generated by the presence of the debris. Therefore,
regularly clearing of the debris at the expansion joints is important for routine
maintenance of the bridge to avoid unintended internal forces due to restricted
thermal movements.
The effects of bearing misalignment were investigated. Small errors in
orienting the bearings do not significantly impact the difference in bridge
movements, support shears and thermal stresses. Detailed examination of
bridge layout based on the rigid body model, it is found that guiding the
bearing along the chord toward Bent 2 is appropriate for most bridges with the
radius larger than 800 ft.

273
Chapter 9 Summary and Conclusions

This investigation in this dissertation includes field measurements, laboratory

tests and finite element thermal analyses and finite element structural analyses. The

bearings of nine bridges in the Houston area have been instrumented and monitored for

more than a year to measure bearing movements due to changes in temperature. A

detailed instrumentation of the steel girders on one of the Houston bridges was made

utilizing thermocouples to measure temperature distribution. In addition, thermocouples

were applied to the steel girders and concrete bridge deck on a simple twin box girder

bridge located at Ferguson Structural Engineering Laboratory.

The description of temperature field over a steel-concrete composite cross section

was first investigated. Based on the structural response of a composite bridge section to

thermal loading, the sectional temperature field was decomposed into four components as

shown in Figure 9-1: effective bridge temperature Te, vertical linear thermal gradient Tv,

transverse linear thermal gradient Th, and residue temperature contributing to self-

equilibrating stresses Tres. If the section is unrestrained, the effective bridge temperature

and linear thermal gradients are associated with the longitudinal displacement and

rotation, respectively. If the section is restrained, the effective bridge temperature and
linear thermal gradients are associated with the axial force and moment, respectively. The

residual temperature is the leftover of subtracting the linear components from the total

temperature distribution. It generally reflects how nonlinear the sectional temperature

distribution is and produces self-equilibrating stresses regardless of whether or not the

section is restrained.

274
Figure 9-1 Decomposition of A Temperature Profile

Following parameterizing the temperature field over a section, both two- and

three-dimensional finite element heat conduction models of composite steel bridges were

developed and validated by the temperature distributions over cross sections at a time

instant, as well as the temperature history at a specified point, obtained from the field

monitoring and laboratory tests. The three-dimensional FE thermal model was used to

examine the impact of bridge curvature on the thermal behavior of composite steel

girders. It was found that the effective bridge temperature and vertical linear temperature

difference showed little change along the bridge axis. The transverse linear temperature

difference did vary significantly along the bridge axis. To simplify the problem, the

extreme value was used to bound the effect for design purposes. The comparisons

between the two- and three-dimensional FE thermal models showed that two-dimensional

FEA thermal analyses are sufficient for studying the thermal behavior of these bridge

systems. To save computational efforts, the two-dimensional FE model was used to

investigate the impact of bridge geometry and orientation on the thermal parameters that

275
were used to describe the temperature field over a cross section. It is found that under

the given weather condition, the most critical thermal loads are achieved under the

following bridge configurations:

N-S bridge orientation;

Shorter lengths of the concrete deck overhang;

Deeper steel girder webs;

Thinner concrete decks;

Wider girder spacing.

It was also observed that the temperature distribution along the width is relatively

uniform, so the vertical linear temperature difference and the residual temperature were

combined into a nonlinear vertical thermal gradient over the depth. A vertical

temperature distribution for heating and another one for cooling were further proposed as

given in Figure 9-2, both of which have good correlation with results from the FEA

studies. For each curve, there are three parameters to be determined: T1, T2 and T3. These

parameters were defined in Section 5.5.2.

276
T1

y
hc
Concrete Deck
T

Depth of Superstructure
T2 Steel Girder

4
y
T = T2 + (T1 T2 ) 1
hc + 8
8

T3

(a) Vertical Temperature Distribution for Heating


T1
hc/2 hc/2

T2
hc

Concrete Deck
8
Depth of Superstructure

Steel Girder

T3

(b) Vertical Temperature Distribution for Cooling


Figure 9-2 Proposed Vertical Temperature Distributions during heating and cooling

Environmental conditions including solar radiation, wind speed and ambient air

temperature have complex and significant effects on the thermal behavior of composite

steel girders. Rational analyses on these environmental factors need to be conducted to


277
reflect their impact on each thermal parameter defined previously. By observing the

random features of the environmental factors, a statistical model based on extreme value

analysis was built to develop design thermal loads. The sample data of each thermal

parameter were obtained from the validated two-dimensional finite element thermal

model with the 45-year long historic weather data in Texas as the input (from 1961 to

2005). The critical section sizes within realistic ranges and critical orientation were

chosen to obtain the hourly sample data for a conservative design. Considering the large

geographical area of Texas, four cities (Austin, Wichita Falls, Brownsville and El Paso)

were used to bound its weather conditions. Extreme value analyses on the sample data

were performed to obtain the relationship between thermal loads and return periods for

each thermal load parameter. The thermal loads with 100 year return period were

compared to the ones suggested by AASHTO. Based upon this comparison the

following conclusions were reached:

The upper limit of effective bridge temperature is very close to that with return

period of 100 years. However, the lower limit given by the AASHTO LRFD

(2004) is equivalent with that with return period of 16 years. The current

AASHTO limit of 0 F should be lowered to -8 F to provide a 100 year return

period.

The positive vertical temperature gradient during heating provided in The

AASHTO LRFD Bridge Design Specification (2004) is conservative because

the effects of reflected solar radiation from the ground are ignored.

Transverse linear temperature gradient is on the same order of the vertical

gradient. The impact on bridge design needs further assessment.

278
A statistical analysis on the weather data over several years can be used to

validate the AASHTO provision or develop new guidelines thereby resulting in reliable

temperature data that bridges will realistically be subjected to. Only the unsurfaced

bridge was addressed and the effect of black-top such as asphalt wearing surfaces was not

considered. According to the research results conducted by Priestley (1976), the black-

top would induce a smaller temperature gradient due to the insulation of the black-top.

The suggested relationship between the design gradient and the black-top thickness can

be used if the black-top is present.

Following thermal analyses, the thermal loads with 100-year return period were

used to investigate the structural responses of composite bridges. The pier deflections and

bearing movements taken from laboratory tests and field monitoring were used to

characterize the behaviors of elastomeric and pot bearings that are most frequently used

in tub girder bridge systems. By using the identified bearing behavior as boundary

conditions, a finite element structural model was developed and also verified by the

thermal movements obtained from field monitoring. The vibrating wire strain gages that

were instrumented on the Intercontinental Airport bridge are significantly affected by

sunshine and the measured stresses are not reliable in the instances where significant

temperature variations occur in the gages (Writer, 2007). Due to the large variations in

temperature in the field, no correlation between the measured stress and the calculated

values was achieved.

The validated finite element structural model was used to evaluate the structural

responses of curved steel bridges with different boundary conditions subjected to

different thermal load combinations that may be experienced during the service life of the

bridges. Based upon the analyses, the following conclusions were reached:

279
The positioning of fixed bearings at a flexible interior bent is a better choice.

The demands of guide force, bridge movements at the ends, support shears and

bottom flange stresses are much smaller than those when fixed bearings are

positioned at the rigid abutment.

Neither tangential orientation nor chorded orientation toward a fixed bearing at

an interior bent is appropriate. The bridges actually experience displacements

towards a stationary point located away from the fixed bearing. When orienting

bearings towards this stationary point, smaller guide forces are achieved

compared to the traditional orientation towards the fixed bearing. A tangential

orientation was also considered. In general the tangential orientation did not

lead to a significant difference than orienting the bearing towards the fixed

bearing; however larger effects may result with wider bridges or girders with a

smaller radius of curvature. In comparing the different orientations, the best

behavior is achieved when the bearing is oriented towards the actual stationary

point on the bridge.

The possibility of guiding both girders as proposed by the AGC Contractors

was examined. The results showed that the expansion in the radial direction

must be accommodated by the rigid guides and large undesirable internal

forces are generated in the guides. Therefore, it is not appropriate to guide both

girders of curved bridges.

Girder movements due to thermal changes are accommodated through a

combination of pier deflection, bearing movement and girder deformations due

to induced thermal stresses. Based on the observation that the fraction of

movements accommodated by girder deformations due to induced thermal

stresses are negligible compared to the other two fractions for most of

280
composite steel bridges, a rigid body model was proposed to estimate the

location of the stationary point as given in Equation 9-1. The symbols in the

equation were defined in Section 8.5.1.

