197
INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler Visscher A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Civil Engineering University of Toronto © Copyright by Brent Tyler Visscher (2008)

INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT

by

Brent Tyler Visscher

A thesis submitted in conformity with the requirementsfor the degree of Master of Applied ScienceGraduate Department of Civil Engineering

University of Toronto

© Copyright by Brent Tyler Visscher (2008)

Page 2: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

ii

Innovative Pre-cast Cantilever Constructed Bridge Concept

Brent Tyler VisscherMaster of Applied ScienceDepartment of Civil EngineeringUniversity of Toronto, Canada2008

ABSTRACT

Minimum impact construction for bridge building is a growing demand in modern urban

environments. Pre-cast segmental construction is one solution that offers low-impact, economical,

and aesthetically pleasing bridges. The standardization of pre-cast concrete sections and segments

has facilitated an improved level of economy in pre-cast construction. Through the development

of high performance materials such as high strength fibre-reinforced concrete (FRC), further

economy in pre-cast segmental construction may be realized. The design of pre-cast bridges using

high-strength FRC and external unbonded tendons for cantilever construction may provide an

economical, low-impact alternative to overpass bridge design.

This thesis investigates the feasibility and possible savings that can be realized for a single cell

box girder bridge with thin concrete sections post-tensioned exclusively with external unbonded

tendons in the longitudinal direction. A cantilever-constructed single cell box girder with a

curtailed arrangement of external unbonded tendons is examined.

Page 3: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

iii

ACKNOWLEGEMENTS

This work has been partially supported by the National Science and Engineering Research

Council of Canada.

I would like to thank my supervisor Dr. Paul Gauvreau for his guidance, support, and

fellowship throughout the course of this design project.

Many thanks are owed to my friends and colleagues, especially those within our bridge

engineering research group, for their helpful discussions and insightful contributions to the

development of this project: Billy Cheung, Jamie McIntyre, Kris Mermigas, Talayah Noshiravani,

Graham Potter, Carlene Ramsay, Jason Salonga, and Jimmy Susetyo. Additional thanks go out to

other graduate students who I have had the pleasure of spending time with throughout my graduate

study: Jeff Erochko, Hyungjoon Kim, Nabil Mansour, Michael Montgomery, and Lydell Wiebe.

Special thanks to Jimmy Susetyo for his generous contributions in concrete material design and for

his laboratory assistance in test specimen preparation. I would like to thank John MacDonald for

his contribution during cylinder testing.

Finally, I thank my family for their constant encouragement and support throughout this task.

Page 4: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

TABLE OF CONTENTS

ABSTRACT. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

ACKNOWLEGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

TABLE OF CONTENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

LIST OF SYMBOLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

1.0 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1

1.1 Statement of Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2 Scope and Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.3 Thesis Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.0 BACKGROUND . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7

2.1 Constitutive Laws for High Performance FRC. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72.1.1 Compressive Stress-Strain Behaviour. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72.1.2 Tensile Stress-Strain Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.2 Typical Modern Highway Overpass Bridge Design Currently in Ontario . . . . . . . . . . . . . . . . . . . . . . . . . . 132.2.1 Standardization of Precast Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.3 Description of the Proposed Segmental Box Girder Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.4 Visual Comparison of Proposed Box Girder with Typical Slab-on-Girder Overpass . . . . . . . . . . . . . . . . . 222.5 Benefits and Drawbacks of the Proposed Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.5.1 Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.5.2 Drawbacks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.0 LONGITUDINAL FLEXURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .28

3.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283.2 Moment-Curvature Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293.3 Behaviour of Unbonded Tendons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.3.1 Reference State of Strain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313.3.2 Long-Term Effective Prestress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.4 Tendon Stress Calculation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333.5 Preliminary Design Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373.6 Preliminary Design Methodology for Cantilever Tendons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.6.1 Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393.6.2 Ultimate Moment Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423.6.3 Sizing of Cantilever Tendons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.7 Change in Structural System. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473.8 Preliminary Design Methodology for Continuity Tendons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

iv

Page 5: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

3.8.1 Ultimate Moment Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503.8.2 Spreading of Forces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513.8.3 Sizing of Continuity Tendons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533.8.4 Secondary Prestress Moment due to Continuity Tendons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.9 Bottom Flange Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583.10 Prestress Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 603.11 Ultimate Limit State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633.12 Serviceability Limit States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3.12.1 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 653.12.2 Cracking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 653.12.3 Global Deflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

