You are on page 1of 426

DESIGN OF HIGHWAY OVERHEAD CANTILEVER-TYPE SIGN SUPPORT

STRUCTURES FOR FATIGUE LOADS

by

IAN E. HOSCH

FOUAD H. FOUAD, COMMITTEE CHAIR

TALAT SALAMA

VIRGINIA SISIOPIKU

HOUSSAM TOUTANJI

NASIM UDDIN

A DISSERTATION

Submitted to the graduate faculty of The University of Alabama at Birmingham,


in partial fulfillment of the requirements for the degree of
Doctor of Philosophy

BIRMINGHAM, ALABAMA

2009
Copyright by
Ian E. Hosch
2009

ii
DESIGN OF HIGHWAY OVERHEAD CANTILEVER-TYPE SIGN SUPPORT
STRUCTURES FOR FATIGUE LOADS

IAN E. HOSCH

DEPARTMENT OF CIVIL, CONSTRUCTION, AND ENVIRONMENTAL


ENGINEERING

ABSTRACT

The 2001 edition of the American Association of State Highway and Transporta-

tion Officials (AASHTO) Standard Specifications for Structural Supports for Highway

Signs, Luminaires and Traffic Signals has been revised in its entirety through a major re-

search project conducted under the auspices of the National Cooperative Highway Re-

search Program (NCHRP 17-10). A major part of the revision included updated provi-

sions and criteria for extreme wind loads and new provisions and criteria on fatigue de-

sign. These provisions differ considerably from those in previous editions of the specifi-

cations, and have remained relatively unchanged in the 2009 edition.

The impact of the fatigue criteria on the design of overhead sign structures has not

been fully evaluated. The fatigue design loads do not adequately reflect the stresses gen-

erated on these structures from wind-induced fatigue loading. In addition, the provisions

do not account for the variety of support structures in design, each with different configu-

ration, sizes, shapes, and material properties that influence vibration behavior. As a re-

sult, the vulnerability of sign support structures to wind-induced fatigue loading is not

fully realized.

The main goal of the project was to conduct theoretical and experimental pro-

grams of study to evaluate the performance of cantilever-type highway overhead sign

support structures subjected to wind induced fatigue loads. A theoretical program was

iii
developed that took into account the variety of sign supports structures in design, as well

as addressed the vulnerability of these structures to wind-induced fatigue loading. The

experimental program was developed to evaluate the accuracy of the theoretical study. A

finite element analysis program was conducted to simulate the wind-induced loading en-

vironment and the response of sign support structures to this environment. Alterations

were made to the model that could not be done experimentally due to costs and time re-

straints. The results of the finite element analysis were compared to the theoretical and

experimental programs of study.

The developed information and criteria on fatigue design of sign support struc-

tures were used to develop fatigue design loads to provide an improved and more reliable

design method. Recommendations are made to update the current specifications to in-

clude more reliable fatigue loads.

Keywords: Cantilever-type sign support structures, fatigue design, fatigue load, natural

wind, truck-induced wind, vibration

iv
DEDICATION

The dissertation is dedicated to friends and family, most importantly to my

mother Sandra J. Haigh who gave me strong will and unconditional support during my

time as a student.

v
ACKNOWLEDGEMENTS

It was important to acknowledge the following people who helped with the re-

search preformed in this project:

 Dr. Fouad H. Fouad who was my mentor and advisor,

 Richard Hawkins and Dr. Jason Kirby who helped with the important technical

experimentation aspects and advice with presentations,

 Daniel Jones and Ricky Love of the Alabama Department of Transportation who

helped with the instrumentation and field testing,

 The Morrison Family, Brett and Dave, who helped specifically for making the ex-

tensions needed for the anemometers during field testing, and

 Andrew Sullivan for setting up and reviewing the traffic recorders required with

the truck testing.

vi
TABLE OF CONTENTS

Page

ABSTRACT....................................................................................................................... iii

DEDICATION .....................................................................................................................v

ACKNOWLEDGMENTS ................................................................................................. vi

LIST OF TABLES ........................................................................................................... xiii

LIST OF FIGURES ........................................................................................................ xvii

LIST OF ABBREVIATIONS ........................................................................................ xxvi

CHAPTER

1 INTRODUCTION .........................................................................................................1

Problem Statement ...................................................................................................1


Objective ..................................................................................................................4
Study Initiatives .......................................................................................................5
Work Plan ................................................................................................................6
Summary ............................................................................................................6
Project Tasks ......................................................................................................7
Significance and Benefits ......................................................................................11
Dissertation Outline ...............................................................................................12

2 LITERATURE REVIEW ............................................................................................16

Chapter Overview ..................................................................................................16


Research Paper Breakdown ...................................................................................16
Research Review ....................................................................................................19
DeSantis and Haig (1996) ...............................................................................19
Davenport (1961) .............................................................................................20
Kaczinski et al (1998) ......................................................................................22
Fouad et al (1997-2005) ..................................................................................23
Dexter and Ricker (2002).................................................................................24
Cook et al (1996) .............................................................................................25

vii
Page

Creamer et al (1979)........................................................................................26

3 FATIGUE PROVISIONS OF THE AASHTO SUPPORT SPECIFICATIONS .........27

Chapter Overview ..................................................................................................27


Fatigue Load due to Natural Wind.........................................................................27
Infinite-Life Approach......................................................................................27
Predicting the Environment .............................................................................29
Structural Excitation ........................................................................................35
Design Fatigue Equation for Natural Wind.....................................................36
Fatigue Load due to Truck-Induced Gust ..............................................................37

4 SIGN STRUCTURE INSTRUMENTATION .............................................................39

Chapter Overview ..................................................................................................39


Cantilever-Type Highway Overhead Sign Support Structure ...............................39
Geometric Properties .......................................................................................40
Material Properties ..........................................................................................42
Instrumentation Process .........................................................................................43
Strain Gauges .........................................................................................................47
Truss Chord Members......................................................................................48
Post Support .....................................................................................................50
Anchor Bolts.....................................................................................................53
Anemometers .........................................................................................................57
Accelerometers ......................................................................................................60
Data Acquisition System........................................................................................63

5 STRUCTURAL TESTING FOR NATURAL WIND


AND TRUCK-INDUCED WIND GUST....................................................................68

Chapter Overview ..................................................................................................68


Natural Wind Gust .................................................................................................68
Pre-Determined Sample Size ...........................................................................68
Testing Day Ranking Schedule ........................................................................69
Test Procedure .................................................................................................71
Truck-Induced Wind Gust .....................................................................................72
Test Procedure .................................................................................................74

6 FATIGUE RESISTANCE ...........................................................................................77

Chapter Overview ..................................................................................................77


Constant-Amplitude Fatigue Thresholds ...............................................................78
Failure Index ..........................................................................................................78
Evaluation of the Fatigue Stresses in the Anchor Bolts.........................................79

viii
Page

General Description of the Anchor Bolts.........................................................80


Fatigue Stress due to Natural Wind Gust ........................................................83
Fatigue Stress due to Truck-Induced Wind Gust .............................................89
Evaluation of the Fatigue Stress in the Post-to-Base-Plate Connection ................93
General Description of the Post-to-Base-Plate Connection ............................93
Fatigue Stress due to Natural Wind Gust ........................................................95
Fatigue Stress due to Truck-Induced Wind Gust ...........................................103
Evaluation of the Fatigue Stress in the Chords ....................................................108
General Description of the Chords ................................................................108
Fatigue Stress due to Natural Wind Gust ......................................................112
Fatigue Stress due to Truck-Induced Wind Gust ...........................................119
Discussion of Results ...........................................................................................123

7 EXPERIMENTAL MODAL ANALYSIS ................................................................128

Chapter Overview ................................................................................................128


Modal Data Utilization ........................................................................................128
Modal Analysis Test Setup ..................................................................................129
Systemizing the Degrees of Freedom ..................................................................134
Spectral Analysis .................................................................................................135
Modal Shapes .......................................................................................................140
Frequency Response Function .......................................................................141
Modal Damping ...................................................................................................149

8 EXPERIMENTAL CALCULATION OF THE FATIGUE LOAD


DUE TO NATURAL WIND GUST..........................................................................156

Chapter Overview ................................................................................................156


Fatigue Load Calculation Approach ....................................................................156
Structural Excitation ............................................................................................158
Collected Sample Size ....................................................................................158
Usable Data Collected ...................................................................................159
Reduction of Structural Excitation Experimental Data .................................161
Structural Response .............................................................................................165
Data Offsetting ...............................................................................................166
Strain Ranges .................................................................................................172
Wind Pressure Back-Calculation .........................................................................174
Theoretical Structural Analysis .....................................................................175
Development of the Stress Element ................................................................184
Wind Pressure Calculation ............................................................................189
Wind Velocity vs. Wind Pressure ........................................................................191
Infinite-Life Approach .........................................................................................195

ix
Page

9 THEORETICAL CALCULATION OF THE FATIGUE LOAD


DUE TO NATURAL WIND GUST..........................................................................197

Chapter Overview ................................................................................................197


Significance of the Theoretical Program .............................................................197
Methodology ........................................................................................................200
Structural Excitation ............................................................................................200
Supports Specifications ..................................................................................201
Experimentally Collected Data ......................................................................207
Comparison between the Supports Specifications
and the Experimental Excitation ....................................................................213
Structural Response .............................................................................................215
Response Power Spectral Density and the Root-Mean-Square .....................215
Vibration Response Spectrum..............................................................................220
Natural Frequency .........................................................................................222
Critical Damping Percentage ........................................................................224
Peak-to-Peak Stress Range ............................................................................226
Infinite-Life Approach .........................................................................................228
Vibration Response Spectrum for the Supports Specifications ......................234

10 EXPERIMENTAL CALCULATION OF THE FATIGUE LOAD


DUE TO TRUCK-INDUCED WIND GUST ............................................................235

Chapter Overview ................................................................................................235


Fatigue Load Calculation Approach ....................................................................235
Collected Sample Size .........................................................................................236
Truck Specimen ...................................................................................................237
Structural Excitation ............................................................................................238
Structural Response .............................................................................................240
Stress Range ...................................................................................................241
Truck-Induced Wind Pressure Back-Calculation ................................................244
Theoretical Structural Analysis .....................................................................245
Development of the Stress Element ................................................................252
Wind Pressure Calculation ............................................................................255

11 THEORETICAL CALCULATION OF THE FATIGUE LOAD


DUE TO TRUCK INDUCED WIND GUST ............................................................258

Chapter Overview ................................................................................................258


Research Significance ..........................................................................................258
Methodology ........................................................................................................261
Structural Excitation ............................................................................................262
Structural Response .............................................................................................265
Shock Response Spectrum ...................................................................................267

x
Page

12 FINITE ELEMENT ANALYSIS ..............................................................................274

Chapter Overview ................................................................................................274


Model Development.............................................................................................275
Geometry ........................................................................................................275
Element Type ..................................................................................................279
Material Definition ........................................................................................279
Finite Element Modal Analysis ...........................................................................280
Comparisons of FEA to Experimental Modal Analysis .................................282
Finite Element Structural Analysis ......................................................................283
Natural Wind Gust .........................................................................................284
Truck-Induced Wind Gust ..............................................................................289
Finite Element Dynamic Fatigue Loading Analysis ............................................293
Natural Wind Gust .........................................................................................294
Truck-Induced Wind Gust ..............................................................................304

13 PROPOSED FATIGUE PROVISIONS ....................................................................311

Chapter Overview ................................................................................................311


Fatigue Load due to Natural Wind Gust ..............................................................311
Discussion of Results .....................................................................................311
General Fatigue Design Equation for Natural Wind Gust ............................315
Detailed Fatigue Design Equation for Natural Wind Gust ...........................316
Fatigue Load due to Truck-Induced Wind Gust ..................................................320
Discussion of Results .....................................................................................321
General Fatigue Design Equation for Truck-Induced Wind Gust .................322
Detailed Fatigue Design Equation for Truck-Induced Wind Gust ................324
Design Examples .................................................................................................327
Supports Specifications Fatigue Design Equations .......................................328
Natural Wind Gust .........................................................................................329
Truck-Induced Wind Gust ..............................................................................335

14 SUMMARY AND CONCLUSIONS ........................................................................338

Chapter Overview ................................................................................................338


Fatigue Resistance ...............................................................................................340
Modal Analysis ....................................................................................................343
Fatigue Load due to Natural Wind Gust ..............................................................344
General Fatigue Design Equation for Natural Wind Gust ............................345
Detailed Fatigue Design Equation for Natural Wind Gust ...........................346
Fatigue Load due to Truck-Induced Wind Gust ..................................................351
General Fatigue Design Equation for Truck-Induced Wind Gust .................351
Detailed Fatigue Design Equation for Truck-Induced Wind Gust ................352
Finite Element Analysis Method .........................................................................356

xi
Page

Contributions to the Practice................................................................................357


Recommendations for Future Research ...............................................................358

LIST OF REFERENCES .................................................................................................359

APPENDIX

A INSTRUMENTATION IDENTIFICATION AND LAYOUTS .........................365

B ANEMOMETER MOUNTING INSTRUCTIONS.............................................374

C ACCELEROMETER MOUNTING INSTRUCTIONS ......................................376

D RANKING SCHEDULE FOR NATURAL WIND TESTING ...........................383

E TRUCK-INDUCED WIND GUST TESTING PROCEDURE ...........................396

xii
LIST OF TABLES

Table Page

2.1 Terrain Coefficients (13) -------------------------------------------------------------------- 21

3.1 Terrain Coefficients (13) -------------------------------------------------------------------- 30

4.1 Material Properties -------------------------------------------------------------------------- 43

6.1 Supports Specifications Constant-Amplitude Fatigue Thresholds (1) --------------- 79

6.2 Anchor Bolt Clearance Lengths ----------------------------------------------------------- 83

6.3 Estimated Microstrain Range at Fatigue Wind for Anchor Bolts --------------------- 87

6.4 Anchor Bolt Stress Range and Failure Index for Natural Wind ---------------------- 88

6.5 Anchor Bolt Microstrain Ranges from the Truck Tests -------------------------------- 91

6.6 Anchor Bolt Stress Range and Failure Index for Truck Tests------------------------- 92

6.7 Estimated Microstrain Range at Fatigue Wind for Section AA --------------------- 100

6.8 Section AA Stress Range and Failure Index for Natural Wind --------------------- 102

6.9 Section AA (Normal) Microstrain Ranges for the Truck Tests --------------------- 104

6.10 Section AA (Shear) Microstrain Ranges from the Truck Tests --------------------- 105

6.11 Section AA Stress Range and Failure Index for Truck Tests ----------------------- 107

6.12 Estimated Microstrain Range at Fatigue Wind for Chords ------------------------- 115

6.13 Chord Stress Range and Category E Failure Index for Natural Wind ------------- 117

6.14 Chord Stress Range and Category B Failure Index for Natural Wind ------------ 118

6.15 Chord 1 Microstrain Ranges from the Truck Tests ----------------------------------- 119

xiii
Table Page

6.16 Chord 2 Microstrain Ranges from the Truck Tests ----------------------------------- 120

6.17 Chord 3 Microstrain Ranges from the Truck Tests ----------------------------------- 120

6.18 Chord 4 Microstrain Ranges from the Truck Tests ----------------------------------- 121

6.19 Chord Stress Range and Category E Failure Index for Truck Tests---------------- 122

6.20 Chord Stress Range and Category B Failure Index for Truck Tests --------------- 123

6.21 Experimental Stress Ranges for Natural Wind Tests --------------------------------- 124

6.22 Experimental Stress Ranges for Truck-Induced Wind Tests ------------------------ 124

7.1 Experimental Natural Frequencies ------------------------------------------------------ 139

7.2 Modal Damping Results ------------------------------------------------------------------ 155

8.1 Collection Dates and Times for Natural Wind ---------------------------------------- 158

8.2 Development of the Wind Directionality Unit Vector ------------------------------- 165

8.3 Exposure Area Breakdown --------------------------------------------------------------- 178

8.4 Drag Coefficients -------------------------------------------------------------------------- 179

8.5 Height Coefficients ------------------------------------------------------------------------ 180

8.6 Effective Area ------------------------------------------------------------------------------ 182

8.7 Material Properties at Section AA and Section BB ----------------------------------- 185

8.8 Relevant Stress and Strain Equations --------------------------------------------------- 187

8.9 Example of Wind Pressure Magnitude Calculation at Location A ----------------- 189

9.1 Terrain Coefficients (13) ----------------------------------------------------------------- 201

9.2 Natural Frequency Comparison --------------------------------------------------------- 224

10.1 Order of Truck Runs and Truck Speed ------------------------------------------------- 236

10.2 Peak-to-Peak Strain Ranges ------------------------------------------------------------- 244

xiv
Table Page

10.3 Exposure Truss Area Breakdown ------------------------------------------------------- 248

10.4 Drag Coefficients ------------------------------------------------------------------------- 249

10.5 Effective Area ----------------------------------------------------------------------------- 250

10.6 Material Properties at Section AA and Section BB ---------------------------------- 253

10.7 Relevant Stress and Strain Equation --------------------------------------------------- 254

10.8 Truck-Induced Wind Gust Pressure ---------------------------------------------------- 256

11.1 Natural Frequency Comparison --------------------------------------------------------- 272

12.1 Material Properties ------------------------------------------------------------------------ 280

12.2 Natural Frequencies of the FEA Model------------------------------------------------ 281

12.3 FEA and Experimental Modal Analysis Comparison ------------------------------- 283

12.4 Normal Strain from FEA at Section AA ----------------------------------------------- 288

12.5 Comparison of FEA with Experimental Strain Range ------------------------------- 288

12.6 Normal Strain from FEA at Section AA ----------------------------------------------- 292

12.7 Comparison of FEA with Experimental Strain Range ------------------------------- 293

12.8 Modal Analysis of the Aluminum Structure ------------------------------------------ 299

12.9 FEA Comparison for Natural Wind Gust --------------------------------------------- 301

12.10 FEA Comparison for Truck-Induced Wind Gust------------------------------------- 310

13.1 Experimental and Theoretical Results for Natural Wind Gust --------------------- 312

13.2 Comparison of Results from Experimental and Theoretical Programs ----------- 321

13.3 Design Case Description for Natural Wind Gust ------------------------------------- 330

13.4 Design Case Results for Natural Wind Gust ------------------------------------------ 331

13.5 Design Case Description for Truck-Induced Wind Gust ---------------------------- 335

xv
Table Page

13.6 Design Case Results for Truck-Induced Wind Gust --------------------------------- 337

14.1 Experimental Stress Ranges for Natural Wind Tests -------------------------------- 341

14.2 Experimental Stress Ranges for Truck-Induced Wind Tests ----------------------- 342

14.3 Results of Back-Calculation Verification --------------------------------------------- 357

xvi
LIST OF FIGURES

Figure Page

1.1 I-65/I-565 interchange, Exit 340B. --------------------------------------------------------- 3

1.2 Failure of cantilever sign support. ---------------------------------------------------------- 3

1.3 Close-up of fractured anchor bolt.---------------------------------------------------------- 4

1.4 Cantilever-type highway overhead sign structure. --------------------------------------- 7

2.1 Research paper breakdown. ---------------------------------------------------------------- 17

2.2 Referenced papers in the research paper breakdown spreadsheet. ------------------- 18

3.1 Wind velocity PSD for annual mean wind velocity 11 mph. -------------------------- 31

3.2 Force PSD for annual mean wind velocity 11 mph. ------------------------------------ 33

3.3 Force PSD using limit-state wind velocity 37 mph. ------------------------------------ 35

4.1 Cantilever-type highway overhead sign support structure. ---------------------------- 40

4.2 Elevation view of the structure configuration. ------------------------------------------ 41

4.3 Plan view of structure configuration. ----------------------------------------------------- 41

4.4 Side view of structure configuration. ----------------------------------------------------- 42

4.5 Strain gauged post support in laboratory. ------------------------------------------------ 44

4.6 Strain gauged truss overhang in laboratory. --------------------------------------------- 45

4.7 Set anchor bolts in structure foundation. ------------------------------------------------- 45

4.8 Truss overhand attachment. ---------------------------------------------------------------- 46

4.9 Sign attachment. ----------------------------------------------------------------------------- 46

xvii
Figure Page

4.10 Finished instrumented structure. ---------------------------------------------------------- 47

4.11 Truss-to-pole support strain gauges. ----------------------------------------------------- 48

4.12 Chord strain gauges. ------------------------------------------------------------------------ 49

4.13 Chord strain gauges on the assembled structure. --------------------------------------- 50

4.14 Pole support strain gauges. ---------------------------------------------------------------- 51

4.15 Rosette and uni-axial strain gauges on shaft. ------------------------------------------- 52

4.16 Instrumented post support.----------------------------------------------------------------- 52

4.17 Anchor bolts before installation. ---------------------------------------------------------- 54

4.18 Anchor bolts and foundation. ------------------------------------------------------------- 54

4.19 Anchor bolt strain gauges. ----------------------------------------------------------------- 55

4.20 Anchor bolt strain gauging preparation. ------------------------------------------------- 55

4.21 Strain gauged anchor bolt. ----------------------------------------------------------------- 56

4.22 Instrumented anchor bolts on assembled structure. ------------------------------------ 56

4.23 WindSonic ultra-sonic wind and direction sensor. ------------------------------------- 57

4.24 Overview of anemometer layout. --------------------------------------------------------- 58

4.25 AN-1 and AN-2 above traffic lane.------------------------------------------------------- 58

4.26 Anemometers AN-3 and AN-4 above post. --------------------------------------------- 59

4.27 Accelerometer locations for cantilever structure. -------------------------------------- 61

4.28 Accelerometer location AC-2. ------------------------------------------------------------ 64

4.29 Accelerometer location AC-3. ------------------------------------------------------------ 64

4.30 AC-2 and AC-3 view. ----------------------------------------------------------------------- 65

4.31 Accelerometer mounting block. ----------------------------------------------------------- 65

xviii
Figure Page

4.32 Mounting block at AC-1. ------------------------------------------------------------------- 66

4.33 Typical test setup with van data acquisition system. ----------------------------------- 66

4.34 Van data acquisition system and computer setup. -------------------------------------- 67

5.1 Truck and driver for truck-induced wind test. ------------------------------------------- 73

5.2 JAMAR Trax Flex HS recorders. --------------------------------------------------------- 75

5.3 Typical truck-induced wind gust runs. --------------------------------------------------- 76

6.1 Anchor bolt connection. -------------------------------------------------------------------- 80

6.2 Anchor bolt layout. -------------------------------------------------------------------------- 81

6.3 Anchor bolt identification. ----------------------------------------------------------------- 82

6.4 All anchor bolt microstrain range vs. wind velocity. ----------------------------------- 84

6.5 Transformed independent variable for AB-8. ------------------------------------------- 86

6.6 Estimated strain range at fatigue wind for AB-8. --------------------------------------- 86

6.7 Estimated Microstrain range at fatigue wind for anchor bolts.------------------------ 87

6.8 Failure index of anchor bolts for natural wind. ------------------------------------------ 89

6.9 Depth chart of anchor bolt strain range in truck tests. ---------------------------------- 90

6.10 Failure index of anchor bolts for truck tests. -------------------------------------------- 92

6.11 Fillet-welded tube-to-transverse plate connection. ------------------------------------- 93

6.12 Post support strain gauges. ---------------------------------------------------------------- 94

6.13 Strain gauge locations at Section AA. --------------------------------------------------- 94

6.14 45° rosette arrangement with coordinate axis. ------------------------------------------ 96

6.15 Normal microstrain range vs. wind velocity. ------------------------------------------- 97

6.16 Rosette microstrain range vs. wind velocity. ------------------------------------------- 98

xix
Figure Page

6.17 Transformed independent variable for SGR-AA-11. ---------------------------------- 99

6.18 Estimated strain range at fatigue wind for SGR-AA-11----------------------------- 100

6.19 Estimated microstrain range at fatigue wind for Section AA. ---------------------- 101

6.20 Failure indexes for natural wind of Section AA. ------------------------------------- 103

6.21 Depth chart of Section AA normal strain in truck tests. ----------------------------- 105

6.22 Depth chart of Section AA shear strain in truck tests. ------------------------------- 106

6.23 Failure index for Section AA truck tests. ---------------------------------------------- 107

6.24 Chord-to-plate weld. ---------------------------------------------------------------------- 108

6.25 Chord-to-column bolted connection. --------------------------------------------------- 109

6.26 Chord instrumentation layout. ---------------------------------------------------------- 110

6.27 Chord labeling. ---------------------------------------------------------------------------- 110

6.28 Chord strain gauge labeling.------------------------------------------------------------- 111

6.29 Chord microstrain ranges vs. wind velocity. ------------------------------------------ 113

6.30 Transformed independent variable for SG-C2-5. ------------------------------------ 114

6.31 Estimated strain range at fatigue wind for SG-C2-5. -------------------------------- 115

6.32 Estimated microstrain range for fatigue wind of chords.---------------------------- 116

6.33 Failure indexes of chord gauges for natural wind. ----------------------------------- 118

6.34 Depth chart of chord strain in truck tests. --------------------------------------------- 121

6.35 Comparison of results between natural wind and truck gusts. --------------------- 125

7.1 Accelerometer locations for cantilever structure. ------------------------------------ 130

7.2 Accelerometer location AC-2. ---------------------------------------------------------- 131

7.3 Accelerometer location AC-3. ---------------------------------------------------------- 131

xx
Figure Page

7.4 Overview of accelerometers AC-2 and AC-3. ---------------------------------------- 132

7.5 Mounting block at AC-1. ---------------------------------------------------------------- 133

7.6 Forced event used for modal analysis. ------------------------------------------------- 137

7.7 Structural response to the forced event. ----------------------------------------------- 138

7.8 Modes 1 and 3 from AC-2-Y. ----------------------------------------------------------- 139

7.9 Modes 2 and 4 from AC-2-Z. ----------------------------------------------------------- 140

7.10 Quadrature picking of mode 1. ---------------------------------------------------------- 143

7.11 Mode 1: torsion about support shaft. --------------------------------------------------- 144

7.12 Quadrature picking of mode 2. ---------------------------------------------------------- 145

7.13 Mode2: vertical rocking. ----------------------------------------------------------------- 145

7.14 Quadrature picking of mode 3. ---------------------------------------------------------- 146

7.15 Mode 3: horizontal truss twist. ---------------------------------------------------------- 147

7.16 Quadrature picking of mode 4. ---------------------------------------------------------- 148

7.17 Mode 4: outward and inward clamping. ----------------------------------------------- 148

7.18 Typical truck transient event from accelerometers. ---------------------------------- 150

7.19 Exponential decay of the transient truck event. -------------------------------------- 152

7.20 Trendline of extracted peak amplitudes. ----------------------------------------------- 153

8.1 Wind rose diagram of all data. ----------------------------------------------------------- 160

8.2 Wind rose diagram of usable data. ------------------------------------------------------ 161

8.3 Coordinate system. ------------------------------------------------------------------------ 164

8.4 Strain values per 0.5 ph wind velocity intervals. -------------------------------------- 167

8.5 Transformed regressor. ------------------------------------------------------------------- 168

xxi
Figure Page

8.6 Parabolic curve fit.------------------------------------------------------------------------- 169

8.7 Parabolic offset. ---------------------------------------------------------------------------- 170

8.8 90 mph projection using offsetting procedure. ---------------------------------------- 170

8.9 Peak-to-peak range. ----------------------------------------------------------------------- 172

8.10 45° rosette arrangement with coordinate axis. ---------------------------------------- 174

8.11 Area breakdown of the front face. ------------------------------------------------------ 176

8.12 Area breakdown of East side face. ----------------------------------------------------- 177

8.13 Area breakdown of West side face. ---------------------------------------------------- 177

8.14 Height coefficient stepped profile. ----------------------------------------------------- 181

8.15 Free body diagram. ------------------------------------------------------------------------ 184

8.16 Cross section of post at Section AA and Section BB. ------------------------------- 186

8.17 Strain gauge locations at Section AA and Section BB.------------------------------ 188

8.18 Typical stress element. ------------------------------------------------------------------- 188

8.19 Wind velocity vs. wind pressure for Section AA. ------------------------------------ 191

8.20 Wind velocity vs. wind pressure for Section BB. ------------------------------------ 192

8.21 Wind velocity vs. wind pressure for Rosettes. ---------------------------------------- 192

8.22 Wind velocity vs. wind pressure for all sections. ------------------------------------ 193

8.23 Transformed wind velocity vs. wind pressure. --------------------------------------- 194

8.24 Wind velocity vs. wind pressure trendline. ------------------------------------------- 195

8.25 Fatigue load due to natural wind. ------------------------------------------------------- 196

9.1 Highway overhead sign support structures. ------------------------------------------- 198

9.2 Wind velocity PDS for annual mean wind velocity. --------------------------------- 202

xxii
Figure Page

9.3 Force PDS for annual mean wind velocity. ------------------------------------------- 204

9.4 Force PDS using limit-state wind velocity 37 mph. --------------------------------- 206

9.5 Experimental wind velocity PDS. ------------------------------------------------------ 209

9.6 Logarithmic transformation of average wind velocity PDS. ----------------------- 210

9.7 Best fit line approximating the average wind velocity PDS. ----------------------- 211

9.8 Theoretical plot of the experimental average wind velocity PDS. ----------------- 212

9.9 Approximation of the experimental average wind pressure PDS. ----------------- 213

9.10 Comparison of experimental to Supports Specifications PDS. --------------------- 214

9.11 Structural response to wind pressure excitation. ------------------------------------- 218

9.12 Response of n SDOF systems to common excitation input. ------------------------ 220

9.13 RMS wind pressure VRS for 1.82% damping. --------------------------------------- 221

9.14 First modal shape equal to 1.61 Hz. ---------------------------------------------------- 222

9.15 RMS wind pressure VRS with damping equal to 0.5%. ---------------------------- 225

9.16 Response PDS with damping equal to 0.5%. ----------------------------------------- 225

9.17 VRS of peak-to-peak amplitude for 1082% damping.------------------------------- 228

9.18 Peak-to-peak VRS plots versus average wind velocity. ----------------------------- 229

9.19 Linear transformation of the peak-to-peak VRS. ------------------------------------- 231

9.20 Fatigue load comparison between experimental and VMS. ------------------------ 231

9.21 Fatigue wind pressure VRS for 1.82% damping. ------------------------------------ 232

9.22 Fatigue wind pressure VRS for damping range. -------------------------------------- 233

9.23 Fatigue wind VRS of Supports Specifications (2% damping). --------------------- 234

10.1 Truck and driver for truck-induced wind test. ---------------------------------------- 237

xxiii
Figure Page

10.2 Anemometer layout and orientation. --------------------------------------------------- 241

10.3 Strain gauge locations at Section AA and Section BB.------------------------------ 242

10.4 Maximum peak-to-peak strain range. -------------------------------------------------- 243

10.5 Underneath exposed horizontal area. -------------------------------------------------- 247

10.6 Exposed underneath truss area breakdown. ------------------------------------------- 248

10.7 Free body diagram. ----------------------------------------------------------------------- 252

10.8 Cross section of strain gauge location Section AA and Section BB. -------------- 254

10.9 Typical stress element. ------------------------------------------------------------------- 255

11.1 Highway overhead sign support structures. ------------------------------------------- 259

11.2 Vertical truck-induced wind gust impulses. ------------------------------------------- 264

11.3 Structural response time history due to the Control impulse. ---------------------- 267

11.4 Response of n SDOF systems to common excitation input. ------------------------ 268

11.5 SRS for vertical truck-induced wind gust. -------------------------------------------- 269

11.6 Second modal shape of cantilever structure. ------------------------------------------ 271

12.1 FEA model of cantilever-type sign support structure. ------------------------------- 276

12.2 Sign-to-truss connection. ---------------------------------------------------------------- 277

12.3 Truss-to-post connection. ---------------------------------------------------------------- 277

12.4 Foundation connection. ------------------------------------------------------------------ 278

12.5 Modal shapes of the first five modes from FEA. ------------------------------------- 281

12.6 Average wind pressure PDS for FEA input. ------------------------------------------ 296

12.7 VRS used for the FEA analysis. -------------------------------------------------------- 298

12.8 VRS for the aluminum FEA model with 0.5% damping. --------------------------- 299

xxiv
Figure Page

12.9 VRS PDS response to wind pressure excitation. ------------------------------------- 302

12.10 SAP2000 response PDS. ----------------------------------------------------------------- 303

12.11 Smaller frequencies of the SAP2000 response PDS. -------------------------------- 303

12.12 Transient excitation function for FEA input. ----------------------------------------- 306

12.13 SRS used for the FEA analysis. --------------------------------------------------------- 308

13.1 VRS for 1.5% damping ratio. ----------------------------------------------------------- 319

13.2 Fatigue wind pressure VRS for damping range. -------------------------------------- 319

13.3 SRS for vertical truck-induced wind gust. -------------------------------------------- 326

13.4 Direction of wind loading for natural wind gust. ------------------------------------- 330

13.5 VRS for Case 1 and Case 2. ------------------------------------------------------------- 331

13.6 VRS for Case 3. --------------------------------------------------------------------------- 332

13.7 VRS for Case 4. --------------------------------------------------------------------------- 332

13.8 VRS for Case 5 and Case 6. ------------------------------------------------------------- 333

13.9 Direction of wind loading for truck-induced wind gust. ---------------------------- 336

14.1 Comparison of results between natural wind and truck gusts. --------------------- 344

14.2 VRS for 1.5% damping ratio. ----------------------------------------------------------- 350

14.3 Fatigue wind pressure VRS for damping range. -------------------------------------- 350

14.5 SRS for vertical truck-induced wind gust. -------------------------------------------- 355

xxv
LIST OF ABBREVIATIONS

AASHTO American Association of State Highway Transportation Officials

AC accelerometer

ALDOT Alabama Department of Transportation

CAFL constant amplitude fatigue limit

DFT Discrete Fourier transform

DIA diameter

DOF degree-of-freedom

DOT Department of Transportation

FEA finite element analysis

FFT fast Fourier transform

FT Fourier transform

FRF frequency response function

NA not available

NCHRP National Cooperative Highway Research Program

PDS power density spectrum

RMS root-mean-square

SDOF single degree-of-freedom

SG strain gauge

SGR strain gauge rosette

xxvi
S-N stress vs. number of cycles

SRS shock response spectrum

UAB University of Alabama at Birmingham

VMS variable message sign

VRS vibration response spectrum

xxvii
1

CHAPTER 1

INTRODUCTION

Problem Statement

Presently, there is a lack of research involving the development of fatigue loads

specifically for sign support structures. Likewise, there is little knowledge and under-

standing of this subject. In addition, there is a lack of accountability for the variety of

sign support structures in design each with different configuration, sizes, shapes, and ma-

terial properties that influence vibration behavior. Sign support structures are highly

flexible with very low damping properties which make them highly susceptible to wind-

induced fatigue loading, and as such the vulnerability of these structures is not com-

pletely realized. As a result, currently published fatigue design equations do not ade-

quately reflect the stresses generated on these structures from wind-induced fatigue load-

ing.

The 2009 AASHTO Standard Specifications for Structural Supports for Highway

Signs, Luminaires and Traffic Signals (1-6) (hereafter referred to as Supports Specifica-

tions) include fatigue design criteria for cantilevered overhead sign structures that could

significantly impact the design of these structures. Fatigue loadings due to four wind

phenomena are generally prescribed in the specifications:

 Natural wind gust,


2

 Truck-induced wind gust,

 Galloping, and

 Vortex shedding.

Fatigue loads due to natural wind and truck-induced wind gusts are applicable to

overhead sign structures and are the focus of this study. Galloping may occur in overhead

sign structures, but 4-chord trusses (which are analyzed in this study) are not susceptible

to galloping. Vortex shedding is not applicable to the overhead sign structures of this

study.

In April 2006, a cantilevered overhead sign structure located at the I-565 and I-65

Interchange, Exit 340B, (Figure 1.1) failed due to fatigue of the anchor bolts. Fatigue

fracture of the anchor bolts due to combined axial stresses and bending in the bolts was

noted. The photographs in Figure 1.2 show the cantilever overhead sign that failed. A

close-up of the bolt fracture is shown in Figure 1.3. The bolt layout was designed using

an earlier version of the Supports Specifications that did not consider fatigue design.

The main goal of this project was to evaluate the performance of cantilever-type

highway overhead sign structures subjected to wind-induced fatigue loads resulting from

natural wind and truck-induced wind gusts, and to develop more reliable fatigue loads for

design. The project involved detailed theoretical as well as experimental programs to ad-

dress this issue.


3

Nashville
I-565
I-65

Huntsville
Decatur

Exit
340B
Birmingham

FIGURE 1.1 I-65/I-565 interchange, Exit 340B.

FIGURE 1.2 Failure of cantilever sign support.


4

FIGURE 1.3 Close-up of fractured anchor bolt.

Objective

The main objectives of this work can be enumerated in the following three initia-

tives:

1. Perform theoretical and experimental studies to evaluate the performance of canti-

lever-type highway overhead sign support structures subjected to wind induced

fatigue loads.

2. Use this information to develop fatigue design loads to provide an improved and

more reliable design method.

3. Compare the developed criteria with the Supports Specifications (1), and recom-

mend new equations for the specifications.


5

Study Initiatives

Study initiatives of this study to accomplish the main objectives are enumerated

as follows:

1. Complete a detailed theoretical fatigue loading analysis program and perform

calculations to develop an accurate generalized fatigue design model to ac-

count for the variety of sign support structures in design.

2. Instrument one cantilever-type highway overhead sign structure with the nec-

essary instrumentation needed to evaluate fatigue loading on the structure, and

take field measurements under different natural wind and truck induced wind

gust conditions.

3. Evaluate the fatigue resistance of the structure to wind loading using the ex-

perimentally obtained measurements. Determine areas of the structure that are

most susceptible to fatigue and evaluate their performance under fatigue con-

ditions, and establish sections of the structure best suitable for calculating fa-

tigue loads from the experimental data.

4. Develop design fatigue load criteria for cantilever-type highway overhead

sign support structures from the experimental measurements.

5. Compare the experimental fatigue load criteria to the results and conclusions

of the theoretical program.

6. Create finite element models and check the developed fatigue load criteria

(experimental and theoretical) for accuracy by evaluating stress at critical lo-


6

cations of the structure. Make alterations to the FEA model and evaluate the

precision of the methodology developed in the theoretical program.

7. Perform fatigue load calculations in accordance with the Supports Specifica-

tions and assess the “accuracy” of Supports Specifications fatigue provisions

with the developed fatigue loading criteria.

8. Propose design recommendations as to fatigue load considerations for over-

head sign structures based on the developed loading criteria.

Work Plan

Summary

A detailed theoretical program to evaluate fatigue loading on sign support struc-

tures was developed. The theoretical program addressed the variety of sign support struc-

ture used in design each with different configurations, sizes, shapes, and material proper-

ties. Finite element models were developed to perform loading simulations of the devel-

oped fatigue criteria and stresses at critical locations were evaluated.

An experimental program was developed to check the accuracy of the loading

provisions developed theoretically. A cantilever-type highway overhead sign structure

(shown in Figure 1.4) was selected for field measurement. The structure was instru-

mented with electric strain gages to determine stresses at critical locations. Accelerome-

ters and anemometers were mounted on the structure to determine structural dynamic

properties and wind velocities behavior. Two types of applied loads were considered in

the testing program:


7

a) Natural wind gust, and

b) Truck induced wind gust.

The experimental program placed specific emphasis on the instrumentation and

measuring of forces in base plate vicinity including the anchor bolts. Recorded data was

analyzed and compared to developed analytical methods, used to verify the procedures of

Supports Specifications provisions, and to propose practical wind-induced fatigue design

recommendations for overhead sign structures.

FIGURE 1.4 Cantilever-type highway sign structure.

Project Tasks

The project involved an extensive analytical program to evaluate the fatigue loads

on sign support structures. An experimental program was performed to check the accu-
8

racy of the theoretical program on these structures. Fatigue loading criteria were devel-

oped and recommendations were proposed to provide a more reliable fatigue design

method. The objectives for this project were accomplished through the following tasks:

Task 1: Analytical Studies

A theoretical model was developed to address vibration behavior of sign support

structures to wind-induced fatigue loading. The model was created in such a way as to

account for the variety of sign support structures, each with different sizes, shapes, con-

figurations and material properties. Emphasis was placed on the vibration of the structure

due to wind-induced fatigue loading. The design fatigue load was determined based on

the dynamic characteristics of the structures. The infinite-life approach to fatigue design

was performed.

The loading criteria were used as input for computer software to check critical

stresses in the main members, and for evaluating the accuracy of the loading criteria. The

SAP 2000 v. 10 finite element analysis (FEA) computer software was used (11). A three-

dimensional full-scale cantilever structure was modeled in the FEA program. The model

was based on the shop drawings provided by ALDOT, and was same structure used in the

experimental program for field measurements and testing. A modal analysis was con-

ducted as well as static and dynamic loading simulations including material alterations.

Task 2: Site Selection

The site selection of the experimentally tested cantilever structure was located at

the same place where the previous one failed (see Figure 1.1).
9

Task 3: Sign Structure Instrumentation

The support structure was instrumented with strain gauges, accelerometers, and

anemometers. Electric strain gages were placed at maximum stress locations on the verti-

cal and horizontal structural members of the overhanging section, and in the vicinity of

the base plate to determine strains under the different loading conditions. Accelerometers

were placed on the post and the cantilever overhang to determine the major dynamic

properties. Wind speed readings were recorded using anemometers placed 4 ft (1.22 m)

above the top of the post (to prevent shielding effects) for ambient readings, and along

the cantilever section for truck induced gusts measurements. Each transducer provided

simultaneously measured time history streamlines.

Task 4: Structural Testing

The overhead sign structure was tested under loading conditions: a) Natural wind

gust, and b) Truck induced wind gust. The natural wind data was taken over an extended

time period in an effort to capture the predominant natural wind gusts. Wind data from

the National Weather Service near Huntsville, AL, in the form of the annual mean wind

velocity for the area, was determined to help schedule testing days, and to compare to the

wind measurements taken to distinguish if the measured results were representative of the

wind environment. A standard semi–trailer vehicle was used to apply the truck induced

wind gusts. Varying truck speeds were used and the corresponding structural response

data was recorded.


10

Task 5: Experimental Data Reduction

The data was analyzed using the critical stresses determined in the structural

members and anchor bolts corresponding to wind velocity excitation. An evaluation of

the fatigue resistance was performed first to understand the distribution of stresses in the

structure, and to determine areas of the structure that were most vulnerable to fatigue

damage. A modal analysis was also performed using the experimental measurements.

Structural dynamic characteristics of the structure were determined such as the modal

frequencies, modal shapes, and damping characteristics. The analysis was used in the de-

velopment of the theoretical program as well as to verify finite element models.

Fatigue loading was developed from the data analysis after accomplishing an un-

derstanding of the structural behavior to wind loading and the distribution of stresses. The

infinite life approach to fatigue design was utilized and applied. The data collected of

natural wind gust were analyzed as a randomly occurring continuous load, whereas the

truck induced data were evaluated based on principles related to transient loading envi-

ronments.

Task 6: Design Recommendations

Design fatigue load recommendations for natural wind and truck-induced gusts

were developed after the completion of Task 5. The comparison of the analytical results

to the experimental data was utilized and fatigue design criteria were developed that en-

compassed the results of these efforts. The recommendations addressed the “accuracy” of

Supports Specifications fatigue provisions for cantilever-type sign structures.


11

Task 7: Design Examples

The effect of the proposed provisions was assessed and explained by performing

fatigue load calculations for design. The examples compared fatigue loads using the fa-

tigue provisions of Supports Specifications and the fatigue loads according to the pro-

posed guidelines of this study of both experimental and theoretical means.

Task 8: Project Report

A report summarizing Tasks 1 through 7 was prepared.

Significance and Benefits

The fatigue provisions of Supports Specifications have not been fully used or

evaluated by most state DOTs, and as such, insufficient information is available on the

design of structural supports using the current fatigue provisions of the specifications.

The provisions were developed primarily through analytical methods, and therefore an

extensive comparison of the provisions to fatigue loading through experimental meas-

urements has not been performed. In addition, the provisions are only applicable to spe-

cific types of support structures, and changes in size, shape, and material properties

would result in significant different vibration behavior and subsequent fatigue load. A

study is needed to develop fatigue loads that apply to all type of support structures, and

reflect reliable loading criteria to be used in design. A methodology is needed that reveals

the vulnerability of sign supports structures to wind-induced fatigue loading, from which

the engineer can reliably make alterations to the structural design.


12

Dissertation Outline

An outline of the dissertation is provided. The outline identifies the enclosed

chapters with a brief description of the work involved. The dissertation is presented in the

following outline:

Chapter 1 Introduction: An introduction to the study is presented. The objectives

of the project are stated with a breakdown of the tasks needed to accomplish the

projects objectives.

Chapter 2 Literature Review: A review of information in the literature related to

the project is provided. The most relevant material is presented in a form as to de-

scribe the brief historical development of the fatigue guidelines for support struc-

tures.

Chapter 3 Fatigue Provisions of the AASHTO Supports Specification: An intro-

duction to the fatigue provisions for natural wind and truck-induced gusts in the

Supports Specifications is provided. A historical description on the development

of the specification is provided.

Chapter 4 Sign Structure Instrumentation: A description of the support structure

instrumentation for experimental testing is described and illustrated.


13

Chapter 5 Structural Testing for Natural Wind and Truck-Induced Gusts: A de-

scription on the development of the testing program is provided. The testing pro-

cedure is explained for natural wind and truck-induced gusts.

Chapter 6 Fatigue Resistance: A detail of the fatigue resistance analysis of the

support structure using the experimentally collected data is provided. The analysis

helped to understand the distribution of stress in the structure, and to determine

the proper approach to calculate the fatigue load from the experimental data.

Chapter 7 Experimental Modal Analysis: A modal analysis on the structure was

performed using the experimentally collected data. The analysis helped to under-

stand the dynamic behavior of the structure needed primarily for the analytical

program.

Chapter 8 Experimental Calculation of the Fatigue Load due to Natural Wind

Gust: The procedure used in the reduction of data collected experimentally for

natural wind gust. The methodology and calculation of the fatigue load due to

natural wind is presented and described.

Chapter 9 Theoretical Calculation of the Fatigue Load due to Natural Wind Gust:

An analytical program was developed to calculate the fatigue load due to natural

wind guts. The method was developed to account for the variety of sign support
14

structures in design, each with different configurations, sizes, shapes, and material

properties.

Chapter 10 Experimental Calculation of the Fatigue Load due to Truck-Induced

Gust: The procedure used in the reduction of data collected experimentally for

truck-induced wind gust. The methodology and calculation of the fatigue load due

to truck-induced gusts is presented and described.

Chapter 11 Theoretical Calculation of the Fatigue Load due to Truck-Induced

Gust: An analytical program was developed to calculate the fatigue load due to

truck-induced wind guts. The method was developed to account for the variety of

sign support structures in design, each with different configurations, sizes, shapes,

and material properties.

Chapter 12 Finite Element Analysis: The finite element analysis program is pre-

sented. The analysis was conducted to verify the structural analysis used to calcu-

late the fatigue load. The analytical programs were also verified by changing the

material properties to address the accuracy of the proposed methodology.

Chapter 13 Proposed Fatigue Provisions: The proposed fatigue design equations

for natural wind and truck-induced gusts are presented. Design examples are pro-

vided to compare with the current provisions in the Supports Specifications and to

clarify the proper use of the proposed design equations.


15

Chapter 14 Conclusions: The developed conclusions are presented based on the

completed work of this project.

List of References: All references used in the dissertation are provided.

Appendix: Other relevant information is attached in the appendices. This included

handouts, instrumentation layouts, and other useful information in the completion

of the project. Ranking schedules are also included that helped in scheduling test-

ing days between ALDOT and the UAB research team.


16

CHAPTER 2

LITERATURE REVIEW

Chapter Overview

An extensive literature review was performed on fatigue of overhead highway

sign support structures. A spreadsheet was developed to categorize the reviewed studies

by breaking down the research into important areas of interests. Interests included rec-

ommendations for future research, vibration analysis, and instrumentation of support

structures. The cataloging helped to develop and systemize the research program. Also

provided in the literature review chapter are descriptions of the most relevant studies that

played a crucial role in the development and progress of the fatigue provisions of the

Supports Specifications, as well as other projects that were considered important to the

research of this project.