4
K ix cos i
x = r cos = i=1 R = K1x cos 1 + K 2 x cos 2 + K 3 x cos 3 + K 4 x cos 4 R
4
K1x + K 2 x + K 3 x + K 4 x
K ix
i=1 (9-1)

4

K iy sin i
i=1 R = K1 y sin 1 + K 2 y sin 2 + K 3 y sin 3 + K 4 y sin 4 R
y = r sin =
4
K1 y + K 2 y + K 3 y + K 4 y
K iy
i=1

By further observing that the stationary point is fairly insensitive to bridge

curvature, a simplified method was proposed analogous to straight bridges as

given in Equation 9-2. The symbols in the equation were defined in Section

8.5.2.

K A S A + K B S B + K C SC + K D S D
S= (9-2)
K A + K B + KC + K D

Both methods were compared to finite element solutions and could accurately

predict both the stationary point and thermal movements at the ends under

different pier flexure, bearing stiffness and bridge curvature. If rigid body

model is used, the chord-orientation toward the calculated stationary point is

recommended since best behavior is achieved. If the simplified method is

preferred, it is recommended the chord-orientation toward the fixed bent is


adopted because the simplified method cannot estimate the offset of the

281
location of stationary point in the radial direction for the following reasons: 1)

if the pier is rigid, the stationary point is actually close to the fixed bent; 2) if

the pier is flexible, most of thermal movements are accommodated by pier

deflection and the bearing orientation becomes less important.

By assuming that all bearings at pier tops are completely locked, the procedure

for calculating the most adverse shear force at pier tops using the both methods

is examined. Both methods give similar estimates.

The maximum magnitudes of thermal stresses were also evaluated under

different possible load conditions that may occur during the service life of

bridges. The results showed that the maximum tensile thermal stress in the

bottom flange may be as high as 6.1 ksi while the maximum compressive stress

is 4.3 ksi under the thermal load with 100-year return period. The 1st principal

stress in concrete deck is 0.61 ksi, which may crack concrete. The effect of

the unintended restraints due to the debris built up at the expansion joints was

also considered, which showed that large guide forces, support shears and

thermal stresses may be generated by the presence of the debris. Therefore,

regularly clearing of the debris at the expansion joints is important for routine

maintenance of the bridge to avoid unintended internal forces due to restricted


thermal movements.

Small errors in orienting the bearings do not significantly impact bridge

movements, support shears and thermal stresses. Detailed examination of

bridge layout based on the rigid body model, it is found that guiding the

bearing along the chord toward Bent 2 is appropriate for most bridges with the

radius larger than 800 ft. However, large errors may result in complete

locking of the rigid guides and inhibit the movement at the abutments similar

282
to the unintended restraints from the debris in the expansion joint. In this

situation, large internal forces may be produced in the guides. Therefore,

large errors in orienting the bearings should be avoided.

283
Appendix A Solar Geometry and Solar Radiation Model

SOLAR GEOMETRY

To calculate solar radiation incident on bridge surfaces, it is necessary to specify

the position of the considered surface relative to the sun. Some basic angles have to be

defined. The position of any point on the earths surface with respect to the suns rays

can be determined uniquely by three basic angles: the solar declination D, the hour angle

H and the latitude L.


Declination Angle D: The angle between the earth-sun line and the equatorial

plane. Declination changes with the date and is independent of the location. The

maximum declination is 23.45 on the summer/winter solstice. The declination will be

positive between March 22 and Sept. 22 approximately and negative at other times.

Equation A-1 adequately approximates the declination angle for engineering calculations

(Goswami et al, 1999). The unit of D in the equation is in degrees.

360(284 + n )
D = 23.45 sin (A-1)
365

The Hour Angle H of a Point (-180H180): the angle between the meridian

passing the point and the one directly in line with the suns rays. In solar radiation

calculations, local solar time (LST) must be used to express the time of a day. Solar time

is based on the apparent angular motion of the sun across the sky (Goswami et al, 1999).

The time when the sun crosses the meridian of the observer is the local solar noon. Solar

time does not necessarily coincide with the standard clock time. The relation between the
solar time and the standard time is as follows (Hsieh, 1986).

284
60 Minute
LST = LCT + (Lstd Lloc ) + ET DT (A-2)
15 Degree

where LCT is local civil time on an hour scale from 0 to 24; Lstd is the longitude of

the standard meridian in the local time zone; Lloc is the local longitude and DT is

modification factor for daylight saving time. DT=1 if daylight savings time is in effect,

and DT=0 otherwise. ET is the equation of time (Hsieh, 1986), which can be

approximated by Equation A-3:

ET = 9.87 sin 2 Z 7.53 cos Z 1.5 sin Z (A-3)

Where Z=360(n-81)/364, n is the day of the year and ET is in minutes. At local

solar noon the hour angle is zero and LST=12. The hour angle expresses the time of a

day with respect to solar noon. One hour of time equals 15 degrees of hour angle. The

hour angle can be defined by Equation A-4:

15 Degree
H= (LST 12) (A-4)
1 Hour

where 0LST24. The magnitude of H is symmetrical about solar noon.

Afternoon hours are designated as positive while morning hours are negative.

The Latitude L of a point (-90L90): the angle between the line from the

center of the earth to the point and the equator. The latitude is positive in the northern

hemisphere while it is negative in the southern hemisphere.

These three basic angles are illustrated in Figure A-1.

285
N
P

L
D

Figure A-1 Three Basic Solar Angles (Redrawn from Hsieh, 1986)

Besides the three basic geometric angles, three derived angles are also important in solar radiation
calculations, which are illustrated in

Figure A-2.

Figure A-2 Three Derived Solar Angles (Redrawn from Hsieh, 1986)

286
The Solar Zenith Angle : the angle between the suns rays and the direction

perpendicular to the earths surface.

The Solar Azimuth Angle : the angle between the suns rays and due south

measured in the horizontal, with westward being designated as positive for the northern
hemisphere.

The Solar Altitude Angle : the angle between the suns rays and the horizontal.

Obviously, += 90.

Figure A-3 shows all six angles in a coordinate system with the Z-axis coincident

with the earths axis and the XY plane coincident with the earths equatorial plane. Point

P represents the position of the observer.

Figure A-3 All Six Solar Angles

The suns zenith, altitude, and azimuth angles can be related to the suns

declination, hour angle and latitude by Equation A-5 (Hsieh, 1986).

287
sin = cos = cos D cos H cos L + sin D sin L (A-5)

The sunrise and sun set hours and the length of a particular day can be determined
by this equation. The hour angles Hs at sunset and Hr at sunrise can be found by solving
this equation when = 0.

H s = cos 1 ( tgD tgL ) (A-6)


H r = H s (A-7)

Hs takes positive values between 0 and 180 at sunset. The sunrise and sunset
time in hours from local solar noon is then equal to Hr /15 and Hs /15, respectively. The
length of the day in hours is given by:

2
cos 1 ( tgD tgL ) (A-8)
15

The azimuth angle of the suns rays can be determined by (Hsieh, 1986):

cos = sec (cos D cos H sin L sin D cos L ) (A-9)


sin = sec cos D sin H (A-10)

For horizontal surfaces, the above three derived angles can uniquely determine
their position with respect to the suns rays. However, for tilted surfaces, three additional
angles need to be defined.
The Tilted Angle : the angle between the surface and the horizontal plane.

288
The Surface Azimuth Angle : the angle between the surface normal and due
south measured in the horizontal plane, with westward designated as positive. For curved
and straight bridges, the azimuth angles of the web surface can be calculated as shown in
Figure A-4, respectively.

(a) Curved Bridge (b) Straight Bridge

Figure A-4 Azimuth Angle of Bridge Surfaces

The Solar Incidence Angle : the angle between the suns rays and the surface
normal. For a horizontal plane, the incidence angle and the zenith angle are the same.

Figure A-5 shows a coordinate system with the Z-axis coincident with the zenith.
The XY plane coincides with the horizontal surface. Point O represents the position of the
observer. The incidence angle can be also expressed in terms of five basic angles: the
declination angle D, the hour angle H, the latitude L, the tilted angle of the surface under
consideration , and the surface azimuth angle .

cos = cos D cos H sin L sin cos sin D cos L sin cos
(A-11)
+ cos D sin H sin sin + cos D cos H cos L cos + sin D sin L cos

289
Figure A-5 Incident Angle of A Tilted Surface

MODEL OF SOLAR RADIATION

The beam component of solar radiation incident on a horizontal surface can be


calculated using the following equations (Duffie and Beckman, 1991).