3.13 Refinement in Tendon Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673.13.1 Ductility of Fibre-Reinforced Concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673.13.2 Deformation Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 693.13.3 Example of Tendon Stress Increase for Negative Bending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703.13.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.14 Global System Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733.14.1 Span-to-Depth Ratio for Constant Depth Box Girder. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733.14.2 Range of Spans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 743.14.3 Alternative Girder Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.0 TRANSVERSE FLEXURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .78

4.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784.2 Typical Transverse Prestressing for Segmental Bridges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794.3 Prestressing Concept for Light Weight Slab Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.3.1 Post-Tensioning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 804.3.2 Pretensioning. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.4 Bond Strength of 15mm Pretensioning Strands in FRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 824.4.1 Mechanisms of Bond . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 824.4.2 Development of Stress in a 15mm Pretensioning Strand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 834.4.3 Improvement of Bond Strength due to High Strength FRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

4.5 Methods of Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 854.5.1 Grillage Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864.5.2 Frame Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

4.6 Comparative Study for Transverse Stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 884.6.1 Cross Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 894.6.2 Material Definitions used in Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 914.6.3 Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 924.6.4 Deflections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

4.7 Prestress Moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 944.8 Service Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 954.9 Cross Section Design and Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

4.9.1 Transverse Rib Proportions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

v

Page 6: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

4.9.2 Web Spacing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 984.10 Design for Barrier Impact Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

5.0 SHEAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .101

5.1 Introduction to Strut-and-Tie Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1015.2 Funicular Load Path for Girders Prestressed with Curtailed Tendons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1035.3 Parallel Chord Truss Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1075.4 Arching Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

5.4.1 Tied Arch Model (Noshiravani, 2007) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1115.5 Preliminary Design Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1165.6 Opening of Joints in Segmental Bridges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1185.7 Design of Web Reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

5.7.1 Comparison to CAN/CSA-S6-06 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

6.0 ANCHORAGE ZONE. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .125

6.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1256.2 Flow of Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1266.3 Strut-and-Tie Model for Local Spreading of Forces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.3.1 Anchorage of External Unbonded Tendons in Slab Blisters (Wollman, 1993) . . . . . . . . . . . . . . . . . . 1276.3.2 Design Model for External Tendon Blisters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.4 Detailing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1336.5 Reinforcement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

7.0 FIXED END ABUTMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .138

7.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1387.2 Bras de la Plaine Bridge, France (Tanis, 2003) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1387.3 Flexible Supports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1407.4 Design Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

7.4.1 Preliminary Design Alternatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1417.4.2 Recommended Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

7.5 Sizing of Flexible Piles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1437.6 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

8.0 MATERIAL UTILIZATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .145

8.1 Reference Bridges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1458.1.1 Windward Viaduct, Interstate Route H-3, Hawaii, USA (Hawaii DOT, 1991) . . . . . . . . . . . . . . . . . . 1458.1.2 Hwy 407 - Islington Avenue Underpass, Toronto, Canada (MTO, 1990) . . . . . . . . . . . . . . . . . . . . . . 147

8.2 Concrete Consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1488.3 Mild Reinforcing Steel Utilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1508.4 Longitudinal Post-Tensioning Utilization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1528.5 Transverse Post-Tensioning Utilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1538.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

9.0 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .155

vi

Page 7: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

9.1 Longitudinal Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1559.2 Transverse Design. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1569.3 Shear Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1569.4 Anchorage Zone Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1569.5 Economy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1579.6 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

10.0 REFERENCES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .159

APPENDIX A - DESIGN DRAWINGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .164

APPENDIX B - MATERIAL TESTING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .175

APPENDIX C - CAN/CSA-S6-06 SHEAR PROVISIONS . . . . . . . . . . . . . . . . . . . . . . . .178

CURRICULUM VITAE. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .180

vii

Page 8: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

LIST OF FIGURES2.0 BACKGROUND ..................................................................................................................7