Research Paper Breakdown

Various research papers and documentations on sign support structures were re-

viewed before developing the research program. Important aspects that had relevancy to

this project were identified during the review process. Each of the documents reviewed

were tagged and categorized regarding the predetermined interests of this study. A

spreadsheet was created that helped to label the reviewed studies with respect to the

tagged aspects. Figure 2.1 displays the spreadsheet showing the properties of interests for
17

this project and the color coded categorizing method. The numbers at the top of the

spreadsheet reference the papers (Figure 2.2) from which the color coded categories were

indentified (7-10, 12, 15-17, 19, 20, 33, 36, 37, 50, 52, 53, 55, 59, 60, 64). Only the most

important references related to supports structures are listed in the figure. When reference

to a particular subject was needed during the project, the spreadsheet was used by utiliz-

ing the legend with respect to the property of interest, and identifying the documentation

that contained information on the subject. This process helped to allow the research to

perform smoothly and efficiently.

Referenced Papers
Property of Interest
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Cantilever-Type Highway Sign Support Structure
Cantilever-Type VMS Support Structure ∆
Bridge-Type Highway Sign Support Structure
Bridge-Type VMS Support Structure ∆
High Mast Structure
Traffic Signal Structure
Luminare Structure
Strain Guage Transducers
Anemometer Transducers
Accelerometer Transducers
Pressure Transducers
Galloping
Vortex Shedding
Anchor Bolt Study
Natural Wind Gusts: Fatigue Loading ∆ ∆ ∆ ∆ ∆
Natural Wind Gusts: Fatigue Resistance ∆
Truck-Induced Wind Gusts: Fatigue Loading ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆
Truck-Induced Wind Gusts: Fatigue Resistance ∆
Horizontal Loading for Truck Gust ∆
Suction Pressure for Truck Gust ∆
Variation in Truck Type
Spectral Analysis
Natural Frequencies of Support Structures ∆
Damping of Support Structures
Modal Shapes of Support Structures ∆
Steel Structure
Aluminum Structure
Design Example

Legend
Involved Experimental Program
Invovled Analytical Program
Involved Experimental & Analytical Programs
∆ Includes Equation

FIGURE 2.1 Research paper breakdown.


18

Referenced Papers
Cook, Ronald A., Bloomquist, D., Agosta, A.M., Taylor, K.F., Wind Load Data for Variable Message Signs . Report
1
Number 0728-9488. Florida Department of Transportation, Research Management Center, April, 1996.
Azzam, D., Fatigue Behavior of Highway Welded Aluminum Light Pole Support Structures. Dissertation, University of Adron,
2
May 2006.
Creamer, B. M., Frank, K. H., Klingner, R. E., Fatigue Loading on Cantilever Sign Support Structures from Truck Wind
3 Gusts. Research Report Number 209-1F. Texas State Department of Highways and Public Transportation, Transportation
Planning Division, April, 1979.
Dexter, R. J., Ricker, M. J., Fatigue-Resistant Design of Cantilevered Signal, Signs, and Light Supports . NCHRP Report
4
469, The Transportation Research Board, Washington D.C. 2002
Zalewski, B., Huckelbridge, A., Dynamic Load Environment of Bridge-Mounted Sign Support Structures. Report No.
5
ST/SS/05-002. Ohio Department of Transportation, Office of Research and Development, September 2005.
Foutch, D.A., Kim, T.W., LaFave, J.M., Rice, J.A. Evaluation of Aluminum Highway Sign Truss Designs and Standards
6 for Wind and Truck Gust Loadings. Research Report N0. 153. Illinois Department of Transportation, Bureau of materials
and Physical Research, December 2006.
Ginal, S. Fatigue Performance of Full-Span Support Structures Considering Truck-Induced Gust and Natural Wind
7
Pressures. Thesis, Marquette University. December 2003.
Kaczinski, M.R., Dexter,R.J., and VanDien, J.P. Fatigue Resistant Design of Cantilevered Sign, Signal and Light
8
Supports. NCHRP Report 412. Transportation Research Board. Washington D.C. 1998.
South, S.M. Fatigue Analysis of Overhead Sign and Signal Structures. Report No. 115. Illinois Department of
9
Transportation, Bureau of Materials and Physical Research. May 1994.
Fouad, F.H., Calvert, E.A., and Nunez, E. Structural Supports for Highway Signs, Luminaires, and Traffic Signals.
10
NCHRP Report 411, Transportation Research Board. Washington D.C. 1998.
Edwards, J.A., and Bingham, W.L. Deflection Criteria for Wind Induced Vibrations in Cantilever Highway Sign
11
Structures. Report No. FHWA/NC/84-001, Center for Transportation Engineering Studies, North Carolina State University,
DeSantis, P.V., and Haig, P.E., Unanticipated Loading Causes Highway Sign Failure. Proceedings of ANSYS Convention,
12
1996.
Albert, M.N., Manuel, L., Frank, K.H., and Wood, S.L., Field Testing of Cantilevered Traffic Signal Structures under
13 Truck-Induced Gust Loads . Report No. FHWA/TX-08/0-4586-2, Center for Transportation Research, University of Texas
at Austin, 2007.
Cali, P., and Covert, E.E., On the Loads on Overhead Sign Structures in Still Air by Truck Induced Gusts. Wright
14
Brothers Facility Report 8-97, Massachusetts Institute of Technology.
Ramy, A.S., Fatigue Resistant Design of Non-Cantilevered Sign Support Structures. Thesis, University of Alabama at
15
Birmingham, 2000.
Fisher, J.W., Nussbaumer, A., Keating, P.B., and Yen, B.T., Resistance of Welded Details Under Variable Amplitude Long-
16
Life Fatigue Loading. NCHRP Report 354, The Transportation Research Board, Washington D.C., 1993.
Irwin, H.P., and Peeters, M. An Investigation of the Aerodynamic Stability of Slender Sign Bridges, Calgary. LTR-LA-
17
246, national Research Council Canada-Aeronautical Establishment, 1980.
McDonald, J.R., Mehta, K.C., Oler, W., and Pulipaka, N., Wind Load Effects on Signs, Luminaires and Traffic Signal
18 Structures. Texas Department of Transportation Report No. 1303-1F, Wind Engineering Research Center-Texas Tech
University, Lubbaock, TX, 1995.
Gilani, A.S., Chavez, J.W., and Whittaker, A.S., Fatigue-Life Evaluation of Chaneable Message Sign Structures, Volume
19 1 - As Built Structures. Report No. UCB/EERC-97/10, Earthquake Engineering Research Center, University of California,
Berkeley, CA, 1997.
Kashar, L., Nester, M.R., Johns, J.W., Hariri, M., and Freizner, S., Analysis of the Catastrophic Failure of the Support
20 Structure of a Changeable Message Sign. Structural Engineering in the 21st Century, Proceedings of the 1999 Structures
Congress, New Orleans, LA, 1115-118, 1999.

FIGURE 2.2 Referenced papers in the research paper breakdown spreadsheet.

Many aspects that were considered important by the researchers were not found in

the literature and were therefore excluded from the spreadsheet. Some aspects were sim-

ply dropped from the spreadsheet as the research progressed because of relevancy,

whereas other aspects were added which resulted in the breakdown shown in the Figure

2.1.
19

The research paper breakdown served as a useful tool during the development of

the research program. Relevant information on gathered literature documents was easily

and quickly identified to extract information. The breakdown also helped to identify areas

where research is needed or is lacking. Only the most prominent papers as they related to

the research performed with this project was included in the breakdown. A complete list

of reviewed papers is provided in the List of References at the end of the dissertation.

Research Review

A description of the reviewed studies that played a crucial role in the development

of the fatigue provisions in the Supports Specifications are listed as follows. Other pro-

jects included in the literature review were considered important to this research and are

described.

DeSantis and Haig (1996)

The fatigue provisions for truck-induced wind gust in the Supports Specifications

were initially based on this study. The research focused on a cantilever-type overhead

VMS support structure. The structure failed due to fatigue loading for which prompted

the study. After the structure was replaced, large deflections were observed due to wind

gust created from passing trucks. In the analysis, the researchers assumed that the veloc-

ity of the wind gusts onto the structure was equal in magnitude to the speed of the truck,

in addition to a gust factor equal to 1.3 to account for head winds. The wind pressure was

calculated using the fundamental fluid mechanics relationship between wind force and

the square of the wind velocity. The resulting wind pressure was doubled to account for
20

the upward deflection of the sign plus the downward deflection due to the pull of gravity.

The concluded value represented a pressure range to be used for fatigue design, and is

shown in Eq. 2.1.

PTG = 36.6Cd (psf) [Eq. 2.1]

where
PTG = design fatigue pressure load due to truck - induced wind gust
C d = drag coefficient

The researchers performed a FEA model of the structure using the software pack-

age ANSYS. The cantilevered end of the structure was observed in the field to deflect

vertically about 1 ft (0.305 m) in length after exposed to wind gust from passing trucks.

Experimentation was not performed to validate the tip deflection other than visual obser-

vation. To help verify the observations, the researchers back-calculated the wind pressure

that would theoretically produce a 1 ft (0.305 m) deflection, and inputted the pressure

into the ANSYS program of the modeled structure. The results matched the wind pres-

sure calculated from Eq. 2.1 and was concluded as the appropriate design load based on

this comparison (16).

Davenport (1961)

Davenport’s research was not focused on overhead sign support structures spe-

cifically, but rather focused on the characteristics of wind velocity. His research has pro-

vided an accurate model for simulating wind velocity behavior. The model was used in
21

the development of the fatigue provisions for natural wind gust in the Supports Specifica-

tions. It simulated the randomness along with the gustiness and turbulence commonly

associated with wind velocity. The model was developed in the form of a power density

spectrum generally used for predicting randomly occurring events. The power density

spectrum was created using experimentally measured wind velocity time histories gath-

ered from sites located around the world. An empirical formulation was developed as a

function of the mean wind velocity, and is shown in Eq. 2.2. Terrain coefficients were

also indentified the formulation process, and are shown in Table 2.1 (13).

4κV102 x 2
Sv ( f ) = 4
[Eq. 2.2]
f (1 + x )
2 3

where
S v ( f ) = wind velocity power spectral density at any height
f = frequency
V10 = mean wind velocity at a stadard height of 10 meters above ground leve
lκ = surface drag coefficient (Table 2.1)
1200 f f
x= 2
with 2 in cycles per meter.
V10 V10

TABLE 2.1 Terrain Coefficients (13)

Type of Surface κ α
Open unobstructed country (e.g., prairie-type grass-
0.005 0.15
land, arctic tundra, desert)
Country broken by low clustered obstructions such as
0.015 - 0.020 0.27 – 0.31
trees and houses (below 10 m high)
Heavily built-up urban centers with tall buildings 0.050 0.43
22

Kaczinski et al (1998)

The research performed by Kaczinski reported in NCHRP Report 412 formed the

framework of the fatigue provisions for galloping, vortex shedding, natural wind gust,

and truck-induced wind gusts in the Supports Specifications. The majority of his work

related to natural wind and truck-induced wind gusts was entirely theoretical. By using

the research performed by DeSantis and Haig, the fatigue provisions for the truck-

induced gust were created. Equation 2.1 was recommended as the appropriate pressure

load to use for truck-induced wind gusts.

Davenport’s wind velocity power density spectrum was used to create the fatigue

provisions for natural wind gust. The velocity spectrum was transformed into a wind

force spectrum. The spectrum was applied as input into an FEA program to generate a

stress response spectrum for four particular overhead sign support structures. The stress

spectrum was formed at different locations in the modeled structure and compared. The

fatigue load was then developed using the infinite-life approach. Equation 2.3 was devel-

oped from the results and was recommended as the appropriate pressure load to use for

natural wind gust.

PNW = 5.2Cd I F (psf) [Eq. 2.3]

where
PNW = design fatigue pressure load due to natural wind gusts
C d = drag coefficient
I F = importance factor
23

Report 412 also focused on developing S-N curves (stress vs. number of cycles)

for anchor bolt connection details. The results were used to create the constant-amplitude

fatigue thresholds currently available in the Supports Specifications. Recommendations

for anchor bolt design and structural analysis were provided. In addition, Report 412 in-

troduced “Importance Factors” to be used with the Supports Specifications. Fatigue de-

sign examples are provided for different types of supports structures including overhead

sign, traffic signals, and luminaires. The examples provided a thorough design procedure

using the proposed fatigue design equations. Stresses at critical details were calculated

and compared to the fatigue thresholds (52).

Fouad et al (1997-2005)

The research by Fouad et al in NCHRP project 17-10 and 17-10(2) (32, 33)

evaluated fatigue design of sign support structures. The research was compiled into a

specification based on the Allowable Stress Design (ASD) methodology, the basis of

which the current Supports Specifications were formed. The impact of the new specifica-

tions was analyzed with significant conclusions. The research brought together important

fatigue concerns such as galloping, vortex shedding, natural wind, and truck-induced

gusts. Many design examples were made which provided the framework for engineers to

use for fatigue evaluations. The work performed by Fouad et al compiled all relevant in-

formation on fatigue of sign support structures into a single stand-alone document. It now

resides as a well established document for fatigue design that is specific to sign support

structures, from which helped to propel future enhancements and innovations to fatigue

design of these structures. The researchers also provided valuable information on trans-
24

forming the specification into a Load and Resistance Factor Design (LRFD) methodology

in the future. The focus was to first develop a much needed stand-alone specification us-

ing the ASD method, and then upgrade to LRFD in the future (21-35).

Dexter and Ricker (2002)

The research performed by Dexter and Ricker reported in NCHRP Report 469

formed the framework of the current Supports Specifications. Their research involved

experimental and analytical programs. They concluded that the fatigue provisions for

natural wind gusts in the Supports Specifications were accurate. However, the truck-

induced gusts were highly conservative as compared to their experimentally determined

values. A reduction in the truck provisions were recommended by Eq. 2.4. The reduction

did not reflect the experimental values gathered in the report because they were signifi-

cantly less the fatigue provisions, and it was believed the committee would not accept

them. The authors were also concerned with the type of structures used in the analysis.

The fatigue load was determined from analysis on variable message signs which are sig-

nificantly heavier than conventional signs. The reduction of the fatigue design equation

for truck-induced gusts shown as Eq. 2.4 was recommended.

PTG = 18 .8C d I f (psf)


[Eq. 2.4]

where
PTG = design fatigue pressure load due to truck - induced wind gust
C d = drag coefficient
I F = importance factor
25

Report 469 also provided a reduction to the fatigue load to accommodate for

height of the structure above the ground. They recommended a linear reduction in wind

pressure starting from 19.7 ft (6.00 m) above ground level to zero at approximately 32.8

ft (10.0 m) above ground level. A recommended area on the support structure to apply the

truck load was also provided. It was proposed to apply a uniformly distributed load to the

12 ft (3.66 m) area of the structure that produces the maximum stress range. This was

recommended as opposed to applying the load to portions of the structure that is not di-

rectly above the traffic lane (16).

Cook et al (1996)

The research performed by Cook et al is indentified in this review due to its rele-

vancy to fatigue design due to truck-induced gusts. Their research determined the truck-

induced wind gusts through direct measurement of wind pressure. Pressure transducers

were placed on a bridge spanning a highway to measure the wind pressure from passing

trucks. Importantly, the bridge was extremely rigid so that the measured values were not

influenced by the type of structure. Twenty-three loading events were measured and re-

corded.

The results provided a clear view of the wind behavior created from passing

trucks. It revealed a biaxial loading event with a vertically applied pressure and a hori-

zontally applied pressure. A suction event was also recorded as the truck passed under-

neath the structure. The results indicated a ramped loading impulse. Variation of the

truck-induced wind gust with respect to height above ground level was also measured. A

10% pressure reduction per foot above 17 ft (5.18 m) above ground level was found (12).
26

Creamer et al (1979)

The work performed by Creamer et al was very influential for future research.

Their research focused on fatigue induced loading due truck wind gusts. Where Cook’s

determined the truck wind pressure directly using pressure transducers, Creamer’s work

determined the truck wind pressure indirectly. They measured strain variations at critical

locations of a cantilevered structure due to truck wind gusts, and back calculated for wind

pressure. A ramped impulse function was developed and inputted into a FEA model of

the tested structure. The loading simulation produced similar strains on the structure as

measured experimentally (10).


27

CHAPTER 3

FATIGUE PROVISIONS OF THE AASHTO SUPPORTS SPECIFICATIONS

Chapter Overview

An overview of the development of the natural wind and truck-induced wind

gusts fatigue provisions within the AASHTO Standards Specifications is provided. This

was believed to be an important and required task so that a complete understanding of the

developed fatigue provisions was obtained before addressing their accuracy. The provi-

sions for natural wind were formed indirectly through theoretical calculations. They were

developed using the infinite-life approach to fatigue design. Spectral analysis formed the

framework of the development procedure. The truck-induced wind gust fatigue provi-

sions were developed through theoretical and experimental observations.

Fatigue Load due to Natural Wind

Infinite-Life Approach

The method used to develop the natural wind provisions was based on spectral

analysis in collaboration with the infinite-life approach to fatigue design. The provisions

are adequate within certain limitations, as they were formed using four particular catego-

rized structural types, each of which were chosen to represent the population of the cate-

gory selected. The structural response of two overhead signal support structures, one can-

tilever-type overhead sign support structure, and one luminaire support structure to natu-
28

ral wind excitation were analyzed, the transmitted stresses of each were averaged, and the

code was developed from the averaged results (52).

The AASHTO fatigue code was based on the constant amplitude fatigue limit

(CAFL). Tests to developed plots of the stress vs. the number of cycles (S-N curves) for

common connection details were used to determine this value. Fatigue stress ranges that

occurred below the CAFL were estimated to have an infinite life. Stresses ranges at criti-

cal locations such as welded or bolted connections must be under the CAFL to be in

compliance with the code. Changes in the design must be made if they are not to lower

the fatigue stresses to fall below the CAFL.

The CAFL tests were based on constant amplitude, whereas amplitudes that occur

due to natural wind are random with variable amplitudes. To account for variable ampli-

tude loading, the infinite-life approach adopted the findings of NCHRP Report 354 where

variable amplitude fatigue tests were performed on a variety of full scale connection de-

tails. From the results, it was determined that failure would still occur if the variable am-

plitudes exceeded the CAFL of the connection detail 0.05% of the time. However, if the

CAFL was exceeded 0.01% of the time, the specimen was said to have an infinite life

(19). In view of the results, the infinite-life approach involved estimating the stress

ranges that would occur onto the structure with a 0.01% exceedence probability, includ-

ing dynamic amplification. These ranges were referred to as the fatigue limit-state load

ranges. The structure was designed such that the limit-state stress ranges did not exceed

the CAFL of the detail in question.


29

Predicting the Environment

The first step in determining the fatigue limit-state load ranges to develop the

AASHTO fatigue provisions was to estimate the structural excitation due to natural wind.

This involved predicting the natural wind environment for which the structure is to be

exposed. This was done using a spectral analysis. Simulating the natural wind force can

be difficult due to the randomness, turbulence and gustiness of natural wind velocity. A

static load that represents the force due to wind can be easily calculated, as well as a dy-

namic load in terms of a periodic wave such as a sine wave. This type of examination

does not account for the gustiness and turbulent nature of the wind itself. In most cases, a

gust factor is used to account for the random nature of wind excitation, specifically for

structural capacity predictions. A. G. Davenport developed a wind velocity power density

spectrum (PDS) curve (13) shown in Eq. 3.1 that simulated the turbulent nature of wind

velocity.

4κV102 x 2
Sv ( f ) = 4
[Eq. 3.1]
f (1 + x )
2 3

where
S v ( f ) = wind velocity power density spectrum at any height
f = frequency
V10 = mean wind velocity at a stadard height of 10 meters above ground level
κ = surface drag coefficient (Table 3.1)
1200 f f
x= 2
with 2 in cycles per meter.
V10 V10
30

TABLE 3.1 Terrain Coefficients (13)

Type of Surface κ α
Open unobstructed country (e.g., prairie-type grass-
0.005 0.15
land, arctic tundra, desert)
Country broken by low clustered obstructions such as
0.015 - 0.020 0.27 – 0.31
trees and houses (below 10 m high)
Heavily built-up urban centers with tall buildings 0.050 0.43

Random excitation is best analyzed over a frequency spectrum. PDS curves are

commonly used for dynamic loading analysis, but more specifically suited to random vi-

bration. In random excitations, there are no periodic systems within the excitation that

can be analyzed at specific frequencies because the excitation is applied randomly over a

spectrum of frequencies. To account for this, the natural wind fatigue analysis involved

an examination of the total energy, or power, in the random excitation over a frequency

band. A typical PDS curve is plotted with the ordinate in units of the parameter (force,

acceleration, velocity, displacement, etc.) squared divided by the bandwidth (e.g., N2/Hz),

and the abscissa in units of frequency (Hz). The peaks on a PDS curve identify the fre-

quency range(s) at which the majority of the energy lies within the particular excitation.

The area under a PDS curve is equal to the mean square value, and the square root of this

area is equal to the root-mean-square (RMS) value: a value of which plays an important

role in determining the magnitude of loads that are transmitted onto a structure from vi-

bration (39, 40).

Davenport developed his wind velocity PDS curve from at least 70 experimental

wind velocity data collections from various locations around the world. His intention was

to develop a model which simulated the turbulence and gustiness of wind velocity. He

developed Eq. 3.1 from the 70 experimental data collections. The equation is a function
31

of wind velocity frequency with respect to a mean wind velocity at a specified height. His

formulation is plotted in Figure 3.1 for frequencies ranging from 0 to 10 Hz, an open ter-

rain (see Table 3.1), and an annual mean wind velocity of 11 mph (5 m/s).

FIGURE 3.1 Wind velocity PDS for annual mean wind velocity 11 mph.

Once the behavior of the wind velocity environment was estimated, the PDS was

transformed into a wind force PDS by using principles related to fluid mechanics. An in-

duced force onto a structure due to natural wind, referred to as drag, is proportional to

wind velocity squared, as shown by the Eq. 3.2

1
FD = ρC d AV 2 [Eq. 3.2]
2
32

where
FD = drag force
kg
ρ = density of air = 1.22
m3
C d = drag coefficient
A = area of exposed surface
V = wind velocity at any height.

By utilizing the proportionality between drag pressure and wind velocity, and accounting

for the turbulent wind velocity and its variance about the mean wind velocity, a wind

pressure PDS was developed from Davenport’s wind velocity PDS shown in Eq. 3.3. The

plotted equation is shown in Figure 3.2 for an annual mean wind velocity of 11 mph (5

m/s) and normalized for exposed area and the drag coefficient.

S F ( f ) = ρ 2Cd2 A2V 2 Sv ( f ) [Eq. 3.3]

The PDS curve accounts for the gustiness and turbulence of wind velocity over a

spectrum of frequencies, based on an averaged wind velocity taken at a specified height

above ground level. Since most support structure are at or around 32.8 ft (10.0 m) in

height, the PDS curve was well suited for these types of structures. Yet, the PDS can be

used at any particular height by using the power law profile for approximating variation

in wind velocity with height, as shown in Eq. 3.4:

V = V10αz α [Eq. 3.4]


33

where
V = wind velocity at height z
α = surface coefficien t in Table 4
z = height above ground.

FIGURE 3.2 Force PDS for annual mean wind velocity 11 mph.

In this case, where the objective was concentrated on formulating a design code

for fatigue wind, the wind velocity variable in the pressure PDS equation was taken at the

standard height of 32.8 ft (10.0 m) above ground level, and kept uniform across the ex-

posed façade. The purpose of which was to provide a simplified design equation for

commercial use. Some conservative formulation exists as the wind velocity typically in-

creases from the ground level upwards (13).


34

Once the natural wind environment was estimated, the next step was to apply the

PDS to the infinite-life approach in determining the limit-state stress ranges. Since the

force spectrum was based primarily on the annual mean wind velocity, the wind velocity

that was exceeded 0.01% of the time was found and referred to as the limit-state wind

velocity. The force spectrum was calculated using the limit-state wind velocity and used

to calculate the structural response to determine the limit-state stress range.

Wind velocity is random in nature, but it can be predicted though statistical rela-

tionships. It has been found through many experiments that the magnitude of the wind

velocity vector will form a Rayleigh distribution (14, 52, 54). By using the Rayleigh dis-

tribution, the 0.01% exceedence probability can be found through the relationship in Eq.

3.5 based on the annual mean wind velocity.

−πv 2

PE (v) = e 4V
2
[Eq. 3.5]

where
PE (v) = probability
v = wind velocity corresponding to the probability
V = mean wind velocity.

An analysis was conducted to determine which annual mean wind velocity to use in

Eq. 3.5 to determine the limit-state wind velocity (wind velocity with a 0.01% ex-

ceedence probability). The annual mean wind velocities of major U.S. cities were ana-

lyzed. It was found that an annual mean wind velocity of 11 mph (5 m/s) was exceeded in

only 19% of the U.S. cities analyzed and was therefore chosen. By plugging in 11 mph in

Eq. 3.5, and solving for the wind velocity corresponding to the 0.01% probability, a limit-
35

state wind velocity was found to be equal to 37 mph (17 m/s). The force spectrum was

then formed using the limit-state wind velocity (see Figure 3.3) and was used as the natu-

ral wind velocity prediction model for structural excitation.

FIGURE 3.3 Force PDS using limit-state wind velocity 37 mph.

Structural Excitation

Once the excitation was determined, the spectral density of the response was cal-

culated. This was done through finite element analysis for the signal, sign, and luminaire

support structures. Each structure had different dynamic characteristics such as natural

frequency and critical damping percentages, all of which are based on the size, shape, and

material of the structure. The response PDS was calculated in units of stress squared di-
36

vided by Hz. The area underneath the response PDS curve is equal to the variance of the

response about the mean. The square root of the area is equal to the RMS. If the mean is

equal to zero, then the RMS is equal to the standard deviation (58, 62). Since the re-

sponse curves of support structures are predominately controlled by a single frequency of

vibration, the developed response PDS curves were narrow-banded at the natural fre-

quency of the structures. Therefore, the stress range was calculated as a constant ampli-

tude in the form of a sinusoid. For any sinusoid, the peak stress amplitude is equal to the

square root of two times the RMS (assuming a zero mean for analysis purposes). The

stress amplitude was then multiplied by two to account for a peak-to-peak stress range.

The RMS was determined from the response PDS for each structure, and the effective

stress range was calculated using Eq. 3.6

S reff = 2.8σ rms


[Eq. 3.6]

where
S reff = effective stress range or limit - state stress range
σ rms = RMS of the response

Design Fatigue Equation for Natural Wind

Equation 3.7 was calculated for the four type of structures analyzed. The analysis

was normalized for the drag coefficient and exposed area. The result ranged from 3.6 to

6.3 psf (170 to 300 Pa) using 1% and 2% damping percentages. An average was taken

and used for the code. The final equation (Eq. 3.7) was multiplied by the drag coefficient,

importance factor, and a wind factor to account for other annual mean wind velocities.
37

2
v 
PNW = 5.2C d I F   [Eq. 3.7]
 11 

where
PNW = design fatigue load due to natural wind, psf (Pa)
C d = drag coefficient
I F = importance factor
v = annual mean wind velocity other than11 mph (5 m/s)

Fatigue Load due to Truck-Induced Gust

The truck-induced gust fatigue provision was developed as an impulse loading

occurring over a finite length of time. The loading was created when semi-trailer trucks

passed underneath the structure causing the structure to vibrate. The vibration generated

stresses in the structure that can potentially accumulate fatigue damage over time.

The Supports Specifications design equations for truck-induced gusts was origi-

nally developed from the work performed by Desantis and Haig (16), and later revised

from the work presented in NCHRP Report 469 (14). It assumes that the wind loading

onto the structure was equivalent to the speed of the passing truck, plus a gust factor of

1.3 to account for head wind. For instance, a truck traveling at 65 mph (29.1 m/s) would

produce a 65 mph (29.1 m/s) wind onto the structure. The result was an 18.3 psf (876 Pa)

magnitude static pressure. The code only accounted for a vertical force applied upward

(ground up) onto the structure. Assuming that the upward motion of the structure result-

ing from the truck gust was equivalent to a proceeding downward motion, the 18.3 psf

(876 Pa) was doubled to account for a peak-to-peak range equaling 36.6 psf (1760 Pa).

NCHRP Report 469 investigated the fatigue load due to truck gusts performed by

researchers on VMS cantilever structures. The research indicated that the current code
38

was too conservative based on their conclusions. A reduction of the loading was recom-

mended by the following equation:

2
V 
PTG = 18.8C d I F   [Eq. 3.8]
 65 

where
PTG = design fatigue load due to truck gust, psf (Pa)
C d = drag coefficient
I F = importance factor
V = truck speed, mph (m/s)

The reduction was determined based on experimental evidence from VMS struc-

tures. Strain measurements on a cantilevered VMS support structure were recorded dur-

ing random truck events. An equivalent static load was calculated that would produce the

same strain range measured experimentally. The resulting load was to be applied verti-

cally to all horizontal areas (underside of the structure) along a 12 ft length (3.7 m), or

equal to the width of a travel lane. A reduction in the load with respect to height of the

structure above ground level was also permitted. It was discovered the pressure load de-

creased linearly with height, with the maximum occurring at 19.7 ft (6 m) above ground

level to zero at 32.8 ft (10 m).


39

CHAPTER 4

SIGN STRUCTURE INSTRUMENTATION

Chapter Overview

A cantilever-type highway overhead sign structure was selected for field meas-

urement. The structure was instrumented with electric strain gages to determine stresses

at critical locations. Accelerometers were mounted on the structure to determine struc-

tural dynamic behavior, and anemometers were attached to measure wind velocity vec-

tors (magnitude and direction). Each transducer provided time history streamlines that

were collected simultaneously using a data acquisition system. A detailed description of

the instrumentation program is provided in this chapter.

Cantilever-Type Highway Overhead Sign Support Structure

A four chord truss cantilever-type highway overhead sign support structure was

chosen for analysis. The structure was located at northbound Exit 340B on highway I-65

near Decatur, Alabama USA. A picture of the structure evaluated for this project is

shown in Figure 4.1.


40

FIGURE 4.1 Cantilever-type highway overhead sign support structure.

Geometric Properties

The truss overhang was made of 3.5 in (88.9 mm) diameter steel pipe supported

by a 24 in (610 mm) steel pipe shaft support. The shaft was support by a 1.25 in (31.8

mm) thick by 35 in (889 mm) diameter base plate with eight 1.5 in (38.1 mm) anchor

bolts connected to a concrete foundation. The structural configurations are shown in Fig-

ure 4.2 through Figure 4.4.


41

14 ft 25 ft

10.5 ft

2 ft Four Chord
Cantilever
Truss

11.5 ft 4 ft

4 ft
Post

Sign 26.75 ft

FIGURE 4.2 Elevation view of structure configuration.

4 ft

3 ft

10.75 in

37.75 ft

FIGURE 4.3 Plan view of structure configuration.


42

3 ft

FIGURE 4.4 Side view of structure configuration.

Material Properties

The post and truss section was made of API-5L-X52 steel pipe. All plates were

structural steel ASTM A572 Gr. 50. The anchor bolts were AASHTO M314-90 Gr. 55

(essentially the same as ASTM F1554 Gr. 55). The W-shape and T-shape sections used

for the sign-to-truss connection were made of A572 Gr. 50 steel. The sign was aluminum

wrought alloy designated as 6061-T6. The concrete pile foundation was conventional

concrete with #9 size rebar Gr. 60. All of the major required material properties of the

designations given are listed in Table 4.1.


43

TABLE 4.1 Material Properties

Modulus of Yield Tensile


Material
Material Members Elasticity Stress Stress
Designation
(psi) (psi) (psi)

Steel Pipe Truss, Post API-5L-X52 29,000,000 52,000 66,000

Truss and ASTM A572


Steel Plate 29,000,000 50,000 65,000
Base plates Gr. 50
AASHTO
Steel Rod Anchor Bolts M314-90 Gr. 29,000,000 55,000 75,000
55

Aluminum Sign 6061-T6 10,000,000 37,000 42,000

4, 000 psi
Pile
Concrete Compressive 3,600,000 NA NA
Foundation
Strength
Pile ASTM A706
Rebar 29,000,000 60,000 80,000
Reinforcement Gr. 60

Instrumentation Process

There was a unique advantage associated with this project. The researchers were

able to instrument the structure before and during assembly, which allowed for a thor-

oughly detailed instrumentation program. Limitations associated with an already assem-

bled structure in the field were avoided as a result. All strain gauges, except on the anchor

bolts, were placed on the structure at the ALDOT laboratory in Montgomery, Alabama

USA before transportation to the assembly site (Figure 4.5 – 4.6). The anchor bolts were

instrumented on site the day after the foundation was poured and the anchor bolts set

(Figure 4.7). The post support was placed on the foundation and bolted to the anchor

bolts after the anchor bolts were gauged. A period of 14 days (28 days after pouring)
44

transpired after the post support was erected for concrete curing of the foundation. After

the curing stage was completed, the truss overhang and sign was attached to the post sup-

port (Figure 4.8 – 4.9). A lift bucket was used to place the accelerometers and anemome-

ters in their appropriate preconceived locations. The finished instrumented structure is

shown in Figure 4.10.

FIGURE 4.5 Strain gauged post support in laboratory.


45

FIGURE 4.6 Strain gauged truss overhang in laboratory.

FIGURE 4.7 Set anchor bolts in structure foundation.


46

FIGURE 4.8 Truss overhang attachment.

FIGURE 4.9 Sign attachment.


47

FIGURE 4.10 Finished instrumented structure.

Strain Gauges

Uni-axial strain gauges and single-plane rectangular strain gauge rosettes (meas-

uring in three primary directions: 0°, 45°, and 90°) were used to measure the strain re-

sponse of the structure. They were placed on the structure at locations where strains were

most critical, and especially applicable to determining the causative load. These locations

included: truss chord members near the truss-to-support connection, bottom of the pole

support member near the base plate, and the anchor bolts. A list of all strain gauges and

their locations along with a schematic detailing these locations are provided in Appendix

A.
48

Truss Chord Members

A total of 16 uni-axial strain gauges were used on the truss chord members lo-

cated near the truss-to-support connection: four gauges per chord placed along the cir-

cumference of the member. These gauges measured strain in a single direction, oriented

along the longitudinal length of the chord member, in an attempt to capture tensile and

compressive strain readings. A schematic of the gauge locations are shown in Figure

4.11, with the actual gauged structure at this location shown in Figure 4.12. A picture of

the gauges on the assembled structure is shown in Figure 4.13.

Placed on each
Chord at 90º
around Chord
Circumference
Centerline
of 1st Truss
Panel

3’- 4.5”

FIGURE 4.11 Truss-to-pole support strain gauges.


49

FIGURE 4.12 Chord strain gauges.


50

FIGURE 4.13 Chord strain gauges on the assembled structure.

Post Support

Uni-axial and rosette strain gauges were used for the post support. The gauges

were attached to the structure in the laboratory before shipment to the assembly site. Two

levels, located nearest to the base plate, were gauged along the circumference of the

member in accordance to the schematic in Figure 4.14. This was done in an attempt to

capture tensile and compressive strains near the base plate. Forty-five degree rectangular

rosettes were used to measure transverse shear and torsion strains. A close up of Section

AA viewing a rosette and Section BB located 4 in (102 mm) above Section AA viewing a

uni-axial strain gauge is shown in Figure 4.15. Two levels were chosen for redundancy

issues in the case of strain gauge measurement failure since the structure was tested out-
51

side, and the testing duration lasted over multiple months. A picture of the assembled in-

strumented post support is shown in Figure 4.16

Section AA Section BB

Uni-axial Strain Gauge

Rosette Strain Gauge

4 in Section BB

Section AA
12 in

Base Plate Weld

FIGURE 4.14 Pole support strain gauges.


52

FIGURE 4.15 Rosette and uni-axial strain gauges on shaft.

FIGURE 4.16 Instrumented post support.


53

Anchor Bolts

Strain gauges were attached to the anchor bolts in the field, after the concrete

foundation concrete was placed. A picture of the anchor bolts before placement into the

concrete foundation is shown in Figure 4.17, and after placement in the foundation before

gauging in Figure 4.18. The material specifications of the anchor bolts are provided in

Table 4.1. The concrete foundation reached 11 ft (3.35 m) into the ground, with the rein-

forcement cage reaching 10.5 ft (3.20 m) deep.

A total of eight anchor bolts were used for the structural foundation connection.

Uni-axial strain gauges were attached to each anchor bolt in the orientation shown in Fig-

ure 4.19, totaling eight strain gauges. The gauges were attached as close as possible to the

foundation due to assembly procedures and leveling processes. The gauges were located

tangent to the outside rim of the anchor bolt group formation, with the sensitive end ori-

ented along the longitudinal length of the bolt. The threads at the gauge location were

shaved (Figure 4.20) and cleaned for gauge attachment. This was done for each bolt. A

picture of a finished instrumented anchor bolt is shown in Figure 4.21, and of the assem-

bled structure in Figure 4.22.


54

FIGURE 4.17 Anchor bolts before installation.

FIGURE 4.18 Anchor bolts and foundation.


55

Support

Base Plate

Strain Gauge

Anchor Bolt
Foundation

Orientation

FIGURE 4.19 Anchor bolt strain gauges.

FIGURE 4.20 Anchor bolt strain gauging preparation.


56

FIGURE 4.21 Strain gauged anchor bolt.

FIGURE 4.22 Instrumented anchor bolts on assembled structure.


57

Anemometers

Anemometers were used to measure wind velocity and direction. The WindSonic

ultra-sonic wind and direction sensor was used for this application (see Figure 4.23). The

anemometer calculated speed and direction by measuring the time needed for generated

sound pulses to travel from one transducer to the other within the air gaseous medium. A

combined total of six anemometers were required for wind measurements of natural wind

and truck induced wind gusts. Two were placed above the post support structure, and four

on the truss section of the support structure. Steel extensions were fabricated to fit the

anemometers to the structure. A list of the anemometers identification and locations are

provided in Appendix A, along with a schematic detailing their locations. An overview of

their locations can also be seen in Figures 4.24 – 4.26. The mounting instructions pre-

pared for ALDOT are given in Appendix B.

FIGURE 4.23 WindSonic ultra-sonic wind and direction sensor.


58

FIGURE 4.24 Overview of anemometer layout.

FIGURE 4.25 AN-1 and AN-2 above traffic lane.


59

FIGURE 4.26 Anemometers AN-3 and AN-4 above post.

Anemometers AN-1 and AN-2 were used to measure the wind created from trav-

eling trucks. They were placed along the centerline of the traffic lane above where the

trucks traveled. The distance was 30 ft (9.14 m) from the centerline of the post support.

The anemometers were attached onto the truss structure using steel extensions as to not
60

be blocked by the sign. Two anemometers were used for redundancy issues, as well as for

providing differing planes of measurement: AN-1 measured the horizontal plane whereas

AN-2 measured the vertical plane. This enabled a three dimensional wind velocity vector

to be measured and used in the data analysis.

Anemometers AN-3 and AN-4 were used to measure the ambient wind environ-

ment. The anemometers were attached to a steel extension that reached 4 ft (1.22 m)

above the structure to avoid any effects for wind dynamics (turbulence, vortices, etc.)

caused by the structure. Two anemometers were used for redundancy as well as for pro-

viding differing measurement planes: AN-3 measured the vertical plane whereas AN-4

measured the horizontal plane. This enabled a three dimensional wind velocity vector to

be measured and used in the data analysis. Each anemometer provided wind velocity and

direction angle measured from an identified North direction marked on the instrument

with measurement recordings in compass bearings. For all anemometers, the North com-

pass on the instrument was directed from the front façade of the sign, opposing the direc-

tion of traffic.

Accelerometers

Accelerometers were used to obtain the acceleration response of the structure.

They were of the piezoelectric type that uses a piezoelectric crystal mounted to a small

mass from which the voltage output is converted to acceleration. Each accelerometer had

a maximum capacity of 96.5 ft/sec2 (3 G). The data was used to determine major dynamic

characteristics of the structure such as:


61

 Natural frequencies and periods

 Modal shapes

 Critical damping percentages

The locations of the accelerometers were strategically chosen, as the accelerome-

ter only gave the natural frequency of the member upon which it rests. A combined total

of six unidirectional accelerometers were required for the structure. They were placed at

particular locations to measure each possible degree of freedom in vibration direction of

the structure. Accelerometer locations are shown in Figure 4.27.

AC-1

AC-2

Centerline
Post

AC-3
Centerline
Truss

FIGURE 4.27 Accelerometer locations for cantilever structure.


62

Three accelerometers were placed at location AC-1 (see Figure 4.27) to measure

the vertical (perpendicular to the direction of traffic), horizontal (parallel to the direction

of traffic), and longitudinal (transverse to the direction of traffic) accelerations of the post

member. Two accelerometers were placed at location AC-2 to measure the vertical and

horizontal directions (Figure 4.28). One accelerometer was placed at location AC-3 to

measure the horizontal direction for indentifying torsion behavior of the overhang truss

(Figure 4.29). An overview of locations AC-2 and AC-3 is shown in Figure 4.30.

The accelerometers were fixed to the mounting surface by means of a two

threaded screws. Electrical insulation between the accelerometers and the test surface was

built into the transducer device. All mounting surfaces were flat to avoid distortion that

may produce strains which could affect the accelerometer’s response. Carefulness was

taken when screwing the accelerometers to the attachment surface as to not overreach the

torque recommended by the manufacturer.

Mounting blocks were used at each location for accelerometer attachment (see

Figure 4.31). For instance at location AC-1 where three accelerometers were needed, all

three were attached to one single attachment block. A picture of this location with the at-

tached mounting block is shown in Figure 4.32. Once the truss section of the structure

was erected, the AC-1 accelerometers were screwed onto the mounting block with the

sensitive ends (measurement direction) oriented in the proper directions. The same type

of block was used at locations AC-2 and AC-3. Due to the round surface at these loca-

tions (steel pipe truss web member), a small flat steel plate was welded to the side of the

members for which the mounting blocks and accelerometers were attached. The size of

the plate was manufactured as small as possible, with enough space to accurately and se-
63

curely mount the accelerometers and the plate to the web member, but without the possi-

bility of creating significant additional wind drag. The initial mounting instructions pre-

pared for ALDOT are given in Appendix C.

Data Acquisition System

All instrumentation was wired into a data acquisition system. The data was con-

verted into engineering units by the data acquisition system, filtered, and stored onto the

hard drive of a portable computer. Data was then saved onto Maxwell CD-R data storage

disks to be distributed after testing. A white van was used to hold the data acquisition

system and computer during testing. It was driven to and parked underneath the sign

structure on the side of the highway for testing. Figure 4.33 shows a typical test setup

with the van and all instrumentation wiring fed through a side hole in the van and hooked

to the acquisition system. A close up of the data acquisition system and computer is

shown in Figure 4.34.

The data acquisition was capable for collecting all data from instrumentation si-

multaneously, which was required for the fatigue tests. The maximum number of chan-

nels used was as follows:

 Strain Gauges: 48 channels

 Anemometers: 8 channels

 Accelerometers: 6 channels

 Total Channels = 62
64

FIGURE 4.28 Accelerometer location AC-2.

FIGURE 4.29 Accelerometer location AC-3.


65

FIGURE 4.30 AC-2 and AC-3 view.

FIGURE 4.31 Accelerometer mounting block.


66

FIGURE 4.32 Mounting block at AC-1.

FIGURE 4.33 Typical test setup with van data acquisition system.
67

FIGURE 4.34 Van data acquisition system and computer setup.

It was important to be specific about the sampling rate of the data acquisition sys-

tem. The vibration behavior of sign support structures are typically around 1 to 10 Hz,

including all modes of vibration. In view of this, data were collected at 60 samples per

second for all instrumentation, which was at least six times greater than the highest an-

ticipated modal frequency of 10 Hz. The Nyquist frequency was therefore 30 Hz, allow-

ing all frequencies recorded below 30 Hz to be accurate, avoiding aliasing and other

prominent sampling errors. A digital low-pass Butterworth filter was used and set at 20

Hz to filter out unwanted high frequencies from the collected data.


68

CHAPTER 5

STRUCTURAL TESTING FOR NATURAL WIND


AND TRUCK-INDUCED WIND GUST

Chapter Overview

The sign support structure was tested under loading conditions: a) Natural wind

gust, and b) Truck induced wind gust. The wind data was taken over an extended time

period in an effort to capture the predominant natural wind gusts. Wind data from the Na-

tional Weather Service near Huntsville, Alabama USA, in the form of the annual mean

wind velocity for the area, was determined to help schedule appropriate testing days be-

tween ALDOT and the UAB research team, and to compared to the wind measurements

taken for accuracy so that a proper representation of the natural wind environment was

established. A standard semi-trailer vehicle was used to apply the truck induced wind

gusts. Varying truck speeds was used and the corresponding load data was recorded.

Natural Wind Gust

Pre-Determined Sample Size

At least 32 hours of wind velocity data was desired for data collection. The inten-

tion was to have the collected data distributed between the following sampling intervals:

 0 to 10 mph (0 – 4.47 m/s): low wind

 10 to 20 mph (4.47 – 8.94 m/s): medium wind (fatigue wind)


69

 20 to 30 mph (8.94 – 13.4 m/s): high wind.

A 20 to 30 mph (8.94 – 13.4 m/s) was considered as high wind according to the

labels. The wind velocity having a 0.01% exceedence probability in accordance to the

infinite-life approach of the Supports Specifications was 38 mph (17 m/s). Capturing this

wind was considered probable, and likely during sampling of the categorized “high wind”

label; however such conditions warranted extreme weather that may be unsuitable for

field testing due to safety issues. The intention was to capture the vibration behavior of

the sign support structure in response to vibratory induced wind velocities, typically

above 10 mph (4.47 m/s), and form a relationship between wind velocity and structural

response stress ranges. Low wind data was wanted in order to establish a wind velocity

limit in which structural vibration was excited.

Testing Day Ranking Schedule

The next issue was determining appropriate testing days for natural wind. A good

distribution of measured wind velocities, and structural response, was needed. Since wind

cannot be produced manually, the days that were scheduled for testing were chosen based

on a trial and error process, which was unavoidable. ALDOT needed at least a two week

advance to schedule times. In addition, testing could not be performed during precipita-

tion.

A ranking schedule was prepared to help with the scheduling decisions. It was

based on a raking scheme over a spectrum of 1 to 10:


70

 1 being the worst day to test for natural wind with respect to fatigue loading, and

 10 being the best day to test for natural wind with respect to fatigue loading.

The ranking system was developed from historical data collected by the National

Weather Service Forecast Office located in Huntsville, Alabama USA. Climatology re-

cords ranging from early 2008 to 1999 were used in its development. It was prepared for

the months of October 2008 to March 2009 for testing. All ranking schedules used are

provided in Appendix D.

A ranking of 10 indicated that, from the historical records, the averaged natural

wind velocity, maximum wind velocity, gusts occurrences and direction were ideal for

testing with respect to this project. A ranking of 1 indicated the opposite. Most rankings

were from 5 and up, which indicated their percentage of the “ideal” event preferred when

testing for natural wind gusts.

The proposed method for application of the raking schedule for field testing was

as follows:

1. Determine the best possible day(s) for which natural wind testing is possible in

the ALDOT work schedule.

2. Look on the UAB ranking schedule associated with the required month of natural

wind testing.

3. Determine day(s) with the best ranking (preferably rank 8 and up) from the UAB

schedule
71

4. Develop a coincidence between the day(s) chosen from the ALDOT work sched-

ule and the day(s) selected from the UAB rank schedule (preferably choosing a

string of days with good rankings).

5. Check local weather updates and forecasts (approaching fronts, storms, etc…).

6. Chose optimal days from the information gathered from steps 1 through 5 to test.

Also included on the schedule was a moving average taken over a three day pe-

riod. This was provided to illustrate a smoother pattern of test day rankings. It helped the

ALDOT officer to choose a string of optimal testing days where the moving average was

high (preferably 8 and up). Even though the early months of 2009 during the projected

project schedule looked favorable, it was recommended to begin testing as soon as possi-

ble as these conditions may not actually exists in real time.

It was important to note that the UAB ranking schedule was only a forecast that

was based on averaged historical records. It showed a historical pattern of ideal wind oc-

currences in relation to the required needs of this project. Yet, the actual wind conditions

of the chosen testing time periods did to a large degree reflect the UAB ranking. There

were however scheduling delays and cancelations during the testing program due to pre-

cipitation.

Test Procedure

The test procedure for determining the fatigue load from natural wind gusts went

as follows. An appropriate testing day was scheduled between UAB and ALDOT with

the help of the ranking schedule. A testing day was typically scheduled at least two weeks
72

in advance. Even though natural wind was highest during storm events, scheduled days

were canceled for extreme weather as precipitation created large errors in the data collec-

tion. In addition, the existence of lightning created hazardous working conditions.