I bh = ktb I 0 (A-12)
k atu
sin( + 5 )
ktb = 0.9 (A-13)
1 0.00012 h h 500
(A-14)
k a = 0.94 0.0001(h 500) 500 < h 3000
0.69 h 3000

ktb is beam transmittance coefficient accounting for attenuation of solar radiation


by the atmosphere and can be determined by Equation A-13 (Dilger and Ghali, 1980). tu
is turbidity factor that is dependent on local weather conditions and also air pollution. It
generally ranges from 2 for a clear-sky condition to 8 in the presence of air pollution (Fu
290
et al, 1990). ka is relative atmosphere pressure at different altitudes h in meter can be
approximated by Equation A-14 [Ibrahim (1995), Kehlbeck (1975)]. is the solar
attitude angle as define in Figure A-2 and Figure A-3 and I0 refers to the extraterrestrial
radiation striking the horizontal surface. Equation A-15 can be used to calculate I0
(Duffie and Beckman, 2006):

360n
I 0 = I sc 1 + 0.033 cos cos (A-15)
365

where Isc is the solar constant and can be estimated as 1367 w/m2 (Iqbal, 1983); n
is the day of the year; and is the zenith angle of the sun as define in Figure A-2 and
Figure A-3. The method for calculating according to the sun-earth geometry was
provided in the previous section of this appendix.
The diffuse component of solar radiation incident on a horizontal surface can be
calculated using Equation A-16 [Fu et al (1990), Ibrahim (1995)].

I dh = ktd I 0 (A-16)

where ktd is diffuse transmittance coefficient which can be calculated by the


empirical Equation A-17 (Liu and Jordan, 1960).

ktd = 0.271 0.294ktb (A-17)

where ktb is beam transmittance coefficient as defined in Equation A-13.

291
Appendix B Validation of ANSYS Model

FSEL BRIDGE

The following figures compare the measured temperature histories at the locations
of thermocouples with those obtained from both the 2-d and 3-d ANSYS models on
August 22, 2006. Only those that were not given in Chapter 4 are given.

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
39 Measured
20 2 8 12
I 18
3 E 7 13 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-1 Temperature History at the Location of Thermocouple 3

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
20 2 8
39
12 18
Measured
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-2 Temperature History at the Location of Thermocouple 4

292
120

Temperature ( F) 100

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
9 11 20 19 34
21 1 10
39 Measured
20 2
E
8 12
I 18
3 7 13 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-3 Temperature History at the Location of Thermocouple 5

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
40 25
26
37
38
31
32

22 27 28 33
21 1 10 9
39
11 20 19 34 Measured
20 2 8 12
I 18
3 E 7 13 17
3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-4 Temperature History at the Location of Thermocouple 6

293
120

Temperature ( F) 100

80
o

60

23 35 29
40 24
25
26
36
37
38
30
31
32

22 27 28 33
21 1 10 9 11 20 19 34 Measured
20 2 8
39
12 18
3 E 7 13
I 17
3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)
Figure B-5 Temperature History at the Location of Thermocouple 7

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
20 2 8
39
12 18
Measured
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-6 Temperature History at the Location of Thermocouple 8

294
120

Temperature ( F) 100

80
o

60

23 35 29
24 36 30
31
40 25
26
37
38 32

22 27 28 33
21 1 10 9
39
11 20 19 34
Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-7 Temperature History at the Location of Thermocouple 9

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9
39
11 20 19 34
Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-8 Temperature History at the Location of Thermocouple 11

295
120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
39 Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-9 Temperature History at the Location of Thermocouple 12

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
9 11 20 19 34
21 1 10
39 Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-10 Temperature History at the Location of Thermocouple 13

296
120

Temperature ( F) 100

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
39 Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-11 Temperature History at the Location of Thermocouple 14

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
39 Measured
20 2 8 12
I 18
3 E 7 13 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-12 Temperature History at the Location of Thermocouple 15

297
120

Temperature (oF) 100

80

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
20 2 8
39
12 18
Measured
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-13 Temperature History at the Location of Thermocouple 16

140

120

100
Temperature ( F)
o

80

60 29
23 35
24 36 30
25 37 31
26 38 32
40
22 27 28 33
21 1 10 9
39
11 20 19 34 Measured
2 8 12 18
20 3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-14 Temperature History at the Location of Thermocouple 17

298
140

120

100
Temperature ( F)
o

80

60 23 29
35
24 36 30
25 37 31
26 38 32
40
22 27 28 33
21 1 10 9 11 20 19 34
2 8
39
12 18
Measured
20 3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-15 Temperature History at the Location of Thermocouple 18

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
20 2 8
39
12 18
Measured
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-16 Temperature History at the Location of Thermocouple 19

299
140

120

100
Temperature ( F)
o

80

60
23 35 29
24 36 30
25 37 31
26 38 32
40
22 27 28 33
9 11 34
21 1 10
39
20 19
Measured
2 8 12 18
20 3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-17 Temperature History at the Location of Thermocouple 23

120

100
Temperature ( F)

80
o

60

23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
10 9 11 20 19 34
21 1
39 Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-18 Temperature History at the Location of Thermocouple 24

300
120

100
Temperature ( F)

80
o

60

23 35 29
24 36 30
31
40 25
26
37
38 32

22 27 28 33
9 11
21 1 10
39
20 19 34
Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-19 Temperature History at the Location of Thermocouple 25

120

100
Temperature ( F)

80
o

60

23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
20 2 8
39
12 18
Measured
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-20 Temperature History at the Location of Thermocouple 26

301
140

120

100
Temperature ( F)
o

80

60
23 35 29
24 36 30
25 37 31
26 38 32
40
22 27 28 33
21 1
2
10 9
39
11
12
20 19 34
Measured
8 18
20 3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-21 Temperature History at the Location of Thermocouple 35

120

100
Temperature ( F)

80
o

60

23 35 29
24 36 30
40 25
26
37
38
31
32

22 27 28 33
9
21 1 10
39
11 20 19 34
Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-22 Temperature History at the Location of Thermocouple 36

302
120

100
Temperature ( F)

80
o

60

23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 10 9 11 20 19 34
1
39 Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-23 Temperature History at the Location of Thermocouple 37

120

100
Temperature ( F)

80
o

60
23 35 29
24 36 30
25 37 31
40 26 38 32

22 27 28 33
21 1 10 9 11 20 19 34
39 Measured
20 2 8 12 18
3 E 7 13
I 17 3d ANSYS Model
4 5 6 14 15 16 2d ANSYS Model
0
0 4 8 12 16 20 24
Time (hr)

Figure B-24 Temperature History at the Location of Thermocouple 38

303
The following figures compare the measured temperature over the instrumented
section with those obtained from 3-d ANSYS models on August 22, 2006. Only those
that were not given in Chapter 4 are given.

Figure B-25 Sectional Temperature Distribution at 0 am

Figure B-26 Sectional Temperature Distribution at 2 am


304
Figure B-27 Sectional Temperature Distribution at 4 am

Figure B-28 Sectional Temperature Distribution at 8 am

305
Figure B-29 Sectional Temperature Distribution at 12 am

Figure B-30 Sectional Temperature Distribution at 14 am

306
Figure B-31 Sectional Temperature Distribution at 16 am

Figure B-32 Sectional Temperature Distribution at 18 am

307
Figure B-33 Sectional Temperature Distribution at 20 am

Figure B-34 Sectional Temperature Distribution at 22 am

308
Figure B-35 Sectional Temperature Distribution at 24 am

INTERNATIONAL AIRPORT BRIDGE

The following figures compare the measured temperature histories at the locations
of thermocouples with those obtained from the 2-d ANSYS models on July 17, 2006.
Only those that were shown given in Chapter 4 are shown.