Figure 2-1. Variation in modulus of elasticity with the ultimate compressive strength of concrete............... 8Figure 2-2. Compressive stress-strain relationship for nominal 80MPa FRC test specimens and simplified

model ...................................................................................................................................... 9Figure 2-3. Photo of intensively cracked compression cylinder after failure and significant post-peak

straining ................................................................................................................................ 10Figure 2-4. a) Dogbone-shaped specimen: dimensions and sensors, b) test set-up of the uniaxial tensile test,

showing the front side of the specimen (adapted from Habel et al., 2006) .......................... 11Figure 2-5. Tensile stress-strain relationship for nominal 80MPa FRC test specimens and simplified model:12Figure 2-6. Typical slab-on-girder highway underpass built in Markham, Ontario on Highway 407.

Photograph courtesy of Scott Steves, 2006........................................................................... 13Figure 2-7. Recommended girder spacing with respect to span length for standard CPCI precast I-girders

(adapted from CPCI, 1996)................................................................................................... 14Figure 2-8. Plan and elevation view of a possible cantilever-constructed overpass bridge.......................... 16Figure 2-9. Cross section and details of the constant depth box girder shown in Figure 2-8 ....................... 17Figure 2-10. Typical standard segment components of proposed box girder concept .................................. 18Figure 2-11. Top flange supported by steel struts, a) example shown of Shibakawa Viaduct, Japan (photo

courtesy of Takashi Kosaka, 2006), b) rendering of proposed 90m overpass structure ....... 19Figure 2-12. Second Severn Bridge, a) total external unbonded prestressing tendons, b) cantilever

construction showing bulkhead face of segment with no continuous bonded steel across the joint and no internal bonded tendons (adapted from Mizon, 1997)...................................... 20

Figure 2-13. Possible tendon arrangement shown for half the span (symmetrical about span centreline) ... 21Figure 2-14. Numbering scheme for segment labels..................................................................................... 22Figure 2-15. Rendered perspective views on typical highway overpass construction using precast CPCI

girders (left) and proposed cantilever-constructed single cell box girder (right) as seen on an overcast day........................................................................................................................... 23

Figure 2-16. Deck cross sections for a) two-span precast CPCI girder overpass and b) proposed cantilever-constructed single cell box girder. ........................................................................................ 23

Figure 2-17. Visual slenderness attained through shadow casting for a superstructure with a a) small deck overhang and b) large deck overhang. .................................................................................. 25

Figure 2-18. Collapse slab-on-girder bridge after being struck by tractor-trailer (adapted from El-Tawil, 2005) ..................................................................................................................................... 26

Figure 2-19. Elevation (1:2500 scale) showing the result of elevating the overpass road grade.................. 27

3.0 LONGITUDINAL FLEXURE..........................................................................................28Figure 3-1. Idealization of box cross section for analysis of longitudinal flexure........................................ 29Figure 3-2. Internal lever arm from tendon elevation to the compression flange, a) for a continuously guided

tendon, b) for a tendon attached only at discrete locations................................................... 30Figure 3-3. Strain Distribution at ultimate limit state (adapted from Naaman, 2006) .................................. 31Figure 3-4. Material strains due to equilibrium in the reference state and in the ultimate state ................... 31Figure 3-5. Family of moment-curvature responses for a specific cross section and varying prestress force

assuming a concentric prestress ............................................................................................ 35Figure 3-6. Iterative method behaviour for tendon stress calculation for a) girder impending plastic hinging,

b) girder experiencing plastic hinging .................................................................................. 36Figure 3-7. Iterative method solution for obtaining tendon stress by incrementing applied load................. 37

viii

Page 9: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 3-8. Typical segment cross section .................................................................................................... 40Figure 3-9. Comparison of live load models defined in CAN/CSA-S6-06 .................................................. 41Figure 3-10. ULS bending moments on continuous girder........................................................................... 42Figure 3-11. Cantilever prestress concepts for haunched and constant depth girders .................................. 44Figure 3-12. Preliminary ULS demand and capacity for cantilever PT........................................................ 45Figure 3-13. Layout of cantilever tendons .................................................................................................... 46Figure 3-14. Redistribution of forces through concrete creep due to change in structural system (adapted from