On the scheduled testing day, ALDOT drove the data acquisition van to the site

and began attaching wiring from the instrumentation to the van. This took typically one

to two hours. The testing began after instrumentation was operated and all transducers

were checked for accuracy. All instrumentation was measured simultaneously at 60 sam-

ples per second. Data recording intervals were set at no smaller than 45 minutes each,

therefore allowing for continuous 45 minute long recorded streamlines. Anemometers

AN-1 and AN-2 were used to measure the ambient wind velocity and direction during

testing. All strain gauges and accelerometers were used to measure the structural re-

sponse. Data was saved onto a CD for storage and distribution, and later used for deter-

mining the fatigue load due to natural wind back in the lab.

After data collection was finished for the scheduled event, all wiring was discon-

nected from the data acquisition system. Instrumentation connections were then stored

properly next to the structure for future testing, and ALDOT researchers left with the data

acquisition van. This procedure continued for each scheduled day until at least 32 hours

of usable data was measured.

Truck-Induced Wind Gust

The data collection event involving truck induced gusts was a staged occurrence.

Random truck events were also considered for possible truck measurements, but due to

the location of the sign, trucks were not able to reach a speed that would create signifi-
73

cant structural response. For this reason, a truck and driver was hired to run underneath

the sign structure at various speeds to measure the fatigue load due to truck induced

gusts. A picture of the truck and driver is shown in Figure 5.1. The height of the semi-

trailer was 13.5 ft (4.11 m) width equal to 8.5 ft (2.59 m), with a 53 ft (16.2 m) long box

trailer.

FIGURE 5.1 Truck and driver for truck-induced wind test.

Several truck types were considered for collection, based on geometric character-

istics of the vehicle:

 Truck cab only

 Truck cab with wind guard

 Cab + trailer

 Cab with wind guard+ trailer


74

 Dumpster truck (full)

 Dumpster truck (empty).

The selected truck (Figure 5.1) was representative of the majority of trucks using the

highway, and was applicable to a fatigue loading study.

Test Procedure

The test procedure for truck induced wind gust analysis was a cyclic event. The

truck drove loops underneath the sign structure a various speeds while the data acquisi-

tion system was recording structural responses from the instrumentation. A range of

speeds from 60 mph to 70 mph was used for the test. Three runs per speed were con-

ducted, totaling nine truck runs. Traffic counters were placed across the road underneath

the sign structure (see Figure 5.2) to record truck speed during each run, and to compare

to odometer readings for accurately identifying the truck speed. The traffic counters were

set to run parallel with the data acquisition system to have simultaneous time readings

during each truck run. They were JAMAR Trax Flex HS recorders with 0.365 in (9.271

mm) tubes spaced at 36 in (91.4 cm), and taped at 12 in (30.5 cm) intervals.

Due to obstacles at the site, ALDOT used traffic control for the testing to prevent

interference from cars during each run. Constant communication between the cab and the

ALDOT officer at the structure site was kept to indicate when the truck was approaching

so that the passing run could be marked on the data and later isolated in the lab for analy-

sis. Detailed instructions of the truck induced gust test procedure prepared for ALDOT

are provided in Appendix E.


75

FIGURE 5.2 JAMAR Trax Flex HS recorders.

The truck tests were scheduled on a low wind day so that the truck gust could be

isolated from the data. This typically was on days with natural wind velocities averaging

less than 10 mph (4.47 m/s). Structural excitation was created for greater natural wind

velocities, and therefore created unreadable truck gust impulses from the data collection

as natural wind generally controlled structural response. A picture of one of the truck

runs is shown in Figure 5.3. Anemometers AN-3 and AN-4 measured wind gust created

from the truck run, and all strain gauges and accelerometers were used to measure the

structural response. The data was later isolated in the lab to determine the fatigue load

due to truck induced wind gusts.


76

FIGURE 5.3 Typical truck-induced wind gust run.


77

CHAPTER 6

FATIGUE RESISTANCE

Chapter Overview

An analysis of the resistance of the support structure was first performed with the

collected experimental data. This helped to understand basic operational principles of the

structure including:

 Structural behavior,

 Distribution of stresses and strains at critical locations,

 Sections of the structure that was most vulnerable to fatigue, and

 Best areas of the structure to developed fatigue loads from the measured data.

Particular attention was concentrated on the connection details such as the anchor bolts,

column-to-base-plate fillet-welded connection, and the truss chord-to-column connection.

The stresses at these critical locations were compared to the constant-amplitude fatigue

thresholds specified in the Supports Specifications to evaluate the structure’s resistance to

service loading. Sections of the structure that were most vulnerable to service loading

were identified. Areas of the structure were identified as the best case for developing fa-

tigue loads from the collected experimental data.


78

Constant-Amplitude Fatigue Thresholds

The constant-amplitude fatigue threshold, also referred to as the endurance limit,

was a stress limitation used in fatigue analysis. The stress occurring at particular details

due to an applied fatigue load was compared to the limitations. The detail was considered

safe to significant fatigue damage if the stress was less than the endurance limit. Design

changes must be made such as material strength or shape changes if the stress is above

the endurance limit.

The endurance limit fatigue provisions provided by the Supports Specifications

were used for the analysis. They were based on stress versus number of cycles (S-N)

curves on a variety of connection details for support structures. The curves were devel-

oped from experimental testing in which a constant-amplitude load was applied to the

particular detail at repeated cycles until failure.

The specifications divided the variety of details into separate stress categories.

The constant-amplitude fatigue thresholds for the relevant stress categories for the tested

support structure in this project are listed in Table 6.1 with reference to the detail (1).

Failure Index

The failure index was calculated at the instrumented locations. This helped to

value the resistance of the structure to fatigue loading. The failure index was defined as

the ratio of the fatigue stress in the structure divided by the constant-amplitude fatigue

threshold for the particular connection detail (Eq. 6.1). The fatigue threshold was ex-

tracted from the Supports Specifications depending on the type of connection under

evaluation.
79

Fatigue Stress
Failure Index = ≤1 [Eq. 6.1]
Constant - Amplitude Fatigue Threshold

TABLE 6.1 Supports Specifications Constant-Amplitude Fatigue Thresholds (1)

Stress Steel Aluminum


Detail Detail Description
Category (ksi) (ksi)
Chord-to-
Net Section of fully-tightened
B Column 16 6
bolted connections
(Bolts)

Anchor bolts or other fasteners


D Anchor Bolts 7 2.5
in tension

Members with axial and bend-


Chord-to- ing loads with fillet-welded end
E Column connections without notches 4.5 1.9
(Weld) perpendicular to the applied
stress

Column-to- Fillet-welded tube-to-transverse


E' 2.6 1.0
Base-Plate plate connections

Evaluation of the Fatigue Stress in the Anchor Bolts

The fatigue stresses were calculated using the experimentally measured strain

values. Calculated three second peak-to-peak strain ranges were used for the natural wind

gust analysis, and the maximum peak-to-peak ranges measured during the truck tests

were used for the truck-induced gusts analysis. A detailed description on the methodol-

ogy in developing the three second peak-to-peak strain ranges and wind velocity averages
80

is provided in Chapter 8: Experimental Calculation of the Fatigue Load due to Natural

Wind Gust.

General Description of the Anchor Bolts

The support structure had eight 1.5 in (38.1 mm) diameter AASHTO M314-90

Gr. 55 anchor bolts (see Figure 6.1). All eight were instrumented with uni-axial strain

gauges oriented along the axial length of the bolt. A layout of the anchor bolts is shown

in Figure 6.2, and the labeling used for the analysis is shown in Figure 6.3.

FIGURE 6.1 Anchor bolt connection.


81

45°

22.5°

1.5 in DIA Anchor Bolts


30 in Bolt Circle

FIGURE 6.2 Anchor bolt layout.

Important to anchor bolt analysis is the clearance distance of the anchor bolt from

the bottom of the base plate to the foundation. This distance is generally between 1 in

(25.4 mm) and 3 in (76.2 mm) in practice. The clearance distance for the tested support

structure was measured in the field and is listed in Table 6.2 for each anchor bolt. General

structural analysis of anchor bolts involves purely axial stresses, but as the clearance

length of the anchor bolt increases, bending stress are distributed onto the bolt which are

not accounted for in the analysis. This in turn would create error in the structural analysis

results by the design engineer, and therefore fatigue related evaluations may be inaccu-

rate without availability of experimental values.


82

Direction of Traffic

Roadway
Front Face x
of Support
Structure y

AB-4 AB-5

AB-3 AB-6

AB-2 AB-7

AB-1 AB-8

FIGURE 6.3 Anchor bolt identification.


83

TABLE 6.2 Anchor Bolt Clearance Length

Anchor Bolt Length (in)


AB-1 3.375
AB-2 1.625
AB-3 1.125
AB-4 1.500
AB-5 1.125
AB-6 0.8125
AB-7 1.3125
AB-8 3.000

The experimentally measured strain values of the anchor bolts were used for this

analysis. Each anchor bolt (totaling eight) was evaluated. The stress range was deter-

mined and plotted versus wind velocity. A regression was performed on the plot to esti-

mate the stress range at the fatigue wind of 38 mph (17 m/s). The stress range at this

value was compared to the endurance limit of the particular detail (stress category D).

Fatigue Stress due to Natural Wind Gust

Strain ranges were determined from the experimentally collected natural wind

data on the anchor bolts. The maximum peak-to-peak strain range was calculated at three

second intervals. The simultaneously occurring three second wind velocity was also cal-

culated to juxtapose the anchor bolt strain ranges. The maximum strain ranges and wind

velocities were plotted and averaged together at 1 mph (0.447 m/s) intervals. This was

done for the total collected data measurements using all eight anchor bolts. The resulting

plot of the averaged maximum peak-to-peak strain range versus the three second wind

velocity at 1 mph (0.447 m/s) intervals is shown in Figure 6.4.


84

FIGURE 6.4 All anchor bolt microstrain range vs. wind velocity.

As seen in Figure 6.4, anchor bolts AB-8 and AB-1 demonstrated the largest

strain ranges during the experimental measurements. As seen in Table 6.2, these anchor

bolts had the largest clearance from the foundation to the base plate, and was believed to

be the reason for the large strain ranges measured experimentally. The wind data was di-

rected predominately on the front face of the structure (see Chapter 8); however anchor

bolts AB-8 and AB-1 show the largest ranges, followed by AB-2.

Based on the Supports Specifications, and the detailed fatigue loading analysis in

Chapter 8, the fatigue wind was determined to be 38 mph (17 m/s). The velocity was de-

termined as the wind velocity with a 0.01% exceedence probability from the annual mean

wind velocity equal to 11 mph (5 m/s) in accordance to the infinite-life approach to fa-

tigue design philosophy. The fatigue wind velocity was not measured during the experi-
85

mental tests, but wind velocities slightly lower than the fatigue wind were measured, and

therefore a regression analysis was performed on the data obtained in Figure 6.4 to esti-

mate the strain occurring at this wind. The estimation process was performed in similar

fashion to the process conducted in Chapter 8 for computing the wind pressure at the fa-

tigue wind.

By using the fundamental fluid mechanics relationship between pressure and the

velocity squared, a regression was performed on the experimental data to form a best fit

line that follows this relationship. The independent variable (x-axis, wind velocity) was

transformed into wind velocity squared and plotted with the microstrain ranges to “lin-

earize” the data. A linear best fit line was produced from the transformed data plot and

recorded (see Figure 6.5). The independent variable was transformed back into its origi-

nal form and the trendline equation was used to estimate the microstrain range at the fa-

tigue wind. The resulting plot is shown in Figure 6.6 for AB-8, the anchor bolt with the

largest recorded microstrain ranges. At the fatigue wind of 38 mph (17 m/s), the strain

range was estimated at 242.065(10-6), or 242.065 microstrain. The estimated strain range

result for all anchor bolts is listed in Table 6.3, and is plotted in a bar graph (Figure 6.7)

to better illustrate the variability in microstrain ranges.


86

FIGURE 6.5 Transformed independent variable for AB-8.

FIGURE 6.6 Estimated strain range at fatigue wind for AB-8.


87

TABLE 6.3 Estimated Microstrain Range at Fatigue Wind for Anchor Bolts

Anchor Bolt Microstrain Range (in/in)


AB-1 183.215
AB-2 107.129
AB-3 65.210
AB-4 103.939
AB-5 85.721
AB-6 78.640
AB-7 93.031
AB-8 242.065

FIGURE 6.7 Estimated microstrain range at fatigue wind for anchor bolts.

The stress range at the fatigue wind was determined from the estimated mi-

crostrain ranges, and compared to the constant-amplitude fatigue thresholds for the an-

chor bolt stress category D. Stress category D is for mechanically fastened connections

such as “anchor bolts or other fasteners in tension” (1). The stress range was calculated
88

for each of the eight anchor bolts tested. It is important to point out that if the fatigue

stress range was above the threshold, the design did not comply with the Supports Speci-

fications fatigue provisions with respect to the limits placed on the fatigue stress ranges

for anchor bolts. This is true regardless if the support structure was designed for fatigue

using the in-place fatigue design equation for natural wind. In such cases, believed to be

due to the large clearance between the foundation and the base plate, an additional bend-

ing stress was developed in the bolt because of an increased influence of out-of-plane

forces at the base plate connection.

The fatigue stress ranges calculated from the experimental microstrain ranges es-

timated at the fatigue wind is listed in Table 6.4. The values are listed with the constant-

amplitude fatigue thresholds of steel anchor bolts for the stress category D.

TABLE 6.4 Anchor Bolt Stress Range and Failure Index for Natural Wind

Fatigue Mi- Fatigue Stress Constant-


Anchor
crostrain Range Range Amplitude Fatigue Failure Index
Bolt
(in/in) (ksi) Threshold (ksi)
AB-1 183.215 5.31 7.00 0.76
AB-2 107.129 3.11 7.00 0.44
AB-3 65.21 1.89 7.00 0.27
AB-4 103.939 3.01 7.00 0.43
AB-5 85.721 2.49 7.00 0.36
AB-6 78.64 2.28 7.00 0.33
AB-7 93.031 2.70 7.00 0.39
AB-8 242.065 7.02 7.00 1.00

The failure index is also provided in Table 6.4, which was equal to the stress

range induced onto the anchor bolt at the fatigue wind divided by the constant-amplitude

fatigue threshold for the category detail (stress category D). A failure index greater than
89

one would indicate a fatigue stress range occurring greater than the allowable threshold.

As listed in the table, the failure index for each anchor bolt was less than one, with the

exception of AB-8 which was equal to one and was border line in the threshold design

check. A visual on the failure index variation with anchor bolt is provided in Figure 6.8.

FIGURE 6.8 Failure index of anchor bolts for natural wind.

Fatigue Stress due to Truck-Induced Wind Gust

The maximum peak-to-peak strain ranges were determined from the experimen-

tally collected data of the truck tests. Each of the eight anchor bolts was evaluated. A de-

tailing of the maximum peak-to-peak strain range procedure is provided in Chapter 10:

Experimental Calculation of the Fatigue Load due to Truck-Induced Gusts. As described

in the chapter, only four of the nine truck tests were usable on account of prevailing natu-
90

ral wind velocities. The truck runs utilized in this analysis was measured on low wind

days. A plot of the truck test strain results is provided in Figure 6.9, showing the relation-

ship between the truck speed and the resulting maximum peak-to-peak strain range in a

depth chart. The microstrain ranges per anchor bolt as measured from the truck tests are

listed in Table 6.5. The maximum range was recorded on AB-7 of truck test Truck 4

equaling 175.713 microstrain, when the truck was traveling at the highest tested speed of

70 mph (31.3 m/s). As seen in the figure, the majority of the strain occurred at the higher

truck speeds.

FIGURE 6.9 Depth chart of anchor bolt strain range in truck tests.
91

TABLE 6.5 Anchor Bolt Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Anchor Bolt Truck Run
(mph) (in/in)
Truck 1 60 34.292
Truck 2 65 32.968
AB-1
Truck 3 70 35.504
Truck 4 70 32.078
Truck 1 60 43.307
Truck 2 65 38.497
AB-2
Truck 3 70 43.268
Truck 4 70 43.685
Truck 1 60 17.606
Truck 2 65 16.168
AB-3
Truck 3 70 18.257
Truck 4 70 18.366
Truck 1 60 32.451
Truck 2 65 24.530
AB-4
Truck 3 70 20.777
Truck 4 70 15.193
Truck 1 60 26.517
Truck 2 65 18.992
AB-5
Truck 3 70 15.995
Truck 4 70 11.874
Truck 1 60 12.060
Truck 2 65 10.796
AB-6
Truck 3 70 10.611
Truck 4 70 13.328
Truck 1 60 57.603
Truck 2 65 55.234
AB-7
Truck 3 70 59.432
Truck 4 70 175.713
Truck 1 60 49.362
Truck 2 65 47.470
AB-8
Truck 3 70 48.529
Truck 4 70 49.843

The maximum peak-to-peak strain range measured from all truck tests was used

to calculate the stress range and determine the failure index for the truck tests. The results

are listed in Table 6.6 and illustrated in the bar chart of Figure 6.10. The results indicate
92

no significant effect from truck gusts on the anchor bolts with respect to the fatigue

threshold of the detail, with the exception of AB-7 for Truck 4.

TABLE 6.6 Anchor Bolt Stress Range and Failure Index for Truck Tests

Fatigue Mi- Fatigue Stress Constant-


Anchor
crostrain Range Range Amplitude Fatigue Failure Index
Bolt
(in/in) (ksi) Threshold (ksi)
AB-1 35.504 1.030 7 0.147
AB-2 43.685 1.267 7 0.181
AB-3 18.366 0.533 7 0.076
AB-4 32.451 0.941 7 0.134
AB-5 26.517 0.769 7 0.110
AB-6 13.328 0.387 7 0.055
AB-7 175.713 5.096 7 0.728
AB-8 49.843 1.445 7 0.206

FIGURE 6.10 Failure index of anchor bolts for truck tests.


93

Evaluation of the Fatigue Stress in the Post-to-Base-Plate Connection

General Description of Post-to-Base-Plate Connection

The column-to-base-plate fillet-welded connection is shown in Figure 6.11. The

post support was instrumented at 12 in (305 mm) at Section AA and 16 in (406 mm) at

Section BB. Only Section AA was evaluated in this chapter because it was closer to the

weld. A total of eight strain gauges at each section was placed, which included 4 rosettes.

A general layout of the gauges is shown in Figure 6.12, and the labeling methodology

shown in Figure 6.13. Each strain gauge on Section AA was evaluated to provide an es-

timation of the fatigue resistance of the column-to-base plate fillet-weld specified as

stress category E' in the Supports Specifications (see Table 6.1).

FIGURE 6.11 Fillet-welded tube-to-transverse plate connection.


94

Section AA Section BB

Uni-axial Strain Gauge

Rosette Strain Gauge

4 in Section BB

Section AA
12 in

Base Plate Weld

FIGURE 6.12 Post support strain gauges.

Strain Gauge

FIGURE 6.13 Strain gauge locations at Section AA.


95

The evaluation was performed in the same fashion as the anchor bolts. The stress

range was determined for natural wind gusts and plotted versus wind velocity intervals

for natural wind gusts. A regression was performed on the plot to estimate the stress

range at the fatigue wind of 38 mph (17 m/s). The maximum stress ranges were deter-

mined for the truck-induced gusts. The stress ranges were compared to the endurance

limit of the particular detail (stress category E') and the failure index was calculated.

Fatigue Stress due to Natural Wind Gust

Strain ranges were determined from the experimentally collected natural wind

data on the post support. The maximum peak-to-peak strain range was calculated at three

second intervals. The simultaneously occurring three second wind velocity was also cal-

culated to juxtapose the strain ranges. The maximum strain ranges and wind velocities

were plotted and averaged together at 1 mph (0.447 m/s) intervals. This was done for all

measured data of the strain gauges on the post support. A total of 12 readings were made:

8 normal stress values, and 4 shear stress values. The shear stress values were determined

from the rosette strain gauges. The gauges were rectangular strain rosettes arranged in a

45° pattern as shown in Figure 6.14. The strain transformation equation, Eq. 6.2, was

used to calculate the experimental shear strain measurement from the rosettes.

ε a = ε x cos2 (θ a ) + ε y sin 2 (θ a ) + γ xy sin(θ a ) cos(θ a )


ε b = ε x cos2 (θb ) + ε y sin 2 (θb ) + γ xy sin(θb ) cos(θb ) [Eq. 6.2]
ε c = ε x cos2 (θc ) + ε y sin 2 (θc ) + γ xy sin(θc ) cos(θc )

With θa equal to -135°, θb equal to -90°, and θc equal to -45°, Eq. 6.2 becomes:
96

ε x = εa − εb + εc
ε y = εb [Eq. 6.3]
γ xy = ε a − ε c

Equation 6.3 was used to calculate the experimental peak-to-peak three second shear

strain ranges.

FIGURE 6.14 45° rosette arrangement with coordinate axis.

The resulting plots of the averaged maximum peak-to-peak microstrain range ver-

sus the three second wind velocity at 1 mph (0.447 m/s) intervals is shown in Figure 6.15

for the uni-axial normal strain values and in Figure 6.16 for the rosette shear strain val-

ues.
97

FIGURE 6.15 Normal microstrain range vs. wind velocity.

As seen in Figure 6.15, strain gauges SGR-AA-11 at location E and SG-AA-5 at

location B demonstrated the largest normal strain ranges during the experimental meas-

urements. Since the loading was directed on the front face of the structure, this outcome

seems reasonable because of the resulting moment about the x-axis. Rosette gauges R2 at

location C and R4 at location G demonstrated the largest shear strain ranges (Figure

6.16). This is because of the torsion created about the post, in addition to the shear cre-

ated by the wind force directed onto the front face of the structure.
98

FIGURE 6.16 Rosette microstrain range vs. wind velocity.

The fatigue wind was determined to be 38 mph (17 m/s) in Chapter 8: Experimen-

tal Calculations of the Fatigue Load due to Natural Wind Gust. It was found as the wind

velocity with a 0.01% exceedence probability from an annual mean wind velocity equal

to 11 mph (5 m/s) in accordance to the infinite-life approach to fatigue design philoso-

phy. This wind velocity was not measured during the experimental tests, and therefore a

regression analysis was performed on the data obtained in Figure 6.15 and Figure 6.16.

The estimation process was performed in similar fashion to the process conducted in the

anchor bolt section, as well as a detailed description provided in Chapter 8 for computing

the fatigue wind pressure at the fatigue wind.

By using the fundamental fluid mechanics relationship between pressure and the

velocity squared, a regression was performed on the experimental data to form a best fit
99

line that followed this relationship. The independent variable (x-axis, wind velocity) was

transformed into wind velocity squared and plotted with the microstrain ranges to “lin-

earize” the data. A linear best fit line was produced from the transformed data plot and

recorded (see Figure 6.17). The independent variable was transformed back into its origi-

nal form and the trendline equation was used to estimate the microstrain range at the fa-

tigue wind. The resulting plot is shown in Figure 6.18 for SGR-AA-11, one of the strain

gauges with the largest recorded microstrain ranges as an example. At the fatigue wind of

38 mph (17 m/s), the strain range was estimated at 46.854 (10-6), or 46.854 microstrain.

The estimation strain range result for all strain gauges is listed in Table 6.7, along with a

bar chart (Figure 6.19) to demonstrate the variability in strain values as it relates to the

circumference of the post members.

FIGURE 6.17 Transformed independent variable for SGR-AA-11.


100

FIGURE 6.18 Estimated strain range at fatigue wind for SGR-AA-11.

TABLE 6.7 Estimated Microstrain Range at Fatigue Wind for Section AA

Microstrain Range
Section AA
(in/in)
SGR-AA-3 41.1571
SG-AA-5 45.921
SGR-AA-7 20.9966
Uni-axial SG-AA-9 36.6934
(Normal Strain) SGR-AA-11 46.8539
SG-AA-13 31.621
SGR-AA-15 20.921
SG-AA-1 30.4554
R1 41.0425
Rosette R2 52.5839
(Shear Strain) R3 42.9721
R4 47.0807
101

FIGURE 6.19 Estimated microstrain range at fatigue wind for Section AA.

The stress range at the fatigue wind was determined from the estimated mi-

crostrain ranges in Table 6.7, and compared to the constant-amplitude fatigue thresholds

for the anchor bolt stress category E'. This was done for each of the strain gauges listed in

the table. If the fatigue stress range was above the threshold, the design did not comply

with the Supports Specifications fatigue provisions with respect to the limits placed on

the fatigue stress ranges for column-to-base-plate fillet-welded connection. This is true

regardless if the support structure was designed for fatigue using the in-place fatigue de-

sign equation for natural wind.

The fatigue stress ranges calculated from the experimental microstrain ranges es-

timated at the fatigue wind is listed in Table 6.8. The values are listed with the constant-

amplitude fatigue thresholds of steel anchor bolts for the stress category E'.
102

TABLE 6.8 Section AA Stress Range and Failure Index for Natural Wind.

Fatigue Mi-
Fatigue Stress Constant-
crostrain
Section AA Range Amplitude Fatigue Failure Index
Range
(ksi) Threshold (ksi)
(in/in)
SGR-AA-3 41.1571 1.194 2.6 0.459
SG-AA-5 45.921 1.332 2.6 0.512
SGR-AA-7 20.9966 0.609 2.6 0.234
SG-AA-9 36.6934 1.064 2.6 0.409
SGR-AA-11 46.8539 1.359 2.6 0.523
SG-AA-13 31.621 0.917 2.6 0.353
SGR-AA-15 20.921 0.607 2.6 0.233
SG-AA-1 30.4554 0.883 2.6 0.340
R1 41.0425 1.190 2.6 0.458
R2 52.5839 1.525 2.6 0.587
R3 42.9721 1.246 2.6 0.479
R4 47.0807 1.365 2.6 0.525

The failure index is also provided in Table 6.8, and is shown graphically in Figure

6.20. A failure index greater than one would indicate a fatigue stress range occurring

greater than the allowable threshold. As listed in the table, the failure index for each

gauge was less than one and subsequently passes the threshold design check. This was

only an estimation of the strain at the base-plate connection as the strain gauges were lo-

cated 12 in (305 mm) above the connection. The failure indexes were approximately

50%, and therefore the passing assumption was justified.


103

FIGURE 6.20 Failure indexes for natural wind of Section AA.

Fatigue Stress due to Truck-Induced Wind Gusts

The maximum peak-to-peak strain ranges were determined from the experimen-

tally collected data from the truck tests. Each of the strain gauges was evaluated. The mi-

crostrain ranges per gauge as measured from the truck tests are listed in Table 6.9 and

Table 6.10. The maximum normal strain range was recorded on SGR-AA-11 of truck test

Truck 4 equaling 12.087 microstrain, when the truck was traveling at 70 mph (31.3 m/s).

The maximum shear strain range was recorded on R2 of 13.337 microstrain for the same

truck test. A depth chart showing the induced strain range with respect to truck speed is

shown in Figure 6.21 for the normal strain values measured by the uni-axial strain

gauges, and Figure 6.22 for the shear strain values measured by the rosette strain gauges.
104

TABLE 6.9 Section AA (Normal) Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Section AA Truck Run
(mph) (in/in)
Truck 1 60 11.371
Truck 2 65 10.438
SGR-AA-3
Truck 3 70 10.510
Truck 4 70 11.085
Truck 1 60 10.636
Truck 2 65 9.324
SG-AA-5
Truck 3 70 10.348
Truck 4 70 9.515
Truck 1 60 8.803
Truck 2 65 7.694
SGR-AA-7
Truck 3 70 6.325
Truck 4 70 4.413
Truck 1 60 10.140
Truck 2 65 8.645
SG-AA-9
Truck 3 70 8.293
Truck 4 70 8.757
Truck 1 60 11.640
Truck 2 65 10.915
SGR-AA-11
Truck 3 70 11.081
Truck 4 70 12.087
Truck 1 60 7.821
Truck 2 65 6.158
SG-AA-13
Truck 3 70 5.938
Truck 4 70 5.669
Truck 1 60 8.040
Truck 2 65 6.648
SGR-AA-15
Truck 3 70 6.070
Truck 4 70 4.289
Truck 1 60 7.400
Truck 2 65 6.990
SG-AA-1
Truck 3 70 7.365
Truck 4 70 6.995
105

TABLE 6.10 Section AA (Shear) Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Section AA Truck Run
(mph) (in/in)
Truck 1 60 10.298
Truck 2 65 10.162
R1
Truck 3 70 10.418
Truck 4 70 10.660
Truck 1 60 12.938
Truck 2 65 12.410
R2
Truck 3 70 12.917
Truck 4 70 13.337
Truck 1 60 9.947
Truck 2 65 9.844
R3
Truck 3 70 10.545
Truck 4 70 9.842
Truck 1 60 11.571
Truck 2 65 11.551
R4
Truck 3 70 12.306
Truck 4 70 12.155

FIGURE 6.21 Depth chart of Section AA normal strain in truck tests.


106

FIGURE 6.22 Depth chart of Section AA shear strain in truck tests.

The maximum range measured for each gauge was used to calculate the stress

range to determine the failure index for the truck tests. The results are listed in Table

6.11. The results indicate no significant effect from truck gusts on the column-to-base-

plate fillet-welded connection with respect to the fatigue threshold of the detail (Figure

6.23). The column was large enough in size to withstand the truck-induced wind loading.

Aluminum material could be used in place of steel provided that the natural wind fatigue

stress ranges were lower than 1 ksi (7 MPa) which was not the case (see Table 6.8)
107

TABLE 6.11 Section AA Stress Range and Failure Index for Truck Tests

Fatigue Mi-
Fatigue Stress Constant-
crostrain
Section AA Range Amplitude Fatigue Failure Index
Range
(ksi) Threshold (ksi)
(in/in)
SGR-AA-3 11.371 0.330 2.6 0.127
SG-AA-5 10.636 0.308 2.6 0.119
SGR-AA-7 8.803 0.255 2.6 0.098
SG-AA-9 10.140 0.294 2.6 0.113
SGR-AA-11 12.087 0.351 2.6 0.135
SG-AA-13 7.821 0.227 2.6 0.087
SGR-AA-15 8.040 0.233 2.6 0.090
SG-AA-1 7.400 0.215 2.6 0.083
R1 10.660 0.309 2.6 0.119
R2 13.337 0.387 2.6 0.149
R3 10.545 0.306 2.6 0.118
R4 12.306 0.357 2.6 0.137

FIGURE 6.23 Failure index for Section AA truck tests.


108

Evaluation of the Fatigue Stress in the Chords

General Description of the Chords

Two areas of concern were evaluated with the chord-to-column connection. The

first concern was analyzing the chord-to-plate weld (Figure 6.24). The next concern was

the net section of the bolted connection (Figure 6.25). The failure index in regards to the

constant-amplitude fatigue thresholds for the connection detail was checked for both. The

connection involved four chords. Each chord was strain gauged with uni-axial gauges

oriented along the axial length of the chord. A schematic of the instrumentation layout is

shown in Figure 6.26. Each chord was labeled one through four (Figure 6.27), and the

individual strain gauges were labeled per chord as shown in Figure 6.28 of a cross section

of the chords as seen from the point of view of the sign.

Chord-to-Plate
Weld

FIGURE 6.24 Chord-to-plate weld.


109

FIGURE 6.25 Chord-to-column bolted connection.


110

Placed on each
Chord at 90º
around Chord
Circumference
Centerline
of 1st Truss
Panel

3’- 4.5”

FIGURE 6.26 Chord instrumentation layout.

Chord 3

Chord 1

Chord 4

Chord 2

FIGURE 6.27 Chord labeling.


111

Traffic

SG-C1-10
SG-C1-2

SG-C1-1
SG-C1-3 Chord 1 Chord 3 SG-C1-9

SG-C1-11

SG-C1-4 SG-C1-12
z
y
SG-C1-6 SG-C1-16

SG-C1-15

SG-C1-8 Chord 2 Chord 4 SG-C1-13

SG-C1-5

SG-C1-7 SG-C1-14

Uni-Axial Strain Gauge

SG-C1-4 = Gauge malfunction

FIGURE 6.28 Chord strain gauge labeling.

The evaluation was performed in the same fashion as the anchor bolts and Section

AA. The stress range was determined for natural wind gusts and plotted versus wind ve-

locity intervals for natural wind gusts. A regression was performed on the plot to estimate
112

the stress range at the fatigue wind of 38 mph (17 m/s). The maximum stress ranges were

determined for the truck-induced gusts. The stress ranges were compared to the endur-

ance limit of the particular detail and the failure index was calculated.

Fatigue Stress due to Natural Wind Gust

Strain ranges were determined from the experimentally collected natural wind

data on the chords. The maximum peak-to-peak strain range was calculated at three sec-

ond intervals. The simultaneously occurring three second wind velocity was also calcu-

lated to juxtapose the chord strain ranges. The maximum strain ranges and wind veloci-

ties were plotted and averaged together at 1 mph (0.447 m/s) intervals. This was per-

formed for all measured data of the strain gauges on the chords. A total of 13 readings

were made: 4 gauges per chord, with 3 gauges not responding during testing (see Figure

6.28). The resulting plots of the averaged maximum peak-to-peak microstrain range ver-

sus the three second wind velocity at 1 mph (0.447 m/s) intervals are shown in Figure

6.29.

As seen in Figure 6.29, strain gauge SG-C2-5 located on the inside of Chord 2

demonstrated the largest strain ranges during the experimental measurements. Most of the

gauges showed similar results. Since the loading was directed on the front face of the

structure, this outcome appeared reasonable. The separation at the higher wind velocities

was due to a lessoned amount of collected data for the velocity intervals.
113

FIGURE 6.29 Chord microstrain ranges vs. wind velocity.

The fatigue wind was determined to be 38 mph (17 m/s) in Chapter 8: Experimen-

tal Calculations of the Fatigue Load due to Natural Wind Gust. It was found as the wind

velocity with a 0.01% exceedence probability from an annual mean wind velocity equal

to 11 mph (5 m/s) in accordance to the infinite-life approach to fatigue design philoso-

phy. This wind velocity was not measured during the experimental tests, and therefore a

regression analysis was performed on the data obtained in Figure 6.29. The estimation

process was performed in similar fashion to the process conducted in Chapter 8 for com-

puting the fatigue wind pressure at the fatigue wind.

By using the fundamental fluid mechanics relationship between pressure and the

velocity squared, a regression was performed on the experimental data to form a best fit

line that followed this relationship. The independent variable (x-axis, wind velocity) was
114

transformed into wind velocity squared and plotted with the microstrain ranges to “lin-

earize” the data. A linear best fit line was produced from the transformed data plot and

recorded (see Figure 6.30). The independent variable was transformed back into its origi-

nal form and the trendline equation was used to estimate the microstrain range at the fa-

tigue wind. The resulting plot is shown in Figure 6.31 for SG-C2-5, the strain gauge with

the largest recorded microstrain ranges as an example. At the fatigue wind of 38 mph (17

m/s), the strain range was estimated at 46.4367(10-6), or 46.4367 microstrain. The estima-

tion strain range result for all strain gauges is listed in Table 6.12. A bar graph is pro-

vided in Figure 6.32 to illustrate the variability of the strain ranges measured by the

gauges.

FIGURE 6.30 Transformed independent variable for SG-C2-5.


115

FIGURE 6.31 Estimated strain range at fatigue wind for SG-C2-5.

TABLE 6.12 Estimated Microstrain Range at Fatigue Wind for Chords

Microstrain Range
Chord Chord Strain Gauge
(in/in)
SG-C1-1 33.3139
1 SG-C1-2 36.5662
SG-C1-3 34.407
SG-C2-5 46.4367
2 SG-C2-6 40.8636
SG-C2-8 39.6164
SG-C3-10 35.5541
3 SG-C3-11 39.496
SG-C3-12 33.5019
SG-C4-13 37.6525
SG-C4-14 39.1607
4
SG-C4-15 41.6753
SG-C4-16 39.7908
116

FIGURE 6.32 Estimated microstrain range for fatigue wind of chords.

The stress range at the fatigue wind was determined from the estimated and the

failure indexes were calculated. Two areas of concern were evaluated:

1. The weld of the chords to the structural plate. The weld ran parallel to the normal

strain in the chord.

2. The bolted connection of the structural plate to the column.

For Concern 1, the stress category E was used because of the welded connection with ax-

ial and bending loads applied parallel to the weld. Stress category B was used for Con-

cern 2 because of the bolted joints. Although the strain gauges were not located directly

onto these locations, it was felt that an estimate could be made of its fatigue resistance
117

criteria. The resulting failure indexes of each strain gauge for Concern 1 and Concern 2 is

listed in Table 6.13 and Table 6.14, respectively. A plot showing their relationship to

each other and the potential impact of the design to fatigue resistance concerning the two

areas is provided in Figure 6.33. The failure indexes in Table 6.13 and Table 6.14 were

approximately 25%, proving that the fatigue stress in the chord-to-column connection

was satisfactory with regards to the fatigue thresholds.

TABLE 6.13 Chord Stress Range and Category E Failure Index for Natural Wind

Fatigue Mi-
Chord Fatigue Stress Constant-
crostrain
Strain Range Amplitude Fatigue Failure Index
Range
Gauge (ksi) Threshold (ksi)
(in/in)
SG-C1-1 33.3139 0.966 4.5 0.215
SG-C1-2 36.5662 1.060 4.5 0.236
SG-C1-3 34.407 0.998 4.5 0.222
SG-C2-5 46.4367 1.347 4.5 0.299
SG-C2-6 40.8636 1.185 4.5 0.263
SG-C2-8 39.6164 1.149 4.5 0.255
SG-C3-10 35.5541 1.031 4.5 0.229
SG-C3-11 39.496 1.145 4.5 0.255
SG-C3-12 33.5019 0.972 4.5 0.216
SG-C4-13 37.6525 1.092 4.5 0.243
SG-C4-14 39.1607 1.136 4.5 0.252
SG-C4-15 41.6753 1.209 4.5 0.269
SG-C4-16 39.7908 1.154 4.5 0.256
118

TABLE 6.14 Chord Stress Range and Category B Failure Index for Natural Wind

Fatigue Mi-
Chord Fatigue Stress Constant-
crostrain
Strain Range Amplitude Fatigue Failure Index
Range
Gauge (ksi) Threshold (ksi)
(in/in)
SG-C1-1 33.3139 0.966 16 0.060
SG-C1-2 36.5662 1.060 16 0.066
SG-C1-3 34.407 0.998 16 0.062
SG-C2-5 46.4367 1.347 16 0.084
SG-C2-6 40.8636 1.185 16 0.074
SG-C2-8 39.6164 1.149 16 0.072
SG-C3-10 35.5541 1.031 16 0.064
SG-C3-11 39.496 1.145 16 0.072
SG-C3-12 33.5019 0.972 16 0.061
SG-C4-13 37.6525 1.092 16 0.068
SG-C4-14 39.1607 1.136 16 0.071
SG-C4-15 41.6753 1.209 16 0.076
SG-C4-16 39.7908 1.154 16 0.072

FIGURE 6.33 Failure indexes of chord gauges for natural wind.


119

Fatigue Stress due to Truck-Induced Wind Gust

The maximum peak-to-peak strain ranges were determined from the experimen-

tally collected data from the truck tests. Each of the strain gauges was evaluated. The mi-

crostrain ranges per gauge as measured from the truck tests are listed in Table 6.15 – 6.19

for each chord respectively. The maximum strain range was recorded on SG-C2-5 of

truck test Truck 3 equaling 12.075 microstrain, where the truck was traveling at 70 mph

(31.3 m/s). A plot depicting the distribution of strain with respect to the truck speed is

provided in Figure 6.34.

TABLE 6.15 Chord 1 Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Strain Gauge Truck Run
(mph) (in/in)
Truck 1 60 7.646
Truck 2 65 7.668
SG-C1-1
Truck 3 70 7.665
Truck 4 70 8.161
Truck 1 60 8.483
Truck 2 65 8.417
SG-C1-2
Truck 3 70 8.288
Truck 4 70 8.802
Truck 1 60 8.196
Truck 2 65 7.871
SG-C1-3
Truck 3 70 8.282
Truck 4 70 8.729
120

TABLE 6.16 Chord 2 Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Strain Gauge Truck Run
(mph) (in/in)
Truck 1 60 11.500
Truck 2 65 11.680
SG-C2-5
Truck 3 70 12.075
Truck 4 70 10.870
Truck 1 60 10.150
Truck 2 65 10.229
SG-C2-6
Truck 3 70 10.505
Truck 4 70 9.724
Truck 1 60 9.912
Truck 2 65 10.018
SG-C2-8
Truck 3 70 10.433
Truck 4 70 9.728

TABLE 6.17 Chord 3 Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Strain Gauge Truck Run
(mph) (in/in)
Truck 1 60 8.379
Truck 2 65 8.150
SG-C3-10
Truck 3 70 8.572
Truck 4 70 8.027
Truck 1 60 9.402
Truck 2 65 9.658
SG-C3-11
Truck 3 70 9.542
Truck 4 70 9.201
Truck 1 60 7.756
Truck 2 65 7.902
SG-C3-12
Truck 3 70 8.016
Truck 4 70 7.517
121

TABLE 6.18 Chord 4 Microstrain Ranges from the Truck Tests

Truck Speed Microstrain Range


Strain Gauge Truck Run
(mph) (in/in)
Truck 1 60 8.274
Truck 2 65 8.429
SG-C4-13
Truck 3 70 8.511
Truck 4 70 8.828
Truck 1 60 8.596
Truck 2 65 8.748
SG-C4-14
Truck 3 70 8.779
Truck 4 70 8.755
Truck 1 60 9.228
Truck 2 65 9.247
SG-C4-15
Truck 3 70 9.123
Truck 4 70 9.690
Truck 1 60 8.897
Truck 2 65 9.022
SG-C4-16
Truck 3 70 8.845
Truck 4 70 8.915

FIGURE 6.34 Depth chart of chord strain in truck tests.


122

The maximum range measured for each gauge was used to calculate the stress

range to determine the failure index for the truck tests. The results are listed in Table 6.19

for Concern 1 (stress category E) and Table 6.20 for Concern 2 (stress category B). The

results indicate no significant effect from truck gusts on the chord connection (welded

and bolted) to the column.

TABLE 6.19 Chord Stress Range and Category E Failure Index for Truck Tests

Chord Fatigue Mi- Fatigue Stress Constant-


Strain crostrain Range Range Amplitude Fatigue Failure Index
Gauge (in/in) (ksi) Threshold (ksi)
SG-C1-1 8.161 0.237 4.5 0.053
SG-C1-2 8.802 0.255 4.5 0.057
SG-C1-3 8.729 0.253 4.5 0.056
SG-C2-5 12.075 0.350 4.5 0.078
SG-C2-6 10.505 0.305 4.5 0.068
SG-C2-8 10.433 0.303 4.5 0.067
SG-C3-10 8.572 0.249 4.5 0.055
SG-C3-11 9.658 0.280 4.5 0.062
SG-C3-12 8.016 0.232 4.5 0.052
SG-C4-13 8.828 0.256 4.5 0.057
SG-C4-14 8.779 0.255 4.5 0.057
SG-C4-15 9.690 0.281 4.5 0.062
SG-C4-16 9.022 0.262 4.5 0.058
123

TABLE 6.20 Chord Stress Range and Category B Failure Index for Truck Tests

Chord Fatigue Mi- Fatigue Stress Constant-


Strain crostrain Range Range Amplitude Fatigue Failure Index
Gauge (in/in) (ksi) Threshold (ksi)
SG-C1-1 8.161 0.237 16 0.015
SG-C1-2 8.802 0.255 16 0.016
SG-C1-3 8.729 0.253 16 0.016
SG-C2-5 12.075 0.350 16 0.022
SG-C2-6 10.505 0.305 16 0.019
SG-C2-8 10.433 0.303 16 0.019
SG-C3-10 8.572 0.249 16 0.016
SG-C3-11 9.658 0.280 16 0.018
SG-C3-12 8.016 0.232 16 0.015
SG-C4-13 8.828 0.256 16 0.016
SG-C4-14 8.779 0.255 16 0.016
SG-C4-15 9.690 0.281 16 0.018
SG-C4-16 9.022 0.262 16 0.016

Discussion of Results

An analysis of the fatigue resistance of the structure was performed. The stress

ranges determined in the structure were compared to the constant-amplitude threshold

provided by the Supports Specifications. A failure index was calculated to provide a rela-

tionship between the gauged locations with a control value. The distribution of the stress

in the structure was evaluated and the most appropriate location for determining the fa-

tigue load was identified.

The largest strain recorded was located on the anchor bolts followed by the post

and chords as expected. The distribution of stress followed a clear and distinct path to the

foundation. The major findings are listed in Table 6.21 and 6.22 for natural wind and

truck-induced gusts. The table lists the stress ranges determined at the average wind ve-
124

locity and the fatigue wind velocity. Maximum ranges at the specified location are listed

for each wind velocity, as well as the average stress range from all gauges at the section.

TABLE 6.21 Experimental Stress Ranges for Natural Wind Tests

Location Stress Range (ksi)


on the Measurement Type Wind Velocity (mph)
Average Maximum
Structure
Average Wind 13 0.200 0.245
Chord Normal
Fatigue Wind 38 1.111 1.347
Average Wind 13 0.184 0.224
Normal
Fatigue Wind 38 1.089 1.359
Post
Average Wind 13 0.210 0.234
Shear
Fatigue Wind 38 1.332 1.525
Anchor Average Wind 13 0.608 0.967
Normal
Bolt Fatigue Wind 38 4.07 7.02

TABLE 6.22 Experimental Stress Ranges for Truck-Induced Wind Tests

Location on the Measurement Truck Speed Stress Range (ksi)


Structure Type (mph) Average Maximum
60 0.241 0.334
Chord Normal 65 0.242 0.339
70 0.253 0.350
60 0.244 0.338
Normal 65 0.215 0.317
70 0.220 0.351
Post
60 0.260 0.375
Shear 65 0.255 0.360
70 0.297 0.387
60 0.880305 1.670487
Anchor Bolt Normal 65 0.788333 1.601786
70 1.044773 5.095677

For the truck tests, the stress ranges on averaged increased with the speed of the

truck. No significant loading was generated from the truck gusts as compared to the natu-
125

ral wind gust measurements. The stress generated from natural wind was as much as four

times greater than truck-induced wind gusts. This is illustrated in the bar graph of Figure

6.35, showing the stress ranges for the fatigue wind as compared to the stress ranges for

the 70 mph (31.3 m/s) truck speed. A significant increase in stress range for both natural

wind and truck gusts was observed for the anchor bolts. This was due to the anchor bolts

with large clearances as compared to the other bolts, resulting in an unbalanced distribu-

tion of stress. The bolt with the largest stress ranges were located closest to the neutral

axis created with respect to the applied load.

FIGURE 6.35 Comparison of results between natural wind and truck gusts.

Significant conclusions were made after the fatigue behavior of the structural

members was evaluated. The results helped to direct the project as to which gauges to use
126

in calculated the fatigue loads due to natural wind and truck-induced wind gusts. The

depth charts helped to distinguish the areas of the structure that were most vulnerable to

fatigue loading. The analysis detailed the stress flow in the structure from the point of

loading to the connections. Large strain ranges were measured in the chords, post, and

anchor bolts, with the largest found in the anchor bolts. Strain values measured in the

chords were slightly less than the strains in the post indicating a smooth stress transfer

through the structure to the foundation.

The stress ranges in the column and chords were not significant enough to create

fatigue damage, however the stress ranges in the anchor bolts did show reasonable evi-

dence of possible fatigue failure. It was a fatigue failure of the anchor bolts that caused

the previous structure to fail due to a combined stress of axial and bending. The problem

areas with this project were associated to the bolts with clearances equal to 3 in (76.2

mm) or more. It was believed that if the clearance lengths were decreased, a better distri-

bution of the stress from the column to the foundation would occur, thereby reducing the

overloading of stress onto one anchor bolt. The stress distribution would better reflect the

structural analysis during the design phase of the support structure, and subsequently in-

crease the accuracy of the fatigue analysis estimation made by the engineer.

One aspect needed in calculation of the fatigue load was a lessoned variability of

the strain readings from one gauge to the next as was found with the anchor bolt meas-

urements. Although the distribution would not be the same because of the location to the

neutral axis from one gauge to the other, a preconceived pattern illustrating a balanced

state should be easily observed. Likewise, strain readings in the post were decided as the

most noteworthy from the other instrumented locations in regards to fatigue load calcula-
127

tions. The strains in the anchor bolts and chords varied significantly from the balanced

state, with extreme differences found in the anchor bolts. In view of the results, it was

concluded that the best location for calculating the fatigue loads from the measured data

was the column sections.