309
140
Measured

120 2d ANSYS Model

100
Temperature (oF)

80

60
20 24
21 25
22 26
40 23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-36 Temperature History at the Location of Thermocouple 3

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-37 Temperature History at the Location of Thermocouple 4

310
120
Measured
2d ANSYS Model
Temperature (oF) 100

80

60
20 24
21 25
40 22
23 28 29
26
27
30
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-38 Temperature History at the Location of Thermocouple 5

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60
20 24
21 25
40 22
23
26
28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-39 Temperature History at the Location of Thermocouple 6

311
120
Measured
2d ANSYS Model
Temperature (oF) 100

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-40 Temperature History at the Location of Thermocouple 7

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
21 25
40 22 26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20 3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-41 Temperature History at the Location of Thermocouple 8

312
120
Measured
2d ANSYS Model
Temperature (oF) 100

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-42 Temperature History at the Location of Thermocouple 9

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-43 Temperature History at the Location of Thermocouple 11

313
120
Measured
2d ANSYS Model
Temperature (oF) 100

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-44 Temperature History at the Location of Thermocouple 12

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-45 Temperature History at the Location of Thermocouple 13

314
120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-46 Temperature History at the Location of Thermocouple 14

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-47 Temperature History at the Location of Thermocouple 15

315
120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-48 Temperature History at the Location of Thermocouple 16

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-49 Temperature History at the Location of Thermocouple 17

316
120
Measured
2d ANSYS Model
Temperature (oF) 100

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-50 Temperature History at the Location of Thermocouple 18

120
Measured
2d ANSYS Model
100
Temperature (oF)

80

60

20 24
40 21
22
25
26
23 28 29 30 27
1 31 32 13 19
7
2 8 12 18
20
3 E 6
C 14 I 17

4 5 9 11 15 16
10
0
0 4 8 12 16 20 24
Time (hr)

Figure B-51 Temperature History at the Location of Thermocouple 19

317
Appendix C Applicability of Proposed Vertical Thermal Gradient

The comparisons between the extreme distributions obtained from the above
ANSYS model and the proposed ones are conducted for different situations.

EFFECT OF BRIDGE ORIENTATION


10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(a) Vertical Temperature Distribution for Heating


10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling


Figure C-1 Effect of Bridge Orientation on Vertical Temperature Distribution (XI=0)

318
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)

(a) Vertical Temperature Distribution for Heating


10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling
Figure C-2 Effect of Bridge Orientation on Vertical Temperature Distribution (XI=30)

319
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)
(b) Vertical Temperature Distribution for Cooling

Figure C-3 Effect of Bridge Orientation on Vertical Temperature Distribution (XI=60)

320
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling

Figure C-4 Effect of Bridge Orientation on Vertical Temperature Distribution (XI=90)


321
EFFECT OF OVERHANG LENGTH OF CONCRETE DECK

10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)

(b) Vertical Temperature Distribution for Cooling

Figure C-5 Effect of Overhang length on Vertical Temperature Distribution (B1=26 in)

322
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)
(b) Vertical Temperature Distribution for Cooling
Figure C-6 Effect of Overhang length on Vertical Temperature Distribution (B1=44 in)

323
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)
(b) Vertical Temperature Distribution for Cooling
Figure C-7 Effect of Overhang length on Vertical Temperature Distribution (B1=62 in)

324
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling


Figure C-8 Effect of Overhang length on Vertical Temperature Distribution (B1=80 in)

325
EFFECT OF THICKNESS OF CONCRETE DECK

10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure C-9 Effect of The Thickness of Concrete Deck (T1=8 in)

326
15

Concrete
0
-10 -5 0 5 10 15 20
Height (inch)

-15

Steel Girder
-30

-45

Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
15
Concrete

0
-15 -10 -5 0 5 10
Height (inch)

-15

Steel Girder

-30

-45

Proposed
ANSYS Data
-60
o
Temperature ( F)
(b) Vertical Temperature Distribution for Cooling

Figure C-10 Effect of The Thickness of Concrete Deck (T1=10 in)

327
15

Concrete
0
-10 -5 0 5 10 15 20
Height (inch)

-15

Steel Girder
-30

-45

Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
15
Concrete

0
-15 -10 -5 0 5 10
Height (inch)

-15

Steel Girder

-30

-45

Proposed
ANSYS Data
-60
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling
Figure C-11 Effect of The Thickness of Concrete Deck (T1=12 in)

328
15

Concrete
0
-10 -5 0 5 10 15 20
Height (inch)

-15

Steel Girder
-30

-45

Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
15
Concrete

0
-15 -10 -5 0 5 10
Height (inch)

-15

Steel Girder

-30

-45

Proposed
ANSYS Data
-60
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling

Figure C-12 Effect of The Thickness of Concrete Deck (T1=14 in)

329
EFFECT OF WIDTH OF CONCRETE DECK

10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(a) Vertical Temperature Distribution for Heating


10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure C-13 Effect of The Width of Concrete Deck (S=144 in)

330
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)

(a) Vertical Temperature Distribution for Heating


10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling

Figure C-14 Effect of The Width of Concrete Deck (S=192 in)

331
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)
(b) Vertical Temperature Distribution for Cooling

Figure C-15 Effect of The Width of Concrete Deck (S=240 in)

332
10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-12 -10 -8 -6 -4 -2 0 2 4 6

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling

Figure C-16 Effect of The Width of Concrete Deck (S=288 in)

333
EFFECT OF DEPTH OF THE GIRDER

10

Concrete
0
-10 -5 0 5 10 15 20

-10
Height (inch)

-20

Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
o
Temperature ( F)

(a) Vertical Temperature Distribution for Heating


10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8

-10
Height (inch)

-20
Steel Girder
-30

-40

-50
Proposed
ANSYS Data
-60
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling

Figure C-17 Effect of The Depth of Steel Girder (D=66 in )

334
10
Concrete
0
-10 -5 0 5 10 15 20
-10

-20
Height (inch)

-30

-40
Steel Girder
-50

-60

-70

-80 Proposed
ANSYS Data
-90
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
0
-10 -8 -6 -4 -2 0 2 4 6 8
-10

-20
Height (inch)

-30

Steel Girder -40

-50

-60

-70

-80 Proposed
ANSYS Data
-90
o
Temperature ( F)
(b) Vertical Temperature Distribution for Cooling

Figure C-18 Effect of The Depth of Steel Girder (D=88 in )

335
10
Concrete
-10 -5 0 5 10 15 20
-10

-30
Height (inch)

-50
Steel Girder

-70

-90
Proposed
ANSYS Data
-110
Temperature (oF)
(a) Vertical Temperature Distribution for Heating
10
Concrete
-8 -4 0 4 8 12
-10

-30
Height (inch)

Steel Girder -50

-70

-90
Proposed
ANSYS Data
-110
Temperature (oF)
(b) Vertical Temperature Distribution for Cooling
Figure C-19 Effect of The Depth of Steel Girder (D=110 in )

336
10
Concrete
-10 -5 -10 0 5 10 15 20

-30
Height (inch)

-50

Steel Girder
-70

-90

-110
Proposed
ANSYS Data
-130
Temperature (oF)

(a) Vertical Temperature Distribution for Heating


10
Concrete
-8 -4 -10 0 4 8 12

-30
Height (inch)

-50
Steel Girder
-70

-90

-110
Proposed
ANSYS Data
-130
Temperature (oF)

(b) Vertical Temperature Distribution for Cooling


Figure C-20 Effect of The Depth of Steel Girder (D=132 in )

337
Appendix D Extreme Value Analysis

VERTICAL LINEAR TEMPERATURE DIFFERENCE (TV)

The plots of the empirical CDFs for vertical linear temperature difference on the
Weibull probability paper are shown in Figure D-1 and Figure D-2.

12

10
Reduced Variate, h

4
-4.47 -4.44 -4.41 -4.38

Reduced Variate, x

Figure D-1 Weibull Plot of Tv for Maxima

10

8
Reduced Variate, h

4
-4.74 -4.72 -4.70 -4.68 -4.66
Reduced Variate, x

Figure D-2 Weibull Plot of Minus Tv for Maxima

338
Both the right tails in Figure D-1 and Figure D-2 agree well with straight lines, so
Weibull type distributions are accepted. By the least square method described above, the
parameters are estimated as listed in Table D-1.

Table D-1 Parameter Estimate for Tv

Parameter Tv for Maxima Tv for Minima


Domain of Attraction Weibull Weibull
The Location 108 129
The Scale 92.86 121.53
The Shape Factor 66.38 65.76

To get how good the predictions are, the accepted distributions are also plotted
against the empirical CDF on an arithmetic scale as shown in Figure D-3 and Figure D-4.

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
21 23 25 27 29

Vertical Linear Temperature Difference (oF)

Figure D-3 Arithmetic Plot of Tv for Maxima

339
1.00

0.99

CDF 0.98

0.97

From Sample
From Least Square Method
0.96
16 18 20 22 24 26 28

Minus Vertical Linear Temperature Difference (oF)

Figure D-4 Arithmetic Plot of Minus Tv for Maxima

In order to make clear in what environmental conditions the extreme vertical linear temperature
differences are achieved, the histograms of month and time for them in the right tails are plotted in

Figure D-5 to Figure D-8.