Podolny et al., 1982) ............................................................................................................. 48Figure 3-15. Time-dependent creep coefficient assuming t0 = 28 days, a) for 75 year period, b) for 5 year

period .................................................................................................................................... 49Figure 3-16. Sectional forces due to prestressing with top and bottom tendons........................................... 51Figure 3-17. Moment development length, la, due to anchorage of continuity tendon (adapted from Menn,

1990) ..................................................................................................................................... 52Figure 3-18. Required moment development length for a girder post-tensioned with unbonded tendons ... 53Figure 3-19. Layout of continuity tendons.................................................................................................... 54Figure 3-20. Preliminary ULS demand and capacity for continuity tendons................................................ 55Figure 3-21. Secondary prestress moment due to prestressed continuity tendon ......................................... 57Figure 3-22. Ultimate state of stress for section with undersized bottom flange.......................................... 59Figure 3-23. Appropriately sized bottom flange thickness for negative bending at ULS............................. 59Figure 3-24. Primary prestress force distribution.......................................................................................... 61Figure 3-25. Girder response for varying prestress force.............................................................................. 63Figure 3-26. Girder response at ULS (half span shown)............................................................................... 64Figure 3-27. Girder response at SLS (half span shown) ............................................................................... 66Figure 3-28. Stress-strain response of FRC and typical normal-weight concrete......................................... 68Figure 3-29. FRC comparison for total moment-curvature response in negative bending ........................... 69Figure 3-30. Plastic hinging in negative moment regions............................................................................. 70Figure 3-31. 90m span with midspan hinge when load is close to failure .................................................... 71Figure 3-32. Stresses in unbonded tendons for curtailed prestressing. a) change in tendon stresses during

construction due to the addition of segments and post-tensioning, b) change in tendon stresses due to the application of increasing uniform load until failure............................................. 72

Figure 3-33. Ultimate behaviour for a continuous girder using unbonded draped tendons (adapted from Muller et al., 1990) ............................................................................................................... 73

Figure 3-34. Material consumption for varying span/depth ratios................................................................ 74Figure 3-35. Concept comparison of haunched girder and constant depth girder elevation......................... 76

4.0 TRANSVERSE FLEXURE ..............................................................................................78Figure 4-1. Possible pretensioning arrangement ........................................................................................... 82Figure 4-2. Bond mechanisms between FRC and 15mm pretensioning strand (adapted from Chao, 2006) 83Figure 4-3. Idealized strand stress profile in a pretensioned strand under applied load (adapted from Kahn,

2002) ..................................................................................................................................... 83Figure 4-4. Generalized frame element properties for grillage analysis (adapted from Menn, 1990).......... 87Figure 4-5. 3-dimensional grillage model ..................................................................................................... 88Figure 4-6. Comparison of top flange thicknesses of AASHTO-PCI-ASBI standard box girder examples

(adapted from Prestress/Precast Concrete Institute, 1997) ................................................... 90Figure 4-7. Cross section dimensions for typical segment............................................................................ 91

ix

Page 10: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-8. Assumed stress-strain material properties for investigation of transverse stiffness................... 91Figure 4-9. Live load models for transverse load analysis............................................................................ 92Figure 4-10. Vertical slab deflection envelope due to live load, AASHTO-PCI-ASBI box girder .............. 93Figure 4-11. Edge beam stiffening ................................................................................................................ 93Figure 4-12. Vertical slab deflection envelope due to live load of proposed box girder: a) thin slab stiffened

with transverse ribs only, b) thin slab stiffened with transverse ribs and longitudinal edge beam...................................................................................................................................... 94