128

CHAPTER 7

EXPERIMENTAL MODAL ANALYSIS

Chapter Overview

A modal analysis of the structure was conducted from the experimental data. The

accelerometers measurements were used for the study. Important dynamic properties

were gathered from the analysis to be used in the development of the fatigue loads due to

natural wind and truck-induced wind gusts.

Modal Data Utilization

A modal analysis was conducted in order to understand the dynamic behavior of

the structural system. It was obtained through an experimental technique using measured

acceleration data obtained from operational dynamic forces. The major parameters that

were obtained from the modal analysis were as follows:

 Identify modal frequencies and periods of vibration,

 Determine modal damping,

 Obtain modal shapes,

 Improve and verify FEA analytical models, and

 Predict response to dynamic excitation


129

The investigation of these aspects helped to devise and develop the fatigue loading due to

natural wind and truck-induced wind gusts.

The data obtained from the accelerometers were used for the analysis. Although

the strain gauges could be used, the only parameters extracted from the analysis would be

the modal frequencies and modal damping values. The modal shapes cannot be deter-

mined from the strain gauge data.

Modal Analysis Test Setup

Accelerometers were used to obtain the acceleration response of the structure.

They were of the piezoelectric type that uses a piezoelectric crystal mounted to a small

mass from which the voltage output is converted to acceleration. Each accelerometer had

a maximum capacity of 96.5 ft/sec2 (3 G). The locations of the accelerometers were stra-

tegically chosen, as the accelerometer only gave the natural frequency of the member

upon which it rests. A combined total of six unidirectional accelerometers were required

for the structure. They were placed at particular locations to measure each possible de-

gree of freedom in vibration direction. Accelerometer locations are shown in Figure 7.1.
130

AC-1

AC-2

Centerline
Post

z
AC-3
y Centerline
Truss
x

FIGURE 7.1 Accelerometer locations for cantilever structure.

Three accelerometers were placed at location AC-1 (see Figure 7.1) to measure

the vertical (perpendicular to the direction of traffic), horizontal (parallel to the direction

of traffic), and longitudinal (transverse to the direction of traffic) accelerations of the post

member. Two accelerometers were placed at location AC-2 to measure the vertical and

horizontal directions (Figure 7.2). One accelerometer was placed at location AC-3 to

measure the horizontal direction for indentifying torsion behavior of the overhang truss

(Figure 7.3). An overview of locations AC-2 and AC-3 is shown in Figure 7.4.
131

FIGURE 7.2 Accelerometer location AC-2.

FIGURE 7.3 Accelerometer location AC-3.


132

FIGURE 7.4 Overview of accelerometers AC-2 and AC-3.

The accelerometers were fixed to the mounting surface by means of a two

threaded screws. Electrical insulation between the accelerometer and the test surface was

built into the transducer device. All mounting surfaces were flat to avoid distortion that

may produce strains which could affect the accelerometer’s response. Carefulness was

taken when screwing the accelerometers to the attachment surface as to not overreach the

torque recommended by the manufacturer (39).

Mounting blocks were used at each location for accelerometer attachment. For

instance at location AC-1 where three accelerometers were needed, all three were at-
133

tached to one single attachment block. A picture of this location with the attached mount-

ing block is shown in Figure 7.5. Once the truss section of the structure was erected, the

AC-1 accelerometers were screwed onto the mounting block with the sensitive ends

(measurement direction) oriented in the proper directions. The same type of block was

used at locations AC-2 and AC-3. Due to the round surface at these locations (steel pipe

truss web member), a small flat steel plate was welded to the side of the members for

which the mounting blocks and accelerometers were attached. The size of the plate was

manufactured as small as possible, with enough space to accurately and securely mount

the accelerometers and the plate to the web member, but without the possibility of creat-

ing significant additional wind drag. The initial mounting instructions prepared for AL-

DOT are given in Appendix C.

FIGURE 7.5 Mounting block at AC-1.


134

Systemizing the Degrees of Freedom

The number of degrees of freedom was chosen to represent the dynamic behavior

of the structural system. They were chosen based on the expected modal shapes of the

system. Five dynamic vibratory shapes were identified as crucial dynamic behavior that

governs the vibratory movement of the structure (51):

1. Torsion of the cantilevered truss and sign about the support shaft (moment about

z-axis)

2. Vertical vibration of the truss and shaft in the plane perpendicular to the ground

(moment about y-axis)

3. Horizontal vibration of the truss and shaft in the plane parallel to the ground

(moment about x-axis)

4. Longitudinal vibration of the truss and shaft along the axial length of the truss

cantilever (moment about y-axis)

5. Torsion vibration of the truss about the longitudinal axis (moment about x-axis)

The placements of the accelerometers were identified based on the five assumed vibra-

tory responses and are shown in Figure 7.1. The associated degrees of freedom and their

alignment were defined as follows:

 AC-1: 1x, 1y, 1z

 AC-2: 2y, 2z

 AC-3: 3y
135

Six degrees of freedom were indentified with the cantilever-type sign support structure of

this study.

Spectral Analysis

A spectral analysis was conducted first to obtain natural frequencies and their or-

der of appearance. These values were used in identifying and determining modal shapes.

Only the first four modes were discovered from the spectral plots. From the analysis of

the plots, the first two modes showed the primary means of vibration.

The data collected for each accelerometer was used in the analysis. The collected

time duration used was 10 minutes. In order for completeness and accurate representation

of the dynamic behavior of the structural system, the chosen event must excite all major

natural frequencies. The event was representative a random vibration occurrence, and as-

sumed as a continuous function. Hanning windows were used to ensure the signal begins

and ends at zero. The Fourier transform (shown as Eq. 7.1) was used to transform the

time domain into a frequency domain:

∫ x (t ) e
− j 2πft
X(f) = dt [Eq. 7.1]
−∞

where
X ( f ) = fourier transform
x(t ) = continous time series
j = −1
f = frequency where - ∞ < f < ∞
t = time

The results were in real and imaginary components.


136

The spectral plot of the forced event causing the structural vibration is shown in

Figure 7.6. The data used for this analysis was taken from the anemometer placed 4 ft

(1.22 m) above the post support, AN-4. The measured data was in velocity units (speed

per second), from which the drag pressure was calculated (Eq. 7.2). According to the ori-

entation of the equation, the value calculated was in the form of the drag pressure divided

by the drag coefficient.

PD 1
= ρV 2 [Eq. 7.2]
CD 2

where
PD = drag pressure
C D = drag coefficient
slug  kg 
ρ = density of air, 0.0023657 3 
1.22 3 
ft  m 
V = wind velocity
137

FIGURE 7.6 Forced event used for modal analysis.

The plot shows significant pressure around the lower frequencies, showing a broadband

spectrum evident of the gustiness and turbulent nature of wind behavior, and drastically

drops of near zero for higher frequencies.

A typical spectral plot of the response of the structure to the forced event, meas-

ured in acceleration units from the accelerometers, is shown in Figure 7.7. The plots

shows a significant broadband spectrum around the lower frequencies evident of the

background turbulence effect of wind pressure, and an extremely narrowband spectrum at

the natural frequencies of the structural system evident of the resonance effect. These

spikes are identified as the natural frequency of vibration of the support structure. The

resulting response of the structure is a combination of the background turbulence and the

resonant vibration.
138

FIGURE 7.7 Structural response to the forced event.

The next step was compiling the modal frequencies of the support structure meas-

ured from each accelerometer during structural vibration. The frequencies were identified

by the spikes on the spectrum plots. Four major natural frequencies were found and are

listed Table 7.1 in their order of appearance along the frequency domain. The spectrum

plots of the frequencies are shown in Figure 7.8 and Figure 7.9 for accelerometers AC-2-

Y and AC-2-Z, located on the top horizontal strut of the cantilevered truss overhang.

The plots demonstrate a dominance of vibration associated with the first two mo-

dal frequencies. Figure 7.8 was from measured acceleration in the y direction (horizontal,

parallel to direction of traffic), showing significant vibration in this direction at a modal

frequency of 1.61 Hz identified as mode 1. Figure 7.9 was from acceleration measured in

the z direction (vertical, perpendicular to the ground), indicating significant vibration in


139

this direction at a modal frequency of 1.64 Hz identified as mode 2. The amplitude of

mode 1 in Figure 7.7 was greater than mode 2 in Figure 7.8, which revealed the majority

of vibration was controlled by mode 1 during the measured sample. Other frequencies of

vibration were excited, but the implication of these vibrations was small.

TABLE 7.1 Experimental Natural Frequencies

Mode Number Frequency (Hz)


1 1.61
2 1.64
3 3.96
4 5.60

FIGURE 7.8 Modes 1 and 3 from AC-2-Y.


140

FIGURE 7.9 Modes 2 and 4 from AC-2-Z.

Modal Shapes

The modal shapes were indentified for each of the four modes of vibration de-

scribed in the spectral analysis section. The modal shapes illustrated the direction and

shape of the vibrating structure. This was done using a system known as Quadrature

Picking using frequency response functions (FRF) formulated from the force and re-

sponse spectral plots. The shapes were found by using the imaginary component of the

Fourier transform (Eq. 7.1) to identify the modal direction and placement of the point on

the structure where one accelerometer with respect to another accelerometer was located

during resonant vibration. The strategic placement of the accelerometers, and the direc-

tion they measure, before testing during the development of the research program was

extremely important for the success of the analysis.


141

Frequency Response Function

The FRF is described as the ratio of the output spectrum to the input spectrum. In

the context of this project, the FRF was the spectral acceleration divided by the spectral

wind pressure. This particular FRF, shown in general terms in Eq. 7.3, is commonly re-

ferred to as accelerance:

X (ω )
H (ω ) = [Eq. 7.3]
F (ω )

where
H (ω ) = accelerance
X (ω ) = spectral acceleration response
F (ω ) = spectral force excitation
ω = frequency

The process required remedial steps before formulating Eq. 7.3 in order to reduce

noise errors in the output spectra. Through using the method of least squares, the result-

ing accelerance was estimated as the cross spectrum of the response and the force excita-

tion divided by the autospectrum of the excitation, as shown in Eq. 7.4 (39). This helped

to reduce noise errors in the spectral calculations.

( F )( X ) GFX (ω)
H 1 (ω) = = [Eq. 7.4]
( F )(F ) GFF (ω)
142

where
H 1 (ω ) = accelerance estimator
F = spectral force excitation input
X = spectral accleration response output
G FX (ω ) = cross spectrum of response and excitation
G FF (ω ) = autospectrum of excitation

The spectral force excitation was in the form of a Fourier transform shown in Fig-

ure 7.6. The spectral response was the Fourier transform of the accelerometer data de-

scribed in Figure 7.8 and 7.9 as an example. The accelerance was calculated for each ac-

celerometer located on the post and truss overhang. Real and imaginary components re-

sulted in the Fourier transform calculation, from which the imaginary component was

used to identify modal shapes in a process known as Quadrature Picking. The accelerance

becomes imaginary at the modal frequencies, whereby the amplitude of the imaginary

component is proportional to the modal displacement of the point on the structure where

the accelerometer was attached. A basic outline of the modal shape was achieved with

each modal frequency by first calculating the accelerance for each accelerometer, identi-

fying the modal frequencies in the accelerance, plotting the imaginary components at

each identified frequency, and then comparing the modal displacements between the ac-

celerometers (39).

Mode 1

Mode 1 equal to 1.61 Hz was measured by accelerometers AC-1-Y on the post

support, and AC-2-Y and AC-3-Y on the truss overhang. The plots of the imaginary

components at the mode indicated a larger modal displacement of the truss in the y direc-
143

tion than the shaft. The plots of the imaginary components are shown together in Figure

7.10 for AC-1-Y and AC-2-Y. Accelerometer AC-3-Y was identical to AC-2-Y. The

comparison of the modal displacements between the two accelerometer locations demon-

strated a torsion vibration of the truss about the support shaft by the relatively large mo-

dal displacement of the truss to the shaft. Each accelerometer plot indicated a negative

modal displacement in the y direction. A general illustration of the outlined modal shape

obtained from experimental data is shown in Figure 7.11. The mode is best described as

torsion about the support shaft with a slight bending of the shaft in the y direction.

FIGURE 7.10 Quadrature picking of mode 1.


144

AC-1-Y

AC-2-Y

z
AC-3-Y

x
y

FIGURE 7.11 Mode 1: torsion about support shaft.

Mode 2

Mode 2 equal to 1.64 Hz was measured by accelerometers AC-1-X on the post

support, and AC-2-Z on the truss overhang. The plots of the imaginary components at the

mode indicated an upward modal displacement of the truss in the positive z direction, fol-

lowed by a backward modal displacement of the shaft in the negative x direction. This

was indicative of a vertical rocking motion of the truss and shaft support. Mode 2 was

slightly picked up by AC-1-Z on the shaft, but was less significant than the other acceler-

ometers. This mode was not measured from the accelerometers oriented in the y direc-

tion, indicating no movement in this direction of this mode. The plots are shown together

in Figure 7.12 for AC-1-X and AC-2-Z. A general illustration of the outlined modal

shape obtained from experimental data is shown in Figure 7.13.


145

FIGURE 7.12 Quadrature picking of mode 2.

AC-1-X

AC-2-Z

x
y

FIGURE 7.13 Mode 2: vertical rocking.


146

Mode 3

Mode 3 equal to 3.97 Hz was measured by accelerometers AC-1-Y on the post

support, and AC-2-Y and AC-3-Y on the truss overhang. The plot of the imaginary com-

ponent for the post showed a significant modal displacement in the positive y direction,

whereas the modal displacement of the truss was opposite in direction and less signifi-

cant. The plots are shown together in Figure 7.14 for AC-1-Y and AC-2-Y. AC-3-Y was

identical to AC-2-Y. The modal shape is best described as a horizontal twist of the truss,

but also described by a horizontal rocking of the post support with the truss displacement

dependent and rocking backward in response. A general illustration of the outlined modal

shape obtained from experimental data is shown in Figure 7.15.

FIGURE 7.14 Quadrature picking of mode 3.


147

AC-1-Y

AC-2-Y

AC-3-Y

x
y

FIGURE 7.15 Mode 3: horizontal truss twist.

Mode 4

Mode 4 equal to 5.63 Hz was picked up by accelerometers AC-1-X on the post

support, and AC-2-Z on the truss overhang. The plots of the imaginary components at the

mode indicate an upward modal displacement of the truss in the positive z direction, fol-

lowed by a forward modal displacement of the shaft in the positive x direction. This was

indicative of an outward and inward clamping motion of the truss and shaft support.

Mode 4 was slightly picked up by AC-1-Z on the shaft, but was less significant than the

other accelerometers. This mode was not measured from the accelerometers oriented in

the y direction, indicating no movement in this direction of this mode. The plots are

shown together in Figure 7.16 for AC-1-X and AC-2-Z. A general illustration of the out-

lined modal shape obtained from experimental data is shown in Figure 7.17.
148

FIGURE 7.16 Quadrature picking of mode 4.

AC-1-X

AC-2-Z

x
y

FIGURE 7.17 Mode 4: outward and inward clamping.


149

Modal Damping

A modal damping analysis was performed from the experimentally collected ac-

celerometer data in order to determine the damping ratio, ξ, of the structure for further

fatigue analysis in this project. The damping ratio was used in developing a dynamic

model of the structure to simulate the response to wind loading events, and to determine

the fatigue load based on the dynamic characteristics. This process is discussed in detail

in Chapter 9: Theoretical Calculation of the Fatigue Load to due Natural Wind Gusts.

There are many ways to calculate the modal damping of the structure. For the

purpose of this project, only the damping of the first two major vibratory modes was

needed. This was because Mode 1 and Mode 2 formed the majority of the vibration, as

seen in Figures 7.8 and 7.9. Their modal frequencies were nearly identical, but this did

not suggest that the rate at which the vibration decays (damping) were identical. It is also

important to point out that, by looking at Figures 7.8 and 7.9, the majority of vibration

was in the first modal frequencies, and therefore the vibration shape was controlled pre-

dominately by these modal frequencies.

The transient events used in the damping analysis were the measured response of

the truss from experimental testing of the fatigue load due to truck gusts. A large truck

was driven periodically under the sign structure and the response of the structure was

measured using anemometers, strain gauges, and accelerometers. This was done on a

relatively low wind day as to not have external effects from natural wind gusts. The result

was a noticeable transient event as seen from the response data. A total of four truck runs

were used for the analysis, from which two damping calculations were made per mode

(mode 1 and mode 2), equaling a total of eight damping calculations per mode, and 16
150

damping calculations total. An average was taken from the eight values for each and used

for the remainder of the project as the damping value of the sign structure for the mode of

interest.

An understanding of the forced excitation was developed before the damping cal-

culations were made. An experimentally measured transient event was isolated and plot-

ted on amplitude versus time graph. An example of a typical truck transient event meas-

ured by the accelerometers, and used for this analysis, is shown in Figure 7.18.

FIGURE 7.18 Typical truck transient event from accelerometer.

The event depicted in Figure 7.18 was representative of a free vibration system of oscilla-

tory motion. The system was underdamped and periodic, since the majority of vibration

was controlled by a single frequency; in this case, Mode 1 equal to 1.61 Hz. Other modal
151

frequencies were present due to the ruggedness of the plot, but were insignificant to the

major modal shape of Mode 1. The duration of the event was roughly 10 seconds. It was

measured using accelerometer AC-2-Y, which measured oscillatory motion in the y direc-

tion (parallel to the direction of traffic).

The displacement of the response was estimated mathematically as an under-

damped, periodic dynamic system. The expression shown in Eq. 7.5 is a representation of

the exponential decay in amplitude with time in conjunction with the periodic motion of

the modal frequency:

x (t ) = A0e −ξω n t cos(ω d t − ϕ 0 ) [Eq. 7.5]

where
x(t ) = amplitude as a function of time (G)
A0 = initial amplitude (G)
ξ = damping ratio
ω n = natural frequency (rad/sec)
ω d = damped frequency (rad/sec)
t = time (sec)
ϕ 0 = initial phase (rad)

The exponential expression in Eq. 7.5 (shown as Eq. 7.6) is representative of the expo-

nential decay of the amplitude peaks with increasing time. The cosine expression repre-

sents the periodic motion of the system. In relation to the measured transient event, Eq.

7.6 (and Eq. 7.5) is better described in Figure 7.19.


152

FIGURE 7.19 Exponential decay of the transient truck event.

y = Ae −ξωt [Eq. 7.6]

where
y = amplitude as a fuction of time of the exponential expression (G)
A = initial amplitude of the exponential expression (G)
ξ = damping ratio
ω = natural frequency (rad/sec)
t = time (sec)

A regression analysis was done to determine the damping ratio. The positive peak

amplitudes of each wavelength was extracted from the time history graphs and plotted

separately. The time was equalized to start from zero so that the amplitude, A0, could be
153

realized. An example of the resulting plot is shown in Figure 7.20. A trendline was fitted

to the plotted peaks and the equation of the trendline was extracted. The exponent of the

trendline equation was set equal to the absolute value exponent of Eq. 7.6. By knowing

the natural frequency of the structure, ω, which was determined in the Spectral Analysis

and Modal Shapes sections of this chapter, the damping ratio, ξ, was calculated. A se-

quence of the calculation events, using the trendline equation in Figure 7.20 as an exam-

ple, is shown in the following system of equations:

FIGURE 7.20 Trendline of extracted peak amplitudes.

y = Ae −ξωt = 0.0447e −0.244 x [Eq. 7.7]


154

where
A = 0.0447 G
-ξωt = −0.244 x
t=x

and with
ω = 1.61 Hz = 10.116 rad/sec (Mode 1)

therefore

− ξωt = −0.244 x
0.244 [Eq. 7.8]
ξ= = 0.02412 = 2.41%
10.116

This process was performed for the upper positive peaks and lower negative peaks

for each transient truck event measured. The above example was for AC-2-Y, measuring

vibration in the horizontal direction of Mode 1. The damping for Mode 2, using AC-2-Z,

was also performed in the same manner. The results of the modal damping analysis are

listed in Table 7.2. The larger damping value for Mode 1 was attributed to an additional

aerodynamic damping of the sign as it vibrated horizontally about the shaft support.
155

TABLE 7.2 Modal Damping Results

Top/Bottom
Transient Damping Ratio
Accelerometer Mode Peak Ampli-
Event (%)
tudes
Top 2.41
AC-2-Y Mode 1
Bottom 2.30
1
Top 1.38
AC-2-Z Mode 2
Bottom 1.63
Top 2.20
AC-2-Y Mode 1
Bottom 2.19
2
Top 0.65
AC-2-Z Mode 2
Bottom 0.67
Top 1.50
AC-2-Y Mode 1
Bottom 1.38
3
Top 0.543
AC-2-Z Mode 2
Bottom 0.495
Top 1.246
AC-2-Y Mode 1
Bottom 1.335
4
Top 0.398
AC-2-Z Mode 2
Bottom 0.922
Average Damping Ratio for Mode 1 (1.61 Hz): Shaft Torsion 1.820
Average Damping Ratio for Mode 2 (1.64 Hz): Vertical Rocking 0.836
156

CHAPTER 8

EXPERIMENTAL CALCULATION OF THE FATIGUE LOAD


DUE TO NATURAL WIND GUST

Chapter Overview

The calculation of the fatigue load due to natural wind from the experimental data

is presented in this chapter. Only collected data with wind velocity greater than 9 mph (4

m/s) and directed onto the front face of the structure was used. The fatigue load due was

based on an upper to lower peak–to-peak stress range. A three second average was used

for data reduction. The data was broken down into two components 1) structural excita-

tion, and 2) structural response. The anemometers made up the excitation, whereas the

accelerometers and strain gauges made up the structural response.

Fatigue Load Calculation Approach

The equivalent static wind load approach was used in determining the fatigue load

due to natural wind. The excitation and response of the structure was measured experi-

mentally. The behavior was dynamic in nature. An equivalent static wind load was back-

calculated from the measured response values that would produce the same dynamic re-

sponse of the structure in terms of the maximum peak-to-peak stress ranges.

The experimental data collected was analyzed in the same fashion as the devel-

opment of the natural wind fatigue provisions in the Supports Specifications, except for
157

using experimental data in place of theoretical values. A general description of the

evaluation process went as follows:

1. Wind velocity directionality unit vectors were developed to describe loading ori-

entations,

2. Maximum peak-to-peak stress ranges were determined,

3. Equivalent static wind pressures were back-calculated using the stress ranges and

loading unit vectors,

4. The pressures were categorized to their corresponding wind velocities and plotted,

5. The wind velocity corresponding to a 0.01% probability of exceedence was de-

termined, and

6. The pressure corresponding to the 0.01% exceedence probability was extracted

from the plot.

The intention for the likeness in analysis to the Supports Specifications was to enable a

direct comparison to be made between the code and the experimental data. This would

allow for a good understanding of the code’s accuracy.

The data collected from the strain gauges form the basis of the analysis. Strain

gauges were placed on the chord members, the supporting shaft, and the anchor bolts,

totaling 44 gauges. It was determined that the most consistent measurements were found

in the gauges located on the post (see Chapter 6: Fatigue Resistance). The usage of the

post gauges also allowed for a more straightforward back-calculation of the equivalent

static wind load.


158

The step-wise procedure for determining the fatigue load generalized in the nu-

meration presented earlier is detailed in the following sections. The detailing describes

the procedure performed by the researcher starting from the experimentally collected raw

data to the back-calculation of the equivalent static wind load performed at the conclu-

sion of the process.

Structural Excitation

Collected Sample Size

A total of 36.75 hours of excitation and response data was recorded for the natural

wind fatigue load program. Since the acquisition system recorded data at a rate of 60

samples per second, the collected hours corresponded to 7,938,000 data points per in-

strument and 492,156,000 total data points of all 62 instruments. The collection dates and

times of the data samples are shown in Table 8.1. The wind data was collected in 45 min-

ute intervals except for data collected on 4/8/2009 and 4/9/2009, which was taken in 25

minute intervals. The collection data on these dates were designated for truck-induced

wind gust testing, however natural wind data was collected between truck runs. A total of

53 data collection intervals were accomplished: 44 intervals of 45 minutes each, and nine

intervals of 25 minutes each.

TABLE 8.1 Collection Dates and Times for Natural Wind

Collection Date Number of Hours


1/14/2009
19.5 hours
1/15/2009
3/3/2009
12.75 hours
3/4/2009
4/8/2009
4.5 hours
4/9/2009
159

Usable Data Collected

The data collected from the anemometers were used to distinguish between usable

and discarded data to apply for wind pressure back-calculation. Data was considered us-

able if it complied with the following criteria:

 Wind velocity was greater than 9 mph (4 m/s), and

 Wind velocity was directed onto the front face of the structure.

The data was broken down into wind velocity and direction. It was discovered from the

collected structural response data that significant structural vibration was only induced

with wind velocities greater than 9 mph (4 m/s). Only wind velocities directed onto the

front of the structure were used for evaluation. The resulting usable data that complied

with the above limitations was three hours and thirty minutes, equaling a total of

46,872,000 data points.

The natural wind data collected was nicely spread between the desired wind ve-

locities. Wind rose diagrams were developed for all data collected to help in determining

usable data directed onto the front face of the structure. The wind rose diagram in Figure

8.1 is of all 36.75 hours collected, whereas the wind rose diagram in Figure 8.2 is of the

data determined usable and used for the analysis. The data demonstrated in the wind rose

diagrams were measured using the ambient wind anemometer AN-4. The directions are

shown in compass bearings. The North end of the anemometer was oriented opposing

traffic, directed away from the front face of the structure.


160

The average wind velocity of the data considered usable was 12.96 mph (5.79

m/s). The measurement was taken from anemometer AN-4 located 4 ft (1.22 m) above

the structure equaling 32.75 ft (9.98 m) above ground level. The wind velocity was ori-

ented onto the front face of the structure as shown in Figure 8.2. A maximum wind veloc-

ity recorded in this orientation was 33 mph (14.8 m/s).

0° (North)

Counts

270° (West) 90° (East)

180° (South)

FIGURE 8.1 Wind rose diagram of all data.


161

0° (North)

Counts
270° (West) 90° (East)

180° (South)

FIGURE 8.2 Wind rose diagram of usable data.

Reduction of Structural Excitation Experimental Data

The collected excitation data determined as usable was further reduced for analy-

sis. The direction of the wind velocity vector was transformed from the compass bearings

used for experimental measurement, to polar bearings for data analysis. A wind direc-

tionality unit vector was developed to use in the back-calculation for determining the

equivalent static wind load.


162

Averaging Time

The data streamlines were averaged every three seconds in order to be in compli-

ance with the drag coefficients and height coefficients available in the Supports Specifi-

cations for wind loading analysis. It was representative of an instantaneous value.

Transformation from Compass Bearings to Polar Bearings

The magnitude and direction of the wind velocity vector was broken down into

North (Y component) and East/West (X component) components. The North direction

was oriented in the direction away from the front face of the structure and opposing the

direction of traffic. An average of the components for every three seconds was taken. The

wind velocity vector was then reformed for each three second window.

The reasoning for this procedure was twofold. First, it solved the cross-over prob-

lem. In compass bearings, 360° was the same as 0° on the direction scale. A simple aver-

age of values without a breakdown into components would not take care of this issue. For

example, the average of 315° and 45° equals 180°, which would indicate the wind was

blowing opposite (South) of the true direction (North). A breakdown into components

solved the cross-over problem. Secondly, low wind velocities did not excite as much vi-

bration in the structure as higher wind velocities. For that reason, the X and Y compo-

nents of the breakdown took into account the magnitude of the wind velocity and not just

direction. The restructuring of the wind velocity vector over the three second averaging

window placed more emphasis on the larger magnitudes which would create more sig-

nificant vibration, and less emphasis on the smaller velocities which had less significant

results.
163

Once the three second window averaging was accomplished, the reformed three

second wind velocity vector was transformed from its compass bearings to polar bear-

ings. This was done by the following equation:

Polar Bearings = (−1)[(Compass Bearings+ 180ο ) − 90 ο ] [Eq. 8.1]

The 180° addition was to account for incoming wind as the anemometer measured the

wind direction along compass bearings from the direction it was blowing.

Equation 8.1 was applied to anemometer AN-4 measuring the horizontal plane.

Anemometer AN-3 measured the vertical plane, and AN-3 was not working for the diag-

nosed usable data. When it was working, the wind velocity direction was primarily be-

tween 350° and 10° (cross-over through zero), and averaged close to zero. In view of this,

it was taken as zero, with no up or down component.

Wind Directionality Unit Vector

After the excitation was transformed from compass bearings to polar bearings, a

wind directionality unit vector was formed. The unit vector was developed into Cartesian

coordinates (i, j, k). This was done in order describe mathematically where the wind ve-

locity, and subsequent wind pressure load, was directed for back-calculation of the

equivalent static wind load. Using the polar coordinates, the unit vector was calculated by

the following system of equations:


164

λ = (λ x i + λ y j + λ z k )
 V   V y   V  
=  x i +   j +  z k  [Eq. 8.2]
 V   V   V  
= [(cos β cos α )i + (cos β sin α ) j + (sin β )k ]

where
λ = wind directionality unit vector
α = wind orientation along the horizontal plane (xy - plane) measured from AN - 4
β = wind orientation along the vertical plane (zy - plane) measured from AN - 3

A schematic of the coordinate system is shown in Figure 8.3 for better understanding of

the unit vector directionality.

FIGURE 8.3 Coordinate system.


165

A generalized example of the transformation from compass bearings to polar bearings,

and the development of the unit vector are shown in Table 8.2.

TABLE 8.2 Development of the Wind Directionality Unit Vector

Compass Bear- Magnitude Polar Bearings Unit Vector (unit-less)


ings (degrees) (mph) (degrees) i j k
AN-4 AN-3 AN-4 AN-3 AN-4 AN-3
-0.918 -0.397 0
66.640 0 17.306 17.306 -156.64 -180

A wind velocity directionality unit vector was formed for each three second averaging

window for all usable data collected. The unit vector represented the direction of the

structural excitation with respect to the coordinate system shown in Figure 8.3. The mag-

nitude of the wind velocity was proportional to the magnitude of the wind pressure to be

back-calculated from the measured structural response.

Structural Response

The structural response analysis involved the strain gauge data corresponding to

the usable wind data (data was measured simultaneously in the field). The same three

second averaging window of the structural excitation was used for the analysis. The ma-

jor outcome was to determine the experimentally measured stress ranges from the meas-

ured structural excitation to back-calculate the equivalent static wind load and subsequent

calculation of the fatigue load.


166

Data Offsetting

The first step involved an offsetting procedure to generate the true strain time his-

tory from the collected raw data. Wind velocity occurs randomly in nature and cannot be

controlled manually. For this reason, the zeroing of the gauges before testing in the field

was not representative of a true no-strain condition. The structure was continually vibrat-

ing due to the wind velocity presence, and the zeroing of the scales occurred during

backward and sometimes forward vibratory movement which was unavoidable. The solu-

tion was to perform an offsetting procedure on the collected data to offset the strain value

to be representative of a non-vibratory state.

Important for this process was the simultaneous collection of strain and wind ve-

locity data. The strain data collected during each run was filtered with respect to its corre-

sponding collected wind velocity. This was done in +/- 0.5 mph (0.224 m/s) wind veloc-

ity magnitude intervals. The strain values that occurred during each interval were deter-

mined and averaged together. The values were plotted on a wind velocity versus strain

diagram (see Figure 8.4). For example, all strain values occurring during a 3.5 to 4.5 mph

(1.56 to 2.01 m/s) interval were averaged and plotted as a 4 mph (1.79 m/s) data point.

The data formed a parabola, which adhered to the fluid mechanics relationship between

wind force and the square of wind velocity.


167

FIGURE 8.4 Strain values per 0.5 mph wind velocity intervals.

A regression analysis was performed to determine a best fit line of the plotted

data. This was done through a transformation regressor linearization process. The inde-

pendent variable (wind velocity on the abscissa axis) was squared. It was plotted versus

its corresponding averaged strain (see Figure 8.5). The linearization of the data proved

the purely parabolic nature of the data. A best fit line was then constructed as a linear

predictor. The intercept of the trendline on the ordinate axis indicated the strain value to

offset. For example, the offset for the data run presented in Figure 8.5 was 1.1501 mi-

crostrain. This means that when the strain gauges were zeroed during the testing proce-

dure, the structure was vibrating due to the continuous wind excitation, and at the mo-

ment of zeroing an approximate magnitude of 1.1501 microstrain was induced onto the

structure. This is evident from Figure 8.5 at the zero squared wind velocity point.
168

FIGURE 8.5 Transformed regressor.

Returning to Figure 8.4, a line was plotted to fit the data based on the information

gathered from the transformation (Figure 8.5). As shown in Figure 8.5, the slope of the

trendline was the slope of the parabola, and the intercept of the trendline was the intercept

of the parabola. Plotting a line on Figure 8.4 using the transformation regressor values is

shown in Figure 8.6.


169

y = 0.0351x 2 + 1.1501

FIGURE 8.6 Parabolic curve fit.

Once the trendline was produced for the parabolic data, the offset (1.1501 for this

data run example) was subtracted from the trendline to produce an accurate zeroed time

history. The offset plot is shown in Figure 8.7. As a side note, the plot in Figure 8.7 can

be used to project strain values for higher wind velocities than that used for fatigue analy-

sis. For instance, Figure 8.8 shows the plot projected to a 90 mph (40.2 m/s) wind, which

is used for capacity calculations and design.


170

y = 0.0351x 2

FIGURE 8.7 Parabolic offset.

FIGURE 8.8 90 mph projection using offsetting procedure.


171

A systemized procedure of the offsetting process is numerated as follows:

1. Filter the collected strain data in +/- 0.5 mph (0.224 m/s) intervals corresponding

to the wind data collection,

2. Strain values from each interval were averaged,

3. Averaged strain values versus the wind velocity interval were plotted to test cor-

respondence with force and wind velocity as a parabola,

4. The data was transformed into a linear plot by squaring the independent variable

(wind velocity),

5. The transformed wind data was plotted with its corresponding averaged strain to

test the linearization process,

6. A regression was performed on the transformed data to generate a linear best fit

trendline,

7. The trendline was projected to determine its intercept with the ordinate axis, and

8. The intercept value was subtracted from the raw strain data time history stream-

line to produce the true strain time history.

The offsetting procedure described was done for each strain gauge time history of

the usable data collection. It produced the true strain values to account for zeroing the

instrumentation during wind velocity excitation. To lessen the magnitude of the offset,

ALDOT was encouraged to zero the instrumentation and begin data collection when the

wind velocity seemed to be low. Since zeroing at exactly a zero wind velocity was practi-

cally impossible, the offset procedure was implemented to all data.


172

Strain Ranges

The next step was analyzing the offset data to determine the strain range. The off-

set data was averaged every three seconds. The maximum and minimum values within

each three second window were determined and subtracted from each other. The result

was the maximum peak-to-peak range within each three second window. This was done

for all usable offset data and strain gauges. Other parameters such as the standard devia-

tion and peak-to-standard deviation ratio were determined for each three second window

for behavioral purposes. An example of a typical peak-to-peak range calculation, repre-

sented in a plotted streamline, is shown in Figure 8.9.

FIGURE 8.9 Peak-to-peak range.


173

Rosette Strain Gauge

The rosette gauges located at Section AA were used to measure shear stresses in

the shaft. Before the peak-to-peak range was determined, a strain transformation was

needed to calculate the shear strain. The gauges were rectangular strain rosettes arranged

in a 45° pattern as shown in Figure 8.10. The strain transformation equation, Eq. 8.3, was

used to calculate the experimental shear strain measurement from the rosettes.

ε a = ε x cos2 (θ a ) + ε y sin 2 (θ a ) + γ xy sin(θa ) cos(θ a )


ε b = ε x cos2 (θb ) + ε y sin 2 (θb ) + γ xy sin(θb ) cos(θb ) [Eq. 8.3]
ε c = ε x cos2 (θc ) + ε y sin 2 (θc ) + γ xy sin(θc ) cos(θc )

With θa equal to -135°, θb equal to -90°, and θc equal to -45°, Eq. 8.3 becomes:

ε x = εa − εb + εc
ε y = εb [Eq. 8.4]
γ xy = ε a − ε c

Equation 8.4 was used to calculate the experimental shear stress in the shaft for which the

stress ranges were determined and the wind pressure magnitude, P, was found.
174

FIGURE 8.10 45° rosette arrangement with coordinate axis.

Wind Pressure Back-Calculation

The equivalent static wind load that would produce the same peak-to-peak stress

range determined experimentally was back-calculated. The three second ranges and wind

directionality unit vectors were used for this calculation. Two measures formed the basis

of the process. First, the stress range was determined from the experimentally measured

strain ranges as described in the Structural Response section of this chapter. Second, a

theoretical structural analysis was performed on the structure to determine the stresses at

the gauged locations for comparison with the experimental values. The calculation in-

volved a combined loading analysis and unsymmetrical bending. A stress element of

normal and shear stress values for each three second interval was developed at the loca-

tion of the gauge to be used in the comparison. Importantly, the stress values were devel-
175

oped with the wind pressure magnitude, P, kept as a variable. The two measures, experi-

mental stress and theoretical stress, were set equal to each other and the wind pressure

magnitude, P, was solved (see Eq. 8.5). The instrumented locations along the shaft sup-

port (Section AA and Section BB) were used exclusively for this evaluation because of

their orientation to in-plane and out-of-plane displacements, as well as the consistency of

the data collected at this location.

Exp. Stress Range = (Theo. Stress Range)( P )


Exp. Stress Range [Eq. 8.5]
P=
Theo. Stress Range

Theoretical Structural Analysis

The structural analysis involved determining exposed areas of the structure to

natural wind velocity and using them to form force vectors along with the wind direction-

ality unit vectors. The wind pressure magnitude, P, was kept as a variable. This produced

point loads with direction at each exposed area. Drag and height coefficients provided by

the Supports Specifications, which were developed using three second wind averages in

previous studies, were determined for each member exposed to wind and used in the

analysis. The idea was to simulate as accurately as possible the wind loading conditions.

Equations of equilibrium were developed and the acting moment vector (resultant) was

solved. The moment vectors were used to develop stress elements in a combined loading

analysis and unsymmetrical bending at each strain gauged location, and Eq. 8.5 was used

to solve for the wind pressure magnitude to conclude the back-calculation.


176

Exposure Area Breakdown

The front and the sides of the structure were exposed to wind pressure from the

wind data determined as usable. Each face was divided into individual segments for

structural analysis. The segmented division depended on the type of member exposed.

The front face of the structure contained three segments, front sign, front truss, and front

shaft. The side of the structure was segmented into two areas: East side face and the West

side face. An illustration of the exposed areas to wind pressure used for the structural

analysis is shown in Figures 8.11 through 8.13. Each exposed area, area centroid, and its

distance to the strain gauges at Section AA and Section BB were calculated. The results

are shown in Table 8.2. The shaft segments were further broken down into three seg-

ments to account in variation of wind pressure with height (detailed in the next section).

FIGURE 8.11 Area breakdown of the front face.


177

FIGURE 8.12 Area breakdown of East side face.

FIGURE 8.13 Area breakdown of West side face.


178

TABLE 8.3 Exposure Area Breakdown

Distance of Centroid to Strain Gauges (in)


Section AA Section BB
Area
Segment Location Label (12 in above Base (16 in above Base
(in2)
Plate) Plate)
i j k i j k
Front
Sign A1 26,108 360 22.6 336 360 22.6 332
Sign
Front Chord A2-1 2,016 145 0 309 145 0 305
Truss Web A2-2 1,118 145 0 309 145 0 305
Bottom A3-1 4,435 0 0 92.4 0 0 90.4
Front
Middle A3-2 2,362 0 0 234 0 0 230
Shaft
Top A3-3 1,195 0 0 308 0 0 304
Bottom A4-1 4,435 0 0 92.4 0 0 90.4
East
Middle A4-2 2,362 0 0 234 0 0 230
Shaft
Top A4-3 1,195 0 0 308 0 0 304
Bottom A5-1 4,435 0 0 92.4 0 0 90.4
West
Middle A5-2 2,362 0 0 234 0 0 230
Shaft
Top A5-3 1,195 0 0 308 0 0 304

Effective Area Breakdown

Effective areas of the areas listed in Table 8.3 were calculated. The effective area

represented the exposed area (Table 8.3) multiplied by the drag coefficient of the mem-

bers that make up the area, and the height coefficient of the area centroid above ground

level. The back-calculation was based on three second ranges and wind directionality unit

vectors, and therefore the drag coefficients and height coefficients in the Supports Speci-

fications were applicable.

Drag coefficients. All drag coefficients were calculated using Table 3-6 of the

Supports Specifications. Fifty year design life was used in the calculation with a velocity

conversion factor equal to 1.00 (no conversion needed). The results are listed in Table
179

8.4. A value of 1.123 was used for the sign, 1.2 was used for the chord members in the

overhanging truss, and 1.1 was used for all web members and shaft support.

TABLE 8.4 Drag Coefficients

Segment Location Label Drag Coefficient, Cd

Front Sign Sign A1 1.123


Chord A2-1 1.2
Front Truss
Web A2-2 1.1
Bottom A3-1 1.1
Front Shaft Middle A3-2 1.1
Top A3-3 1.1
Bottom A4-1 1.1
East
Middle A4-2 1.1
Shaft
Top A4-3 1.1
Bottom A5-1 1.1
West
Middle A5-2 1.1
Shaft
Top A5-3 1.1

Height coefficients. All height coefficients were calculated based on the Table 3-5

of the Supports Specifications. Exposure condition C (open terrain with scattered obstruc-

tions) was used for the calculation. The height coefficient represents a change in turbu-

lence of wind pressure, becoming less turbulent and more stable as the height above

ground level increases. The coefficient used in the Supports Specifications was defined

using the power law of the wind velocity profile with respect to height above ground

level shown in Eq. 8.6. The curved profile with height defined by Eq. 8.6 was simplified

into a stepped profile by the Supports Specifications and is shown in Figure 8.14 demon-

strating the values used for this structure. The exposure definition only applied to the
180

shaft support as the centroid of the truss and sign was located between 24.6 ft (7.50 m)

and 32.8 ft (10 m) where the coefficient was equal to 1.0. All height coefficients used in

the structural analysis for each area segment is listed in Table 8.5.

α
 h 
vw = v10   [Eq. 8.6]
 h10 

where
v w = wind velocity
v10 = wind velocity at a reference height of 32.8 ft (10 m); normalized at 1.0
h = height above ground level
h10 = reference height of 32.8 ft (10 m)
α = terrain constant equal to 0.16 for neutral air above flat open coast.

TABLE 8.5 Height Coefficients

Segment Location Label Height Coefficient, Kz

Front Sign Sign A1 1.0


Chord A2-1 1.0
Front Truss
Web A2-2 1.0
Bottom A3-1 0.87
Front Shaft Middle A3-2 0.94
Top A3-3 1.0
Bottom A4-1 0.87
East
Middle A4-2 0.94
Shaft
Top A4-3 1.0
Bottom A5-1 0.87
West
Middle A5-2 0.94
Shaft
Top A5-3 1.0
181

FIGURE 8.14 Height coefficient stepped profile.

Effective area calculation. The effective area was defined as the exposed area of

the segment multiplied by the drag coefficient of the members that make up the area

segment, and the height coefficient of the centroid location of the area above ground

level. The results of the calculation are listed in Table 8.6.


182

TABLE 8.6 Effective Area

Drag Effective
Area Height Co-
Segment Location Label Coefficient, Area, Ae,i
(in2) efficient, Kz
Cd (in2)
Front
Sign Ae,1 26,108 1.123 1.0 29,320
Sign
Front Chord Ae,2-1 2,016 1.2 1.0 2,419
Truss Web Ae,2-2 1,118 1.1 1.0 1,230
Bottom Ae,3-1 4,435 1.1 0.87 4,244
Front
Middle Ae,3-2 2,362 1.1 0.94 2,442
Shaft
Top Ae,3-3 1,195 1.1 1.0 1,315
Bottom Ae,4-1 4,435 1.1 0.87 4,244
East
Middle Ae,4-2 2,362 1.1 0.94 2,442
Shaft
Top Ae,4-3 1,195 1.1 1.0 1,315
Bottom Ae,5-1 4,435 1.1 0.87 4,244
West
Middle Ae,5-2 2,362 1.1 0.94 2,442
Shaft
Top Ae,5-3 1,195 1.1 1.0 1,315

Acting Moment Vector Formulation

The effective areas and the wind directionality unit vectors were used to develop

acting force vectors for each three second interval. Equations of equilibrium were

formed. Using the position vectors listed in Table 8.3, representing the distance of the

centroid of the effective area to the strain gauge locations, the active moment at the strain

gauge location was calculated. This process is described by the following system of equa-

tions.

Acting force vector. The acting pressure vector was defined as the wind direction-

ality unit vector times the pressure magnitude, P, as shown in Eq. 8.7.

P = λ P = (λ x P i + λ y P j + λ z P k ) psi [Eq. 8.7]


183

Knowing the pressure is equal to the force divided over an area, the force vector was de-

fined in Eq. 8.8 in terms of the effective area calculated in Table 8.6, and keeping the

wind pressure magnitude, P, variable.

F = λ (ΣAe ,i P)
[Eq. 8.8]
= [λ x (ΣAe ,i P )i + λ y (ΣAe ,i P) j + λ z (ΣAe ,i P)k ] lb

The components of the force vector in Eq. 8.8 were therefore:

Fx = λ x (ΣAe ,i P ) lb
F y = λ y (ΣAe ,i P ) lb
Fz = λ z (ΣAe ,i P ) lb

Equations of equilibrium. Once the force vectors were developed, equations of

equilibrium were created from which the acting moment equations were solved. A typical

free body diagram of the wind loading on the structure can be generalized as shown in

Figure 8.15 with all force components shown in their positive sense. The acting moment

vector (resultant in Figure 8.15) was calculated from the equations of equilibrium formed

using the free body diagram. The pressure magnitude, P, was kept variable during the

calculation process. This was done for each three second interval for all usable data

streamlines.
184

FIGURE 8.15 Free body diagram.

Development of the Stress Element

A combined loading and unsymmetrical bending analysis was used to determine

the stress and strain values at the strain gauge locations. Section AA and Section BB were

primarily used for this calculation due to the placement of the strain gauges with respect

to in-plane and out-of-plane displacement. The stress element development involved the

combination of the following loading scenarios:


185

 Bending moment stress,

 Torsion stress,

 Normal stress, and

 Transverse shear stress.

Of the four loading conditions bulleted, the analysis involved almost exclusively

bending moment and torsion. The bending moment was analyzed as an unsymmetrical

bending. It was broken down into two components, a drag moment (moment about the x-

axis) and lift moment (moment about the y-axis). Since no vertical variability in wind di-

rection was measured using anemometer AN-3, the normal stress component of the com-

bination was not present. Transverse shear stress was present and used in the wind pres-

sure magnitude, P, evaluation with the rosette strain gauges at Section AA.

Material Properties

The material properties needed for development of the stress element are listed in

Table 8.7. A cross section of the analysis location is shown in Figure 8.16.

TABLE 8.7 Material Properties at Section AA and Section BB.

Material Property Value


Diameter, dshaft 24 in
Area, Ashaft 20.9 in2
Thickness, tshaft 0.281 in
Moment of Inertia, IXX, shaft 1,473 in4
Moment of Inertia, IYY, shaft 1,473 in4
Polar Moment of Inertia, JZZ, shaft 2,945 in4
Modulus of Elasticity, E 29,000,000 psi
Modulus of Rigidity, G 11,000,000 psi
186

FIGURE 8.16 Cross section of post at Section AA and Section BB.

Relevant Stress and Strain Equations

The stress and strain equations relevant to the combined loading analysis are

listed in Table 8.8. The stress element derived from the structural analysis was developed

using these equations.


187

TABLE 8.8 Relevant Stress and Strain Equations

Stress Stress Equation Strain Equations


F σ
Normal σN = z εN = N
A E
M ρ τ
Torsion τT = z γT = T
J ZZ G
M y M yx σ
Unsymmetrical Bending σM = − x + εM = M
I XX I YY E
FQ 2 F τ
Transverse Shear* τS = = γS =
It A G
*
for maximum transverse shear stress at neutral axis for thin walled pipes

Strain Gauge Locations

The stress element was formed for each strain gauged location. For Section AA

and Section BB, the strain gauges were placed circumferentially on the outer surface of

the shaft. The gauges were spaced at 45° from each other for each section. The strain

gauge locations and labeling used for the analysis are illustrated in Figure 8.17. The uni-

axial gauges were oriented along the longitudinal axis (z-axis) of the post. The middle

gauges of the rosettes were oriented along the same direction. The stress at each lo-

cation using the relevant stress equations listed in Table 8.8 was calculated. The values

were placed on a stress element and added together to determine the normal and shear

stresses at the gauged location (see Figure 8.18). The values were calculated with the

wind pressure magnitude, P, kept as a variable.