19%

16%
15%

11%

7%
7%
6%
5% 5%

3%
3%
2%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-5 Histogram of Month for Maximum Tv

340
33%

23%

8%

5% 5% 5% 5%
3% 3%
1%
0% 1% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-6 Histogram of Time for Maximum Tv

31%

29%

19%

9%
8%

3%
1%
0% 0% 0% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-7 Histogram of Month for Minimum Tv

341
29%
27%

14%

11% 10%
8%

0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 1% 0% 0% 0% 0% 0% 0% 0%

Figure D-8 Histogram of Time for Minimum Tv

As can be seen from the figures, it is hard to detect the trend for the maximum
vertical linear temperature difference since its extremes may occur during the whole year.
However, the time column graph seems to show that the maximum is achieved when the
ambient air temperature drops rapidly after a days intense sunshine. Steel girders
become cool due to its high conductivity while the concrete deck still holds a lot of heat.
Most of the minimum vertical linear temperature difference occurs in winter (Dec., Jan.,
and Feb.) when the web of the steel girder, directly exposed to sunshine, gets heat up
quickly and the concrete deck is still cool. It is reasonable because the webs are sunlit at 2
pm to 3 pm when solar radiation occurs is most intense due to the low altitude of sun in
winter.

342
TRANSVERSE LINEAR TEMPERATURE DIFFERENCE (TH)

The plot of the empirical CDFs for transverse linear temperature difference on the
Gumbel probability paper is shown in Figure D-9.

10
Reduced Variate, h

4
14 15 16 17 18
Transverse Linear Temperature Difference (oF)

Figure D-9 Gumbel Plot of Th for Maxima

It is clear that the plot looks like a straight line, so the Gumbel distribution is
accepted. By the least square method described above, the parameters are estimated as
listed in Table D-2.

Table D-2 Parameter Estimate for Th

Parameter Th for Maxima


Domain of Attraction Gumbel
The Location 10.71
The Scale 0.78
343
To get indication of how good the predictions are, the accepted distributions are
also plotted against the empirical CDF on an arithmetic scale as shown in Figure D-10. In
order to make clear in what environmental conditions the extreme transverse linear
temperature difference is achieved, the histograms of month and time for it in the right
tail are plotted in Figure D-11 to Figure D-12.

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
14 15 16 17 18
Transverse Linear Temperature Difference (oF)

Figure D-10 Arithmetic Plot of Th for Maxima

As can be seen from the figures, an overwhelming majority of extreme transverse


linear temperature difference occurs at 9 am in the morning during daily thermal cycle
because the steel webs are directly sunlit and the incident angle is smallest at these hours.
During annual thermal cycle, the extremes are likely to occur in seasons with intense

344
solar radiation, low ambient temperature, and low attitude of the sun leading to a smaller
incident angle.

34%

27%

13%

9%

4% 3% 3% 4%

1% 1% 1%
0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-11 Histogram of Month for Maximum Th

63%

14%
13%

5% 5%

0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-12 Histogram of Time for Maximum Th


345
MAXIMUM RESIDUAL TEMPERATURE

The empirical CDFs for maximum steel and concrete residual temperatures (Tres,s
and Tres,c) on the Gumbel probability paper are plotted in Figure D-13 to Figure D-14.

10
Reduced Variate, h

4
30 32 34 36 38
Maximum Steel Residual Tempeature (oF)

Figure D-13 Gumbel Plot of Tres,s for Maxima

10
Reduced Variate, h

4
31 33 35 37 39
Concrete Residual Temperature (oF)

Figure D-14 Gumbel Plot of Tres,c for Maxima

346
Both the right tails in Figure D-13 and Figure D-14 are not convex, so a Gumbel
type distribution is accepted. By the least square method described above, the parameters
are estimated as listed in Table D-3.

Table D-3 Parameter Estimate for Tres,s and Tres,c

Parameter Tres,s for Maxima Tres,c for Maxima


Domain of Attraction Gumbel Gumbel
The Location 24.05 25.47
The Scale 1.41 1.29

To get how good the predictions are, the accepted distributions are also plotted
against the empirical CDF on an arithmetic scale as shown in Figure D-15 and Figure
D-16.

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
30 32 34 36 38
Maximum Steel Residual Tempeature (oF)

Figure D-15 Arithmetic Plot of Tres,s for Maxima

347
1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
31 33 35 37 39
Concrete Residual Temperature (oF)

Figure D-16 Arithmetic Plot of Tres,c for Maxima

In order to make clear in what environmental conditions the maximum residual


temperatures are achieved, the histograms of month and time for the maximum concrete
and steel residual temperatures in the right tail are plotted in Figure D-17 and Figure
D-20.

37%

32%

13%

5% 5%
3% 2% 3%
1% 1% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

348
Figure D-17 Histogram of Month for Maximum Tres,s

54%

43%

1%
0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 1% 1% 0% 0% 0% 0% 0% 0%

Figure D-18 Histogram of Time for Maximum Tres,s

27%

22%
21%

12%

8%
6%

3%
1%
0% 0% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-19 Histogram of Month for Maximum Tres,c


349
53%

39%

3% 3%
1% 1%
0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-20 Histogram of Time for Maximum Tres,c

As can be seen from the figures, the maximum steel residual temperature most
probably occurs in seasons (March and April) with intense solar radiation and low
ambient temperature during annual thermal cycles. During daily thermal cycles, the
maximum occurs mainly at 9 am and 10 am when the steel webs are directly exposed to
sunshine. In addition, the maximum steel residual temperature occurs at the sunlit corner
of bottom flanges.
The maximum concrete residual temperature most probably occurs in seasons
(April, May and June) with intense solar radiation and medium ambient temperature
during annual thermal cycles. During daily thermal cycles, the maximum occurs mainly
at 11 am and 12 pm when the top surface of the concrete deck is directly exposed to
sunshine. The maximum concrete residual temperature occurs at top concrete surface.

350
VERTICAL TEMPERATURE GRADIENT FOR HEATING

In Chapter 4, the vertical temperature gradient for heating is parameterized using


three parameters: T1, T1-T2 and T3-T2. The plots of the empirical CDFs for these three
parameters on the Gumbel probability paper are shown in Figure D-21 to Figure D-23.

10
Reduced Variate, h

4
22 23.5 25 26.5 28
T1 (oF)

Figure D-21 Gumbel Plot of T1 for Maxima

10
Reduced Variate, h

4
30.0 31.5 33.0 34.5 36.0
T 1-T 2 (oF)

351
Figure D-22 Gumbel Plot of T1- T2 for Maxima

10
Reduced Variate, h

4
3.8 4.1 4.4 4.7 5.0
T 3-T 2 (oF)

Figure D-23 Gumbel Plot of T3- T2 for Maxima

All the three figures look like a straight line, so the Gumbel distribution is
accepted. By the least square method described above, the parameters are estimated as
listed in Table D-4.

Table D-4 Parameter Estimate for T1, T1-T2 and T3-T2

Parameter T1 for Maxima T1-T2 for Maxima T3-T2 for Maxima


Domain of Attraction Gumbel Gumbel Gumbel
The Location 19.07 26.06 2.83
The Scale 0.87 1.07 0.81

To get how good the predictions are, the accepted distributions are also plotted against the empirical
CDFs on an arithmetic scale as shown in

Figure D-24 to Figure D-26.

352
1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
23 24 25 26 27 28
T1 (oF)

Figure D-24 Arithmetic Plot of T1 for Maxima

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
31 32 33 34 35 36
T 1-T 2 (oF)

Figure D-25 Arithmetic Plot of T1-T2 for Maxima

353
1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
3.8 4.2 4.6 5.0
T 3-T 2 (oF)

Figure D-26 Arithmetic Plot of T3-T2 for Maxima

The magnitude of T1 does not matter since it just translates the gradient to the
right or left. The magnitude of vertical temperature gradient for heating depends on T1-T2
and T3-T2. In order to make clear in what environmental conditions the extremes are
achieved, the histograms of month and time for the maximum (T1-T2) and (T3-T2) in the
right tail during heating are plotted in Figure D-27 to Figure D-30.

354
26%
25%

21%

11%
9%
9%

0% 0% 0% 0% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-27 Histogram of Month for Maximum (T1-T2)

52%

34%

14%

0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-28 Histogram of Time for Maximum (T1-T2)

355
23%

20%

17%

15%

13%

5% 5%

0% 0% 0% 0% 0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-29 Histogram of Month for Maximum (T3-T2)

46%

37%

9%
8%

0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-30 Histogram of Time for Maximum (T3-T2)

356
It can be seen that during annual thermal cycle, both thermal parameters most
probably achieve their maximum from April to August in which solar radiation is most
intense. During the daily thermal cycles, the time column graphs show solar radiation has
a great influence on positive vertical temperature gradients. Both of them dominantly
reach extremes at 1 pm and 2 pm when solar radiation is most intense during daily
thermal cycle. Therefore, the positive vertical temperature gradients are mainly
determined by solar radiation.

VERTICAL TEMPERATURE GRADIENT FOR COOLING

In Chapter 5, the vertical temperature gradient for cooling is parameterized using


three parameters: T1, T2-T1 and T2-T3. The plots of the empirical CDFs for these three
parameters on the Gumbel probability paper are shown in Figure D-31 to Figure D-33.

10
Reduced Variate, h

4
3 4 5 6
Minus T 1 (oF)

Figure D-31 Gumbel Plot of T1 for Minima

357
10

Reduced Variate, h

4
14 16 18 20 22
T 2-T 1 (oF)

Figure D-32 Gumbel Plot of T2- T1 for Maxima

10
Reduced Variate, h

4
21 23 25 27 29 31
T 2-T 3 (oF)

Figure D-33 Gumbel Plot of T2- T3 for Maxima

358
The convex right tail in Figure D-31 suggests a Weibull type distribution for T1
for minima. To make sure that the Gumbel type is not wrongly rejected, the empirical
CDF is also plotted on the Weibull probability paper as shown in Figure D-34.
Figure D-34 looks more like a straight line than Figure D-31, so the Weibull
distribution is accepted. By the least square method described above, the parameters are
estimated as listed in Table D-5.

10
Reduced Variate, h

4
-2.1 -2.0 -1.9 -1.8 -1.7 -1.6
Reduced Variate, x

Figure D-34 Weibull Plot of T1 for Minima

Table D-5 Parameter Estimate for T1, T2-T1 and T2-T3

Thermal Parameter T1 for Minima T2-T1 for Maxima T2-T3 for Maxima
Domain of Attraction Weibull Gumbel Gumbel
The Location 11.0 7.86 14.01
The Scale 11.25 1.49 1.70
Shape Factor 11.42 N/A N/A

359
To get how good the predictions are, the accepted distributions are also plotted against the empirical
CDF on an arithmetic scale as shown in

Figure D-35 to Figure D-37.

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
3 4 5 6

Minus T1 (oF)

Figure D-35 Arithmetic Plot of T1 for Minima

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
14 16 18 20 22
T 2-T 1 (oF)

Figure D-36 Arithmetic Plot of T2-T1 for Maxima

360
1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
21 23 25 27 29 31
T 2-T 3 (oF)

Figure D-37 Arithmetic Plot of T2-T3 for Maxima

The magnitude of T1 is immaterial since it just translates the gradient to the right
or left. The magnitude of vertical temperature gradient for cooling depends on T2-T1 and
T2-T3. In order to make clear in what environmental conditions the extremes are
achieved, the histograms of month and time for the maximum (T2-T1) and (T2-T3) during
cooling in the right tail are plotted in Figure D-38 to Figure D-41.

361
19%

16%

13%

10% 10%

6% 6% 6%
5%
3%
3% 3%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-38 Histogram of Month for Maximum (T2-T1)

39%

14%

7% 8%
6% 6%

2% 2%
1% 1% 1% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-39 Histogram of Time for Maximum (T2-T1)

362
17%
17%

11% 11%
9%

6% 6%
5% 5%
5%
4% 4%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-40 Histogram of Month for Maximum (T2-T3)

55%

13%

5% 5% 4%
4%
2% 1% 2%
1% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 1% 0% 1% 0%

Figure D-41 Histogram of Time for Maximum (T2-T3)

363
As can be seen from the figures, during annual thermal cycle, it is hard to detect
the trend for both since their extremes may occur during the whole year. The time column
graphs show that the extremes are likely to occur at 1 am during daily thermal cycle when
the top concrete surface cools down while the concrete core still holds a lot of heat.

MINIMUM RESIDUAL TEMPERATURE

The empirical CDFs for minimum steel and concrete residual temperatures (Tres,s
and Tres,c) on the Gumbel probability paper are plotted in Figure D-42 and Figure D-43.

10
Reduced Variate, h

4
9 10 11 12 13
Minus Steel Residual Temperature (oF)

Figure D-42 Gumbel Plot of minus Tres,s for Maxima

364
10

Reduced Variate, h

4
14 16 18 20 22
Minus Concrete Residual Temperature (oF)

Figure D-43 Gumbel Plot of minus Tres,c for Maxima

Both the right tails in Figure D-42 and Figure D-43 are not convex, so a Gumbel
type distribution is accepted. By the least square method described above, the parameters
are estimated as listed in Table D-6.

Table D-6 Parameter Estimate for Tres,s and Tres,c

Parameter Tres,s for Minima Tres,c for Minima


Domain of Attraction Gumbel Gumbel
The Location 6.72 8.82
The Scale 0.58 1.25

To get how good the predictions are, the accepted distributions are also plotted
against the empirical CDF on an arithmetic scale as shown in Figure D-44 and Figure
D-45.

365
1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
9 10 11 12 13
Minus Steel Residual Temperature (oF)

Figure D-44 Arithmetic Plot of minus Tres,s for Maxima

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
14 16 18 20 22
Minus Concrete Residual Temperature (oF)

Figure D-45 Arithmetic Plot of minus Tres,c for Maxima

366
In order to make clear in what environmental conditions the minimum residual
temperatures are achieved, the histograms of month and time for the minimum concrete
and steel residual temperatures in the right tail are plotted in Figure D-46 to Figure D-49.

25%

21%
20%

10%

7%
6%

3% 3%
1% 1%
1% 1%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-46 Histogram of Month for Minimum Tres,s

42%

38%

6%
3% 4% 4%
2%
0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 1% 0% 0% 0% 0% 0%

Figure D-47 Histogram of Time for Minimum Tres,s


367
21%

14%
13%

11%

9%
8%

5% 5%
4% 4% 4%
3%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-48 Histogram of Month for Minimum Tres,c

47%

22%

5%
3% 3% 3%
1% 1% 1%
0% 0% 1% 0% 0% 0% 0% 0% 0% 0% 1% 0% 0% 1% 0%

Figure D-49 Histogram of Time for Minimum Tres,c


368
As can be seen from the figures, the minimum steel residual temperature may
occur in any month during annual thermal cycles. It is hard to detect the trend and what
weather conditions lead to extremes. They seem to be daily behaviors rather than
seasonal behavior. During daily thermal cycles, the minimum occurs mainly at 8 am and
9 am when one side of steel webs, directly exposed to sunshine, achieves its maximum
and the other shaded achieves its minimum.
Similar to the minimum steel residual temperature, the minimum concrete
residual temperature may occur in any month during annual thermal cycles. During daily
thermal cycles, the maximum most probably occurs at 1 am when the top surface of the
concrete deck cools down and the core is still warm.

TEMPERATURE DIFFERENCE BETWEEN SUNLIT AND SHADED WEBS (TWEB)

The plot of the empirical CDFs for the average temperature difference between
sunlit and shaded webs on the Weibull probability paper is shown in Figure D-50.

369
10

Reduced Variate, h

4
-2.8 -2.6 -2.4 -2.2 -2.0
Reduced Variate, x

Figure D-50 Gumbel Plot of Tweb for Maxima

The plot looks like a straight line, so the Weibull distribution is accepted. By the
least square method described above, the parameters are estimated as listed in Table D-7.

Table D-7 Parameter Estimate for Tweb

Parameter Th for Maxima


Domain of Attraction Weibull
The Location 49.9
The Scale 31.14
Shape Factor 6.31

To get indication of how good the predictions are, the accepted distributions are
also plotted against the empirical CDF on an arithmetic scale as shown in Figure D-51. In
order to make clear in what environmental conditions the maximum is achieved, the

370
histograms of month and time for it in the right tail are plotted in Figure D-52 and Figure
D-53.