188

Strain Gauge

FIGURE 8.17 Strain gauge locations at Section AA and Section BB.

FIGURE 8.18 Typical stress element.


189

Wind Pressure Calculation

The wind pressure magnitude, P, was solved using Eq. 8.5. Strain gauges at Sec-

tion AA involved uni-axial and rosette gauges. Both types of gauges were used for nor-

mal stress values, whereas the rosette gauges were used for shear stress values. The ro-

settes were placed at locations A, C, E, and G. Section BB had only uni-axial gauges and

were used for the normal stress values. An example calculation for the rosette gauge at

Section AA placed on location A is shown in Table 8.9.

TABLE 8.9 Example of Wind Pressure Magnitude Calculation at Location A

Wind Directionality
Experimental
Unit Vector [Theoretical
Wind Pressure
Max. Max. Stress
Section Magnitude, P
Strain Stress Value]*P
i j k (psf)
Range Range (psi)
(µin/in) (psi)
AA 12.351 358.179 73,613.22 0.701
BB -0.682 -0.731 0 13.717 397.793 72,638.24 0.789
Rosette 14.420 158.620 32,695.94 0.699

The example is a snap-shot of a typical three second interval of the diagnosed us-

able data. A summary of the back-calculation procedure to determine the values in Table

8.9 are described by the following five step process:

1. The wind directionality unit vector was formed using the measured anemometer

compass bearings. The compass bearings were transformed into polar bearings

and the directionality unit vector was developed through trigonometric composi-

tions.
190

2. Theoretical structural analysis using combined loading and unsymmetrical bend-

ing was performed to determine the stress element at the strain gauged location.

The theoretical value was calculated using the segmented areas and equations of

equilibrium with the wind pressure magnitude, P, kept as a variable. The wind di-

rectionality unit vector (determined experimentally) was used to establish the di-

rection of the wind pressure magnitude to correlate with the measured strain val-

ues.

3. The resulting theoretical normal stress and shear stress of the element from a

combination of stress values from unsymmetrical bending moments for Section

AA and Section BB, and torsion and transverse shear for the rosettes, was

summed.

4. The experimentally measured maximum strain range within the three second in-

terval was found and multiplied by the modulus of elasticity for Section AA and

Section BB (normal stress), and the modulus of rigidity for the rosette (shear

stress).

5. The wind pressure magnitude, P, was solved by dividing the experimental stress

range in Step 4 by the theoretical stress range in Step 3.

The five step process to determine the equivalent static wind pressure was performed for

all three second intervals of the three hours and thirty minutes of usable data. A wind di-

rectionality unit vector was formed for each interval. This resulted in a total of 4,200

wind pressure magnitude values per gauge.


191

Wind Velocity vs. Wind Pressure

The maximum wind pressure magnitude, P, occurring within each three second

interval was calculated for every stain gauge at Section AA and Section BB, including

the rosettes. The calculation was performed in the manner described by the five step

process listed. The wind pressure magnitudes resulting from the five steps were plotted

versus the corresponding three second wind velocity (average wind for the three second

duration) and are shown in Figure 8.19 through Figure 8.21 for Section AA, Section BB,

and rosettes, respectively; and shown together in Figure 8.22.

FIGURE 8.19 Wind velocity vs. wind pressure for Section AA.
192

FIGURE 8.20 Wind velocity vs. wind pressure for Section BB.

FIGURE 8.21 Wind velocity vs. wind pressure for Rosettes.


193

FIGURE 8.22 Wind velocity vs. wind pressure for all sections.

A profound curve can be seen from the figures, characteristic to a parabola which

coincides with the fundamental fluid mechanics relationship of proportionality between

pressure and the velocity squared. The spread observed at higher velocities was due to the

limited number of data points measured at the velocity range.

A regression analysis was performed to determine a best fit line of the plotted data

to simulate the parabolic curve. This was done through a transformation regressor lineari-

zation process similar to the process performed for the offsetting procedure. An average

was taken of the data points for all sections and gauges. The independent variable (wind

velocity on the abscissa axis) was squared and was plotted versus its corresponding wind

pressure magnitude (see Figure 8.23).


194

FIGURE 8.23 Transformed wind velocity vs. wind pressure.

The linearization of the transformation proved the parabolic nature of the data. A

best fit line was then constructed as a linear predictor to acquire a linear equation of the

transformed data. Reversing the transformation, and using the best fit line equation, a

parabolic trendline of the data was produced and is shown in Figure 8.24. The slope of

the developed trendline was equal to 0.0025, with a y-intercept equal to 0.402.
195

FIGURE 8.24 Wind velocity vs. wind pressure trendline.

Infinite-Life Approach

The infinite-life approach was used in the same fashion as the Supports Specifica-

tions. The wind velocity that was exceeded only 0.01% of the time was calculated using a

Rayleigh distribution density function (Eq. 8.9).

−πv 2
2
PE (v) = e 4v [Eq. 8.9]

where
PE (v) = probability
v = wind velocity
v = mean wind velocity.
196

The wind velocity that exceeds the mean wind velocity 0.01% of time was calculated us-

ing Eq. 8.9, and referred to as the fatigue wind velocity (referred to as the limit-state wind

velocity in the Supports Specifications. Using the same annual mean wind velocity equal

to 11 mph (5 m/s) as the Supports Specifications, the fatigue wind velocity calculated

with Eq. 8.9 was found to be 38.0 mph (17 m/s). As shown in Figure 8.25, the equivalent

static wind pressure magnitude corresponding to a 38 mph (17 m/s) wind velocity was

equal to 4.01 psf (192 Pa).

FIGURE 8.25 Fatigue load due to natural wind.

It was concluded from the results of this project that the fatigue load due to natural wind

gust determined from experimental analysis was equal to 4.01 psf (192 Pa).
197

CHAPTER 9

THEORETICAL CALCULATION OF THE FATIGUE LOAD


DUE TO NATURAL WIND GUST

Chapter Overview

A description of the theoretical program for evaluation of the fatigue load due to

natural wind is presented in this chapter. The fatigue load due to natural wind was calcu-

lated and compared to the experimental value determined in Chapter 8: Experimental

Calculation of the Fatigue Load due to Natural Wind Gust. The theoretical program was

developed as a hybrid of experimental and theoretical data. The process was developed

similar to the Supports Specifications natural wind fatigue provisions with adaptations to

account for the variety of sign support structures in design, each with different configura-

tions, materials, and dynamic behavioral properties.

Significance of the Theoretical Program

The objective was to provide a unified design method for fatigue loading on

highway overhead sign support structures due to natural wind gust. Overhead sign sup-

port structures are highly flexible with low damping properties, which makes them sus-

ceptible to vibratory induced fatigue loading. The magnitude of this load is dependent on

the dynamic behavior and characteristics of the structure itself. The intention was to de-

velop a relationship between fatigue loading and the dynamic response in terms of ran-

dom vibration analysis as a single degree-of-freedom (SDOF) system. Examples include


198

cantilever- and bridge-type overhead sign support structures, as well as variable message

sign (VMS) structures (Figure 9.1). For analysis purposes, these structures were ap-

proximated as a SDOF system because the modes of vibration are significantly separated

such that the vibration in response to randomly applied wind loading is controlled pre-

dominately by a single modal shape. For that reason, the modal shapes were estimated as

vibrating independently from each other in single global directions (10, 20).

Bridge-Type Overhead Sign Support

Cantilever-Type
Overhead
Sign Support

Bridge-Type Overhead VMS Support

FIGURE 9.1 Highway overhead sign support structures.


199

The fatigue provisions for natural wind in the Supports Specifications are ade-

quate within certain limitations. They were developed based on four particular catego-

rized structural types. The structural response of one overhead signal support structure,

one cantilever-type overhead sign support structure, and two luminaire support structures

to natural wind excitation were analyzed. The transmitted stresses of each structure were

averaged, and the fatigue provisions were developed from the averaged results (52). The

Supports Specifications are therefore only applicable to structures of the type mentioned

with the same dynamic properties, most importantly the natural frequency and critical

damping percentage (a.k.a. damping ratio). Differences in these properties, such as the

case with bridge-type sign support structures and VMS support structures, are not ac-

counted for nor addressed in the Supports Specifications.

This study provides a detailed approach to handle cases that have different dy-

namic properties than those used to develop the Supports Specifications. Cantilever-type

sign support structures can have different configurations and made with various materials

and cross sectional shapes. These parameters will dramatically affect the magnitude of

the fatigue load. Thus, a method that incorporates the specific dynamic properties of the

structure is needed for estimating the fatigue load. Bridge-type sign support structures

and VMS support structures, which are not covered by the Supports Specifications, can

also be addressed with the proposed design method. The primary differences in these

structures in comparison to conventional cantilever-type structures are related to stiffness

and mass, which are directly related to the natural frequency and damping of the struc-

ture.
200

Methodology

The development of the fatigue guidelines in the Supports Specifications utilized

the Davenport natural wind velocity power density spectrum (PDS) curve for simulating

natural wind excitation in conjunction with the infinite-life approach for fatigue loading

(13, 52). The method presented employed the same PDS excitation and infinite-life ap-

proach, as well as a PDS excitation developed from the experimental wind velocity data

collected. However, the response of the structure due to this excitation was evaluated dif-

ferently in this research than used in the Supports Specifications. The analysis of the re-

sponse was based on principles of random vibration in utilization of the vibration re-

sponse spectrum (VRS) (39, 40-49, 56). This was done in order to account for the

uniqueness and individuality of sign structures regarding their dynamic properties, which

has significant affect on stresses generated from natural wind fatigue loading. The ap-

proach is equivalent to determining an equivalent static wind load, which produces the

same response on the structure as a randomly applied dynamic wind load. The fatigue

design equivalent static wind load is chosen from the VRS in terms of the natural fre-

quency and damping ratio of the structure. The proposed method can be used as a tool to

determine the appropriate design fatigue load for the particular structure in question dur-

ing the design phase of the project.

Structural Excitation

A comparison was made between the structural excitation developed for the Sup-

ports Specifications and the data collected with this project. Descriptions on the devel-

opment of both approaches are provided and compared.


201

Supports Specifications

The method used to development the natural wind provisions was based primarily

on spectral analysis in collaboration with the infinite-life approach to fatigue design. The

estimation of the structural excitation due to natural wind involved predicting the natural

wind environment for which the structure was to be exposed. This was done using a spec-

tral analysis based on A.G. Davenport’s wind velocity power density spectrum shown in

Eq. 9.1 (13):

4κV102 x 2
Sv ( f ) = 4
[Eq. 9.1]
f (1 + x )
2 3

where
S v ( f ) = wind velocity power spectral density at any height
f = frequency
V10 = mean wind velocity at a stadard height of 10 meters above ground level
κ = surface drag coefficient (Table 9.1)
1200 f f
x= 2
with 2 in cycles per meter.
V10 V10

TABLE 9.1 Terrain Coefficients (13)

Type of Surface κ α
Open unobstructed country (e.g., prairie-type grass-
0.005 0.15
land, arctic tundra, desert)
Country broken by low clustered obstructions such as
0.015 - 0.020 0.27 – 0.31
trees and houses (below 10 m high)
Heavily built-up urban centers with tall buildings 0.050 0.43
202

Davenport developed the wind velocity PDS curve from 70 experimental wind

velocity data collections from various locations around the world. His intention was to

develop a model which simulated the turbulence and gustiness of wind velocity. He de-

veloped Eq. 9.1 from the 70 experimental data collections. The equation is a function of

wind velocity frequency with respect to a mean wind velocity at a specified height. His

formulation is shown in Figure 9.2 for frequencies ranging from 0 to 10 Hz, an open ter-

rain (see Table 9.1), and an annual mean wind velocity of 11 mph (5 m/s):

FIGURE 9.2 Wind velocity PDS for annual mean wind velocity.

Once the behavior of the wind velocity environment was estimated, the PDS was

transformed into a wind force PDS by using principles related to fluid mechanics. The

drag force induced onto a structure due to natural wind is proportional to wind velocity

squared, as shown by the Eq. 9.2:


203

1
FD = ρCd AV 2 [Eq. 9.2]
2

where
FD = drag force
kg
ρ = density of air = 1.22
m3
C d = drag coefficient
A = area of exposed surface
V = wind velocity at any height.

By utilizing the proportionality between drag pressure and wind velocity, a wind pressure

PDS was developed from Davenport’s wind velocity PDS shown in Eq. 9.3. The plotted

equation is shown in Figure 9.3 for an annual mean wind velocity of 11 mph (5 m/s) and

normalized for exposed area and the drag coefficient.

S F ( f ) = ρ 2 C d2 A 2V 2 S v ( f ) [Eq. 9.3]
204

FIGURE 9.3 Force PDS for annual mean wind velocity.

The PDS curve accounts for the gustiness and turbulence of wind velocity over a

spectrum of frequencies, and was based on an averaged wind velocity taken at a specified

height above ground level. Since most support structure are at or around 32 ft (10 m) in

height, the PDS curve is well suited for these types of structures. Yet, the PDS can be

used at any particular height by using the power law profile shown in Eq. 9.4 for ap-

proximating variation in wind velocity with height:

V = V10αz α [Eq. 9.4]

where
V = wind velocity at height z
α = surface coefficien t in Table 8.1
z = height above ground.
205

In this case, where the objective was concentrated on formulating a design code

for fatigue wind, the wind velocity variable in the pressure PDS equation was taken at the

standard height of 32 ft (10 m) above ground level, and kept uniform across the wind ex-

posed façade of the structure. The purpose of which was to provide a simplified design

equation for commercial use. Some conservative formulation exists as the wind velocity

typically increases from the ground level upwards (13).

Once the natural wind environment was estimated, the next step was to apply the

PDS to the infinite-life approach. Since the force spectrum was based primarily on the

annual mean wind velocity, the wind velocity that was exceeded 0.01% of the time was

found and referred to as the limit-state wind velocity, or fatigue wind. The force spectrum

was then calculated using the limit-state wind velocity.

Wind velocity is random in nature, but it can be predicted though statistical rela-

tionships. It has been found through many experiments that the magnitude of the wind

velocity vector will follow a Rayleigh distribution (54). By using the Rayleigh distribu-

tion, the wind velocity that has a probability of exceedence equal to 0.01% was found

through the relationship in Eq. 9.5 as a function of the annual mean wind velocity.

−πv 2

PE (v) = e 4V
2
[Eq. 9.5]

where
PE (v) = probability
v = wind velocity corresponding to the probability
V = mean wind velocity.
206

An analysis was conducted to determine which annual mean wind velocity to use in

Eq. 9.5 to determine the limit-state wind velocity (wind velocity with a 0.01% ex-

ceedence probability). The annual mean wind velocities of 59 major U.S. cities were ana-

lyzed. It was found that an annual mean wind velocity of 11 mph (5 m/s) was exceeded in

only 19% of the U.S. cities analyzed and was therefore chosen. By plugging in 11 mph in

Eq. 9.5, and solving for the wind velocity corresponding to the 0.01% probability, a limit-

state wind velocity was found to be equal to 38 mph (17 m/s). The force spectrum was

then formed using the limit-state wind velocity (see Figure 9.4) and was used as the natu-

ral wind velocity prediction model for structural excitation for the natural wind fatigue

provisions of the Supports Specifications.

FIGURE 9.4 Force PDS using limit-state wind velocity 37


mph.
207

Experimentally Collected Data

A PDS was developed using the wind velocity data collected with this research.

Only the data considered usable (wind data directed on the front face of the structure)

was used for the analysis, and was the same excitation and response data used to deter-

mine the fatigue load from the experimentally collected data presented in Chapter 8. A

velocity PDS of each data collection event was developed and a best fit trendline was

created of the curves. The best fit line was then transformed into a wind pressure PDS

using the same procedure used for the Supports Specifications excitation. The trendline

was compared to the Supports Specifications and both were used for the structural re-

sponse analysis involving construction of the VRS.

Power Density Spectrum

Each wind velocity time history of the usable collected data was developed into a

PDS. The time domain was transformed into the frequency domain through the Fourier

transform (Eq. 9.6). The PDS was calculated by taking the Fourier transform and multi-

plying it by its conjugate, dividing by its period, and then taking the limit as the period

approaches infinity (Eq. 9.7) (39, 42).

∫ x(t )e
− j 2πft
X(f ) = dt [Eq. 9.6]
−∞
208

where
X ( f ) = Foureir transform
x(t ) = time history
f = frequency
t = time
j = imaginary

lim1
S( f ) = X ( f )X *( f ) [Eq. 9.7]
T →∞T

where
S ( f ) = Power density spectrum
X ( f ) = Fourier transform
f = frequency
T = Period
X * ( f ) = Complex conjugate of the Fourier transform

Power spectral density curves are particularly useful for this application. They are ideal

for random vibration analysis due to the inherent statistical properties of the time history

that can be extracted in relation to the vibratory nature of the structure. The area under

the PDS curve is equal to the mean square value. The square root of the mean square

value is equal to the root-mean-square (RMS). For cases where the mean is equal to zero,

the RMS is equal to the standard deviation (18, 39, 42). The developed PDS curves using

Eq. 9.6 and Eq. 9.7 are shown in Figure 9.5, along with the average PDS.
209

FIGURE 9.5 Experimental wind velocity PDS.

Approximation of the Experimental Wind Velocity PDS

A best fit line was developed that approximated the PDS average curve in Figure

9.5. A mathematical expression was needed that followed the curvature of the average

PDS curve. On the log-to-log plot in Figure 9.5, the average PDS curve was viewed as bi-

linear. The magnitude of the ordinate (y-axis) and abscissa (x-axis) data points were

transformed into a log-to-log format so that the ordinate and abscissa axes would be a

linear relationship on a standard linear plot without altering the curvature of the plot. This

was done by taking the logarithm with a base 10 of the ordinate and abscissa values. The

resulting plot is shown in Figure 9.6. The transformed plot (Figure 9.6) was subdivided

into two sections that were observed to be linear. A linear trendline was fit to each sec-

tion. The equation of the trendline was extracted and used as an approximation of the bi-
210

linear curvature of the PDS plot. The logarithmic ordinate and abscissa axes were then

transformed back to its original values using the logarithmic identity in Eq. 9.8.

y = log 10 x ⇔ x = 10 y [Eq. 9.8]

FIGURE 9.6 Logarithmic transformation of average wind velocity PDS.

The next step was plotting the trendline equations onto logarithmic axes. Equiva-

lent power equations, that represented the linear equations on a logarithmic axis, were

developed of the two sectioned plot. The resulting plot is shown in Figure 9.7. The power

equations in Figure 9.7 as a result describe the true curvature of the plot on both a loga-

rithmic and linear axis, and can be used as a mathematical expression approximating the

average wind velocity PDS. The approximation was capped at the lower frequencies by a
211

constant distribution (see Figure 9.8) and used for the remainder of the analysis. As

shown in the figure, the theoretical simulation of the experimental wind velocity PDS av-

erage closely followed the curvature and was viewed as a close approximation of the ex-

perimental curve.

FIGURE 9.7 Best fit line approximating the average wind velocity PDS.
212

FIGURE 9.8 Theoretical plot of the experimental average wind velocity PDS.

Conversion into the Wind Pressure PDS

The theoretical wind velocity PDS approximation was converted into a wind pres-

sure PDS in the same fashion as the Supports Specifications force excitation PDS. Equa-

tion 9.3 was used for the wind velocity-to-pressure conversion, resulting in the plot

shown in Figure 9.9.

It is important to point out that the plot in Figure 9.9 represented the average wind

pressure PDS and not the 0.01% exceedence probability wind pressure PDS used in the

infinite-life approach to fatigue design. It simulated the wind gustiness and turbulence of

an average wind velocity equal to 12.96 mph (5.79 m/s). Based on the experimental na-

ture of the plot, and the process in its development, the 0.01% exceedence probability

was not found exclusively using the average excitation plot in Figure 9.9, but rather
213

found by creating wind pressure PDS curves for each natural wind event measured ex-

perimentally and approximated much like the process used to find the experimental fa-

tigue load. The distribution of each wind pressure PDS was found and plotted according

to the corresponding wind velocity. The distribution was simulated and the value at the

0.01% exceedence probability was found. A more detailed description of this process is

provided in the Structural Response section of this chapter.

FIGURE 9.9 Approximation of the experimental average wind pressure PDS.

Comparison between the Supports Specifications and the Experimental Excitation

A comparison was made between the wind pressure excitation used in the Sup-

ports Specifications and the one developed in this project. The average wind velocity was

used for the comparison. The code used an average wind velocity equal to 11 mph (5
214

m/s) whereas the experimental excitation was based on an average wind velocity equal to

12.96 mph (5.79 m/s). A plot of the comparison is shown in Figure 9.10.

FIGURE 9.10 Comparison of experimental to Supports Specifications PDS.

The comparison plot shows the experimental PDS was greater than the excitation

used in the Supports Specifications. This should be the case since the average wind veloc-

ity of the experimental PDS was greater. However, the fatigue load was based on the

wind velocity with a 0.01% exceedence probability and not the average. Based on the re-

sults obtained later in the analysis of this chapter, the PDS equation for the Supports

Specifications excitation generated a PDS for the fatigue wind that was greater than the

experimental PDS, which explained the slightly larger fatigue load due to natural wind.

The experimental PDS was based on an approximation, found through a fitted curve of
215

experimental results that estimated what the PDS would be for a 0.01% exceedence prob-

ability fatigue wind.

Structural Response

Response Power Spectral Density and the Root-Mean-Square

Stresses are induced onto a structure when it vibrates, and because of the high

flexibility and low damping properties of sign support structures, the vibratory stresses

are enhanced as a direct result of the structure’s dynamic characteristics. In the structural

response analysis, the excitation on the structure from the randomly applied load, and the

subsequent random vibration response of the structure, were approximated through basic

principles of structural dynamics in utilization of the VRS. The VRS in this context was a

tool for determining the load transmitted onto the structure from the vibratory response

created by the natural wind excitation described in the Structural Excitation section of

this chapter.

The first step in the formulation of the VRS was determining the response of the

structure from the wind pressure PDS excitation. The vibration behavior of sign support

structures when excited were approximated as a SDOF system in each of their major

global directions. This was because the majority of vibration was in the first two modes

of vibration as shown in Chapter 7 Experimental Modal Analysis. Importantly, modal

shapes of the first two modes were in distinct individual directions, which behaved inde-

pendently from each other (mode 1 horizontal, and mode 2 vertical). In addition, the am-

plitude of the first modal frequency, the horizontal vibratory motion of the truss which

created torsion about the shaft support, was much larger than the amplitude of the second
216

modal frequency, and therefore controlled much of the vibration in response to the natu-

ral wind loading condition. This is not to imply that the second modal shape did not play

a part, or had no influence in the response behavior, but the significance of this mode as

well as the other higher modes of vibration were small in comparison to the first. In view

of this, the dynamic behavior of the system can be approximated as a SDOF.

The implication of developing the VRS was to account for the different types of

sign support structures, each with differing configurations, sizes and shapes, and materi-

als that have a direct influence on the dynamic characteristics of the structure. In the VRS

development, the dynamic characteristics of the structure such as the natural frequency

and damping ratio were kept variable in the calculation. In line with the directive needed

in the VRS development, the response of a SDOF system due to the wind pressure PDS

excitation developed in the Structural Excitation section of this chapter was calculated

using Eq. 9.9. (39, 41)

  f  
2

1 +  2ξ  
  fn  

Uˆ PPSD ( f ) = YˆPPSD ( f )
 2 2 2
  f    f   [Eq. 9.9]
1 −    +  2ξ  
  f n    fn  

217

where
Uˆ PPSD ( f ) = response pressure PSD as a function of frequency, Pa 2 /Hz (psf 2 /Hz)
Yˆ ( f ) = excitation pressure PSD as a function of frequency, Pa 2 /Hz (psf 2 /Hz)
PPSD

f = frequency, Hz
ξ = damping ratio
f n = natural frequency of the structure, Hz

Equation 9.9 was used to calculate the response pressure PDS for individual natu-

ral frequencies and damping ratios of the particular structure of interest. It was formed

using fundamental Structural Dynamics related to a SDOF system. The divisor term in

Eq. 9.9 is referred to as the transfer function. It was multiplied by the excitation PDS

equaling the response PDS. The units were in pressure. It represented the pressure load in

the form of a PDS curve that was transmitted onto the structure resulting from the vibra-

tory response created by the wind pressure PDS excitation. Dynamic amplification of the

structure due to the applied loading was therefore inherent in Eq. 9.9, accounting for the

dynamic properties that govern the response behavior. A plot of the excitation and re-

sponse calculated using Eq. 9.9 is shown in Figure 9.11, using a the natural frequency of

the first mode equal to 1.61 Hz and the modal damping of the first mode equal to 1.82%

(found in Chapter 7: Experimental Modal Analysis).


218

FIGURE 9.11 Structural response to wind pressure excitation.

The spike shown in the response curve of Figure 9.11 was located at the natural

frequency used in the Eq. 9.9. Other frequency and damping values can be used in the

equation to account for other support structures with differing dynamic properties. The

results would have different spikes than the one depicted in Figure 9.11 of a 1.61 Hz

natural frequency. In addition, the damping value influences the width and height of the

spike. The spike would subsequently shrink for higher damping values, and increase for

lower values, altering the area under the response curve.

The area under the response curve calculated using Eq. 9.9 is equal to the mean

square value. The square root of the area curve is referred to as the root-mean-square

(RMS). The RMS was a value used in this study for determining peak amplitudes within

a random vibration structural response. It is representative of the variance of vibration


219

amplitudes about a mean value in response to the turbulence and gustiness of the natural

wind excitation. The RMS of the response pressure PDS for individual natural frequen-

cies and damping ratios was determined by integrating the response pressure PDS over

the frequency domain, and then taking the square root, as shown by Eq. 9.10 (39, 412).

 
   f  
2

 1 +  2ξ   

   fn  
 
U PRMS ( f n , ξ ) = ∫  YˆPPSD ( f ) df [Eq. 9.10]
0  
2 2
 f   
2
 f   
 1 −    +  2ξ   
   f n    fn   
  

where
U PRMS ( f n , ξ ) = overall response pressure RMS as a function of the natural frequency
and damping ratio, Pa (psf)

For the response curve in Figure 9.11, using a natural frequency equal to 1.61 Hz and

critical damping percentage equal to 1.82%, the square root of the area under the curve

(RMS) calculated using Eq. 9.10, was equal to 0.284 psf (13.6 Pa). The calculated RMS

value is commonly referred to as the overall level of the structural response PDS. The

overall level was calculated for SDOF dynamic systems with individual natural frequen-

cies and damping percentages and plotted to form the VRS that was used to calculate the

fatigue load.
220

Vibration Response Spectrum

The VRS is a plot of the RMS of the response (Eq. 10) versus the range of natural

frequencies used in the equation. The resulting VRS provides the transmitted pressure

RMS onto the structure due to the dynamic amplification of the structure in relation to its

specific dynamic characteristics. Take for example Figure 9.12. The same base input was

applied to n number of structures. However, each structure had different mass and stiff-

ness properties, which gave it different natural frequencies of vibration, fn, and therefore

each individual structural response to the base excitation was different. Meaning, there

were different response PDS curves for each natural frequency (calculated from Eq.

9.10). For each response curve, the RMS was determined, and then plotted along with its

corresponding natural frequency to form the pressure VRS shown in Figure 9.13.

ü1 ü2 ün

m1 m2 mn

k1 c1 k2 c2 kn cn ÿ

fn,1 fn,2 fn,n

FIGURE 9.12 Response of n SDOF systems to common excitation input.


221

FIGURE 9.13 RMS wind pressure VRS for 1.82% damping.

The VRS in Figure 9.13 was from the excitation developed from the experimen-

tally collected measurements, and used a critical damping percentage of 1.82% represent-

ing the tested structure. The plot shows a decrease in RMS wind pressure as the natural

frequency increase in large part due to the lessoned proximity of the resonant frequency

of the structure to the peak amplitude of the excitation PDS. The proper use of the VRS is

to select the natural frequency of the structure of the first modal shape with vibratory mo-

tion in the direction of the structure. For the tested structure, the correct frequency to use

was the first modal frequency equal to 1.61 Hz. The corresponding ordinate value was

equal to 0.284 psf (13.6 Pa) as calculated before.


222

Natural Frequency

Sign support structures have a variety of modal shapes, but because of the large

separation between modes with vibration in the direction of the loading, they vibrate pre-

dominately independent from each other in distinct single directions. The appropriate

natural frequency to use in the VRS must correspond to the modal shape that has motion

in the direction of the natural wind loading. The most critical loading scenario for natural

wind is directed normal to the plain of the sign (in the direction of traffic), most com-

monly referred to as the horizontal modal shape (see Figure 9.14). For sign support struc-

tures, the horizontal modal shape is generally around 1 to 3 Hz, which typically corre-

sponds to the first modal shape for cantilever-type structures, and the second modal shape

for bridge-type structures.

Tip of Post Support

Direction of
Wind Loading Base Plate

Vibratory Modal Shape in the Di-


rection of Loading (Horizontal)

FIGURE 9.14 First modal shape equal to 1.61 Hz.


223

Finite element software (i.e., SAP2000) can be used to estimate the appropriate

modal shapes and their associated natural frequencies to use with the VRS curves. If FEA

software is not available, fundamental structural dynamics of a SDOF system (Eq. 9.11)

can be used for estimating these values (10, 17, 33, 56, 57, 61, 63).

1 K
fn = [Eq. 9.11]
2π M

where
f n = natural frequency, Hz
K = generalize d stiffness, N/m (lb/ft)
M = generalize d mass, kg (slug)

A recommended methodology in estimating natural frequencies and their associated mo-

dal shapes for overhead sign support structures using Eq. 9.11 can be found in the work

performed by Creamer et al. (10), which also contains useful calculation examples. It is

understood that Eq. 9.11 would be predominately used by engineers in commercial appli-

cations, as FEA is not commonly exercised in industry.

For example purposes, the first two modal frequencies of the structure were calcu-

lated using Eq. 9.11. The values were compared to the natural frequency values deter-

mined from the experimental modal analysis in Chapter 7: Experimental Modal Analysis.

The results of the comparison are shown in Table 9.2.


224

TABLE 9.2 Natural Frequency Comparison

Mode Natural Frequency Percent Difference


Number Direction Eq. 9.11 Experimental from Experimental
Horizontal
1 1.73 Hz 1.61 Hz -7.45%
(Torsion)
Vertical
2 1.82 Hz 1.64 Hz -10.98%
(Rocking)

The values are similar, proving that Eq. 9.11 can be used as an estimation of these

values if FEA software is not available. In most cases, the modal natural frequencies with

horizontal and vertical vibratory directions will typically be around 2.0 to 3.0 Hz for can-

tilever- and bridge-type sign support structures, with the frequency of the horizontal

mode slightly lower than the vertical mode.

Critical Damping Percentage

Damping also had an effect on the response of the structure as it vibrates. Take for

example the VRS plot in Figure 9.15, with the same excitation but a damping value de-

creased to 0.5%. With natural frequency equal to 1.61 Hz, the RMS wind pressure be-

came equal to 0.412 psf (19.7 Pa), a significant increase. This was true for any structural

vibration frequency. As a result to a decreased damping value (worse case as opposed to

an increased value), the spike in the response PDS increased in height and width, creating

additional area under the curve as seen in Figure 9.16 as compared to Figure 9.11, and

subsequently increasing the RMS.


225

FIGURE 9.15 RMS wind pressure VRS with damping equal to 0.5%.

FIGURE 9.16 Response PDS with damping equal to 0.5%.


226

A structure with no damping will theoretically vibrate forever. Apply damping,

and the structure will slowly stop vibrating after a period of time. The length of time de-

pends on the damping value, and the vibration frequency. During this time, stress is in-

duced onto the structure. For structures with low damping ratios, longer periods of vibra-

tion at high amplitudes will result, during which time potentially damaging stress is

transmitted onto the structure. If the stress is higher than the endurance limit of the con-

nection detail, the fatigue life of the structure is decreased with each cycle of vibration.

With higher natural frequencies of vibration, the transmitted load is decreased; but more

importantly, damping becomes less important. Damping becomes more of a factor as the

natural frequency of the structure decreases.

Damping is especially relevant to sign support structures because of their rela-

tively high flexibility and subsequent low natural frequencies (1 to 3 Hz). What’s more,

their damping ratios are mostly below 2.0%. A low damping will allow the structure to

vibrate longer at high amplitudes, and thus produce more stress that could potentially

cause fatigue damage. Damping ratios can vary depending on the structural material, and

therefore actual values can be obtained from experimental data of comparable structures

if available.

Peak-to-Peak Stress Range

The design equation for fatigue due to natural wind loading must be representa-

tive of the amplitude peak-to-peak stress range that is induced on the structure during

common everyday vibration. The majority of the structural vibration in response to natu-

ral wind excitation is controlled predominantly by a single modal frequency, the basis for
227

which validates approximating the response as a SDOF system. As a result, the response

wind pressure PDS resembled a narrow-banded spectrum concentrated about the natural

frequency. The RMS value of the spectrum embodies the variance of the vibration ampli-

tudes from a mean, which symbolizes the response of the structure due to the turbulence

and gustiness of natural wind. During a single transient event from a common everyday

natural wind gust, the peak-to-peak ranges of the response will initiate at its largest value

and then decay at rate indicative of the damping ratio. The design fatigue equation must

exemplify the largest range created on the outset of this event. In view of this, given that

the response is predominately controlled by a single modal frequency, and the gustiness

and turbulence of natural wind is typified by the RMS value, the averaged initial peak-to-

peak range of vibration was estimated as a constant-amplitude sinusoid (52). Therefore,

the initial peak amplitude was approximated as the square root of two times the RMS

value, and then doubled to form the initial peak-to-peak range, as shown by Eq. 9.12.

Pressure Range = 2 × 2 × U PRMS ( f n , ξ )


[Eq. 9.12]
= 2.8U PRMS ( f n , ξ )

where
Pressure Range = peak - to - peak pressure range, Pa (psf)
U PRMS ( f n , ξ ) = overall response pressure RMS as a function of the natural frequency
and damping ratio calculated using Equation 3, Pa (psf)

The plot in Figure 9.13 represents the wind pressure in terms of the RMS values

only, and does not represent peak-to-peak ranges. To predict the peak-to-peak pressure

range for design considerations, the VRS chart of RMS values (Figure 9.13) was factored
228

by 2.8 (Eq. 9.12). The resulting plot is shown in Figure 9.17. The tested structure with a

natural frequency equal to 1.61 Hz, and a critical damping percentage equal to 1.82%

would require a wind pressure representing the peak-to-peak amplitude equal to 0.800 psf

(38.3 Pa).

FIGURE 9.17 VRS of peak-to-peak amplitude for 1.82% damping.

Infinite-Life Approach

The infinite-life approach to fatigue design requires the wind pressure from an

average wind velocity that has a probability of exceedence equal to 0.01%. This value

was found by using an approximation through a best fit line of individual VRS of differ-

ing average wind velocities. The VRS in Figure 9.17 was of an average wind velocity

equal to 12.96 mph (5.79 m/s). This average represents all usable collected data. To find
229

the 0.01% VRS, a range of VRS curves were developed of differing average wind veloci-

ties. This was done by dividing the usable data into 10 minute sections, each with ranging

average wind velocities, and forming the peak-to-peak amplitude VRS (as in Figure 9.17)

for each 10 minute segment. A plot was made of the wind pressure extracted from the

VRS for the 1.61 Hz natural frequency and 1.82% damping tested structure versus the

corresponding average wind velocity of the segment used to develop the VRS. The result

of the plot is shown in Figure 9.18 along with the trendline developed from experimen-

tally collected strain data detailed in Chapter 8: Experimental Calculation of the Fatigue

load due to Natural Wind Gust.

FIGURE 9.18 Peak-to-peak VRS plots versus average wind velocity.


230

The peak-to-peak VRS points in Figure 9.18 follow the experimental trendline

developed from the strain gauges in Chapter 8. They mostly lie underneath experimental

trendline which was to be expected because the VRS points are based on a SDOF,

whereas the experimental trendline was based on actual structural response measurements

of a true multiple degree-of-freedom system. The proximity of the VRS points show that

approximating the structure as a SDOF is justified. The VRS point in black represents the

average VRS described in detail and shown in Figure 9.17. The variation in ordinate val-

ues of the VRS points was due to the accuracy of the wind velocity PDS regression

analysis.

The scatter closely followed the experimental trendline for the experimentally

made response measurements. The wind velocity with a 0.01% exceedence probability

was found to equal to 38 mph (17 m/s) in Chapter 8, which was equivalent to a 4.01 psf

(192 Pa) fatigue wind pressure load. In applying the same trendline procedure in Chapter

8, by transforming the domain by squaring the wind velocity to “linearize” the data,

forming a linear trendline to the transformed data, and extracting the trendline equation

(Figure 9.19), the resulting wind pressure at a 38 mph (17 m/s) wind velocity was found

to be equal to 3.80 psf (182 Pa) (see Figure 9.20).


231

FIGURE 9.19 Linear transformation of the peak-to-peak VRS.

FIGURE 9.20 Fatigue load comparison between experimental and VMS.


232

The theoretical VMS calculated value of 3.80 psf (182 Pa) as compared to the ex-

perimental calculated value of 4.01 psf (192 Pa) was very close with a 5.24 percent dif-

ference (calculated from the experimental to the theoretical). The proximity reveals the

justification of approximating the response of the structure as a SDOF. The 3.80 psf (182

Pa) VMS fatigue load using the 0.01% exceedence probability is a factor of 4.75 times

the peak-to-peak RMS value from Figure 9.17. To account for other support structures

with differing dynamic properties, the peak-to-peak RMS amplitude VMS in Figure 9.17

was factored by 4.75. The result is the fatigue load VMS (Figure 9.21) that is in compli-

ance with the 0.01% exceedence probability and the infinite life approach.

FIGURE 9.21 Fatigue wind pressure VRS for 1.82% damping.


233

The fatigue load is extracted from the VRS plot in Figure 9.21 and used for de-

sign. The VRS plot in Figure 9.21 represents fatigue loads from an annual fatigue wind

velocity equal to 38 mph (17 m/s). By knowing the natural frequency of the structure, the

fatigue load can be determined that is based on the dynamic characteristics of the struc-

ture. The plot is of a 1.82% critical damping percentage. Vibration response spectrums

for other damping values ranging from 0.1% to 2% were created and are shown in Figure

9.22. The trend shows damping to become more influential as the natural frequency de-

creases. The fatigue values vary greatly and were therefore important to obtain a true

damping value in the fatigue analysis.

FIGURE 9.22 Fatigue wind pressure VRS for damping range.


234

Vibration Response Spectrum for the Supports Specifications

The procedure described in detail in the previous Vibration Response Spectrum

section was based on the wind excitation developed from this study. The same process

was done for the current Support Specifications wind excitation as well. The resulting

fatigue load VRS is shown in Figure 9.23. The plot was made for a 2.0% critical damping

percentage as was used in the development of the fatigue load. The plot also shows varia-

tion with annual mean wind velocity.

The code does not specify fatigue loading alteration with the dynamic characteristics of

the structure. This process was specific and original to the researchers of this project.

10000
2% Critical Damping

1000
Pressure Range (Pa)

100

1 m/s
10 2 m/s
3 m/s
4 m/s
5 m/s
6 m/s
1 7 m/s
8 m/s
9 m/s
10 m/s
0.1
0.001 0.01 0.1 1 10
Natural Frequency (Hz)

FIGURE 9.23 Fatigue wind VRS of Supports Specifications (2% damping).


235

CHAPTER 10

EXPERIMENTAL CALCULATION OF THE FATIGUE LOAD


DUE TO TRUCK-INDUCED WIND GUST

Chapter Overview

A detailed description on the processes and results of determining the fatigue load

due to truck induced gusts is presented in this chapter. The measurements made experi-

mentally from the truck gusts tests were used in the analysis. The method for data reduc-

tion is presented, along with an analysis of the results. The data was broken down into

two components: 1) structural excitation, and 2) structural response. The fatigue load was

based on a stress range developed from the evaluation.

Fatigue Load Calculation Approach

The fatigue load was determined from the experimentally measured data collected

during the truck gust tests. Only data collected during low wind days, generally less than

8 mph (3.58 m/s), was used. The strain gauge instrumentation on the shaft support Sec-

tion AA and Section BB was used for the fatigue analysis. This was because of the

placement of the gauges and their orientation along the geometry of the structure in rela-

tion to the truck induced load. The measured strain ranges were used to back-calculate

the required wind pressure to produce the measured range. The maximum load calculated

was determined as the fatigue load.


236

Collected Sample Size

A total of nine truck-induced wind gust tests were performed within two days

time. Each test run was at a specific truck speed. The allocation of test runs with respect

to speed is shown below:

 Three runs at 60 mph (26.8 m/s),

 Three runs at 65 mph (29.1 m/s), and

 Three runs at 70 mph (31.3 m/s).

The individual truck runs specific to truck speed, and the order at which they were per-

formed is provided in Table 10.1. The table lists the truck speed as read on the speedome-

ter in the truck cab, and by JAMAR Trax Flex HS recorders using 0.365 in (9.27 mm)

diameter tubes spaced at 36 in (914 mm) that were placed across the road at the sign loca-

tion.

TABLE 10.1 Order of Truck Runs and Truck Speed

Truck Speed (mph) Average


Test Truck Wind
Test Day In
Number Run Recorder 1 Recorder 2 Speed
Truck
(mph)
1 Truck 1 60 58 60 10.11
2 Truck 2 70 70 70 13.41
1 3 Truck 3 65 65 66 16.79
4 Truck 4 60 64 64 15.86
5 Truck 5 65 65 65 14.56
6 Truck 6 60 61 60 6.07
7 Truck 7 65 65 65 5.29
2
8 Truck 8 70 71 70 5.91
9 Truck 9 70 *** 71 7.99
***Device did not record truck speed on this pass. Could have been due to lane positions
237

It was discovered from the truck data that the natural wind velocity on Day 1 of

testing was too strong, and transient truck events could not be isolated from the data.

Ironically, this day was one of the highest natural wind speed days, and therefore the col-

lected data was used in the fatigue load due to natural wind analysis. Day 2 proved to be

successful due to the low wind velocity which made transient truck events easy to isolate.

In view of this, only data collected on Day 2 was used for the fatigue load due to truck

gusts analysis. This amounted to four truck runs deemed usable with one run at a truck

speed of 60 mph (26.8 m/s), one run at a truck speed of 65 mph (29.1 m/s), and two runs

at a truck speed of 70 mph (31.3 m/s).

Truck Specimen

A semi-trailer was used for the experiment. A picture of the truck and driver is

shown in Figure 10.1. The height of the semi-trailer was 13.5 ft (4.11 m) width equal to

8.5 ft (2.59 m), with a 53 ft (16.2 m) long box trailer.

FIGURE 10.1 Truck and driver for truck-induced wind test.


238

Structural Excitation

The major projects analyzing fatigue loading due to truck-induced gusts have

been gathered and reported in NCHRP Report No. 469 (14) which comprised the basis of

the current Supports Specifications fatigue provisions for truck-induced gusts. Impor-

tantly, the conclusions on truck gust fatigue in Report 469 were based on experimental

evaluation of cantilever-type VMS support structures, and not sign support structures.

Many specific observations related to truck gusts on support structures but were inde-

pendent to the type of structure were made during in Report 469. These observations

were used in this project.

Report 469 concluded that the full strength of the truck gusts pressure of 18.8 psf

(900 Pa) was produced on the underneath exposed area of the structure at 19.7 ft (6 m) or

less above the roadway. The pressure decreased linearly to zero at a height of 32.8 ft (10

m) above the roadway. The pressure load applied in the horizontal direction was found to

be insignificant compared to the fatigue load due to natural wind, and was excluded. The

underneath exposed area to truck gusts was determined to cover 12 ft (3.7 m) in length or

the length of the traffic lane, whichever was greater. The provisions indicated that the

truck gust pressure should be applied over the exposed length located where the force

would create the worst-case scenario (largest moment arm). It was specified for most

cases to be the outermost 12 ft (3.7 m) length of the structure (14).

The objective of the truck gust tests performed in this project was to evaluate the

accuracy of the truck gust provisions in the Supports Specifications as applied to cantile-

ver-type sign support structures. Many of the observations specific to the truck gusts
239

structural excitation reported in Report 469 were used in this project. These observations

are summarized as follows:

 Fatigue pressure was applied onto the underneath horizontal exposed area of the

structure,

 The exposed horizontal underneath area equaled 12 ft (3.7 m) in length,

 The exposed area was directly above the traffic lane used by the truck, and

 The gust was assumed maximum at the bottom of the sign and decreases linearly

to zero at 32.8 ft (10 m) above the roadway.

The above excitation specifics were used in the evaluation of the truck gusts on cantile-

ver-type sign supports structures of this project. These observations had no influenced, or

were influenced in their discovery, by the structural response of the structure tested when

the observations were made, and were therefore independent on the type of support struc-

ture. In view of this, these observations were assumed as factual for this project, and were

used in the evaluation of the cantilever sign structure to truck-induced wind gusts.

Anemometers AN-1 and AN-2 were placed on the structure above the traveling

lane to evaluate the speed and direction of the truck gusts. The purpose of the measure-

ments was to reinforce the accuracy of the gust pressure back-calculated from the strain

gauge measurements. AN-1 was oriented to measure the wind speed and direction on the

horizontal plane (xy plane), whereas AN-2 was oriented to measure the wind speed and

direction on the vertical plane (zy plane). Their locations are shown in Figure 10.2.

Transient truck gust events were significant and obvious in measurements made
240

from the strain gauges and accelerometers. The fatigue truck gust pressure was deter-

mined from these measurements. Analysis of the truss anemometers AN-1 and AN-2 in-

dicated no significant transient event. This may be due to the short pulse of the truck

transient and the sampling frequency of the anemometers. The instruments were set at 60

samples per second in line with the strain gauges and accelerometers. The anemometer

had a maximum sampling frequency inherent in the instrument of 4 Hz, and therefore the

remaining samples were repeated until the sampling period was renewed. The limitation

was not discovered by the researchers until after the anemometers were bought and

placed on the structure. This is viewed as the main contributor to the failure of the in-

strument to pick up the transient events.

Structural Response

The measurements made from the strain gauges comprised the majority of the

structural response. Strain gauges located on the shaft support were used primarily be-

cause of their location in line with the symmetry of the structure, and their orientation to

the truck excitation. The evaluation involved exclusively the response to a vertically ap-

plied pressure onto the exposed horizontal area on the underneath section of the structure.

The exposed are was equal to 12 ft (3.7 m) located directly above the traffic lane used by

the truck specimen. The analysis involved a back-calculation from the strain gauge meas-

urements to determine a lift moment from which the truck gust pressure was determined.

The strain gauge measurements encompassed a strain range created form the truck tran-

sient, representing a maximum peak-to-peak range from the collected data.


241

AN-4

AN-3

y
AN-2
x AN-1

Traffic
Direction

FIGURE 10.2 Anemometer layout and orientation.

Stress Range

The truck-induced gust pressure was back-calculated from a peak-to-peak stress

range determined from the strain gauge measurements. Only the working truck tests col-

lected on a low wind day were used for the analysis. This amounted to four tests of one at

60 mph and 65 mph (29.1 m/s), and two at 70 mph (31.3 m/s). The strain gauges at loca-

tion C and G (see Figure 10.3) for Section AA and Section BB were the sole gauges used

for the analysis. This was based on their orientation to the loading event, and their place-

ment on the geometric symmetry of the structure.


242

Strain gauges at the other locations, particularly locations A and E were not used

because of the unknown direction of truck gusts (because of the anemometer failure) and

the unknown gradient of the pressure with height. These two factors dramatically influ-

enced the drag moment (moment about the x axis, see Figure 10.2) that was needed in

calculating the unsymmetrical bending needed for the other gauges not oriented perpen-

dicularly along the x-axis. The back-calculation with strain gauges located at C and G did

not require the determination of the drag moment along the x axis and allowed for a more

straightforward calculation.

Strain Gauge

FIGURE 10.3 Strain gauge locations at Section AA and Section BB.