1.00

0.99
CDF

0.98

0.97

From Sample
From Least Square Method
0.96
34 36 38 40 42 44
Temperature Difference Between Sunlit and Shaded Webs (oF)

Figure D-51 Arithmetic Plot of Tweb for Maxima

37%

29%

11%

8%

4% 3% 3% 2%
1% 1% 1%
0%

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure D-52 Histogram of Month for Maximum Tweb

371
67%

17%
11%

2% 3%
0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Figure D-53 Histogram of Time for Maximum Tweb

As can be seen from the figures, similar to the maximum transverse linear
temperature difference, an overwhelming majority of maximum temperature difference
occurs at 9 am in the morning during daily thermal cycle because the steel webs are
subjected to side sun shine and the incident angle is smallest at these hours. During
annual thermal cycle, the extremes are likely to occur in seasons with intense solar
radiation, low ambient temperature, and low attitude of the sun leading to a smaller
incident angle.

372
Reference

AASHTO, (1994), LRFD Bridge Design Specifications, American Association of State


Highway and Transportation Officials, First Edition, Washington, D.C.
AASHTO, (1998), LRFD Bridge Design Specifications, American Association of State
Highway and Transportation Officials, Second Edition, Washington, D.C.
AASHTO, (2004), LRFD Bridge Design Specifications, American Association of State
Highway and Transportation Officials, Third Edition, Washington, D.C.
AASHTO, (2007), LRFD Bridge Design Specifications, American Association of State
Highway and Transportation Officials, Fourth Edition, Washington, D.C.
AASHTO, (1992), Standard Specifications for Highway Bridges, American Association
of State Highway and Transportation Officials, 15th Edition, Washington, D.C.
AASHTO, (1996), Standard Specifications for Highway Bridges, American Association
of State Highway and Transportation Officials, 16th Edition, Washington, D.C.
AASHTO, (2002), Standard Specifications for Highway Bridges, American Association
of State Highway and Transportation Officials, 17th Edition, Washington, D.C.
AASHTO, (2003), Guide Specifications of Horizontally Curved Highway Bridges,
American Association of State Highway and Transportation Officials,
Washington, D.C.
ANSYS Theory Manual for Version 5.7, Swanson Analysis Systems, Inc., 1992
ANSYS Online Help (2007), www.ansys.com
Arikan Y. (2006), Field Monitoring of Thermal Effects in Steel Box Girder Bridges,
Masters Thesis, The University of Houston
Barber E.S. (1957), Calculation of Maximum Pavement Temperatures from Weather
Reports, Highway Research Board Bulletin 168, pp. 1-8
Berman S.M. (1971), Mathematical Statistics: an Introduction Based on The Normal
Distribution, Scranton, Intext Educational Publisher
Berwanger C., and Symko Y. (1975), Thermal stresses in steel-concrete composite
bridges. Canadian Journal of Civil Engineering, Vol. 2 (1): pp. 66-84
Berwanger C. (1983), Transient thermal behavior of composite bridges. Journal of
Structural Engineering, ASCE, Vol. 109 (10): pp. 2325-2339
Bobba S. (2003), Field Measurements of Diaphragm and Cross-Frame Stresses in Steel
Box Girder Bridge with Skewed Support, Masters Thesis, The University of
Houston

373
Bradberry T.E., Cotham J.C, and Medlock R.D (2005), Elastomeric Bearings for Steel
Trapezoidal Box Girder Bridges, Transportation Research Board, pp. 27-38
British Standard BS 5400 (1978), Steel, Concrete and Composite Bridges, Part I, General
Statement, British Standards Institution, Crowthorne, Berkshire, England
Budin R., and Budin L. (1982), A Mathematical Model for Shading Calculations, Solar
Energy, Vol. 29, No. 4, pp. 339-349
Campbell Scientific Inc., www.campbellsci.com
Castillo E. (1988), Extreme Value Theory in Engineering, Academic Press, Boston
Dilger W., Beauchamp J.C., Cheung M.S., Ghali A. (1981), Field Measurements of
Muskwa River Bridge, Journal of Structural Division, ASCE, Vol. 107, No. ST11,
pp. 2147-2161
Dilger W.H., Ghali A., Cheung M.S. and Maes M.A. (1983), Temperature Stresses in
Composite Box girder Bridges, Journal of Structural Engineering, ASCE, Vol.
109, No. 6, pp. 1460-1478
D.S. Brown Company, www.dsbrown.com
Duffie J.A. and Beckman W.A (1980), Solar Engineering of Thermal Processes, 2nd
Edition, John Wiley & Son, Inc., Hoboken, New Jersey
Duffie J.A. and Beckman W.A (2006), Solar Engineering of Thermal Processes, 3rd
Edition, John Wiley & Son, Inc., Hoboken, New Jersey
Elbadry M.M. and Ghali A. (1983), Temperature Variation in Concrete Bridges, Journal
of Structural Engineering, ASCE, Vol. 109, No. 10, pp. 2355-2374
Elbadry M.M. and Ghali A. (1983), Nonlinear Temperature Distribution and its Effects
on Bridges, IABSE Periodica, No. P66/83, pp. 169-191
Emanuel J.H., and Hulsey J.L. (1978), Temperature Distributions in Composite Bridges,
Journal of Structural Division, ASCE, Vol. 104, No. ST1, pp. 65-78
Emanuel J.H., Lewis D. B. (1981), Abutment-Thermal Interaction of Composite Bridge,
Journal of Structural Division, ASCE, Vol. 107, No. ST11, pp. 2111-2127
Emerson M. (1976), Bridge Temperatures Estimated From the Shaded Temperatures,
TRRL Supplementary Report 696, Department of the Environment, Department
of Transport, Crowthorne, Berkshire, England
Emerson M. (1979), Bridge Temperatures for Setting Bearing and Expansion Joints,
TRRL Supplementary Report 479, Department of the Environment, Department
of Transport, Crowthorne, Berkshire, England
Espinoza O. (2007), Measurements of Deformations and Stresses Due to Plate Out-Of-
Flatness in a Steel Twin Box Girder Bridge System, Masters Thesis, The
University of Texas at Austin
374
Eurocode 1 (2003), Actions on Structures, Part 1-5: General actionsThermal actions,
ENV 1991-1-5:2003, Comite European de Normalization (CEN), Brussels,
Belgium.
Fisher R.A. (1970), Statistical Methods for Research Workers, 14th, Hafner, New York
Fu H.C., Ng S.F., Cheung M.S. (1990), Thermal Behavior of Composite Bridges, Journal
of Structural Engineering, ASCE, Vol. 116, No. 12, pp. 3302-3323
Galambos J., Lechner J., and Simiu E. (1994), Extreme Value Theory and Applications,
Springer, New York
Greenberg M.D. (1998), Advanced Engineering Mathematics, 2nd Edition, Prentice Hall,
Englewood Cliffs, New Jersey
Grisham G.P (2005), Field Measurements of Bearing Displacements in Steel Girder
Bridges, Masters Thesis, The University of Houston
Helwig T.A. and Fan Z.F. (2000), Field and Computational Studies of Steel Traperzoidal
Box Girder Bridges, TxDOT Research Report 1395-3, The University of Houston
Helwig T.A., Herman R.S. and Li D.W. (2004), Behavior of Traperzoidal Box Girders
with Skewed Supports, TxDOT Research Report 0-4148-1, The University of
Houston
Hirst M.J.S. (1981), Solar heating of bridges, Australian road research Board, Vol. 11 (2):
pp. 28-36.
Hilti PD32 (2007), www.lasersquare.com/hilti/pd32.asp
Ho D., Liu C.H. (1989), Extreme Thermal Loadings in Highway Bridges, Journal of
Structural Engineering, ASCE, Vol. 115, No. 7, pp. 1681-1696
Hsieh J.S. (1985), Solar Energy Engineering, Prentice Hall, Englewood Cliffs, New
Jersey
Hulsey J.L., and Powell D.T. (1993), Rational weather model for highway structures,
Transportation research record 1393, National Research Council, Washington
D.C., pp. 54-64.
Hunt B., Cooke N. (1975), Thermal Calculations for Bridge Design, Journal of the
Structural Division, ASCE, Vol. 101, No. ST9, pp. 1763-1781
Ibrahim A.M.M. (1995), Three-Dimensional Thermal Analysis of Curved Concrete Box
Girder Bridges, Masters Thesis, Concordia University
Imbsen R.A., Vandershaf E.E., Schamber R.A., and Nutt R.V. (1985), Thermal Effects in
Concrete Bridge Superstructures, Transportation Research Board, National
Cooperative Highway Research Program Report 276, Washington D.C.
Leadbetter M.R., Lindgren G., and Rootzen H. (1983), Extremes and Related Properties
of Sequences and Processes, Springer, New York
375
Leonardt F., Kolbe G. and Peter J. (1965), Termperaturunterschiede Gefahrden
Spannbetonbrucke, Beton-und Stahlbetonbau, Vol. 60, No. 7, pp. 157-163
Lopez M.G. (1999), Thermally-induced deformation and stresses in a steel trapezoidal
twin box girder bridge, Masters Thesis, Department of Civil Engineering, The
University of Texas at Austin, TX, USA
Lucas J.M., Berred A., and Louis C. (2003), Thermal Actions on a Steel Box Girder
Bridge, Structures and Buildings, Vol. 156, Issue 2, pp. 175-182
Kehlbeck F. (1975), Einfluss der Sonnenstrahlung bei Bruckenbauwrken, Technishe
Unversitat Hannover, Werner-Verlag, Dusseldorf
Kennedy J.B., and Soliman M.H. (1987), Temperature distribution in composite bridges.
Journal of Structural Engineering, ASCE, Vol. 113 (3): pp. 475-482.
Kim H.J. (2007), Thermal Effects on Modular Maglev Steel Guideways, Dissertation,
The University of Texas at Austin
Kipp & Zonen, www.kippzonen.com
Kreith F. and Kreider J.F (1978), Principles of Solar Engineering, McGraw-Hill, New
York
Maes M.A., Dilger W.H. and Ballyk P.D. (1992), Extreme Values of Thermal Loading
Parameters in Concrete Bridges, Canadian Journal of Civil Engineering, Vol. 19,
pp. 935-946
Massicotte B., Picard A., Gaumond Y., and Ouellet C. (1994), Strengthening of a Long
Span Prestressed Segmental Box Girder Birdge, PCI Journal, Vol. 39, No. 3, pp.
52-65
Moorty S. (1990), Thermal Movements in Bridges, Ph.D Dissertation, Department of
Civil Engineering, The University of Washington, USA
Moorty S., Roeder C.W. (1992), Temperature-Dependent Bridge Movement, Journal of
Structural Engineering, ASCE, Vol. 118, No. 4, pp. 1090-1105
Muzumdar P. (2003), Field Measurements of Girder and Top Lateral Truss Stresses in
Box Girder Bridge with Skewed Support, Masters Thesis, The University of
Houston
Narouka M., Hirai I., and Yamaguti T.(1957), Measurement of the Temperature of the
Interior of the Reinforced Concrete Slab of the Shigita Bridge and Presumption of
Thermal Stress, Proceedings of the Symposium on the Stress Measurements for
Bridge and Structures, Japan Society for the Promotion of Science, Tokyo, Japan,
pp. 106-115
National Renewable Energy Laboratory, www.nrel.gov
National Climatic Data Center, www.ncdc.noaa.gov