Transient Events

The evaluation of the transient events used for the truck gust analysis involved

determining the maximum peak-to-peak strain range. A typical transient is shown in Fig-
243

ure 10.4, showing in-plane vibration (mode 2, rocking) to the vertical upward force cre-

ated by the truck gusts as measured from strain gauge SGR-AA-7 located on Section AA

at location C. The length of the transient event was inconsequential for determining the

fatigue load. Due to the nature of the load, its repetition, and randomness, the fatigue load

must represent the maximum peak-to-peak stress range of the structural response. As

long as the endurance limit of the detail is greater than the maximum measured peak-to-

peak range, then significant fatigue deterioration due to truck gust is avoided during the

lifespan of the structure. The peak-to-peak strain range for each truck event is listed in

Table 10.2.

FIGURE 10.4 Maximum peak-to-peak strain range.


244

TABLE 10.2 Peak-To-Peak Strain Ranges

Strain Gauge Peak-To-Peak


Strain Gauge Truck Test Truck Speed
Location on Strain Range
Section Run (mph)
Section (microstrain)
C 8.768
Truck 6 60
G 7.729
C 7.367
Truck 7 65
G 6.405
Section AA
C 6.027
Truck 8 70
G 5.418
C 4.287
Truck 9 70
G 3.897
C 9.245
Truck 6 60
G 8.069
C 7.896
Truck 7 65
G 6.786
Section BB
C 6.444
Truck 8 70
G 5.450
C 4.562
Truck 9 70
G 3.870

The fatigue load due truck gusts was back-calculated from the measured response ranges

listed in Table 10.2.

Truck-Induced Wind Pressure Back-Calculation

The equivalent static wind load that would produce the same maximum peak-to-

peak stress range determined experimentally was back-calculated. Two measures formed

the basis of the process. First, the stress range was determined from the experimentally

measured strain ranges as described in the Structural Response section of this chapter.

Second, a theoretical structural analysis was performed on the structure to determine the

stresses at the gauged locations for comparison with the experimental values. A stress

element of normal values was developed at the location of the gauge to be used in the
245

comparison. Importantly, the stress values were developed with the wind pressure magni-

tude, P, kept as a variable. The two measures, experimental stress and theoretical stress,

were set equal to each other and the wind pressure magnitude, P, was solved (see Eq.

10.1). The instrumented sections along the shaft support (Section AA and Section BB) at

locations C and G (Figure 10.3) were used exclusively for this evaluation because of their

orientation to in-plane displacements with respect to the vertically applied truck gust.

Exp. Stress Range = (Theo. Stress Range)( P )


Exp. Stress Range [Eq. 10.1]
P=
Theo. Stress Range

Theoretical Structural Analysis

The structural analysis involved determined exposed areas and using them to form

the moment vectors needed in determining the load. The wind pressure magnitude, P,

was kept as a variable throughout the process, and solved for using Eq. 10.1 at the end of

the procedure. Point loads with direction at each exposed area was developed with drag

coefficients provided by the Supports Specifications for each member exposed to wind

and used in the analysis. The idea was to simulate as accurately as possible the wind

loading conditions. Equations of equilibrium were developed and the acting moment vec-

tor (resultant) was solved. The moment vectors were used to develop stress elements at

each strain gauged location. Equation 10.1 was used to solve for the wind pressure mag-

nitude.
246

Exposure Area Breakdown

The wind pressure was applied onto a horizontal area located on the underneath

portion of the truss overhang, in compliance with the research collected in NCHRP Re-

port 469 (14) and used in the Supports Specifications. The exposed underneath area

equaled 12 ft (3.7 m) in length. It was located directly above the traffic lane used by the

truck, as shown in Figure 10.5. The truck ran directly underneath the truss anemometers

AN-1 and AN-2, which were placed along the centerline of the second panel. The hori-

zontal area therefore included all exposed horizontal surfaces of the truss 6 ft (1.83 m) on

each side of the anemometer location. This included the two chords, one horizontal strut,

and three horizontal diagonal struts. The sign was not included in the calculation. A

breakdown of the exposed area used in the analysis is shown in Figure 10.6, with values

listed in Table 10.3.


247

Exposed
Area

Anemometers
AN-1 and AN-2
12 ft

FIGURE 10.5 Underneath exposed horizontal area.


248

12 ft (3 Panels @ 4 ft)
Chord

3 ft

I I
Horizontal Diagonal Exposed
Strut Strut Area

FIGURE 10.6 Exposed underneath truss area breakdown.

TABLE 10.3 Exposure Truss Area Breakdown

Distance of Centroid to Strain Gauges (in)


Section AA Section BB
Area
Segment Location Label (12 in above Base (16 in above Base
(in2)
Plate) Plate)
i j k i j k

Chord A1 1,008

Bottom Horizontal
A2 47.34
Truss Strut
381 0 309 381 0 305
Horizontal
Diagonal A3 298.8
Strut

Total Area 1,354.14


249

Effective Area Breakdown

The effective areas of the exposed areas listed in Table 10.3 were calculated. The

effective area represented the exposed area multiplied by the drag coefficient of the

members that made up the area.

Drag coefficients. All drag coefficients were calculated using Table 3-6 of the

Supports Specifications. Fifty year design life was used in the calculation with a velocity

conversion factor equal to 1.00 (no conversion needed). The worst case scenario was ap-

plied by maximizing the wind speed and member diameter. A maximum wind speed of

70 mph (31.293 m/s) was used, assuming the wind gust equals the speed of the truck, and

the largest diameter member equal to 3.5 in (88.9 mm). The results of the drag coefficient

calculation are listed in Table 10.4.

TABLE 10.4 Drag Coefficients

Segment Location Label Drag Coefficient, Cd

Chord A1 1.10

Bottom Truss Horizontal Strut A2 1.10


Horizontal Diago-
A3 1.10
nal Strut

Effective area calculation. The effective area was defined as the exposed area of

the segment multiplied by the drag coefficient of the members that make up the area

segment. The results of the calculation are listed in Table 10.5.


250

TABLE 10.5 Effective Area

Drag Effective
Area
Segment Location Label Coefficient, Area, Ae,i
(in2)
Cd (in2)
Chord A1 1,008 1.10 1,108.8
Horizontal
A2 47.34 1.10 52.074
Bottom Strut
Truss Horizontal
Diagonal A3 298.8 1.10 328.68
Strut
Total Effective Area 1489.554

Acting Moment Vector Formulation

The effective areas were used to develop acting moment vectors representing a vertically

applied truck gust, and equations of equilibrium were formed. Using the position vectors

listed in Table 10.3 of the distance of the centroid of the effective area to the strain gauge

locations, the active moment at the strain gauge location was calculated. This process is

described by the following system of equations.

Acting force vector. The acting pressure vector was defined as a wind directional-

ity unit vector times the pressure magnitude, P, as shown in Eq. 7.7.

P = λ P = (λ x Pi + λ y P j + λ z P k ) psi [Eq. 10.2]

Knowing the pressure is equal to the force divided over an area, the force vector was de-

fined in Eq. 10.3 in terms of the effective area calculated in Table 10.5, and keeping the

wind pressure magnitude, P, variable.


251

F = λ (ΣAe,i P)
[Eq. 10.3]
= [λx (ΣAe,i P)i + λy (ΣAe,i P) j + λz (ΣAe,i P)k ] lb

The components of the force vector in Eq. 10.3 were as follows, with the only z-

component (vertically applied force) applicable to this analysis as the x and y components

were not needed in the calculation of strain gauge locations C and G.

Fx = λ x (ΣAe ,i P ) lb
F y = λ y (ΣAe ,i P ) lb
Fz = λ z (Σ Ae ,i P ) lb

Equations of equilibrium. Once the force vectors were developed, equations of

equilibrium were created from which the acting moment equations were solved. A typical

free body diagram of the wind loading on the structure can be generalized as shown in

Figure 10.7 with all force components shown in their positive sense. The acting moment

vector (resultant in Figure 10.7) was calculated from the equations of equilibrium formed

using the free body diagram. The pressure magnitude, P, was kept variable during the

calculation process.
252

Reaction
MR
FR
Fz

Fx Equal & Opposite


Fy FR
MR
Section AA
z Resultant Section BB

x
y

FIGURE 10.7 Free body diagram.

Development of the Stress Element

A combined loading and unsymmetrical bending analysis was used to determine

the stress and strain values at the strain gauge locations. Locations C and G at Section

AA and Section BB was used for this calculation due to the placement of the strain

gauges with respect to in-plane displacement. The stress element development involved

the combination of the bending stress caused by the upward force and its moment arm,

and the normal stress caused by the upward force.


253

The bending moment was analyzed as an unsymmetrical bending. It was broken

down into two components, a drag moment (moment about the x-axis) and lift moment

(moment about the y-axis). Only the lift moment was needed because strain gauges used

in the calculation and their location and orientation on the structure. Other gauges were

not used because of the uncertainty of the horizontal pressure (applied on the vertical face

of the sign) distribution. The vertical pressure however was calculated with a high level

of accuracy.

Material Properties

The material properties needed for development of the stress element are listed in

Table 10.6. A cross section of the analysis location is shown in Figure 10.8.

TABLE 10.6 Material Properties at Section AA and Section BB.

Material Property Value


Diameter, dshaft 24 in
Area, Ashaft 20.9 in2
Thickness, tshaft 0.281 in
Moment of Inertia, IXX, shaft 1,473 in4
Moment of Inertia, IYY, shaft 1,473 in4
Modulus of Elasticity, E 29,000,000 psi
254

FIGURE 10.8 Cross section of strain gauge location Section AA and Section BB.

Relevant Stress and Strain Equations

The stress and strain equations relevant to the combined loading analysis are

listed in Table 10.7. The stress element derived from the theoretical structural analysis

was developed using these equations.

TABLE 10.7 Relevant Stress and Strain Equations

Stress Stress Equation Strain Equations


F σN
Normal σN = z εN =
A E
M y M x σM
Unsymmetrical Bending σM = − x + y εM =
I XX I YY E
255

Stress Element

A stress element was developed using the equations in Table 10.7. The stress

elements were formed with the wind pressure, P, as a variable. The values were placed on

a stress element and added together to determine the normal stresses at the gauged loca-

tion (see Figure 10.9).

FIGURE
FIGURE 10.9 Typical
10.9 Typical stress element.
stress element.

Wind Pressure Calculation

The stress elements, formed with the wind pressure, P, kept as a variable, were

compared to the experimentally measured normal stress ranges depicted in Table 10.2.

The wind pressure, P, was calculated using Eq. 10.1. The results of the calculation are

listed in Table 10.8.


256

TABLE 10.8 Truck-Induced Wind Gust Pressure

Strain Gauge
Strain Gauge Truck Test Truck Speed Wind Pressure
Location on
Section Run (mph) (psf)
Section
C 7.917
Truck 6 60
G 6.979
C 6.652
Truck 7 65
G 5.783
Section AA
C 5.442
Truck 8 70
G 4.892
C 3.871
Truck 9 70
G 3.519
C 8.348
Truck 6 60
G 7.286
C 7.130
Truck 7 65
G 6.127
Section BB
C 5.819
Truck 8 70
G 4.921
C 4.119
Truck 9 70
G 3.494

The results in Table 10.8 represent the equivalent static wind load to create an

equivalent stress range on the structure as a result of dynamic amplification and vibration

of the structure in response to the truck-induced gust impulse. The maximum pressure

was found to be equal to 8.348 psf (400 Pa) resulting from a truck speed of 60 mph (26.8

m/s). The maximum pressure was considered as the fatigue load, as it was representative

of the maximum stress range measured experimentally from trucks traveling underneath

the structure.

The pressure values calculated from the applicable strain gauges indicated that

trucks traveling at 70 mph (31.3 m/s) created slightly less response than trucks traveling

at 60 mph (26.8 m/s). This occurrence was only measured with the post gauges, as the 70

mph (31.3 m/s) truck speed controlled on the anchor bolts and chords, and therefore more
257

tests may have been needed. Alternatively, this may be due to the natural frequency of

the structure as compare to the impulse frequency of the truck-induced gusts. Trucks

traveling slower may produce a truck gust impulse with a frequency closer to the resonant

frequency of the structure than trucks traveling at faster speeds. In view of this, the re-

sults would indicate a relationship between the back-calculated wind pressure and the dy-

namic characteristics of the structure. A detail analysis of this phenomenon is provided in

Chapter 11 Theoretical Calculation of the Fatigue Load due to Truck-Induced Wind

Gust.
258

CHAPTER 11

THEORETICAL CALCULATION OF THE FATIGUE LOAD


DUE TO TRUCK-INDUCED WIND GUST

Chapter Overview

This chapter focused on determining a theoretical procedure for determining the

design fatigue load due to truck-induced wind gusts on highway overhead sign support

structures. Truck-induced wind gust is a loading generated on the structure from wind

gust created when trucks travel underneath them at highway speeds. Semi-trailer com-

mercial trucks were the focus of the research. It was considered a fatigue load due to the

large number of trucks using the highway system, and the significant wind load generated

on the overhead sign structure with each passing truck. The theoretical study primarily

involved analysis of sign support structures in relation to mechanical vibration and utili-

zation of the shock response spectrum (SRS). The fatigue load due to truck gust was ex-

tracted from the SRS as an equivalent static wind load and used for design.

Research Significance

The fatigue load is determined primarily from the structural dynamic characteris-

tics of the particular structure in question. Operational overhead sign support structures

are highly flexible with low damping properties, which make them susceptible to vibra-

tory induced fatigue loading caused by wind gust impulses generated from passing

trucks. The magnitude of this load is dependent on the dynamic behavior and characteris-
259

tics of the structure. The analysis becomes especially crucial when the frequency of the

impulse matches the natural frequency of vibration of the structure. Frequencies of vibra-

tion and the frequencies of the truck gust impulses are similar, typically around 1 to 3 Hz,

and therefore resonance issues are a major cause of fatigue damage. A method that incor-

porates the response of the structure based on its dynamic properties is needed for esti-

mating the fatigue load. The intention of the study was to develop a relationship between

fatigue loading and the structural dynamic response in terms of mechanical vibration as a

single-degree-of-freedom (SDOF) system. Examples include cantilever- and bridge-type

overhead sign support structures (Figure 11.1).

Bridge-Type Overhead Sign Sup-

Cantilever-
Type
Overhead

Bridge-Type Overhead VMS

FIGURE 11.1 Highway overhead sign support structures.


260

The fatigue provisions for truck gusts within the Supports Specifications were se-

lected from the National Cooperative Highway Research Program (NCHRP) Report 412

(1, 16, 52), and later modified due to the research detailed in the NCHRP Report 469 (14,

16). It was stated in Report 412 that the design fatigue load in regards to truck gusts was

based on the response of a single particular cantilever-type highway overhead variable

message sign (VMS). The Supports Specifications are only applicable to highway signs

with the same structural excitation and dynamic response characteristics reported. Differ-

ences in these characteristics, such as the case with other cantilever- and bridge-type sign

support structures, cannot be accounted for nor are addressed in the current Supports

Specifications.

This chapter details a universal approach to fatigue design to handle cases that

have different structural dynamic properties than those used to develop the current Sup-

ports Specifications. Cantilever-type sign support structures can have different configura-

tions and are made with various materials and cross sectional shapes. These parameters

will dramatically affect the behavior of the structure and subsequent magnitude of the

fatigue load. A method for estimating the fatigue load that incorporates the specific dy-

namic properties of the structure can as a result account for the variety of sign structures

in design. For that reason, the proposed method was considered universal. It is applicable

to any type of sign support structure of varying configuration, material, strength, mass,

and stiffness whereby its response approximated as a SDOF system. Bridge-type sign

support structures, which are not covered by the Supports Specifications, can also be ad-

dressed with the proposed design method. The primary differences in these structures in

comparisons to conventional cantilever-type structures are related to stiffness and mass.


261

Since all of these dynamic features are directly related to the natural frequency of the

structure, the method described in this paper can be used to determine the design fatigue

load due to truck gusts in specific reference to the dynamic characteristics of the particu-

lar structure in question.

Methodology

The framework of the developed method can be broken down into two major

components: 1) the truck gust excitation of the structure, and 2) the dynamic response of

the structure. Research data collected from the literature was used to simulate the truck

gust excitation for this chapter. A typical truck gust pressure loading was developed in

the form of a time history streamline. Different design truck speeds were employed in the

time history simulation, resulting in truck gust impulses occurring at different frequen-

cies, which in turn would excite different vibratory responses in the structure.

The response of the structure and consequent fatigue load due to the truck gust

excitation was evaluated differently than the methods described in the literature. The pro-

posed method was entirely new and unique with respect to overhead sign support struc-

tures. The structural response was theoretically calculated based on principles of me-

chanical vibration, and the fatigue load was determined through utilization of the SRS

(39, 40-49, 56). The load is extracted from the SRS in terms of the natural frequency of

the structure to account for the variety of sign support structures in design, and the esti-

mated speed of the truck to account for different loading impulses. The extracted value is

in the form of an equivalent static wind load which produces the same response onto the

structure as the dynamic impulse. Dynamic amplification was included within the fatigue
262

load chosen from the SRS, since the SRS was developed based on the frequency of vibra-

tion of the structure in response to the loading event. The proposed method is used as a

tool to easily determine the appropriate design fatigue load due to truck gust for the par-

ticular structure in question.

Structural Excitation

An accurate depiction of the truck gust wind excitation without influence from the

type of structure and its dynamic behavior was needed in the development of the theoreti-

cal procedure. Experimental measurements made with this project were not applicable as

the strategically placed anemometers (AN-1 and AN-2) did not pick truck wind gusts.

This was believed to be due to the sampling frequency of the anemometers. The research

performed by Cook et al, in their collaboration with the Florida Department of Transpor-

tation (12) was considered by the researchers as the best candidate for replacement.

Cook’s research involved experimentally determining the truck gust pressure load

through direct measurement of wind pressure from passing trucks. The same process was

conducted by the researchers, but to measure wind velocity. With Cook’ research, the

truck wind pressure was evaluated at varying heights above ground level using pressure

transducers placed on a rigid road bridge spanning over the highway. Time history wind

pressure impulses from 23 truck gust loading events were recorded. The findings in

Cook’s research were closely similar to other experimental research efforts on the subject

performed by different investigators in other locations (10, 14, 17), especially with

NCHRP Report 469 used to developed the current Supports Specifications.


263

Cook’s research indicated a biaxial fatigue loading event. The results indicated a

horizontal and vertical component of the wind gust pressure impulses. The horizontal

component was projected onto the facade within the vertical plane that was oriented per-

pendicular to the ground. The vertical component was projected onto the facade within

the horizontal plane that was oriented parallel to the ground (12). Only the vertical pres-

sure impulses was used with this study, as the natural wind fatigue loading criteria con-

trols wind loading in the horizontal direction. All strain values measured experimentally

were significantly higher (nearly twice as much) with natural wind gusts than measured

for the truck-induced wind gusts.

Some small liberties were taken with this study in order to simplify the pressure

impulses recorded by Cook so that they could be functional for numerical applications.

The time history impulses were transformed into a linear composition by enveloping the

samples by the triangular single cycle loading shown in Figure 11.2. This was done simi-

lar to the research performed by Ginal (37), except impulses from trucks traveling at

other speeds were also employed. The loading simulation in Figure 11.2 represented a

linear increase in wind pressure as the truck approaches, followed by an equivalent (in

magnitude) suction as the truck passes and separates from the structure. This is indicative

a wind pressure directed upward onto the underside facade of the structure as the truck

approached, followed by a suction pressure directed downward as the truck separated

from the structure (12, 37). The impulse in Figure 11.2 represents the measured values

taken at 17 ft (5.2 m) above ground level (12).


264

FIGURE 11. 2 Vertical truck-induced wind gust impulses.

The wind pressure impulse resulting from a truck traveling at 70 mph (31.3 m/s)

was used as an impulse Control (dashed line in Figure 11.2). The Control reflects an en-

velopment procedure of the recorded time history samples of trucks that traveled at this

speed. It was developed as a prediction model representative of the exposure environment

from truck gusts defined at a 90% confidence level that the actual maximum absolute

value of pressure and impulse duration will be equal to or below the Control at least 95%

of the time (12). The envelopment resulted in a linear increase and subsequent decrease

to the maximum impulse pressure equal to 1.50 psf (71.8 Pa) as shown in Figure 11.2.

These values were used for both the approach pressure and the separation pressure of the

Control for symmetrical simplification of the simulated impulse. The duration of the

Control determined by the envelopment procedure was equal to 0.80 seconds.


265

Trucks traveling at other speeds generated wind gust impulses with different fre-

quencies and durations than the Control (see Figure 11.2), and would as a consequence

excite different responses on the structure. Impulses generated at other truck speeds were

identified and employed in this method to account for a variation in excitation frequen-

cies. The impulses were normalized based on the Control impulse at 70 mph (31.3 m/s)

using the mechanics of fluids relationships between wind pressure and the square of wind

velocity, as well as Newtonian dynamic continuous motion kinematics of a particle with a

constant acceleration.

Structural Response

The structural response from truck gust loads described in the previous section

(see Figure 11.2) was approximated through basic principles of structural dynamics. For

analysis purposes, the dynamic behavior of highway overhead sign support structures was

approximated by a SDOF system. This was because the modes of vibration measured in

Chapter 7: Experimental Modal Analysis were significantly separated such that the vibra-

tion in response to transient loading was controlled predominately by a single modal

shape in independent directions. For that reason, the modal shapes were estimated as vi-

brating independently from each other in single global directions (10 ,20, 39). In this

study, the structural response of a SDOF system due to the truck gust load was calculated

using Eq. 11.1 shown in index notation, i (39, 47):


266

Pi = 2e −ξωn ∆t cos(ω d ∆t ) Pi −1
− (e − 2ξωn ∆t ) Pi − 2
+ 2ξω n ∆tpi [Eq. 11.1]
 ω  
+ ω n ∆te −ξωn ∆t  n (1 − 2ξ 2 ) sin(ω d ∆t ) − 2ξ cos(ω d ∆t ) pi −1
 ω d  

where
Pi = response pressure as a function of natural frequency and critical damping percentage, psf (Pa)
pi = truck induced wind gust pressure impulse excitation, psf (Pa)
ωn = undamped natural frequency of the structure, rad/sec
ωd = ωn 1 − ξ 2 = damped natural frequency of the structure, rad/sec
ξ = percent critical damping
t = time, sec

Eq. 11.1 calculates the response pressure time history for individual natural frequencies

and critical damping percentages of the particular structure of interest. It can be seen as

an equivalent pressure load that is transmitted onto the structure from the vibration re-

sponse created by the truck-induced excitation. The equation embodied full dynamic am-

plification based on the dynamic characteristics of the structure.

An example of the vibratory response time history of a typical sign support struc-

ture calculated using Eq. 11.1 is shown in Figure 11.3. The 70 mph (31.3 m/s) Control

impulse was used as the excitation for this example. The natural frequency of the struc-

ture was chosen with a modal shape in the direction of the vertical loading, and was equal

to 1.64 Hz found in Chapter 7: Experimental Modal Analysis. This modal shape is best

described as a rocking vibratory motion in the direction perpendicular to the ground (par-

allel to the vertical component loading). The plot shows response time histories of the

structure for an undamped case, and for a critical damping percentage equal to 1.0%.
267

FIGURE 11.3 Structural response time history due to the Control impulse.

Sign support structures typically have critical damping percentages around 1.0%

or less. Even for 1.0% damping however, the difference in magnitude of the maximum

pressure range observed between the two response cases (Figure 11.3) was very small,

and significantly lessens as the damping percentage decreases. This revealed that the

maximum peak response was not sensitive to structural damping, and therefore the un-

damped case was used in the SRS development, which adheres to other common SRS

development case studies (39, 47). For such instances, the damped natural frequency pa-

rameter, ωd, in Eq. 11.1 became equal to the undamped natural frequency, ωn.

Shock Response Spectrum

The SRS in this context was a tool for determining the maximum load transmitted

onto the structure from the vibration response created by truck gusts. The SRS was a plot
268

of maximum peak values determined from the response time histories calculated using

Eq. 11.1 for a range of SDOF systems (see Figure 11.4) with varying undamped natural

frequencies under a common loading (39, 47). The SRS plot for the vertically directed

truck gust impulse is shown in Figure 11.5. The SRS shown in Figure 11.5 is a plot of the

maximum response pressure ranges versus natural frequencies. The natural frequency

domain extends from 0 to 10 Hz, and individual SRS curves for truck speeds ranging

from 50 mph to 100 mph (22.4 m/s to 44.7 m/s) are plotted based on the impulses pre-

sented in Figure 11.2.

ü1 ü2 ün

m1 m2 mn

k1 c1 k2 c2 kn cn ÿ

fn,1 fn,2 fn,n

FIGURE 11.4 Response of n SDOF systems to common excitation input.


269

FIGURE 11.5 SRS for vertical truck-induced wind gust.

The pressure range, and subsequent stress range, was of interest when dealing

with fatigue analysis. It is important to clarify the curves in Figure 11.5 represent the ini-

tial range of the positive to negative peaks (see Figure 11.3) for each impulse. The stress

generated from the vibration caused by this pressure range will be induced onto the struc-

ture with each passing truck, with little influence from the damping properties of the

structure. The damping will allow the structure to slowly stop vibrating over a period of

time by gradually decreasing the magnitude of this range, as shown in Figure 11.3 for the

1% damping case. However, if the initial pressure range occurring at the onslaught of the

gust impulse generates a stress that is greater than the endurance limit of the material and

connection detail, then the fatigue life of the structure is greatly decreased with each
270

passing truck. And therefore, the design fatigue load must be representative of the initial

range of values plotted in Figure 11.5 in terms of the natural frequency of vibration of the

structure which, as indicated in the figure, has a significant effect on the magnitude of the

initial range.

The fatigue load is extracted from the SRS as the ordinate value corresponding to

the natural frequency of vibration with a modal shape in the direction of the excitation.

Sign support structures have a variety of modal shapes, but because of the large separa-

tion between modes with vibration in the direction of the loading, they vibrate predomi-

nately independent from each other in distinct single directions. The appropriate natural

frequency to use in the SRS must correspond to the earliest modal shape that has motion

in the direction of the truck-induced wind loading. In this case, the truck-induced load

was directed vertically onto the underneath facade of the structure. The appropriate mo-

dal shape would be a vertical vibratory motion (perpendicular to the direction of traffic)

generally associated with the second modal shape of typical cantilever-type sign struc-

tures (see Figure 11.6), and the third modal shape of typical bridge-type sign support

structures.
271

Tip of Post Support

Base Plate
Direction of
Wind Loading

Vibratory Modal Shape in the Di-


rection of Loading (Vertical)

FIGURE 11.6 Second modal shape of cantilever structure.

Finite element software (i.e., SAP2000) can be used to estimate the appropriate

modal shapes and their associated natural frequencies to use with the VRS curves. If FEA

software is not available, fundamental structural dynamics of a SDOF system (Eq. 11.2)

can be used for estimating these values (10, 17, 33, 56, 57, 61, 63).

1 K
fn = [Eq. 11.2]
2π M

where
f n = natural frequency, Hz
K = generalize d stiffness, N/m (lb/ft)
M = generalize d mass, kg (slug)
272

A recommended methodology in estimating natural frequencies and their associated mo-

dal shapes for overhead sign support structures using Eq. 11.2 can be found in the work

performed by Creamer et al. (10), which also contains useful calculation examples.

For example purposes, the first two modal frequencies of the structure were calcu-

lated using Eq. 11.2. The values were compared to the natural frequency values deter-

mined from the experimental modal analysis in Chapter 7 Experimental Modal Analysis.

The results of the comparison are shown in Table 11.12.

TABLE 11.1 Natural Frequency Comparison

Mode Natural Frequency Percent Difference


Number Direction Eq. 8.11 Experimental from Experimental
Horizontal
1 1.73 Hz 1.61 Hz -7.45%
(Torsion)
Vertical
2 1.82 Hz 1.64 Hz -10.98%
(Rocking)

The results were similar, proving that Eq. 11.2 can be used as an estimation of

these values if FEA software is not available. In most cases, the modal natural frequen-

cies with horizontal and vertical vibratory directions will typically be around 2.0 to 3.0

Hz for cantilever- and bridge-type sign support structures, with the frequency of the hori-

zontal mode slightly lower than the vertical mode.

The SRS plots in Figure 11.5 have varying peaks which are dependent on the du-

ration and frequency of the impulse excitation. It would be advisable to design structures

that have natural frequencies outside of the frequencies plotted at these peaks. For the 70

mph (31.3 m/s) Control, the maximum possible structural response from this impulse

would occur onto a structure that has a natural frequency with a vertical vibratory modal
273

shape around 1.60 Hz which is extremely close the structure resonant frequency of the

tested structure at 1.64 Hz. The transmitted pressure range from the response of the im-

pulse for this particular example structure was equal to 8.33 psf (399 Pa), taken as the

ordinate value from the SRS in Figure 11.5.

The load is to be applied as a uniformly distributed load in the manner specified

in the Supports Specifications (1, 14). Accommodation for height above ground level is

also provided in the 2002 revisions. The height provisions conservatively match the find-

ings in Cook’s report (1, 12, 14). Whereas Cook’s findings indicated a 10% reduction in

wind pressure per foot increase starting from 17 ft (5.2 m) above ground level (12), the

Supports Specifications allow for a linear reduction in wind pressure starting from 19.7 ft

(6 m) above ground level to zero at 32.8 ft (10 m) (1, 14). For that reason, the 2002 revi-

sions on pressure reduction due to height above ground level were considered applicable

to the method presented in this chapter. The design traveling speed of the truck to use in

the SRS is generally taken as the speed limit. A Control speed of 70 mph (31.3 m/s) was

chosen as an appropriate design recommendation. Other speeds are provided in Figure

11.5 for design choice, but also to illustrate the loading patterns in the SRS.
274

CHAPTER 12

FINITE ELEMENT ANALYSIS

Chapter Overview

Finite element analysis was performed on the cantilever-type highway overhead

sign support structure. The FEA software package SAP2000 v. 10 was used for the study

(11). The analysis was performed to address three primary areas of interests:

1. Perform a modal analysis,

2. Check the accuracy of the natural wind and truck-induced fatigue loading back-

calculations, and

3. Compare the developed equivalent static wind load to its dynamic load counter-

part for natural wind and truck-induced wind gusts.

The first process was performed to test the validity of the FEA model. The last two proc-

esses were performed for both the natural wind loading and truck-induced wind loading.

The FEA served as a model for determining stresses in locations that were not experi-

mentally instrumented. Validation of the FEA models concentrated on those areas that

were instrumented and tested experimentally. Changes in the models, such as materials,

sizes and shapes, sign area, lumped mass, damping, etc. were also evaluated that would

otherwise not be performed experimentally due to the time and cost of such projects.
275

Model Development

A full scale, three-dimensional model was developed of the experimentally tested

sign structure. The model was created from the shop drawings of the structure provided

by the Alabama Department of Transportation (ALDOT).

Geometry

The basic geometry of the model is shown in Figure 12.1. Body constraints were

used to connect the sign to the W-shape sections, and the W-shape sections to the truss.

Body constraints cause the constrained joints to move together as one three-dimensional

object (11). A close up of the connection is shown in Figure 12.2. Stiffeners were used

for the sign to reduce its flexibility needed in the modal analysis so that the global modal

behavior of the structure as a whole could be evaluated. The truss-to-post connection was

an extremely rigid bolted plate connection. It was modeled using steel plates connected to

the chord ends at three points spaced symmetrically along the weld. Body constraints

were used to connect the plates to the post support to simulate the rigid action so that all

joints cannot displace relative to each other (11). A close up of this connection is shown

in Figure 12.3.
276

FIGURE 12.1 FEA model of cantilever-type sign support structure.


277

FIGURE 12.2 Sign-to-truss connection.

FIGURE 12.3 Truss-to-post connection.


278

The base connection of the post support to the foundation used eight anchor bolts

to attach the post base plate to the 70 in (1,778 mm) diameter by 6 ft (1.83 m) long con-

crete pile foundation. In the FEA model development of this connection, the support post

was first connected directly to the pile. The pile was restrained in all translation and rota-

tional degrees-of-freedom at the end as the only base support for the structure. However,

this connection was too rigid when comparing the FEA modal analysis to the experimen-

tal modal analysis detailed in Chapter 7: Experimental Modal Analysis. The connection

was then modified by modeling the anchor bolts and base plate. Body constraint connec-

tions were used to connect the post support to the base plate. The anchor bolts were con-

nected to the base plate at nodal locations. The anchor bolts were modeled with a 2 in

(50.8 mm) clearance from the bottom of the base plate to the top of the concrete surface.

Rigid links were used to connect the anchor bolts to the pile foundation. This modeled

connection gave it the flexibility that better simulated the existing structural behavior as

determined in Chapter 7. A close up of the connection is shown in Figure 12.4.

FIGURE 12.4 Foundation connection.


279

Element Type

Beam and shell elements were used in the model. Beam elements were used for

the truss and support post, as well as for the anchor bolts, pile foundation, sign W-shape

connections, and sign stiffeners. Shell elements were used for the aluminum sign, W-

shapes to truss connections, truss-to-pole connection plates, and base plate.

Material Definition

The primary material of the structure was steel. The sign was made of aluminum

and the foundation was made of concrete. The post and truss section was made of API-

5L-X52 steel pipe. The plates were made of structural steel ASTM A572 Gr. 50. The an-

chor bolts were made of AASHTO M314-90 Gr. 55 (essentially the same as ASTM

F1554 Gr. 55). The W-shape and T-shape sections used for the sign-to-truss connection

were made of A572 Gr. 50 steel. The sign was made of aluminum wrought alloy desig-

nated as 6061-T6. The concrete pile was made of conventional concrete with #9 size re-

bar Gr. 60. All of the major required material properties of the designations given are

listed in Table 12.1.


280

TABLE 12.1 Material Properties

Modulus of Yield Tensile


Assigned Material
Material Elasticity Stress Stress
Elements Designation
(psi) (psi) (psi)
Steel Pipe
Truss, Post API-5L-X52 29,000,000 52,000 66,000
Truss and ASTM A572
Steel Plate 29,000,000 50,000 65,000
Base plates Gr. 50
AASHTO
Steel Rod
Anchor Bolts M314-90 Gr. 29,000,000 55,000 75,000
55
Aluminum
Sign 6061-T6 10,000,000 37,000 42,000
4, 000 psi
Concrete Pile Founda-
Compressive 3,600,000 NA NA
tion
Strength
Pile Rein- ASTM A706
Rebar 29,000,000 60,000 80,000
forcement Gr. 60

Finite Element Modal Analysis

A modal analysis was conducted in order to detect proper dynamic behavior of

the model, and to determine the natural frequencies of the structure and their correspond-

ing modal shapes. The natural frequencies were compared to the values determined ex-

perimentally for accuracy. The first five modal shapes for the model are shown in Figure

12.5, and the natural frequencies, natural periods, and a description of the modal shapes

are listed in Table 12.2.


281

Mode 1 Mode 2 Mode 3

x
y

Mode 4 Mode 5

FIGURE 12.5 Modal shapes of the first five modes from FEA.

TABLE 12.2 Natural Frequencies of the FEA Model

Period Frequency
Mode Mode Shape
(sec) (Hz)
1 Shaft Torsion 0.63577 1.57
2 Vertical Rocking 0.61645 1.62
3 Horizontal 0.25484 3.92
4 Longitudinal 0.19125 5.23
5 Truss Torsion 0.13408 7.46

There was a distinct difference in modal shapes and frequencies for each mode,

which demonstrates an accurate dynamic behavioral simulation. The x-axis was defined

as oriented along the longitudinal length of the structure, the y-axis oriented opposite the

direction of traffic, and the z-axis oriented vertically along the axial length of the post

support. The first modal shape was indicative of a horizontal action of the sign and truss
282

creating a torsion behavior about the z-axis. The second modal shape was indicative of a

vertical rocking of the truss in the plane perpendicular to the ground about the y-axis. The

third modal shape was indicative of a horizontal rocking of the sign, truss, and shaft in

the direction of traffic about x-axis. The forth modal shape was indicative of a longitudi-

nal rocking of the sign about the y-axis. And the fifth modal shape was indicative of a

torsion behavior of the truss about the x-axis.

Comparisons of FEA to Experimental Modal Analysis

Only four of the five of the indentified modes of vibration were induced in the

field. The fifth mode, torsion of the truss overhang, was not activated when analyzing the

experimental data from wind induced excitations. The primary vibration was in the first

two modal shapes, with the first mode predominately controlling the vibration with larger

amplitudes. The modal shapes in each direction were significantly separated from each

other such that the structure vibrated primarily in single directions, independently from

each other.

A comparison between the modal shapes determined from the experimental and

FEA analyses is provided in Table 12.3. A close comparison between theoretical and ex-

perimental results is shown. The modal shapes were identical in shape, however the am-

plitudes of vibration between the FEA and experimental results was not compared. The

first two modal shapes are the most important with respect to fatigue loading because the

majority of the vibration of the structure was within these modal shapes with the largest

amplitudes of vibration, and therefore inducing the largest fatigue stress on the structure.
283

TABLE 12.3 FEA and Experimental Modal Analysis Comparison

Modal Frequency (Hz) Percent Difference


Mode from Experimental
Experimental FEA
(%)
Mode 1 1.61 1.57 2.48
Mode 2 1.64 1.62 1.22
Mode 3 3.96 3.92 1.01
Mode 4 5.60 5.23 6.61
Mode 5 *** 7.46 NA
*** = Not measured in the field
NA = Not available

Finite Element Structural Analysis

A FEA structural analysis was conducted to test back-calculation procedures in

determining the equivalent static wind load for natural wind and truck-induced gusts. The

process was performed by inputting into the program the equivalent static pressure load

determined for natural wind and truck-induced gusts, and then extracting from the output

the strain at the locations of the strain gauges used to calculate the load. The output strain

from the FEA was compared to the experimentally obtained values, and the accuracy of

the back-calculation was evaluated. The analysis was seen as a structural analysis “dou-

ble check” to test the accuracy of the back-calculation performed, as well as the accuracy

of the FEA model. A successful comparison would indicate that the equivalent static

wind load determined for natural wind and truck-induced gusts was correct, and the FEA

model was accurate in depicting structural behavior, in addition to the modal analysis

comparison.
284

Natural Wind Gust

The equivalent static wind load determined from Chapter 8: Experimental Calcu-

lation of the Fatigue Load due to Natural Wind Gust was inputted into the SAP 2000

program. The FEA output was extracted at locations that were used to calculate the

equivalent static wind load from the experimental data. The output was compared to the

strain determined experimentally at these locations, and the accuracy of the back-

calculation program was evaluated. A generalization of the procedure can be enumerated

as follows:

1. Collect the equivalent static wind load developed from Chapter 8: Experimental

Calculation of the Fatigue Load due to Natural Wind Gust.

2. Collect the strain range value that corresponds to the equivalent static wind load

gathered in Step 1 from the experimental data at the particular strain gauge loca-

tion used for the comparison.

3. Input the equivalent static wind load into the FEA model as the static structural

excitation.

4. Run a linear-static solution of the FEA model with the equivalent static wind load

excitation.

5. Extract the strain value from the FEA solution at the strain gauge location deter-

mined in Step 2.

6. Compare the FEA strain value determined in Step 5 with the experimental value

gathered in Step 2.
285

Equivalent Static Wind Load

The equivalent static wind load determined from Chapter 8: Experimental Calcu-

lation of the Fatigue Load due to Natural Wind Gust was found to be equal to 4.01 psf

(192 Pa). The load was determined through a regression analysis to predict the fatigue

wind load for a 38 mph (17 m/s) wind velocity in compliance with the infinite-life ap-

proach to fatigue design for support structures. This wind velocity however, did not occur

during the experimental testing, and no experimental stress was available to complete the

comparison. A three second wind velocity and direction equal to approximately 14 mph

(6.26 m/s) was arbitrarily chosen in its place, which is slightly larger the average three

second wind velocity of the total collection. By using Eq. 12.1, developed from the re-

gression in Chapter 8 for determining the equivalent static wind pressure at the fatigue

wind, the equivalent static wind pressure for a 14 mph (6.26 m/s) was calculated to be

equal to 0.892 psf (42.7 Pa).

y = 0.0025x 2 + 0.402 [Eq. 12.1]

where
y = dependent variable equal to the equivalent static wind pressure, psf (Pa)
x = independent variable equal to the wind velocity, mph (m/s)

Experimental Strain Range

The average maximum three second strain range corresponding to the three sec-

ond wind velocity equal to 14 mph (6.26 m/s) was determined to be equal to 7.90 (10-6).

Strain gauge SGR-AA-3 at Section AA located 12 in (305 mm) above the base plate weld
286

was used because of its location along the symmetry of the structure, and its orientation

to the in-plane displacements due to the wind load.

FEA SAP2000 Loading Input

An equivalent static wind pressure load equal to 0.892 psf (42.7 Pa) was used as

input into the SAP2000 FEA program. The load was applied to the sign, post support, and

truss including the exposed chords, vertical struts and diagonal struts of the web. The

pressure load for each member was multiplied by the drag coefficient of the particular

member it was applied. The same drag coefficients used to develop the load and Eq. 12.1

was used for this analysis.

The resulting equivalent static wind load was applied horizontally as a uniformly

distributed load to each member. It was applied as a uniformly distributed pressure to the

sign, and was applied as a uniformly distributed linear load (force per length) to the

chords, vertical struts, diagonal struts, and post support. Only the members exposed on

the vertical plane as seen from the elevation view of the structure applied.

FEA Linear-Static Solution

A linear-static solution was conducted in the SAP2000 program on the developed

model of the support structure. The location on the model that corresponded to the strain

gauge used for the comparison (SGR-AA-3) was on the front face (exposed face) of the

support post at 12 in (305 mm) above the base plate. The post was modeled as a beam

element in the FEA model, and therefore the moment at this location was extracted from

the solution. The stress value was calculated using Eq. 12.2, which accounted for the
287

normal stress at the section when the dead weight of the structure is excluded, and the

force was applied purely horizontally. The normal strain was calculated using Eq.12.3

assuming the stress remains within the elastic limit of the steel material. The results of the

calculations are shown in Table 12.4.

Mc
σ= [Eq. 12.2]
I

where
σ = bending moment stress, psi (Pa)
M = bending moment along the x - axis extracted from the FEA, lb - ft (N - m)
c = distnce from the neutral axis to the outer most fiber located at SGR - AA - 3, in (mm)
I = moment of inertia about the x - axis at Section AA, in 4 (mm 4 )

σ
ε= [Eq. 12.3]
E

where
ε = normal strain, in/in, (mm/mm)
σ = normal stress, psi (Pa)
E = modulus of elasticity of steel, psi (Pa)
288

TABLE 12.4 Normal Strain from FEA at Section AA

Extracted from
Calculated Values from FEA Bending Moment
FEA
σ ε
Bending Moment c I E
Eq. 12.2 Eq. 12.3
(lb-ft) (in) (in2) (psi)
(psi) (in/in)

2,324.57 12 1472.7068 227.294 29,000,000 7.84(10-6)

Comparison of FEA with Experimental

A comparison was made with the FEA strain value calculated in Table 12.4 with

the experimentally measured strain range at the specific strain gauge location used for the

analysis. The results of the comparison are provided in Table 12.5. The comparison re-

sulted in a 0.759% difference between the two values. The closeness of the two values

indicate the back-calculation to determine an equivalent static wind load that would pro-

duced the strain range measured experimentally was accurate. The results also further

prove the accuracy of the FEA model with the measured structural behavior of the sup-

port structure.

TABLE 12.5 Comparison of FEA with Experimental Strain Range

Experimental Strain Range Percent Difference from


FEA Strain (in/in)
(in/in) Experimental (%)
7.84(10-6) 7.90 (10-6) 0.759
289

Truck-Induced Wind Gust

The same comparison procedure for natural wind gust was used for truck-induced

wind gust. The equivalent static wind load determined from Chapter 10: Experimental

Calculation of the Fatigue Load due to Truck-induced Wind Gust was inputted into the

SAP 2000 program. The FEA output extracted at locations that were used to calculate the

equivalent static wind load from the experimental data. The output was compared to the

strain determined experimentally at these locations, and the accuracy of the back-

calculation program was evaluated. A generalization of the procedure can be enumerated

as follows:

1. Collect the equivalent static wind load developed from Chapter 10: Experimental

Calculation of the Fatigue Load due to Truck-induced Wind Gust.

2. Collect the strain range value that corresponds to the equivalent static wind load

gathered in Step 1 from the experimental data at the particular strain gauge loca-

tion used for the comparison.

3. Input the equivalent static wind load into the FEA model as the static structural

excitation.

4. Run a linear-static solution of the FEA model with the equivalent static wind load

excitation.

5. Extract the strain value from the FEA solution at the strain gauge location deter-

mined in Step 2.

6. Compare the FEA strain value determined in Step 5 with the experimental value

gathered in Step 2.
290

Equivalent Static Wind Load

The equivalent static wind load determined from Chapter 10: Experimental Cal-

culation of the Fatigue Load due to Truck-induced Wind Gust was found to be equal to

8.348 psf (400 Pa). The load was applied to the end 12 ft (3.66 m) of the overhang truss.

The pressure was applied to the horizontal area on the underneath portion of the truss.

Experimental Strain Range

The experimental strain range used to back-calculate the 8.348 psf (400 Pa) was

equal to 9.245(10-6). The range was measured on the post support at Section AA located

12 in (305 mm) above the base plate. Strain gauge SGR-AA-7 was used because of its

location along the symmetry of the structure, and its orientation to the in-plane displace-

ment caused by the vertically applied wind load.

FEA SAP2000 Loading Input

An equivalent static wind pressure load equal to 8.348 psf (400 Pa) was used as

input into the SAP2000 FEA program. The load was applied to the exposed horizontal

area on the underneath façade of the truss overhang along the end 12 ft (3.66 m) section.

The load was applied vertically upward (away from ground) as a uniformly distributed

linear load (force per length). Only the truss section including the chords, horizontal strut,

and horizontal diagonal struts along the end 12 ft (3.66 m) section of the truss were

loaded. The pressure load for each member was multiplied by the drag coefficient of the

particular member it was applied. The drag coefficients used in Chapter 10 were applied.
291

FEA Linear-Static Solution

A linear-static solution was conducted in the SAP2000 program on the developed

model of the support structure. The location on the model that corresponded to the strain

gauge used for the comparison (SGR-AA-7) was on the side face of the support post at 12

in (305 mm) above the base plate, measuring the in-plane strain of the post due to the ver-

tically applied truck gust. The post was modeled as a beam element in the FEA model,

and therefore only the moment at this location was extracted from the solution. The stress

value was calculated using Eq. 12.4. The equation accounts for the normal stress at the

section with the dead weight of the structure excluded, and the force applied purely verti-

cally. The normal strain was calculated using Eq.12.5 assuming the stress remains within

the elastic limit of the steel material. The results of the calculations are shown in Table

12.6.

Mc P
σ= + [Eq. 12.4]
I A

where
σ = bending moment stress, psi (Pa)
M = bending moment along the x - axis extracted from the FEA, lb - ft (N - m)
c = distnce from the neutral axis to the outer most fiber located at SGR - AA - 3, in (mm)
I = moment of inertia about the x - axis at Section AA, in 4 (mm 4 )
P = axial force, lb (N)
A = cross sectional area, in 2 (mm 2 )

σ
ε= [Eq. 12.5]
E
292

where
ε = normal strain, in/in, (mm/mm)
σ = normal stress, psi (Pa)
E = modulus of elasticity of steel, psi (Pa)

TABLE 12.6 Normal Strain from FEA at Section AA

Calculated Values from FEA Bending Moment and Axial


Extracted from FEA
Force
Bending Axial σ ε
c I A E
Moment Force Eq. 12.4 Eq. 12.5
(in) (in2) (in2) (psi)
(lb-ft) (lb) (psi) (in/in)
9.49
2,772.45 87.87 12 1472.7068 20.9 275.292 29,000,000
(10-6)

Comparison of FEA with Experimental

A comparison was made with the FEA strain value calculated in Table 12.6 with

the experimentally measured strain range at the specific strain gauge location used for the

analysis. The results of the comparison are provided in Table 12.7. The comparison re-

sulted in a 2.53% difference between the two values. The closeness of the two values in-

dicate the back-calculation to determine an equivalent static wind load that would pro-

duced the strain range measured experimentally was accurate. The results also further

prove the accuracy of the FEA model with the measured structural behavior of the sup-

port structure.
293

TABLE 12.7 Comparison of FEA with Experimental Strain Range

Experimental Strain Range Percent Difference from


FEA Strain (in/in)
(in/in) Experimental (%)
9.49(10-6) 9.245(10-6) -2.53

Finite Element Dynamic Fatigue Loading Analysis

An evaluation of the theoretical methods developed for this project was per-

formed using FEA. The VRS for natural wind gusts and SRS for truck-induced wind

gusts were developed using principles related to Structural Dynamics. A SDOF was used

in their formulation. The spectrums (VRS and SRS) provide an equivalent static wind

load based on the dynamic characteristic of the structure that would produce an equiva-

lent stress range as a dynamic loading.

To test the accuracy of the spectrums, dynamic loading was inputted into the FEA

SAP2000 program and critical response values were extracted from the solution. Using

the same loading conditions and dynamic characteristics of the structure, the equivalent

static wind load was extracted from the VRS for natural wind gust, and SRS for truck-

induced wind gust, and was applied as input into the FEA program. A linear-static solu-

tion was solved, and the same response values extracted from the previously run dynamic

solution was performed for the linear-static solution and compared. The comparison

helped to determine the accuracy of the VRS and SRS methods developed with this pro-

ject. In addition, structural parameters such as material properties were altered and the

comparison processes were performed again. The comparison procedure therefore in-

cluded two separate models:


294

1. Cantilever-type support structure (Figure 12.1) made of structural steel material

and all other material labeled in Table 12.1.

2. Cantilever-type support structure (Figure 12.1) made with the same material

properties in Table 12.1, but substituting wrought alloy Aluminum designation

6061-T6 for the truss and post members only.

By changing the material properties, the dynamic characteristics of the structure were al-

tered, and a true test of the VRS and SRS was performed and evaluated without addition

experimental testing which can be time consuming and costly.

Natural Wind Gust

A dynamic loading FEA solution was conducted to test the hypothesis that the use

of the VRS for natural wind gusts is an accurate approach in determining an equivalent

static wind load for fatigue. Details on the development of the VRS are provided in Chap-

ter 9: Theoretical Calculation of the Fatigue Wind Load due to Natural Wind Gust. The

comparison served as a mathematical check on the dynamic response calculations used to

develop the VRS. A dynamic load was inputted into the program as a power spectral den-

sity function, based on the wind pressure power spectral density excitation developed

from the experimentally collected data, and used to develop the VRS in Chapter 9. The

response of the FEA model for deflection at the tip of the post support and the moment at

the base plate was analyzed and recorded. This response was then compared to the linear-

static response of the equivalent static wind load taken from the VRS. A generalized de-

scription of comparison procedure as it was performed is numerated as follows:


295

1. Input the dynamic wind pressure power spectral density function developed in

Chapter 9: Theoretical Calculation of the Fatigue Wind Load due to Natural Wind

Gust as the structural excitation in the FEA SAP2000 model.

2. Run a spectral solution in the FEA program and extract the displacement of the tip

of the post support and the bending moment at the base plate.

3. Gather the dynamic characteristics of the structure from Chapter 7: Experimental

Modal Analysis including the modal frequencies and critical damping percent-

ages.

4. Use the values gathered in Step 3 to extract the equivalent static wind load from

the VRS developed in Chapter 9: Theoretical Calculation of the Fatigue Wind

Load due to Natural Wind Gust.

5. Input the equivalent static wind load determined in Step 4 as a static pressure load

into the FEA SAP2000 program.

6. Run a linear-static solution in the FEA program and extract the displacement of

the tip of the post support and the bending moment of the base plate.

7. Compare static solution values determined in Step 6 with the dynamic solution

values in Step 2 and evaluate the accuracy of the VRS.

8. Change the model from steel to aluminum and perform steps 1 through 7 again.

With exception to Step 3, perform a FEA modal analysis to determined the modal

frequencies of the new structure, and assume a critical damping percentage for the

remainder of the analysis.


296

Wind Pressure Power Density Spectrum

The loading (dynamic and static) was applied horizontally to the front façade of

the structure. The wind pressure power density spectrum developed in Chapter 9: Theo-

retical Calculation of the Fatigue Wind Load due to Natural Wind Gust is shown in Fig-

ure 12.6, and was used as input in the FEA program for the dynamic loading. The spec-

trum in the figure was developed from the average wind velocity equal to 12.96 mph

(5.79 m/s). Likewise, the average peak-to-peak range VRS was used for the comparison

and not the VRS developed from the regression analysis to simulate the wind pressure

from the 38 mph (17 m/s) fatigue wind.

FIGURE 12.6 Average wind pressure PDS for FEA input.


297

The spectrum was inputted into the SAP program as a dynamic power density

spectrum and applied to the exposed structural members on the front façade of the struc-

ture. This included the sign, post support, truss chords, and truss vertical and diagonal

struts.

Wind Pressure Equivalent Static Wind Load

Steel model. The equivalent static wind load was extracted from the VRS (Figure

12.7) and inputted into the FEA program. The load was extracted from the VRS based on

the earliest modal frequency with a modal shape in the direction of the wind loading. This

corresponded to mode 1 with a modal frequency of 1.57 Hz. The modal frequencies

solved by the FEA program were used instead of the experimentally obtained values to be

consistent with the dynamic power density spectrum solution. The dynamic solution will

use the modal frequencies solved by the program to determine resonance areas in the

spectrum. A damping ratio value of 0.015 (1.5%), which was slightly less than the ex-

perimentally determined value of 1.86%, was used in the VRS as well as in the FEA.

The extracted VRS equivalent static wind load for the modal frequencies and

damping criteria was equal to 0.83 psf (39.7 Pa). The VRS load was multiplied by the

drag coefficient of the member it was applied, and loaded horizontally as a uniformly dis-

tributed load on the front face of the exposed members to the wind load. This included

the sign, truss members outside of the sign, and the post support.
298

FIGURE 12.7 VRS used for the FEA analysis.

Aluminum model. The equivalent static wind load was extracted from the VRS

(Figure 12.8) and inputted into the FEA program. A FEA modal analysis was performed

on the model to determine the modal frequencies. The modal frequencies did signifi-

cantly change (see Table 12.8), which was good for the comparison, but the modal shapes

remained the same.


299

FIGURE 12.8 VRS for the aluminum FEA model with 0.5% damping.

TABLE 12.8 Modal Analysis of the Aluminum Structure

Modal Frequency (Hz)


Mode Mode Shape
Aluminum Steel
1 Shaft Torsion 1.127 1.57
2 Vertical Rocking 1.212 1.62
3 Horizontal 3.217 3.92
4 Longitudinal 3.943 5.23
5 Truss Torsion 5.219 7.46

The load was extracted from the VRS based on the earliest modal frequency with

a modal shape in the direction of the wind loading. This corresponded to mode 1 with a

modal frequency of 1.127 Hz for the aluminum model. The damping ratio was changed
300

from 0.0015 (1.5%) used in the steel model to a value of 0.005 (0.5%). This value was

obtained from the literature on damping analysis of aluminum support structures (20).

The extracted VRS equivalent static wind load for the modal frequencies and

damping criteria was equal to 1.332 psf (63.8 Pa). The VRS load was multiplied by the

drag coefficient of the members it was applied, and was loaded horizontally as a uni-

formly distributed load on the front face of the exposed members to the wind load. This

included the sign, truss members outside of the sign, and the post support.

Comparison of Results

The results between the dynamic power spectral density analysis and the static

VRS analysis were compared. The comparison involved the displacement at the tip of the

post support in the y-axis direction (see Figure 12.5 for axis orientation), and the bending

moment at the base plate about the x-axis. The results are listed in Table 12.9. As shown

in the table, the results from the PDS dynamic solution and the VRS linear-static solution

show very close agreement, which substantiate the hypothesis that the equivalent static

wind load extracted from the VRS based on the dynamic characteristics of the structure

will produced the same response on the structure as a dynamic random excitation due to

wind pressure.
301

TABLE 12.9 FEA Comparison for Natural Wind Gust

Critical
Material Natural Tip Base
Damping Analysis Case
Model Frequency Deflection Moment
Percentage
PDS Dynamic
0.09352 in 76.3 kip-in
Solution
1.57 Hz 1.5%
Steel VRS Static
0.0869 in 72.0 kip-in
Solution
Percent Difference from Dynamic FEA 7.08 % 5.64 %
PDS Dynamic
0.3388 in 114.6 kip-in
Solution
1.127 Hz 0.5%
Aluminum VRS Static
0.3344 in 115.6 kip-in
Solution
Percent Difference from Dynamic FEA 1.30 % -0.873 %

The small difference in results between the static VRS and the dynamic PDS was

due to the development of the VRS, and the method of the FEA PDS response calcula-

tion. The VRS was based on a simplification of the structure as a SDOF system, and as a

result had only one resonance spike along the frequency domain within the calculated

response PDS curve (see Figure 12.9). The FEA SAP2000 program calculated the re-

sponse PDS curve using a multiple degree-of-freedom system, which generated multiple

modal spikes.

For example, A SAP2000 output response PDS spectrum of the response bending

moment of the post support member is shown in Figure 12.10. From this view, the

smaller frequency spikes cannot be seen because they are significantly smaller than the

major response of the structure in the form of the first modal frequency. When zooming

into at a smaller window of the spectrum, as shown in Figure 12.11, the smaller spikes at

other modal frequencies are seen but are very small in amplitude as compared to the ma-

jor mode 1 spike in Figure 12.10. Subsequently, with more than one spike, the dynamic
302

FEA had a larger area under the spectrum curve to calculate the root-mean-square value,

which is proportional to the magnitude of the response values. The dynamic FEA indi-

cated a larger variance of vibration amplitudes because of the increased number of ex-

cited modes. As a result, the resulting response values (tip deflection and base plate mo-

ment) of the dynamic FEA were larger the VRS static values developed using a SDOF

system.

FIGURE 12.9 VRS PDS response to wind pressure excitation.


303

FIGURE 12.10 SAP2000 response PDS.

FIGURE 12.11 Smaller frequencies of the SAP2000 response PDS.


304

The small difference in results further justified the simplification of the problem

into a SDOF system for these structures. The resonance spikes in the dynamic FEA re-

sponse PDS were significantly separated from each other such that the vibration modes of

the structure in the y-direction were nearly independent. The largest resonance spike was

associated to the modal shape in the direction of the loading, and therefore was the pre-

dominate mode of vibration. Even though there were other activated modal shapes within

the vibration, they subsequently had very little effect on the overall response. In view of

this, approximating the response as a SDOF system using the VRS was validated. Proper

use of the VRS in the analysis is to select the earliest natural frequency that has a modal

shape in the direction of the excitation, and read off the corresponding equivalent static

wind load ordinate.

Truck-Induced Wind Gust

A dynamic loading FEA solution was conducted to test the hypothesis that the use

of the SRS for truck-induced wind gust is an accurate approach in determining an equiva-

lent static wind load for fatigue. Details on the development of the SRS are provided in

Chapter 11: Theoretical Calculation of the Fatigue Wind Load due to Truck-Induced

Wind Gust. The comparison served as a mathematical check on the dynamic response

calculations used to develop the SRS. A dynamic load was inputted into the program as

transient time history function, the same function used to develop the SRS. The response

of the FEA model for deflection at the tip of the post support and the moment at the base

plate was analyzed and recorded. This response was then compared to the linear-static
305

response of the equivalent static wind load taken from the SRS. A generalized description

of comparison procedure as it was performed is numerated as follows:

1. Input the dynamic transient time history function developed in Chapter 11: Theo-

retical Calculation of the Fatigue Wind Load due to Truck-Induced Wind Gust as

the structural excitation in the FEA SAP2000 model.

2. Run a dynamic solution in the FEA program and extract the displacement range

of the tip of the post support and the bending moment range at the base plate.

3. Gather the dynamic characteristics of the structure from Chapter 7: Experimental

Modal Analysis including the modal frequencies and critical damping percent-

ages.

4. Use the values gathered in Step 3 to extract the equivalent static wind load from

the SRS developed in Chapter 11: Theoretical Calculation of the Fatigue Wind

Load due to Truck-Induced Wind Gust.

5. Input the equivalent static wind load determined in Step 4 as a static pressure load

into the FEA SAP2000 program.

6. Run a linear-static solution in the FEA program and extract the displacement of

the tip of the post support and the bending moment of the base plate. Note: the

SRS is based on a ranged value.

7. Compare static solution values determined in Step 6 with the dynamic solution

values in Step 2 and evaluate the accuracy of the VRS.


306

8. Change the model from steel to aluminum and perform steps 1 through 7 again.

With exception to Step 3, perform a FEA modal analysis to determine the modal

frequencies of the new structure.

Wind Pressure Transient Time History Function

The loading (dynamic and static) was applied vertically to the horizontal area 12

ft (3.66 m) from the end of the truss structure. The wind pressure transient time history

function developed in Chapter 11: Theoretical Calculation of the Fatigue Wind Load due

to Truck-Induced Wind Gust is shown in Figure 12.12, and was used as input in the FEA

program for the dynamic loading. The 70 mph (31.3 m/s) was used for the steel model,

whereas the 60 mph (26.8 m/s) was used for the aluminum model.

FIGURE 12.12 Transient excitation function for FEA input.


307

The transient time history function was inputted into the SAP program as a dy-

namic time history and was loaded in 0.01 second time steps. The loading was applied to

the exposed horizontal area on the underneath face of the structure. This included the

truss chords, one horizontal strut, and three horizontal diagonals.

Wind Pressure Equivalent Static Wind Load

Steel model. The equivalent static wind load was extracted from the SRS (Figure

12.13) and inputted into the FEA program. The load was extracted from the SRS based

on the earliest modal frequency with a modal shape in the direction of the wind loading.

This corresponded to mode 2 with a modal frequency of 1.62 Hz. The modal analysis

solved by the FEA program was used instead of the experimentally obtained values to be

consistent with the FEA dynamic solution. The dynamic solution will use the modal fre-

quencies solved by the program to determine resonance. The steel model used the 70 mph

(31.3 m/s) time history plot. Damping was not considered in the analysis as the initial

peak-to-peak range of the response was needed.

The extracted SRS equivalent static wind load for the modal frequencies and truck

speed criteria was equal to 8.33 psf (399 Pa). The SRS load was multiplied by the drag

coefficient of the member it was applied, and loaded vertically as a uniformly distributed

load on the underneath horizontal face of the exposed members to the wind load. This

included the chords, horizontal strut, and horizontal diagonal members.


308

FIGURE 12.13 SRS used for the FEA analysis.

Aluminum model. The equivalent static wind load was extracted from the SRS

(Figure 12.13) and inputted into the FEA program. A FEA modal analysis was performed

on the model to determine the modal frequencies. The modal frequencies did signifi-

cantly change (see Table 12.8), which was good for the comparison, and the modal

shapes remained the same.

The load was extracted from the SRS based on the first modal frequency with a

modal shape in the direction of the wind loading. This corresponded to mode 2 with a

modal frequency of 1.212 Hz for the aluminum model. The 60 mph (26.8 m/s) time his-

tory was used for the aluminum model to test the accuracy of the SRS with changing

speeds. The extracted SRS equivalent static wind load for the modal frequency was equal
309

to 5.89 psf (282 Pa). The SRS load was multiplied by the drag coefficient of the members

it was applied, and was loaded vertically as a uniformly distributed load on the under-

neath horizontal area of the exposed members to the wind load. This included chords,

horizontal strut, and horizontal diagonal members.

Comparison of Results

The results between the dynamic time history analysis and the static SRS analysis

were compared. The comparison involved the displacement at the tip of the post support

in the x-axis direction (see Figure 12.5 for axis orientation), and the bending moment at

the base plate about the y-axis. Evaluation of the time history solution involved extracting

the maximum peak-to-peak range of displacement and moment values to use in the com-

parison because the SRS was based on a peak-to-peak range. Damping was not consid-

ered in the analysis. The results are listed in Table 12.10.

As shown in the table, the results from the time history (TH) dynamic solution

and the SRS linear-static solution show close agreement, which substantiate the hypothe-

sis that the equivalent static wind load extracted from the SRS based on the dynamic

characteristics of the structure will produced the same response on the structure as a dy-

namic random excitation due to wind pressure. In view of this, approximating the re-

sponse as a SDOF system using the SRS was validated. Proper use of the SRS in the

analysis is to select the first modal frequency that has a modal shape in the direction of

the excitation, and read off the corresponding equivalent static wind load ordinate.
310

TABLE 12.10 FEA Comparison for Truck-Induced Wind Gust

Material Natural Truck Tip Base


Analysis Case
Model Frequency Speed Deflection Moment
TH Dynamic
0.06097 in 27.9 kip-in
Solution
1.62 Hz 70 mph
Steel SRS Static So-
0.0565 in 32.0 kip-in
lution
Percent Difference from Dynamic FEA 7.33 % -12.8 %
TH Dynamic
0.110 in 18.8 kip-in
Solution
1.212 Hz 60 mph
Aluminum VRS Static
0.102 in 23.6 kip-in
Solution
Percent Difference from Dynamic FEA 7.27 % -25.5 %
311

CHAPTER 13

PROPOSED FATIGUE PROVISIONS

Chapter Overview

A complete overview of the research performed with this project was completed

and is presented. A discussion of the results from the project is provided in this chapter,

followed by proposed fatigue design equations for natural wind and truck-induced wind

gusts. The review encompassed both the experimental and theoretical results, and a de-

sign methodology was developed. Design examples are provided to show the use of the

proposed fatigue provisions, and comparisons are made with the Supports Specifications.

Fatigue Load due to Natural Wind Gust

The proposed fatigue design equations for natural wind gusts were developed

from the experimental and theoretical programs with this project. The recommendations

are presented in two forms. The first form is a more generalized design equation, whereas

the second equation is a more detailed design equation that accounts for the dynamic

characteristics of the structure.

Discussion of Results

The evaluation of the results involved primarily the comparison between the

equivalent static wind loads developed from the experimental and theoretical programs.
312

The equivalent static wind load was determined for a fatigue wind equal to 38 mph (17

m/s). The fatigue wind was determined using the infinite-life approach to fatigue design

of support structures as the wind velocity with a 0.01% exceedence probability from an

annual mean wind velocity of 11 mph (5 m/s). The wind pressure at the fatigue wind was

determined in both programs. The results of the analysis are listed in Table 13.1.

TABLE 13.1 Experimental and Theoretical Results for Natural Wind Gust

Equivalent Static Wind


Program Wind Velocity (mph)
Load (psf)
Experimental 38 4.01
Theoretical 38 3.80
Percent Difference from Experimental 5.24%

The results from the two programs were similar with a percent difference from the

experimental equal to 5.24%. The experimental was larger than the theoretical because

the VRS was developed using a SDOF to simulate the structural response. The experi-

mental was based on the real structure, which was a multiple degree of freedom dynamic

system. The closeness of the two values indicated that approximating the structure as a

SDOF was justified.

The resistance analysis performed in Chapter 6: Fatigue Resistance revealed areas

on the structure that were vulnerable to fatigue. A complete distribution of stress within

the structure from the wind loading was established. Large variances in measured strain

ranges were realized in the chord members and anchor bolts taking into account the bal-

anced state with respect to the neutral axis created from the loading. This was especially

true for the anchor bolts, and was believed to be due to the clearance differences between
313

anchor bolts and the base plate. The strain measurements made on the post support at

Sections AA and BB were more consistent and a more straightforward calculation was

accomplished.

The fatigue load due to natural wind gust was determined from the measurements

made on the post support. The location of these gauges, and the orientations to in-plane

and out-of-plane vibrations made the calculation straightforward. The results of the back-

calculation for the wind pressures were consistent with each gauge located along the cir-

cumference of the post. A wind pressure value of 4 psf (191 Pa) was determined from the

analysis. This represented an equivalent static wind load that produced the same stress

range on the structure as measured experimentally from dynamic vibration due to ran-

domly applied natural wind loading. The value was determined at the fatigue wind of 38

mph (17 m/s) in compliance with the infinite-life approach to fatigue design for support

structures.

The same methodology in calculating the 4 psf (191 Pa) wind pressure was per-

formed with this project as with the development of the fatigue load due to natural wind

adopted by the Supports Specifications. The difference was attributed to the attainment of

the load through experimental versus theoretical means. The fatigue load of the Supports

Specifications was accomplished through theoretical computations. Finite element analy-

sis was used, that was based on a multiple degree-of-freedom dynamic system. The same

method of computation was performed with this project, but by using experimentally

measured strain range values at critical locations on the structure.

The formulation of the VRS in calculating the fatigue load due to natural wind

load resulted in a successful comparison with the experimental results. The methodology
314

was a hybrid using both experimental and theoretical applications. The wind velocity data

collected experimentally was used to develop wind pressure power density spectrum ex-

citation by the same procedure performed by the Supports Specifications. The difference

was in the structural response analysis. The VRS was comprehensive and universal in

that the method can be applied to any support structure of different configuration, size,

shape, and material properties. The fatigue load is extracted from the VRS based on the

dynamic characteristics of the structure, which was directly related to the stiffness and

mass of the structure.

The results of the VRS approach produced an equivalent static wind load that was

slightly less than the experimentally obtained value. This was due to the formulation of

the VRS as a SDOF. The response of the structure was approximated as a SDOF,

whereas the real structure is a multiple degree-of-freedom system. As proved by the FEA

dynamic loading comparisons and the close similarity between the VRS value and the

experimental value, the approximation of the structure as a SDOF was justified.

The design fatigue load provisions were developed with this project in response to

the overview and discussion of the results. The established fatigue design equation was

separated into two approaches:

1. A general fatigue design equation

2. A detailed fatigue design equation

The general fatigue design equation was based primarily on the experimental results. A

more detailed approach was developed using the VRS, which was considered accurate
315

due to the close comparison between the results of the experimental and theoretical re-

search program of this project. This approach can be used if the dynamic characteristics

of the structure, such as the modal frequencies and shapes, as well as critical damping

percentages are known by the engineer.

General Fatigue Design Equation for Natural Wind Gust

The general fatigue design equation for natural wind gusts is shown as Eq. 13.1. It

is to be applied horizontally to the exposed façade of the structure as seen on an elevation

view. The load is applied as a uniformly distributed load to be easily applied by the engi-

neer. It is recommended for cantilever-type sign support structures; specifically for sup-

port structures with a modal frequency of 1.61 Hz, and a critical damping percentage of

1.82%.

2
v 
PNW = 4C d I F   [Eq. 13.1]
 11 

where
PNW = design fatigue load due to natural wind gust, psf (Pa)
C d = drag coefficient
I F = importance factor
v = annual mean three second wind velocity, mph (m/s)

The equation takes the form similar to the Support Specifications for easy impli-

cation. The drag coefficients and importance factors are not changed, and are provided in
316

the Supports Specifications. Accommodation for other annual mean wind velocities other

than 11 mph (5 m/s) is also provided in the equation by a simplified ratio calculation.

Detailed Fatigue Design Equation for Natural Wind Gust

A more detailed design equation was developed to account for the variety of sign

support structures in design, each with different configurations, cross sectional shapes,

and material properties. Since these factors have significant influence on the dynamic

characteristics of the structure, and because sign supports structures are highly flexible

with low damping properties, a method was needed in determining the fatigue load based

on the vibration behavior of these structures. The approach presented in this section ac-

counted for the dynamic behavior of the structure in terms of the modal frequency and

critical damping percentages, and the fatigue load was determined based on these proper-

ties. The approach was considered a unified design approach to fatigue

of natural wind gusts in that it accounts for the variety of support structures in design.

Vibration response spectrums were developed that presented the fatigue load

based on the dynamic properties of the structure. The load is extracted from the VRS in

terms of the earliest modal frequency of the structure with a modal shape in the direction

of the loading. The most critical loading scenario for natural wind is directed normal to

the plain of the sign (in the direction of traffic), exciting the modal shape most commonly

referred to as the horizontal modal shape. For sign support structures, the horizontal mo-

dal shape is generally around 1 to 3 Hz, which typically corresponds to the first modal

shape for cantilever-type structures, and the second modal shape for bridge-type struc-

tures. Finite element software (i.e., SAP2000) can be used to estimate the appropriate
317

modal shapes and their associated natural frequencies to use with the VRS curves. If FEA

software is not available, fundamental structural dynamics of a SDOF system (Eq. 13.2)

can be used for estimating these values (10, 17, 33, 56, 57, 61, 63).

1 K
fn = [Eq. 13.2]
2π M

where
f n = natural frequency, Hz
K = generalized stiffness, N/m (lb/ft)
M = generalized mass, kg (slug)

A conservative estimate would be to assume a modal frequency of 1 Hz to use with the

VRS, as the equivalent static wind load increases as the modal frequency decreases be-

cause of the proximity to the broadband amplitude of the wind pressure PSD excitation

used in its development.

The critical damping percentage is especially relevant to sign support structures

because of their relatively high flexibility and subsequent low natural frequencies (1 to 3

Hz). What’s more, their damping ratios are mostly below 2.0%. A low damping will al-

low the structure to vibrate longer at high amplitudes, and thus produce more stress that

could potentially cause fatigue damage. It is recommended for design purposes to assume

a damping ratio equal to 0.015 (1.5%). However, damping ratios can vary depending on

the structural material, and therefore actual values can be obtained from experimental

data of comparable structures if available.


318

Fatigue Design Procedure

A systematic procedure for determining the fatigue design load due to natural

wind based on the proposed detailed approach may be given as follows.

Step 1: annual mean wind velocity. The first step involves determining the annual

mean wind velocity of the site. The National Weather Service Offices near the site can be

used to determine this value. It is recommended to use a value of 11 mph (5 m/s), how-

ever other values can be used if needed.

Step 2: modal analysis. The second step involves a modal analysis of the design-

ing support structure. The earliest modal frequency of the modal shape in the direction of

the natural wind loading is needed, as well as the critical damping percentage. Finite ele-

ment analysis software can be helpful in this step. If FEA is not available, the modal fre-

quency and modal shape can be estimated using Eq. 13.2. If experimental data of compa-

rable structures are not available, it is recommended to use a damping ratio estimated at

0.015 (1.5%).

Step 3: vibration response spectrum. The third step involves the VRS shown in

Figure 13.1 and Figure 13.2. A VRS constant, PVRS, is extracted from the spectrum as the

ordinate value corresponding to the natural frequency and damping ratio determined in

Step 2. If using a damping ratio of 1.5% (recommended), use the VRS in Figure 13.1.

The VRS in Figure 13.2 is provided for other damping ratios if needed.
319

FIGURE 13.1 VRS for 1.5% damping ratio.

FIGURE 13.2 Fatigue wind pressure VRS for damping range.


320

Step 4: fatigue design equation. The final step is plugging all information gath-

ered from Steps 1-3 into the fatigue design equation for natural wind gusts based on the

detailed approach. Input the VRS constant, PVRS, determined from Step 3 into Eq. 13.3

for each member along the facade of the structure exposed to natural wind.

2
v 
PNW = PVRS C d I F   [Eq. 13.3]
 11 

where
PNW = design fatigue load due to natural wind gust, psf (Pa)
PVRS = transmitted pressure constant extracted from the VRS, psf (Pa)
C d = drag coefficient
I F = importance factor
v = annual mean three second wind velocity, mph (m/s)

Fatigue Load due to Truck-Induced Wind Gust

The proposed fatigue design equations for truck-induced wind gust were devel-

oped from the experimental and theoretical programs with this project. The recommenda-

tions are presented in two forms. The first form is a generalized design equation, whereas

the second equation is a more detailed design equation that accounts for the dynamic

characteristics of the structure.


321

Discussion of Results

The evaluation of the results involved primarily the comparison of the results

from the experimental and theoretical programs. In the experimental program, the truck-

induced gust fatigue load was developed from measured strain ranges collected during

the truck tests. An equivalent static wind load was developed that would produce an

equivalent strain range as measured experimentally. The analytical program involved de-

velopment of the SRS, where the structure was approximated to behave dynamically as a

SDOF system. The equivalent static wind load was extracted from the SRS based on the

dynamic characteristics of the structure. A comparison of the results is shown in Table

13.2.

TABLE 13.2 Comparison of Results from Experimental and Theoretical Programs

Equivalent Static Wind


Program Truck Speed (mph)
Load (psf)
Experimental 60 8.348
Theoretical 70 8.33
Percent Difference from Experimental 0.216 %

The percent difference of the theoretical value to the value developed from the

experimental results was 0.216%, indicated good similarity between the two programs.

The theoretical program and SRS was seen as a good approximation for determining the

fatigue load. There was very little variation in the experimental value with respect to

truck speed. The maximum range at the location used for the back-calculation was ob-

served at 60 mph (26.8 m/s), yet values occurring from the higher speeds tested were ex-

tremely close. The majority of analysis of data at the other instrumented locations on the
322

structure including the anchor bolts and chord members showed higher range values oc-

curring at the highest speed of 70 mph (31.1 m/s). In view of this, the design equation

was developed using the maximum range measured on the post member, and since the

measured values were similar with truck speed, the 70 mph (31.3 m/s) was used as the

reference speed for this load. As a result, accounting for higher speeds would not unnec-

essarily increase the fatigue load.

The fatigue resistance analysis indicated that the most suitable location to calcu-

late the fatigue load was from strain measurements on the post member. This was due to

the consistency of the data at this location, the orientation to in-plane vibration, and also

the straightforwardness of the calculation. This allowed for a direct back-calculation of

the causative truck gust loading through combined loading analysis.

A vertical load equal to 8.348 psf (400 Pa) resulting from a truck speed of 60 mph

(26.8 m/s) was calculated. There was evidence in the data of a horizontal load, but the

distribution of this load with height above ground level was not experimentally measured.

The maximum height of the pressure load significantly influences the magnitude of the

calculated load. The recorded strain ranges due to a horizontal load were much larger

during the natural wind tests than for the truck-induced wind tests. As a consequence, no

experimental value was obtained in the analysis for a horizontal component, assuming

natural wind will control in design for this direction.

General Fatigue Design Equation for Truck-Induced Wind Gusts

The general fatigue design equation for truck-induced wind gusts is shown as Eq.

13.4. It is to be applied vertically to the horizontal area on the underneath façade of the
323

structure. It is to be applied directly to a 12 ft (3.66 m) in length portion (or width of the

traffic lane, whichever is greater) located directly over the traffic lane. The load is applied

as a uniformly distributed load to be easily applied by the engineer. It is at maximum at a

height of 19.7 ft (6.00 m) above the roadway, and decreases linearly to zero at a height of

32.8 ft (10 m). Equation 13.4 is recommended for cantilever-type sign support structures;

specifically for support structures with a modal frequency of 1.64 Hz for vertical rocking

modal shapes in-plane with the vertically applied load.

2
 v 
PTG = 8.4C d I F   [Eq. 13.4]
 70 

where
PTG = design fatigue load due to truck - induced wind gust, psf (Pa)
C d = drag coefficient
I F = importance factor
v = traveling speed of the truck, mph (m/s)

The equation takes the form similar to the Support Specifications for easy impli-

cation. The drag coefficients and importance factors are not changed, and are provided in

the Supports Specifications. Accommodation for other traveling speeds other than 70

mph (31.3 m/s) is provided in the equation if needed, however it is not recommended use

speeds below 70 mph (31.3 m/s).


324

Detailed Fatigue Design Equation for Truck-Induced Wind Gusts

A more detailed design equation was developed to account for the variety of sign

support structures in design, each with different configurations, cross sectional shapes,

and material properties. Since these factors have significant influence on the dynamic

characteristics of the structure, and because sign supports structures are highly flexible

with low damping properties, a method was needed in determining the fatigue load based

on the vibration behavior of these structures. The approach presented in this section ac-

counts for the dynamic behavior of the structure in terms of the modal frequency. The

fatigue load is determined based on these properties. The approach was considered a uni-

fied design approach to fatigue design for support structures in that it accounts for the

variety of structures in design.

Shock response spectrums were developed that presents the fatigue load based on

the dynamic properties of the structure. The load is extracted from the SRS in terms of

the earliest modal frequency of the structure with a modal shape in the direction of the

loading. This shape is commonly referred to as the vertical rocking modal shape and has

vibratory motion in-plane to the vertically applied truck load. For sign support structures,

the vertical modal shape is close to the horizontal modal shape, and is generally around 1

to 3 Hz, which typically corresponds to the second modal shape for cantilever-type struc-

tures, and the third modal shape for bridge-type structures. Finite element software (i.e.,

SAP2000) can be used to estimate the appropriate modal shapes and their associated

natural frequencies to use with the SRS curves. If FEA software is not available, funda-

mental structural dynamics of a SDOF system (Eq. 13.5) can be used for estimating these

values (10, 17, 33, 56, 57, 61, 63). A conservative approach would be to use the peak
325

amplitude from the SRS that corresponds to the truck speed regardless of the natural fre-

quency of the structure.

1 K
fn = [Eq. 13.5]
2π M

where
f n = natural frequency, Hz
K = generalized stiffness, N/m (lb/ft)
M = generalized mass, kg (slug)

Fatigue Design Procedure.

A systematic procedure for determining the fatigue design load due to truck-

induced wind gusts based on the proposed detailed approach may be given as follows.

Step 1: design speed of the truck. The first step involves determining the design

speed of the truck. A speed of 70 mph (31.3 m/s) is recommended however other speeds

can be used if needed. It is not recommended to use speeds less than 70 mph (31.3 m/s).

Step 2: modal analysis. The second step involves a modal analysis of the support

structure. The earliest modal frequency with a modal shape in the direction of the truck

gust loading is needed. The damping ratio is not required for truck-induced gusts. Finite

element analysis software can be helpful in this step. If FEA is not available, the modal

frequency and modal shape can be estimated using Eq. 13.5.


326

Step 3: shock response spectrum. The third step involves the SRS shown in Figure

13.3. A SRS constant, PSRS, is extracted from the spectrum as the ordinate value corre-

sponding to the natural frequency determined Step 2.

FIGURE 13.3 SRS for vertical truck-induced wind gust.

Step 4: fatigue design equation. The next step is plugging all information gathered

from Steps 1-3 into the fatigue design equation for truck-induced gusts based on the de-

tailed approach. Simply input the SRS constant, PSRS, determined from Step 3 into Eq.

13.6 for each member along the facade of the structure exposed to truck-induced wind

gusts. It is to be applied in the same fashion as the general design equation for truck-

induced gusts, Eq. 13.4.


327

PTG = PSRS Cd I F [Eq. 13.6]

where
PTG = design fatigue load due to truck - induced wind gust, psf (Pa)
PSRS = transmitted pressure constant extracted from the SRS, psf (Pa)
C d = drag coefficient
I F = importance factor

Step 5: height accommodation. Accommodations for height can be made to the

values determined from Eq. 13.6 for structures higher than 19.7 ft (6.00 m). The truck

gust pressure is assumed maximum at a height of 19.7 ft (6.00 m) above the roadway, and

decreases linearly to zero at a height of 32.8 ft (10 m).

Design Examples

Examples for determining the fatigue load are presented for natural wind and

truck-induced wind gusts. The general and detail approaches were used and compared to

each other as well as to the design fatigue equation in the Supports Specifications. The

comparison tested the accuracy of the Supports Specifications with respect to the experi-

mentally determined values, as well as the theoretical approaches developed with this

project. A variety of case studies of different types of structures were used in the com-

parisons, each with different structural dynamic properties that match real structures in

operation.
328

Supports Specifications Fatigue Design Equations

The fatigue design equations for natural wind and truck-induced gusts are pro-

vided in Eq. 13.7 and Eq. 13.8 (1). The fatigue load was calculated using these equations

for the case studies presented, and the results were compared to the fatigue design equa-

tions developed with this project.

Natural Wind Gust

2
v 
PNW = 5.2C d I F   [Eq. 13.7]
 11 

where
PNW = design fatigue load due to natural wind, psf (Pa)
C d = drag coefficient
I F = importance factor
v = annual mean wind velocity other than11 mph (5 m/s)

Truck-Induced Wind Gust

2
V 
PTG = 18.8C d I F   [Eq. 13.8]
 65 

where
PTG = design fatigue load due to truck gust, psf (Pa)
C d = drag coefficient
I F = importance factor
V = truck speed, mph (m/s)
329

The above equations were used for the comparison, and were applied in the manner

specified in the Supports Specifications fatigue provisions.

Natural Wind Gust

Six different design cases were performed to illustrate the developed design ap-

proaches for calculating fatigue loads due to natural wind, and more importantly to com-

pare the results to the Supports Specifications. A variety of overhead sign support struc-

tures with different dynamic properties and exposure conditions are provided by the six

design cases listed in Table 13.3. The dynamic properties reflect sign support structures

that were evaluated in previous studies (10, 20, 38, 52). The design cases were chosen to

illustrate the potential for underestimation or overestimation of the fatigue provisions in

the Supports Specifications. This was done primarily by altering the natural frequencies

and damping ratios, which represents a practical scenario as these structures are not exact

in size, shape, stiffness, mass, and material.

The natural wind loading for the six cases was directed onto the exposed front fa-

cade of the structure (normal to the front plain of the sign and support members, in the

direction of traffic), as shown in Figure 13.4. A fatigue pressure due to natural wind was

applied to the post support only for this example. A drag coefficient equal to 1.10 was

used for the 2.0 ft (0.61 m) diameter member, and an importance factor equal 1.0. The

results of the analysis using the general and detailed approach (Eq. 13.1 and Eq. 13.3)

and the Supports Specifications (Eq. 13.7) are given in Table 13.4. The VRS used for

each case is provided in Figure 13.5 through Figure 13.8.


330

TABLE 13.3 Design Case Description for Natural Wind Gust

Design
Case Description
Case
 Annual mean wind velocity equal to 11 mph
Case 1  Structural natural frequency of 2.00 Hz in the direction of loading
 0.02 (2%) damping ratio
 Annual mean wind velocity equal to 13 mph
Case 2  Structural natural frequency of 2.00 Hz in the direction of loading
 0.02 (2%) damping ratio
 Annual mean wind velocity equal to 11 mph
Case 3  Structural natural frequency of 0.802 Hz in the direction of loading
 0.02 (2%) damping ratio
 Annual mean wind velocity equal to 11 mph
Case 4  Structural natural frequency of 2.00 Hz in the direction of loading
 0.005 (0.5%) damping ratio
 Annual mean wind velocity equal to 11 mph
Case 5  Structural natural frequency of 0.802 Hz in the direction of loading
 0.005 (0.5%) damping ratio
 Annual mean wind velocity equal to 13 mph
Case 6  Structural natural frequency of 0.802 Hz in the direction of loading
 0.005 (0.5%) damping ratio

Direction of
Wind Loading

Vibratory Modal Shape in the Di-


rection of Loading (Horizontal)

FIGURE 13.4 Direction of wind loading for natural wind gust.


331

TABLE 13.4 Design Case Results for Natural Wind Gust

Project Approach Percent


AASHTO (psf) Difference from
Supports AASHTO
Design Percent
Specifications (%)
Case General Detailed Diff. from
Eq. 13.7
(Eq. 13.1) (Eq. 13.2) General
(psf) General Detailed
(%)
Case 1 5.720 4.400 3.902 11.3 23.1 31.8
Case 2 7.989 6.145 5.449 11.3 23.1 31.8
Case 3 5.720 4.400 4.891 -11.2 23.1 14.5
Case 4 5.720 4.400 5.536 -25.8 23.1 3.22
Case 5 5.720 4.400 8.105 -84.2 23.1 -41.7
Case 6 7.989 6.145 11.320 -84.2 23.1 -41.7

FIGURE 13.5 VRS for Case 1 and Case 2.


332

FIGURE 13.6 VRS for Case 3.

FIGURE 13.7 VRS for Case 4.


333

FIGURE 13.8 VRS for Case 5 and Case 6.

The frequencies listed in Table 13.3 represent the earliest modal frequency with a

modal shape in the direction of the wind loading (see Figure 13.4). The equivalent static

wind loads extracted from the VRS figures were based on these frequencies.

Comparisons between the general and detail fatigue design equation was per-

formed, as well as a comparison with the Supports Specifications. Case 1 was considered

a control case because it closely reflected the conditions in the development of the Sup-

ports Specifications. The development of the natural wind fatigue provisions in the Sup-

ports Specifications used structures with a natural frequency of 2.00 Hz, and a critical

damping percentage equal to 2%. The structural dynamic characteristics specified for

Case 1 were considered representative of the support structures in operation. The results
334

in Table 13.4 indicated the Supports Specifications overestimated the fatigue load as

compared to the experimentally obtained values.

Large differences resulted between the general and detailed equations developed

with this research, and with the Supports Specifications. The comparison was especially

sensitive to damping percentages. Changes in the damping percentages, such as Case 4

through Case 6, resulted in very large differences. The general equation was based on the

tested structure with a damping percentage equal to 1.82%, which was close to 2.00%.

The case studies with a damping percentage equal to 0.5% resulted in an increased value

with the detailed equation, due to an increased flexibility of the structure.

Changes in damping percentages and natural frequencies resulted in extreme dif-

ferences between the approaches. Structures with a lower natural frequency encroached

closer to the broadband peak frequency of the wind, and therefore resulted in higher

resonant behavior. This along with a reduced damping set up an environment where the

structure vibrated with higher frequencies with less resistance.

As seen in Figure 13.1, the design fatigue load calculated using the Supports

Specifications was conservative for structures with a natural frequency greater than 0.65

Hz. This value will increase for critical damping percentages less than 1.5%. With struc-

tures with damping percentages equal to 2%, the Supports Specifications was conserva-

tive for natural frequencies greater than 0.5 Hz. Likewise, for increasing damping per-

centages, the Supports Specifications becomes more conservative at lower structural fre-

quencies. In general, by increasing the natural frequency of the structure and the damping

percentage, the vulnerability of the structure to fatigue damage is considerably decreased.


335

The frequency of the structure becomes further away from the peak amplitude of the

wind pressure frequency, and resonance becomes less influential to the response.

Truck-Induced Wind Gust

Three different design case scenarios were performed to illustrate the proposed

design approach for calculating fatigue loads due to truck gusts, and to compare the re-

sults to the Supports Specifications. A description of the design cases are listed in Table

13.5. The pressure loading calculated using Eq. 13.4 and Eq. 13.6 along with Figure 13.3

of the detailed approach was used for the comparison to the pressure load calculated us-

ing Eq. 13.8 of the Supports Specifications.

TABLE 13.5 Design Case Description for Truck-Induced Wind Gust

Design
Case Description
Case
 Truck speed equal to 65 mph
Case 1
 Structural natural frequency of 2.00 Hz in the direction of loading
 Annual mean wind velocity equal to 65 mph
Case 2
 Structural natural frequency of 0.802 Hz in the direction of loading
 Annual mean wind velocity equal to 70 mph
Case 3
 Structural natural frequency of 2.00 Hz in the direction of loading
 Annual mean wind velocity equal to 70 mph
Case 4
 Structural natural frequency of 0.802 Hz in the direction of loading

The same sign support structures depicted in the natural wind gust design example

section (Table 13.3) was used for the comparison. A four-chord truss overhang was as-

sumed for the case scenarios. The chords were 3.5 in (88.9 mm) in diameter and were a

distance of 20 ft (6.1 m) above ground level. This height was at the reduction threshold
336

according to the Supports Specifications (1), and therefore no accommodation for height

was calculated.

The calculations reflect the design fatigue load to be applied vertically to the hori-

zontal area on the underneath portion of the overhead truss above the traffic lane (see

Figure 13.9). A drag coefficient equal to 1.10 (calculated using the Supports Specifica-

tions provisions for drag coefficients) and an importance factor equal to 1.0 was used for

the load calculation. A summary of the results of the comparison are listed in Table 13.6.

Direction of
Wind Loading

Vibratory Modal Shape in the Di-


rection of Loading (Vertical)

FIGURE 13.9 Direction of wind loading for truck-induced wind gust.


337

TABLE 13.6 Design Case Results for Truck-Induced Wind Gust

Project Approach Percent


AASHTO (psf) Difference from
Supports AASHTO
Design Percent
Specifications (%)
Case General Detailed Diff. from
Eq. 13.8
(Eq. 13.4) (Eq. 13.6) General
(psf) General Detailed
(%)
Case 1 20.680 7.967 6.479 18.7 61.5 68.7
Case 2 20.680 7.967 4.543 43.0 61.5 78.0
Case 3 23.984 9.240 8.272 10.5 61.5 65.5
Case 4 23.984 9.240 4.719 48.9 61.5 80.3

The results from both the general and detailed design equations were significantly

less than the Supports Specifications, as the equations of this project were developed

from experimental evidence on cantilever sign support structures. When comparing the

general and detailed design equations to each other, very interesting conclusions are

gathered. The general equation is recommended for structures with the earliest modal

frequency with a modal shape in the direction of loading equal to 1.64 Hz. As seen in the

SRS of Figure 13.3, the fatigue load due to truck-induced gusts will be significantly re-

duced for structures with natural frequencies as furthest away, greater or smaller, from

approximately 1.6 Hz for trucks traveling at 70 mph (31.3 m/s). Structures with frequen-

cies at 1.6 Hz will see the largest vibration response due to resonance with the truck in-

duced wind pulse. In view of the results gathered with natural wind gusts, it is advisable

to design structures with modal frequencies of the first two modes of vibration higher

than 1.60 Hz to comply with both natural wind and truck-induced wind gust loading.
338

CHAPTER 14

SUMMARY AND CONCLUSIONS

Chapter Overview

A summary of the findings of this project is presented in this chapter. It is divided

into sections corresponding to the major tasks performed in this project including:

 Fatigue resistance,

 Modal analysis,

 Fatigue load due to natural wind gust,

 Fatigue load due to truck-induced wind gust, and

 Finite element analysis method.

Specific conclusions drawn from this project are enumerated as follows:

1. The theoretical models using the VRS for natural wind gust and SRS for truck-

induced wind gust were accurate in determining the fatigue load for the experi-

mentally tested structure with the specific dynamic properties of vibration.

2. General and detailed fatigue design equations were developed from the theoretical

and experimental programs for natural wind. The general equation is recom-

mended for structures with a modal frequency of the earliest modal shape in the
339

direction of loading equal to 1.61 Hz or greater, and a critical damping percentage

equal to 1.82% or greater. The detailed equation is recommended for structures

with natural frequencies less than 1.61 Hz and critical damping percentages less

than 1.82%. The detailed equation can also be used for natural frequencies and

damping percentages greater than 1.82% as the general equation is slightly con-

servative for these structures.

3. General and detailed fatigue design equations were developed from the theoretical

and experimental programs for truck-induced gust. The general equation is rec-

ommended for structures with a modal frequency of the earliest modal shape in

the direction of loading equal to 1.64 Hz or greater. The detailed equation is rec-

ommended for structures with natural frequencies significant greater or less than

1.64 Hz.

4. The Supports Specifications when compared to the developed equations for natu-

ral wind can be conservative for structures with natural frequencies greater than

0.50 Hz. The limiting frequency begins to increase for structures with damping

percentages less than 2.0%.

5. The Supports Specifications when compared to the developed equations for truck-

induced gust were highly conservative for all structures.

6. The fatigue stress generated on the tested cantilever structure from natural wind

gust was as much as four times greater than the fatigue stress generated from

truck-induced wind gust.


340

7. Support structures with natural frequencies greater than 1.60 Hz are less vulner-

able to fatigue loading. The higher the natural frequency of the structure, the less

susceptible the structure becomes to wind induced fatigue loads.

8. Structural vibration became momentous for wind speeds greater than 9 mph (4

m/s). Vibration was minimal for lower wind speeds.

9. The anchor bolts with clearances 3 in (76.2 mm) or larger had stress ranges over

the constant-amplitude threshold specified by the Supports Specifications. This

was due to an increased significance of the bending moment stress in the anchor

bolt, and was especially critical during natural wind loading where high torsion

exists in the post and is transferred to the base plate.

Fatigue Resistance

An analysis of the fatigue resistance of the structure was performed. The stress

ranges determined in the structure were compared to the constant-amplitude threshold

provided by the Supports Specifications. A failure index was calculated to provide a rela-

tionship between the gauged locations with a control value. The distribution of the stress

in the structure was evaluated and the most appropriate location for determining the fa-

tigue load was identified.

The largest strain recorded was located on the anchor bolts followed by the post

and chords, as expected. The major findings are listed in Table 14.1 and 14.2 for natural

wind and truck-induced gusts. The table lists the stress ranges determined at the average

wind velocity and the fatigue wind velocity. Maximum ranges at the specified location

are listed for each wind velocity, as well as the average stress range from all gauges at the
341

section. The stress range was defined as the peak-to-peak range from the measured re-

sponse time histories.