376
Noda N., Hetnarski R.B., and Tanigawa Y. (2000), Thermal Stresses, 1st Edition, Taylor
& Francis, New York
Noda N., Hetnarski R.B., and Tanigawa Y. (2003), Thermal Stresses, 2nd Edition, Taylor
& Francis, New York
Ni Y.Q., Hua X.G., Wong K.Y., Ko J.M. (2007), Assessment of Bridge Expansion Joints
Using Long-term Displacement and Temperature Measurement, Journal of
Performance of Constructed Facilities, ASCE, Vol. 21, No. 2, pp. 143-151
Omega Engineering Technical Reference, www.omega.com
Orgill J.F., and Hollands K.G.T. (1977), Correlation Equation for Hourly Diffuse
Radiation on a Horizontal Surface, Solar Energy, Vol. 19, pp. 357-359
Priestley M.J.N. (1972), Temperature Gradients in Bridges Some Design
Consideration, New Zealand Engineering, Vol. 27, Part 7, pp. 228-233
Priestley M.J.N. (1976), Design Thermal Gradient for Concrete Bridges, New Zealand
Engineering, Vol. 31, Part 9, pp. 213-219
Priestley M.J.N. and Nigel M.J. (1978), Design of Concrete Bridge for Temperature
Gradients, ACI, Vol. 75, No. 5, pp. 209-217
Rahman F., and George K.P. (1979), Thermal Stress Analysis of Continuous Skew
Bridges, Journal of the Structural Division, ASCE, Vol. 105, No. ST7, pp. 1525-
1541
Resnick S. (1987), Extreme Values, Regular Variation, and Point Process, Springer-
Verlag, New York
Reynolds J. and Emanuel J.H. (1974), Thermal Stresses and Movements in Bridges,
Journal of Structural Division, ASCE, Vol. 100, No. ST1, pp. 63-79
Roeder C. (2003), Proposed Design Method for Thermal Bridge Movements, Journal of
Bridge Engineering, Vol. 8, No. 1, pp. 12-19
Sandepudi K.S (1991), Thermal Responses in Florida Bridges, Masters Thesis, Florida
Atlantic University
Silveira A.P., Branco F.A. and Castanheta M. (2000), Statistical Analysis of Thermal
Actions for Concrete Bridge Design. Structural Engineering International, Vol.
10: pp. 33-38.
Soliman M., and Kennedy J.B. (1986), Simplified method for estimate thermal stresses in
composite bridges. Transportation Research Record 1072, National Research
Council, Washington D.C., pp. 23-31.
Soukhov D. (1994), Two Methods for Determination of Linear Temperature Differences
in Concrete Bridge with the Help of Statistical Analysis, Darmstadt Concrete,
Vol. 9: pp. 193-210

377
Soukhov D. (2000), Representative Values of Thermal Actions for Concrete Bridges,
Progress in Structural Engineering and Materials, Vol. 2: pp. 495-501
Sutton J. (2007), Evaluating the Redundancy of Steel Bridges: Effect of a Bridge Haunch
on the Strength and Behavior of Shear Studs under Tensile Loading, Masters
Thesis, The University of Texas at Austin
Tindal T.T. and Yoo, C.H. (2003), Thermal Effects on Skewed Steel Highway Bridges
and Bearing Orientation, Journal of Bridge Engineering, ASCE, Vol. 8, No. 8, pp.
57-65
Tong M., Tham L.G., Au F.T.K., and Lee P.K.K. (2001), Numerical modelling for
temperature distribution in steel bridges. Computers and Structures, Vol. 79 (6):
pp. 583-593.
Tong M., Tham L.G., and Au F.T.K. (2002), Extreme Thermal Loading on Steel Bridges
in Tropical Region. Journal of Bridge Engineering, Vol. 7 (6): pp. 583-593.
VSL International Ltd., www.vsl.com
Writer E.T. (2007), Field Study of Thermal Effects in Steel Plate and Box Girder
Bridges, Masters Thesis, The University of Houston
Yahoo!Map (2005), map.yahoo.com
Yik F.W.H., Chung T.M. and Chan K.T. (1995), A Method to Estimate Direct and
Diffuse Radiation in Hong Kong and its Accuracy, HKIE Transactions, Vol. 2
(1): pp. 9-23
Yura J., Kuma A., Yakut A., Topkaya C., Becker E., and Collingwood J. (2001),
Elasteric Bridge Bearings: Recommended Test Methods, NCHRP Report 449,
Transportation Research Board
Zichner T. (1981), Thermal Effects on Concrete Bridges, C.E.B., Enlarged Meeting
Commission 2, Pavia, pp. 292-313
Zuk, W. (1961), Thermal and Shrinkage Stresses in Composite Beams, Journal of the
ACI, Vol. 58, No. 3, pp. 327-340
Zuk, W. (1965), Thermal Behavior of Composite Bridges Insulated and Uninsulated,
Highway Research Record 76, National Research Council, pp. 231-253

378
Vita

Quan Chen was born in Hubei, China on July 28, 1977, the son of Wanxiang
Chen and Zhengying Chen. After completing his work at No. 1 High School of
Zhongxiang, Hubei in 1995, he received the degree of Bachelor in Science in 1999 and
then the degree of Master in Science in the department of Civil Engineering at Tsinghua
University, China, in 2002. He was admitted the graduate school at the University of
Houston in 2004. In 2005, he entered the graduate school at the University of Texas at
Austin for his Ph.D study.

Permanent address:
Book Distribution Section, Zhongxiang Education Committee
Zhongxinag, Hubei, 431900
P.R. China

This dissertation was typed by the author.

379

You might also like