TABLE 14.1 Experimental Stress Ranges for Natural Wind Tests

Location Stress Range (ksi)


on the Measurement Type Wind Velocity (mph)
Average Maximum
Structure
Average Wind 13 0.200 0.245
Chord Normal
Fatigue Wind 38 1.111 1.347
Average Wind 13 0.184 0.224
Normal
Fatigue Wind 38 1.089 1.359
Post
Average Wind 13 0.210 0.234
Shear
Fatigue Wind 38 1.332 1.525
Anchor Average Wind 13 0.608 0.967
Normal
Bolt Fatigue Wind 38 4.07 7.02

The evaluation of the fatigue resistance results provided an understanding of the

stress distribution throughout the structure. Strains measured in the chords near the

chord-to-column connection matched well with the measured strains on the post. The

strains on the post were slightly larger in proportion to the additional exposed wind area

at these locations. The chords recorded loads created from the exposed area of the sign

and truss, whereas the post recorded loads from the sign, truss, and post members.

The structure as a whole was within the threshold limitations, with the exception

of the anchor bolts. The stress distribution between anchor bolts from the base plate to the

concrete foundation was not consistent. The anchor bolts with large clearances of 3 in

(76.2 mm) or more were heavily loaded, much more than the other bolts. The majority of

stress was concentrated in these bolts and controlled the overall stability of the structure.
342

This was believed to be due to the larger influence of out-of-plane forces induced onto

the bolt from the base plate, which created large bending moments in the bolt.

TABLE 14.2 Experimental Stress Ranges for Truck-Induced Wind Tests

Location on the Measurement Truck Speed Stress Range (ksi)


Structure Type (mph) Average Maximum
60 0.241 0.334
Chord Normal 65 0.242 0.339
70 0.253 0.350
60 0.244 0.338
Normal 65 0.215 0.317
70 0.220 0.351
Post
60 0.260 0.375
Shear 65 0.255 0.360
70 0.297 0.387
60 0.880 1.670
Anchor Bolt Normal 65 0.788 1.602
70 1.045 5.096

The most consistent instrumented location on the structure was the post member.

The best strain results were produced at these locations, showing the balanced state of

strain with respect to the proximity of each gauge to the neutral axis created from the

magnitude and direction of the load. A high degree of variability in stress ranges were

found in the anchor bolts due to the inconsistent clearances between the bolts. A degree

of variability was also observed in the chords. In view of the results, the fatigue load

back-calculation methodology was performed using the strain gauge readings on the post

member. The calculation was more straightforward, and provided consistent results.

For the truck tests, the stress ranges on averaged increased with the speed of the

truck. As compared to the natural wind gust measurements, no significant loading was
343

generated from the truck gusts. This is illustrated in the bar graph of Figure 14.1, showing

the stress ranges for the fatigue wind as compared to the stress ranges for the 70 mph

(31.3 m/s) truck speed. A significant increase in stress range for both natural wind and

truck gusts was observed for the anchor bolts. This was due to the anchor bolts with large

clearances as compared to the other bolts, resulting in an unbalanced distribution of

stress. The values shown in the figure represent the maximum stress ranges recorded at

each section.

Modal Analysis

Experimental and analytical modal analysis was conducted on the tested structure.

Modal frequencies, modal damping, and modal shapes were indentified from the analysis.

The results of the two programs (experimental and analytical) matched well with each

other, showing good modeling simulation of dynamic structural behavior.

The experimental program involved forming FRF from the structural excitation

and response measurements. The accelerometer data proved to be the best instrumenta-

tion to use for the response measurements. Modal frequencies and damping values can be

determined from the strain gauge data, of which matched the accelerometer data, how-

ever corresponding modal shapes cannot be determined using strain gauges.

The analytical program for modal analysis involved FEA using the software

package SAP2000. The modal shapes and modal frequencies were close to the measured

values, indicating an accurate modal analysis with both approaches. This helped to vali-

date the modal was performing accurately in depicting the structural behavior under load.
344

FIGURE 14.1 Comparison of results between natural wind and truck gusts.

Fatigue Load due to Natural Wind Gust

The fatigue load due to natural wind gusts was determined. A general design

equation and a more detailed design equation using the VRS were developed. Compari-

sons were made between the two approaches and with the Supports Specifications fatigue

design equation. The results indicated Supports Specifications were conservative for cer-

tain situations.

It was concluded from the analysis that the design fatigue load calculated using

the Supports Specifications was conservative for structures with a natural frequency

greater than 0.50 Hz. This value increased for critical damping percentages less than

1.82%. Likewise, for increasing damping percentages, the Supports Specifications be-

came more conservative at lower structural frequencies. By increasing the natural fre-
345

quency of the structure and the damping percentage, the vulnerability of the structure to

fatigue damage was decreased. Stress is induced onto the structure when it vibrates;

thereby increasing the frequency and damping values will result in smaller vibration am-

plitudes. More importantly, the modal frequency of the structure becomes further away

from the resonant frequency of excitation and response, and therefore resonance has a

lessoned influence.

General Fatigue Design Equation for Natural Wind Gust

The general fatigue design equation for natural wind gusts is shown as Eq. 14.1. It

is to be applied horizontally to the exposed façade of the structure as seen on an elevation

view. The load is applied as a uniformly distributed load to be easily applied by the engi-

neer. It is recommended for cantilever-type sign support structures; specifically for sup-

port structures with a modal frequency of 1.61 Hz, and a critical damping percentage of

1.82%.

2
v 
PNW = 4C d I F   [Eq. 14.1]
 11 

where
PNW = design fatigue load due to natural wind gust, psf (Pa)
C d = drag coefficient
I F = importance factor
v = annual mean three second wind velocity wind velocity, mph (m/s)
346

The equation takes the form similar to the Support Specifications for easy impli-

cation. The drag coefficients and importance factors are not changed, and are provided in

the Supports Specifications. Accommodation for other annual mean wind velocities other

than 11 mph (5 m/s) is provided in the equation.

Detailed Fatigue Design Equation for Natural Wind Gust

A more detailed design equation was developed to account for the variety of sign

support structures in design, each with different configurations, cross sectional shapes,

and material properties. Since these factors have significant influence on the dynamic

characteristics of the structure, and because sign supports structures are highly flexible

with low damping properties, a method was needed in determining the fatigue load based

on the vibration behavior of these structures. The approach presented in this section ac-

counted for the dynamic behavior of the structure in terms of the modal frequency and

critical damping percentages, and the fatigue load was determined based on these proper-

ties. The approach was considered a unified design approach to fatigue of natural wind

gusts in that it accounts for the variety of structures in design.

Vibration response spectrums were developed that presented the fatigue load

based on the dynamic properties of the structure. The load is extracted from the VRS in

terms of the first modal frequency of the structure with a modal shape in the direction of

the loading. The most critical loading scenario for natural wind is directed normal to the

plain of the sign (in the direction of traffic), which excites the modal shape most com-

monly referred to as the horizontal modal shape. For sign support structures, the horizon-

tal modal shape is generally around 1 to 3 Hz, which typically corresponds to the first
347

modal shape for cantilever-type structures, and the second modal shape for bridge-type

structures. Finite element software (i.e., SAP2000) can be used to estimate the appropri-

ate modal shapes and their associated natural frequencies to use with the VRS curves. If

FEA software is not available, fundamental structural dynamics of a SDOF system (Eq.

14.2) can be used for estimating these values (10, 17, 33, 56, 57, 61, 63).

1 K
fn = [Eq. 14.2]
2π M

where
f n = natural frequency, Hz
K = generalized stiffness, N/m (lb/ft)
M = generalized mass, kg (slug)

A conservative estimate would be to assume a modal frequency of 1 Hz to use with the

VRS, as the equivalent static wind load increases as the modal frequency decreases be-

cause of the proximity to the broadband amplitude of the wind pressure PSD excitation

used in its development.

The critical damping percentage is especially relevant to sign support structures

because of their relatively high flexibility and subsequent low natural frequencies (1 to 3

Hz). What’s more, their damping ratios are mostly below 2.0%. A low damping will al-

low the structure to vibrate longer at high amplitudes, and thus produce more stress that

could potentially cause fatigue damage. It is recommended for design purposes to assume

a damping ratio equal to 0.015 (1.5%). However, damping ratios can vary depending on
348

the structural material, and therefore actual values can be obtained from experimental

data of comparable structures if available.

Fatigue Design Procedure

A systematic procedure for determining the fatigue design load due to natural

wind based on the proposed detailed approach may be given as follows.

Step 1: annual mean wind velocity. The first step involves determining the annual

mean wind velocity of the site. The National Weather Service Offices near the site can be

used to determine this value. It is recommended to use a value of 11 mph (5 m/s), how-

ever other values can be used if needed.

Step 2: modal analysis. The second step involves a modal analysis of the design-

ing support structure. The earliest modal frequency of the modal shape in the direction of

the natural wind loading is needed, as well as the critical damping percentage. Finite ele-

ment analysis software can be helpful in this step. If FEA is not available, the modal fre-

quency and modal shape can be estimated using Eq. 14.2. If experimental data of compa-

rable structures are not available, it is recommended to use a damping ratio estimated at

0.015 (1.5%).

Step 3: vibration response spectrum. The third step involves the VRS shown in

Figure 14.2 and Figure 14.3. A VRS constant, PVRS, is extracted from the spectrum as the

ordinate value corresponding to the natural frequency and damping ratio determined in
349

Step 2. If using a damping ratio of 1.5% (recommended), use the VRS in Figure 14.2.

The VRS in Figure 14.3 is provided for other damping ratios if needed.

Step 4: fatigue design equation. The final step is plugging all information gath-

ered from Steps 1-1 into the fatigue design equation for natural wind gusts based on the

detailed approach. Simply input the VRS constant, PVRS, determined from Step 3 into Eq.

14.3 for each member along the facade of the structure exposed to natural wind.

2
v 
PNW = PVRS C d I F   [Eq. 14.3]
 11 

where
PNW = design fatigue load due to natural wind gust, psf (Pa)
PVRS = transmitted pressure constant extracted from the VRS, psf (Pa)
C d = drag coefficient
I F = importance factor
v = annual mean three second wind velocity, mph (m/s)
350

FIGURE 14.2 VRS for 1.5% damping ratio.

FIGURE 14.3 Fatigue wind pressure VRS for damping range.


351

Fatigue Load due to Truck-Induced Wind Gust

The fatigue load due to truck-induced wind gusts was determined. A general de-

sign equation and a more detailed design equation using the SRS were developed. Com-

parisons were made between the two approaches and with the Supports Specifications

fatigue design equation. The results indicated Supports Specifications were highly con-

servative.

It was concluded from the analysis that the fatigue load due to truck-induced gusts

was significantly reduced for structures with natural frequencies as furthest away, greater

or smaller, from approximately 1.6 Hz for trucks traveling at 70 mph (31.3 m/s). Struc-

tures with frequencies at 1.6 Hz experienced the largest vibration response due to reso-

nance with the truck induced wind pulse. In view of the result with natural wind gusts, it

is advisable to design structures with modal frequencies of the first two modes of vibra-

tion higher than 1.60 Hz to comply with both natural wind and truck-induced wind gust

loading.

General Fatigue Design Equation for Truck-Induced Wind Gust

The general fatigue design equation for truck-induced wind gusts is shown as Eq.

14.4. It is to be applied vertically to the horizontal area on the underneath façade of the

structure. It is to be applied directly to a 12 ft (3.66 m) in length portion (or width of the

traffic lane, whichever is greater) located directly over the traffic lane. The load is applied

as a uniformly distributed load to be easily applied by the engineer. It is maximum at a

height of 19.7 ft (6.00 m) above the roadway, and decreases linearly to zero at a height of

32.8 ft (10 m). Equation 14.4 is recommended for cantilever-type sign support structures;
352

specifically for support structures with a modal frequency of 1.64 Hz for vertical rocking

modal shapes in-plane with the vertically applied load.

2
 v 
PTG = 8 . 4C d I F   [Eq. 14.4]
 70 

where
PTG = design fatigue load due to truck - induced wind gust, psf (Pa)
C d = drag coefficient
I F = importance factor
v = traveling speed of the truck, mph (m/s)

The equation takes the form similar to the Support Specifications for easy impli-

cation. The drag coefficients and importance factors were not changed, and are provided

in the Supports Specifications. Accommodation for other traveling speeds other than 70

mph (31.3 m/s) is provided in the equation if needed, however it is not recommended use

speeds below 70 mph (31.3 m/s).

Detailed Fatigue Design Equation for Truck-Induced Wind Gust

A more detailed design equation was developed to account for the variety of sign

support structures in design, each with different configurations, cross sectional shapes,

and material properties. Since these factors have significant influence on the dynamic

characteristics of the structure, and because sign supports structures are highly flexible

with low damping properties, a method is needed in determining the fatigue load based

on the vibration behavior of these structures. The approach presented in this section ac-
353

counts for the dynamic behavior of the structure in terms of the modal frequency. The

fatigue load is determined based on these properties. The approach is considered a unified

design approach to fatigue design for support structures in that it accounts for the variety

of structures in design.

Shock response spectrums were developed that presents the fatigue load based on

the dynamic properties of the structure. The load is extracted from the SRS in terms of

the earliest modal frequency of the structure with a modal shape in the direction of the

loading. This shape is commonly referred to as the vertical rocking modal shape and has

vibratory motion in-plane to the vertically applied truck load. For sign support structures,

the vertical modal shape is close to the horizontal modal shape, and is generally around 1

to 3 Hz, which typically corresponds to the second modal shape for cantilever-type struc-

tures, and the third modal shape for bridge-type structures. Finite element software (i.e.,

SAP2000) can be used to estimate the appropriate modal shapes and their associated

natural frequencies to use with the SRS curves. If FEA software is not available, funda-

mental structural dynamics of a SDOF system (Eq. 14.5) can be used for estimating these

values (10, 17, 33, 56, 57, 61, 63).

1 K
fn = [Eq. 14.5]
2π M

where
f n = natural frequency, Hz
K = generalized stiffness, N/m (lb/ft)
M = generalized mass, kg (slug)
354

A conservative approach would be to use the peak amplitude from the SRS that corre-

sponds to the truck speed regardless of the natural frequency of the structure.

Fatigue Design Procedure

A systematic procedure for determining the fatigue design load due to truck-

induced wind gusts based on the proposed detailed approach may be given as follows.

Step 1: design speed of the truck. The first step involves determining the design

speed of the truck. A speed of 70 mph (31.3 m/s) is recommended however other speeds

can be used if needed. It is not recommended to use speeds less than 70 mph (31.3 m/s).

Step 2: modal analysis. The second step involves a modal analysis of the design-

ing support structure. The earliest modal frequency with a modal shape in the direction of

the truck gust loading is needed. The damping ratio is not required for truck-induced

gusts. Finite element analysis software can be helpful in this step. If FEA is not available,

the modal frequency and modal shape can be estimated using Eq. 14.5.

Step 3: shock response spectrum. The third step involves the SRS shown in Figure

14.4. A SRS constant, PSRS, is extracted from the spectrum as the ordinate value corre-

sponding to the natural frequency determined Step 2.


355

1 mph = 0.447 m/s


1psf = 47.9 Pa

FIGURE 14.5 SRS for vertical truck-induced wind gust.

Step 4: fatigue design equation. The next step is plugging all information gathered

from Steps 1-3 into the fatigue design equation for truck-induced gusts based on the de-

tailed approach. Simply input the SRS constant, PSRS, determined from Step 3 into Eq.

14.6 for each member along the facade of the structure exposed to truck-induced wind

gusts. It is to be applied in the same fashion as the general design equation for truck-

induced gusts, Eq. 14.4.

PTG = PSRS Cd I F [Eq. 14.6]


356

where
PTG = design fatigue load due to truck - induced wind gust, psf (Pa)
PSRS = transmitted pressure constant extracted from the SRS, psf (Pa)
C d = drag coefficient
I F = importance factor

Step 5: height accommodation. Accommodations for height can be made to the

values determined from Eq. 14.6 for structures higher than 19.7 ft (6.00 m). The truck

gust pressure is assumed maximum at a height of 19.7 ft (6.00 m) above the roadway, and

decreases linearly to zero at a height of 32.8 ft (10 m).

Finite Element Analysis Method

The FEA model of the support structure was used to double check the back-

calculations for natural wind and truck-induced wind gusts. It was also used to compare

dynamic loading to the equivalent static wind load gathered from the VRS and SRS.

Verification of the model was established by comparing the experimental modal analysis

with the FEA. Alterations as needed primarily concerned the connection of the structure

to the foundation was performed to calibrate the model.

It was found that the back-calculation from the measured strain values to the wind

pressure load was accurate. The wind pressure developed from the experimental data was

inputted in the FEA as a uniformly distributed load applied to the exposed portion of the

structure to wind gusts. The strain at the instrumented location was extracted from the

linear-static solution and compared to the strain values measured experimentally. The re-

sults are listed in Table 14.3. The loading produced from the average wind velocity was

used in the natural wind comparison because experimental strain at the fatigue wind was
357

not measured. The results showed close agreement, indicating the structural analysis used

in the back-calculation for the wind pressure was accurate.

TABLE 14.3 Results of Back-Calculation Verification

Analysis Program Percent Difference


Wind Loading Experimental Strain from Experimental
FEA Strain (in/in)
Range (in/in) (%)
Natural Wind Gusts 7.84(10-6) 7.90 (10-6) 0.759
Truck-Induced
9.49(10-6) 9.245(10-6) -2.53
Wind Gusts

The FEA was also used to verify the use of the VRS and SRS as an accurate ap-

proach in determining the fatigue load based on the dynamic characteristics of the struc-

ture. The methods were based on approximating the dynamic behavior of the support

structure as a SDOF. Material properties of the model were changed and the analysis was

performed again. The results indicated the VRS and SRS were accurate in determining

the fatigue load.

Contributions to the Practice

The results from this study were seen as a positive step towards more reliable

loads used in the design of sign support structures. The theoretical program was devel-

oped with the intention for application with the variety of sign supports structures used in

design. An experimental program was produced to check the accuracy of this program.

The results indicate that the developed methodology can be used to design sign supports
358

structures using on more realistic design loads. The impact of the new criteria would

greatly enhance the design community to ensure the safety as well as economy in design.

The applicability of the developed methodology to the design community was

also a major consideration. The VRS and SRS for natural wind and truck-induced wind

gusts respectively can be developed into a table format that can be easily and readily used

by the design engineer to determined design fatigue loads based on the particular struc-

ture in design. The table format would account with variability in sign supports structures

including natural frequency, damping properties, as well as alterations in annual mean

wind velocity and truck speed.

Recommendations for Future Research

Addressing the accuracy of the theoretical program including the VRS and SRS

was the major recommendation for future research. The idea was to developed a universal

methodology that could be used for the variety of sign supports structures in design in-

cluding bridge-type sign supports structures as well as VMS support structures. As this

research proved, the methodology was accurate for application with cantilever-type sign

supports structures that can be approximated as a SDOF. However, future aspirations

were to apply this method to the full variety of sign supports structures. The method was

believed to be applicable to any structure that can be approximated as a SDOF and is ex-

posed to natural wind and truck-induced wind loading. Research is needed to accurately

and scientifically prove that bridge structures and VMS structures actually fall into this

category.
359

LIST OF REFERENCES

1 AASHTO, Standard Specifications for Structural Supports for Highway Signs,


Luminaires and Traffic Signals, 5th Edition, American Association of State High-
way and Transportation Officials, Washington, D.C., 2009.

2 AASHTO, Standard Specifications for Structural Supports for Highway Signs,


Luminaires and Traffic Signals. Fourth Edition, American Association of State
Highway and Transportation Officials, Washington, D.C., 2001.

3 AASHTO, Standard Specifications for Structural Supports for Highway Signs,


Luminaires and Traffic Signals. Third Edition, American Association of State
Highway and Transportation Officials, Washington, D.C., 1994 78.

4 AASHTO (American Association of State Highway and Transportation Officials).


Standard Specifications for Structural Supports for Highway Signs, Luminaires
and Traffic Signals. Washington, D.C., 1985.

5 AASHTO, AASHTO LRFD Bridge Design Specifications, Customary U.S. Units,


4th Edition with 2008 and 2009 U.S. Edition Interims, and AASHTO LRFD Bridge
Design Specifications, SI Units, 4th Edition, American Association of State High-
way and Transportation Officials, Washington, D.C., 2007.

6 AASHTO, Standard Specifications for Transportation Materials and Methods of


Sampling and Testing, 29th Edition and AASHTO Provisional Standards, 2009
Edition, American Association of State Highway and Transportation Officials,
Washington, D.C., 2009.

7 Albert, M.N., Manuel, L., Frank, K.H., and Wood, S.L., Field Testing of Cantile-
vered Traffic Signal Structures under Truck-Induced Gust Loads. Report No.
FHWA/TX-08/0-4586-2, Center for Transportation Research, University of Texas
at Austin, 2007.

8 Azzam, D.; Fatigue Behavior of Highway Welded Aluminum Light Pole Support
Structures. Dissertation, University of Akron, May 2006.
360

9 Cali, P., and Covert, E.E., On the Loads on Overhead Sign Structures in Still Air
by Truck Induced Gusts. Wright Brothers Facility Report 8-97, Massachusetts In-
stitute of Technology.

10 Creamer, B. M., Frank, K. H., Klingner, R. E. Fatigue Loading on Cantilever Sign


Support Structures From Truck Wind Gusts. Research Report Number 209-1F.
Texas State Department of Highways and Public Transportation, Transportation
Planning Division, April, 1979.

11 CSI Analysis Reference Manual. Computers and Structures, Inc., Berkeley, Cali-
fornia, January, 2007.

12 Cook, Ronald A., Bloomquist, D., Agosta, A.M., Taylor, K.F., Wind Load Data
for Variable Message Signs. Report Number 0728-9488. Florida Department of
Transportation, Research Management Center, April, 1996.

13 Davenport, A. G. The Spectrum of Horizontal Gustiness Near the Ground in High


Winds. Quarterly Journal, Royal Meteorological Society, Vol. 87, London 1961.

14 Dexter, R. J., and Ricker, M. J. Fatigue-Resistant Design of Cantilevered Signal,


Signs, and Light Supports. NCHRP Report 469, The Transportation Research
Board, Washington, DC, 2002.

15 Dexter, R. J., Johns, K. W. Development of Fatigue Design Load Ranges for Can-
tilevered Sign and Signal Support Structures, Journal of Wind Engineering and
Industrial Aerodynamics, 77&78, 1998, pp. 315-326.

16 DeSantis, P.V., and Haig, P.E., Unanticipated Loading Causes Highway Sign
Failure. Proceedings of ANSYS Convention, 1996.

17 Edwards, J.A., and Bingham, W.L. Deflection Criteria for Wind Induced Vibra-
tions in Cantilever Highway Sign Structures. Report No. FHWA/NC/84-001, Cen-
ter for Transportation Engineering Studies, North Carolina State University, 1984.

18 Fackler, W.C., Pusey, H.C., Volin, R.H., Schell, E.H. Equivalence Techniques for
Vibration Testing. The Shock and Vibration Information Center, Naval Research
laboratory, Code 6020, Washington D.C., 1972.

19 Fisher, J. W., Keating, P. B., Nussbaumer, A., Yen, B. T. Resistance of Welded


Details Under Variable Amplitude Long-Life Fatigue Loading. NCHRP Report
354, Transportation Research Board, Washington D. C., 1993.

20 Foutch, D. A., Kim, T. W., LaFave, J. M., Rice, J. A. Evaluation of Aluminum


Highway Sign Truss Designs and Standards for Wind and Truck Gust Loadings.
Research Report Number 153. Illinois Department of Transportation, Bureau of
Materials and Physical Research, Physical, December, 2006.
361

21 Fouad, Fouad H.; and Calvert, Elizabeth A., New Wind Design Criteria for Traffic
Signal Support Structures, UTCA Report Number 04219, University Transporta-
tion Center for Alabama, Tuscaloosa, AL (August 31, 2005) 49pp.,
http://utca.eng.ua.edu/projects/final_reports/04219fnl.pdf.

22 Fouad, Fouad H., and Calvert, Elizabeth. “Design of Cantilevered Overhead Sign
Supports,” Transportation Research Record, Issue No. 1928, Transportation Re-
search Board, Washington, D.C., 2005.

23 Fouad, Fouad H.; and Calvert, Elizabeth A., AASHTO 2001 Design of Overhead
Cantilevered Sign Supports, UTCA Report Number 02216, University Transporta-
tion Center for Alabama, Tuscaloosa, AL, March 30, 2004,
http://utca.eng.ua.edu/projects/final_reports/02216fnl.pdf.

24 Fouad, Fouad H.; Calvert, Elizabeth; Davidson, James S.; and Delatte, Norbert.
“Proposed Revisions to AASTHO 2001 Supports Specifications,” Compendium of
Papers CD-ROM, Transportation Research Board, 83rd Annual Meeting, Wash-
ington, DC, Jan. 11-15, 2004.

25 Fouad, Fouad H., and Calvert, Elizabeth. “Wind Load Provisions in the 2001
AASHTO Supports Specifications,” Transportation Research Record, Issue No.
1845, Transportation Research Board, Washington, D.C., 2003.

26 Fouad, Fouad H.; Davidson, James S.; Delatte, Norbert; Calvert, Elizabeth A.;
Chen, Shen-En; Nunez, Edgar; and Abdalla, Ramy, “Structural Supports for
Highway Signs, Luminaires, and Traffic Signals.” NCHRP Report 494, Transpor-
tation Research Board, Washington, D.C., 2003.

27 Fouad, Fouad H.; Davidson, James S.; Delatte, Norbert; Calvert, Elizabeth A.;
Chen, Shen-En; Nunez, Edgar; and Abdalla, Ramy, Structural Supports for High-
way Signs, Luminaires, and Traffic Signals: Draft Final Report, NCHRP Project
17-10(2), Prepared for National Cooperative Highway Research Program, Trans-
portation Research Board, National Research Council, The University of Alabama
at Birmingham, Birmingham, AL, May 2002.

28 Fouad, Fouad H. and Elizabeth Calvert. “Impact Of The New Wind Load Provi-
sions On The Design Of Structural Supports,” Compendium of Papers CD-ROM,
Transportation Research Board, 82nd Annual Meeting, Washington, DC, Jan. 12-
16, 2003.

29 Fouad, Fouad H. and Elizabeth Calvert. “Wind Load Provisions in the 2001 Sup-
ports Specifications,” Compendium of Papers CD-ROM, Transportation Research
Board, 82nd Annual Meeting, Jan. 12-16, 2003.
362

30 Fouad, Fouad H.; and Calvert, Elizabeth A., Evaluating the Design Safety of
Highway Structural Supports, UTCA Report Number 00218, University Transpor-
tation Center for Alabama, Tuscaloosa, AL, August 2001,
http://utca.eng.ua.edu/projects/final_reports/00218report.htm.

31 Fouad, Fouad H.; Nunez, Edgar; and Calvert, Elizabeth A. “Proposed Revisions
to AASHTO Standard Specifications for Structural Supports for Highway Signs,
Luminaires, and Traffic Signals,” Transportation Research Record, Issue No.
1656, Transportation Research Board, Washington, D.C., 1999.

32 Fouad, Fouad H.; Mehta, Kishor C.; and Calvert, Elizabeth A., Wind Loads Re-
port: Final Draft. NCHRP Project 17-10(2), Prepared for National Cooperative
Highway Research Program, Transportation Research Board, National Research
Council, The University of Alabama at Birmingham, Birmingham, AL, Sept.
1999.

33 Fouad, Fouad H.; Calvert, Elizabeth A.; and Nunez, Edgar, “Structural Supports
for Highway Signs, Luminaires and Traffic Signals.” NCHRP Report 411, Trans-
portation Research Board, Washington, D.C., 1998.

34 Fouad, Fouad H.; Calvert, Elizabeth A.; and Nunez, Edgar, Structural Supports for
Highway Signs, Luminaires, and Traffic Signals: Draft Final Report. NCHRP
Project No. 17-10, Department of Civil and Environmental Engineering, The Uni-
versity of Alabama at Birmingham, Birmingham, AL, Sept. 1997.

35 Fouad, Fouad H.; Calvert, Elizabeth A.; and Nunez, Edgar, Structural Supports for
Highway Signs, Luminaires, and Traffic Signals: Draft Final Specification.
NCHRP Project No. 17-10, Department of Civil and Environmental Engineering,
The University of Alabama at Birmingham, Birmingham, AL, Sept. 1997.

36 Gilani, A., and Whittaker, A., “Fatigue-Life Evaluation of Steel Post Structures I:
Background and Analysis.” Journal of Structural Engineering, ASCE, New York,
NY, 2000.

37 Ginal, A.S., Chavez, J.W., and Whittaker, A.S., Fatigue-Life Evalutation of


Changeable Message Sign Structures, Volume 1 – As Built Structures. Report No.
UCB/EERC-97/10, Earthquake Engineering Research Center, University of Cali-
fornia, Berkeley, CA, 1997.

38 Gray, B.D., Fatigue Effects on Traffic Signal Structures. Thesis, University of


Wyoming, December, 1999.
363

39 Harris, C. M. Shock and Vibration Handbook, 4th Ed. McGraw-Hill Companies,


Inc., 1996.

40 Irvine, T. An Introduction to Random Vibration. Vibration Data Publications, Oc-


tober 26, 2000. www.vibrationdata.com.

41 Irvine, T. An Introduction to the Vibration Response Spectrum. Vibration Data


Publications, June 14, 2000. www.vibrationdata.com.

42 Irvine, T. An Introduction to Spectral Functions. Vibration Data Publications,


March 3, 2000. www.vibrationdata.com.

43 Irvine, T. P95/50 Rule—Theory and Applications. Vibration Data Publications,


June 20, 1996. www.vibrationdata.com.

44 Irvine, T. Methods for Converting a Power Spectral Density to a Shock Response


Spectrum. Vibration Data Publications, November 16, 2007.
www.vibrationdata.com.

45 Irvine, T. Sine Function Identification and Removal. Vibration Data Publications,


August 14, 2000. www.vibrationdata.com.

46 Irvine, T. Enveloping Data Via the Vibration Response Spectrum. Vibration Data
Publications, October 11, 1999. www.vibrationdata.com.

47 Irvine, T. An Introduction to the Shock Response Spectrum. Vibration Data Publi-


cations, May 24, 2002. www.vibrationdata.com.

48 Irvine, T. Equivalent Static Loads for Random Vibration. Vibration Data Publica-
tions, March 13, 2008. www.vibrationdata.com.

49 Irvine, T. Random Vibration Fatigue. Vibration Data Publications, August 29,


2003. www.vibrationdata.com.

50 Irwin, H.P., and Peeters, M. An Investigation of the Aerodynamic Stability of


Slender Sign Bridges, Calgary. LTR-LA-246, national Research Council Canada-
Aeronautical Establishment, 1980.

51 James, M.L., Smith, G.M., Wolford, J.C., Whaley, P.W., Vibration of Mechanical
and Structural Systems: With Microcomputer Applications. Harper & Row, Pub-
lishers, Inc., 1989.

52 Kaczinski, M. R., Dexter, R. J., and Van Dien, J. P. Fatigue Resistant Design of
Cantilevered Sign, Signal and Light Supports, NCHRP Report 412, Transportation
Research Board, Washington, D.C., 1998.
364

53 Kashar, L., Nester, M.R., Johns, J.W., Hariri, M., and Freizner, S., Analysis of the
Catastrophic Failure of the Support Structure of a Changeable Message Sign.
Structural Engineering in the 21st Century, Proceedings of the 1999 Structures
Congress, New Orleans, LA, 1115-118, 1999.

54 Liu, H. Wind Engineering: A Handbook for Structural Engineers. Prentice-Hall,


Inc., 1991.

55 McDonald, J.R., Mehta, K.C., Oler, W., and Pulipaka, N., Wind Load Effects on
Signs, Luminaires and Traffic Signal Structures. Texas Department of Transpor-
tation Report No. 1303-1F, Wind Engineering Research Center-Texas Tech Uni-
versity, Lubbaock, TX, 1995.

56 Paz, M. Structural Dynamics: Theory and Computation, 4th Ed. Chapman & Hall,
1997.

57 Scanlan, R. H., Simiu, E. Wind Effects on Structures: An Introduction to Wind En-


gineering. John Wiley & Sons, Inc., 1978.

58 Sheskin, D. J. Handbook of Parametric and Nonparametric Statistical Proce-


dures, 4th Ed. Taylor & Francis Group, LLC, 2007.

59 South, S. M. Fatigue Analysis of Overhead Sign and Signal Structures. Report


Number 115. Illinois Department of Transportation, Bureau of Materials and
Physical Research, Physical, May, 1994.

60 Ramy, A.S., Fatigue Resistant Design of Non-Cantilevered Sign Support Struc-


tures. Thesis, University of Alabama at Birmingham, 2000.

61 Rao, S.S. Mechanical Vibrations, 3rd Ed. Addison-Wesley Publishing Company,


Inc., 1995.

62 Walpole, R.E., Myers, R.H., Myers, S.L., Ye, K., Probability & Statistics for En-
gineers& Scientists, 7th Ed. Prentice Hall, Inc., 2002.

63 Yang, C. Y. Random Vibration of Structures. John Wiley & Sons, Inc., 1986.

64 Zalewski, B., Huckelbridge, A., Dynamic Load Environment of Bridge-Mounted


Sign Support Structures. Report No. ST/SS/05-002. Ohio Department of Transpor-
tation, Office of Research and Development, September 2005.
365

APPENDIX A

INSTRUMENTATION IDENTIFICATION AND LAYOUTS


366

A detail of the instrumentation layout and identification are provided. They were used to

help correspond instrument location with the collected data.

Strain Gauges
367
368
369
370
371

Accelerometers
372

Anemometers
373
374

APPENDIX B

ANEMOMETER MOUNTING INSTRUCTIONS


375

 The support tube requires three 3 equally spaced holes, tapped M5, 7.5 mm from

the top of the tube.

 Pass the cable (fitted with the 9 way Clipper plug) through the tube.

 Note: the customer must fit appropriate strain relief to the cable.

 Connect the plug by twisting it whilst pushing it gently into the socket on the

WindSonic.

 When it locates, twist the outer sleeve clockwise to connect and lock the plug.

 Fix the WindSonic to the tube using the 3 stainless steel screws provided (Maxi-

mum mounting screw torque 4 Nm.)

 It is the responsibility of the customer to ensure that the WindSonic is mounted in

a position clear of any structure, which may obstruct the airflow or induce turbu-

lence.

 Do NOT mount the WindSonic in close proximity of high-powered radar or radio

transmitters. A site survey may be required if there is any doubt about the strength

of external electrical noise.


376

APPENDIX C

ACCELEROMETER MOUNTING INSTRUCTIONS


377

Stud Mounting

 The accelerometer is fixed to the test surface by means of a threaded metal screw

 In preparing the stud mounting, the test surface must be drilled and tapped

 Distortion of the accelerometer as mounted may produce strains that affect the ac-

celerometer’s response. Therefore, it is important:

 To ensure that the test surface is very flat (which can be done by grinding or lap-

ping)

 To prevent the mounting stud from bottoming in the transducer case—this can

lead to strain

 To screw the accelerometer onto the test surface using the torque recommended

by the manufacturer

 For location B, C (attachment to the truss web members), and D (attachment to

post support), where the structural surface is rounded, a solid mounting block can

be fabricated which is rounded to this same contour on one side and flat on the

other side for mounting the accelerometer

 The size of the mounting plate (made of steel preferably) should be as small as

possible, with enough space to accurately and securely mount to the web member,

but not to create significant additional wind drag


378

Accelerometer Mounting Location and Direction

Location A

Location B

Location C

Location D
379

Location A

Line y

Line x
1 2

Accelerometer 1 1

 Sensitive end of accelerometer should be oriented parallel to Line x (perpendicu-

lar to Line y) along the length of the truss

 Accelerometer should be placed as close as possible to the center of the truss

mounting plate (next to accelerometer 2)


380

Accelerometer 2 2

 Sensitive end of accelerometer should be oriented parallel to Line y (perpendicu-

lar to Line x) transverse to the length of the truss

 Accelerometer placed as close as possible to the center of the truss mounting plate

(next to accelerometer 1)

Location B and C

Line z

Line y

5
Line x

 Accelerometers mounted to flat surface of the mounting plate (with one side fab-

ricated to fit the curvature of the web member)


381

 Attachment of mounting plate should be done securely (i.e., welding)

 Mounting plate width should be oriented parallel to the Line z

Accelerometer 3 3
 Accelerometer 3 placed at the mid-span of web member (in line with Accelerome-

ters 1 and 2)

 Sensitive end oriented parallel to line z

Accelerometer 4 4

 Accelerometer 4 placed at the mid-span of web member (in line with accelerome-

ters 1 and 2)

 Sensitive end oriented parallel to line y

Accelerometer 5 5

 Accelerometer 5 placed at the mid-span of the web member (in line with acceler-

ometers 1 and 2)

 Sensitive end oriented parallel to line y


382

Location D

Line z

2 ft

Accelerometer 6 6

 Accelerometer 6 mounted 2 ft from top of post along the centerline of the post

support (Line z) in line with accelerometers 1, 2, 3, 4, 5, and 6.

 Sensitive end oriented parallel to Line z

 Accelerometer mounted to flat surface of the mounting plate (with one side

fabricated to fit the curvature of the post support)

 Attachment of mounting plate should be done securely (i.e., welding)


383

APPENDIX D

RANKING SCHEDULE FOR NATURAL WIND TESTING


384

The ranking schedule used between UAB and ALDOT for determining appropri-

ate days for natural wind testing is provided below.

October 2008 Forecast (1999 -2007)


Average Maximum Peak Wind
Direction Ranking
Day Wind Speed Wind Speed Gusts
(Degrees) (1 – 10)
(mph) (mph) (mph)
1 5.72 14.57 19.57 107 4
2 5.67 14.57 19.00 174 6
3 5.13 11.29 14.00 176 5
4 5.53 13.43 17.29 160 5
5 6.58 17.14 22.00 151 6
6 7.62 16.71 21.86 54 6
7 6.77 14.86 19.29 71 4
8 6.63 15.29 18.71 147 6
9 6.42 15.14 18.14 106 4
10 6.95 15.00 18.43 141 6
11 8.70 16.00 19.86 103 5
12 5.13 12.86 18.71 284 3
13 6.03 13.00 16.43 150 5
14 6.70 17.57 22.14 250 5
15 6.18 14.57 18.50 205 6
16 8.48 17.71 23.43 247 6
17 6.75 18.00 22.29 166 7
18 6.70 15.00 18.43 136 6
19 5.85 15.29 19.14 239 4
20 4.20 10.86 14.29 240 3
21 5.55 14.86 18.86 233 4
22 7.05 14.57 18.29 179 7
23 7.80 19.86 24.86 224 7
24 6.63 17.14 22.86 171 7
25 6.57 14.71 17.57 197 6
26 6.65 16.00 19.29 251 4
27 6.35 15.57 20.43 224 6
28 5.78 14.29 18.43 117 4
29 4.28 13.00 16.14 161 4
30 6.02 13.57 16.14 150 5
31 5.18 15.00 18.40 168 5
385
386

November 2008 Forecast (1999 -2007)


Average Maximum Peak Wind
Direction Ranking
Day Wind Speed Wind Speed Gusts
(Degrees) (1 – 10)
(mph) (mph) (mph)
1 6.52 14.00 17.29 154 6
2 4.82 14.88 19.75 198 5
3 6.15 14.25 17.75 151 6
4 7.68 15.75 21.00 129 6
5 8.00 15.63 21.00 127 6
6 7.85 15.13 21.43 163 7
7 5.52 12.75 16.71 149 5
8 6.58 14.00 19.29 196 6
9 6.27 17.25 22.71 190 7
10 6.50 14.75 20.50 140 6
11 7.87 15.25 19.17 173 7
12 7.87 14.13 19.33 138 7
13 7.35 16.13 21.00 216 7
14 6.22 16.38 20.00 157 6
15 8.28 16.88 21.63 175 8
16 6.50 14.25 17.63 228 5
17 4.78 11.88 15.57 116 3
18 5.57 13.00 17.00 158 5
19 6.27 14.38 18.50 169 6
20 5.80 14.63 18.43 164 6
21 6.16 15.67 57.50 146 5
22 7.63 15.44 57.25 190 6
23 8.11 16.56 48.50 134 6
24 8.33 19.00 47.00 136 6
25 5.83 14.33 17.63 174 6
26 7.06 15.44 19.50 184 7
27 9.10 16.78 22.13 190 8
28 8.81 17.44 58.25 196 7
29 9.03 19.00 51.00 174 7
30 8.64 18.56 57.25 143 7
387
388

December 2008 Forecast (1999 -2007)


Average Maximum Peak Wind
Direction Ranking
Day Wind Speed Wind Speed Gusts
(Degrees) (1 – 10)
(mph) (mph) (mph)
1 6.94 17.89 22.88 185 7
2 6.07 14.78 19.00 188 6
3 8.56 17.78 21.67 211 8
4 7.37 18.63 23.43 174 8
5 8.23 17.56 21.44 212 7
6 5.69 13.33 16.00 183 6
7 6.29 13.67 17.50 215 6
8 6.60 15.00 18.50 93 3
9 7.29 16.56 20.56 132 6
10 7.06 16.89 21.67 172 7
11 6.63 15.89 20.88 249 5
12 8.27 17.22 20.75 149 7
13 8.21 22.11 27.11 180 8
14 8.56 18.56 23.00 164 8
15 5.74 16.78 20.33 176 6
16 8.09 17.67 24.63 244 6
17 7.10 16.44 20.88 244 5
18 7.54 17.22 22.50 223 7
19 9.16 19.44 24.89 226 8
20 7.47 17.33 22.33 209 7
21 7.79 16.67 21.25 198 7
22 10.01 19.67 25.50 223 8
23 8.41 18.56 23.63 230 7
24 6.15 16.88 20.50 144 6
25 7.77 15.89 20.63 195 7
26 5.76 16.00 19.44 200 6
27 4.86 12.11 15.75 213 4
28 8.37 18.67 22.67 213 8
29 7.10 15.44 19.75 256 5
30 7.16 15.11 19.75 168 7
31 6.81 14.89 19.38 155 6
389
390

January 2008 Forecast (2000 -2008)


Average Maximum Peak Wind
Direction Ranking
Day Wind Speed Wind Speed Gusts
(Degrees) (1 – 10)
(mph) (mph) (mph)
1 8.29 15.67 20.14 199 8
2 8.57 17.56 23.13 168 8
3 7.21 18.22 22.33 207 7
4 7.31 15.11 18.78 208 7
5 9.54 18.89 23.63 238 7
6 8.84 17.11 23.00 163 8
7 7.56 16.00 20.13 174 7
8 8.90 17.56 22.88 238 7
9 7.23 17.11 21.22 212 7
10 8.63 18.67 23.11 179 8
11 6.56 16.56 21.13 203 6
12 6.61 14.56 17.75 180 6
13 7.56 17.78 22.44 262 5
14 8.59 17.56 21.88 220 7
15 5.91 15.00 18.75 265 3
16 9.04 16.78 21.86 109 6
17 8.91 17.11 21.88 134 7
18 7.24 15.67 20.13 183 7
19 9.19 17.67 21.89 253 6
20 6.70 16.22 20.44 223 6
21 7.09 15.33 19.00 105 5
22 9.09 18.44 23.38 179 8
23 8.27 17.11 21.75 230 7
24 6.63 16.00 20.67 212 6
25 7.06 17.11 21.67 228 6
26 6.46 15.00 18.00 118 5
27 8.49 16.00 19.67 140 7
28 9.84 18.33 23.13 186 9
29 9.50 21.89 28.22 192 9
30 7.83 19.11 23.38 200 8
31 7.59 16.89 22.88 189 8
391
392

February 2008 Forecast (2000 -2008)


Average Maximum Peak Wind
Direction Ranking
Day Wind Speed Wind Speed Gusts
(Degrees) (1 – 10)
(mph) (mph) (mph)
1 8.24 16.89 20.63 179 8
2 9.09 18.22 24.50 179 8
3 8.00 18.33 22.78 220 7
4 9.73 20.22 26.11 276 6
5 8.43 17.11 21.63 124 7
6 8.79 17.56 21.75 169 8
7 7.86 14.67 18.56 236 6
8 5.04 12.67 16.38 196 5
9 7.06 17.33 22.38 208 7
10 7.73 18.25 22.25 181 8
11 6.36 15.22 18.75 154 6
12 8.07 17.22 21.13 211 7
13 8.26 16.00 22.56 223 7
14 7.93 15.56 18.88 221 7
15 9.06 18.78 23.50 185 8
16 8.57 19.22 23.75 165 8
17 10.60 20.11 24.63 260 6
18 7.40 16.22 20.78 186 7
19 8.16 18.22 21.89 193 8
20 8.99 18.67 23.38 143 8
21 7.54 17.22 21.88 200 7
22 8.23 17.78 22.13 205 8
23 7.57 16.56 20.00 209 7
24 8.96 19.67 24.11 151 8
25 8.94 20.89 26.67 119 7
26 9.17 18.89 25.38 129 7
27 7.13 16.89 22.63 215 7
28 6.18 15.00 18.13 208 6
393
394

March 2008 Forecast (2000 -2008)


Average Maximum Peak Wind
Direction Ranking
Day Wind Speed Wind Speed Gusts
(Degrees) (1 – 10)
(mph) (mph) (mph)
1 9.30 18.44 23.44 227 7
2 7.80 17.22 21.33 234 6
3 9.69 19.56 25.13 206 8
4 10.60 21.44 26.63 245 7
5 8.15 17.22 22.00 271 5
6 7.50 16.43 20.00 240 6
7 7.26 18.25 24.57 186 7
8 8.84 19.00 24.38 224 7
9 10.43 21.75 27.00 186 9
10 6.67 15.38 18.14 137 6
11 8.30 18.75 23.88 216 7
12 8.33 17.75 21.50 211 7
13 8.86 17.13 22.00 156 8
14 9.41 17.75 20.25 168 8
15 7.86 16.38 20.88 169 8
16 10.49 20.50 25.38 239 7
17 8.81 18.25 22.75 133 7
18 6.94 18.00 21.75 103 5
19 9.06 20.13 23.88 123 7
20 9.89 20.50 26.57 197 9
21 10.60 19.50 25.43 271 6
22 8.29 18.25 22.63 85 5
23 5.83 15.13 18.57 251 4
24 7.13 16.13 19.38 249 5
25 7.83 18.50 23.00 219 7
26 7.83 16.38 21.71 189 8
27 8.13 17.38 22.75 203 8
28 9.19 19.63 25.75 225 8
29 7.16 16.63 20.29 113 5
30 8.44 18.63 21.86 209 8
31 7.76 19.38 24.29 254 6
395
396

APPENDIX E

TRUCK INDUCED GUST TEST PROCEDURE


397

Rented Truck Route

Loop Instructions

Starting from I-65 North Exit 334

1. Continue on I-65 North (over Tennessee River) and take Exit 340: 6 miles

2. Get in right lane and Travel Underneath Sign at Required Speed (see fig-

ure below)

3. After passing underneath sign, continue straight (do not take ramp) and

merge back onto I-65 North

4. Continue onto I-65 North and take Exit 351: 11 miles

5. Take left after exiting and travel over bridge crossing I-65

6. Take 1st left and merge onto I-65 South

7. Continue on I-65 South (past sign) over Tennessee River and take Exit

334: 17 miles

8. Take left and travel under bridge (I-65 overpass)

9. Take immediate left and merge onto I-65 North

10. Begin Step 1

Total Distance: 34 miles/loop


398

Nashville

Sign
Decatur

Huntsville

North

Birmingham Exit 340B


399

Truck Test Procedure

1. Meet Truck Driver Arlene at BP at I-65 Exit 322

2. Fill up truck tank with gas

3. Wait for ALDOT Daniel OK call

4. Data acquisition system turned on

5. Get in truck with driver and leave car at BP

6. Call ALDOT Daniel when arriving at Exit 334 (call once for OK, twice to answer

because of problem)

7. ALDOT Daniel readies traffic control

8. Call ALDOT Daniel when arriving at the beginning of Tennessee bridge (call

once for OK, twice to answer because of problem)

9. ALDOT Daniel OK’s traffic control to stop traffic

10. Run loop at required speed

11. Mark on record time of passing

12. Travel back to Exit 334

13. Call ALDOT Daniel when leaving Exit 334 (call once for OK, twice to answer

because of problem)

14. Call ALDOT Daniel when arriving at the beginning of Tennessee bridge (call

once for OK, twice to answer because of problem)

15. Run loop at required speed

16. Mark on record time of passing

17. Call ALDOT Daniel after three loops to discuss

18. Stop testing at 3:30 PM

You might also like