Figure 4-13. Moment distribution due to non-concentric prestress applied at the transfer length ............... 95Figure 4-14. Total transverse bending moments for transverse flexure ........................................................ 96Figure 4-15. SLS top slab stress ranges at the extreme fibre due to service loading.................................... 97Figure 4-16. Longitudinal section for a typical segment .............................................................................. 98Figure 4-17. Web spacing requirements........................................................................................................ 99Figure 4-18. Truss model for basis of barrier reinforcement ........................................................................ 99Figure 4-19. PL-2 barrier detail for impact loading .................................................................................... 100

5.0 SHEAR..............................................................................................................................101Figure 5-1. 45 degree truss model (adapted from Ritter, 1899) .................................................................. 102Figure 5-2. Possible truss models for girders with unbonded and bonded prestressing steel (adapted from

Gauvreau, 1993).................................................................................................................. 102Figure 5-3. Alternative girder designs to resist applied load Q .................................................................. 104Figure 5-4. Effect of prestress arrangement on funicular compression chord ............................................ 105Figure 5-5. Funicular shape of compression spine for continuous cantilever girder with curtailed tendons106Figure 5-6. Strut-and-tie model for a fully prestressed flanged section. a) simplified model; b) through d)

detailed model of web, top flange and bottom flange, respectively (adapted from Schlaich et al., 1989) ............................................................................................................................. 108

Figure 5-7. Variable angle truss model (adapted from Collins et al., 1997) ............................................... 109Figure 5-8. Arch spreading of forces for a rectangular section (adapted from Schlaich et al., 1989) ........ 110Figure 5-9. Comparison of tied arch model for a rectangular section and a flanged section (adapted from

Noshiravani, 2006).............................................................................................................. 112Figure 5-10. Girder dimensions (adapted from Noshiravani, 2007) ........................................................... 113Figure 5-11. Girder cross section and web reinforcement (adapted from Noshiravani, 2007) ................... 113Figure 5-12. Tied arch model for load stage corresponding to flexural cracking (adapted from Noshiravani,

2007) ................................................................................................................................... 113Figure 5-13. Tied arch model for load stage corresponding to ULS (adapted from Noshiravani, 2007) ... 114Figure 5-14. Crack pattern and alternative parallel chord model for load stage corresponding to flexural

cracking............................................................................................................................... 115Figure 5-15. Crack pattern and alternative parallel chord model for load stage corresponding to ULS .... 116Figure 5-16. Truss model for shear design. a) fully prestressed state for continuous beam; b) modified model

to account for flange decompression (formation of plastic hinge) ..................................... 117Figure 5-17. Prestress forces applied to truss model in Figure 5-16 for each applied anchorage force ..... 117Figure 5-18. Incorrect hanger reinforcement model for shear stresses crossing an open joint (adapted from

Virlogeux, 1993) ................................................................................................................. 118Figure 5-19. ULS Combination 1 maximum shear demands...................................................................... 119Figure 5-20. ULS transverse steel demand and chosen capacity ................................................................ 120Figure 5-21. Web reinforcement ................................................................................................................. 120

x

Page 11: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-22. Shear panel tests for normal reinforced concrete and FRC concrete (adapted from Susetyo, 2007) ................................................................................................................................... 122

Figure 5-23. Shear parameter predictions according to CAN/CSA-S6-06 ................................................. 123

6.0 ANCHORAGE ZONE.....................................................................................................125Figure 6-1. Top and bottom corner anchorage blisters................................................................................ 126Figure 6-2. Flow of forces at an intermediate anchorage............................................................................ 127Figure 6-3. Strut-and-tie model for corner blister (Adapted from Wollman, 1993) ................................... 129Figure 6-4. Strut-and-tie model for corner blister satisfying static equilibrium; a) 3d isometric view, b) front

view, c) side elevation......................................................................................................... 130Figure 6-5. Jack clearances necessary for installation (adapted from VSL, 2007)..................................... 132Figure 6-6. Minimum anchorage eccentricities to box girder corner.......................................................... 132Figure 6-7. Anchorage blister lengths ......................................................................................................... 133Figure 6-8. Strut-and-tie anchorage model for a bottom tendon................................................................. 133Figure 6-9. Tension ties linking deviation force to box reinforcement (adapted from Beaupre et al., 1990)134Figure 6-10. Flow of forces for different reinforcement details ................................................................. 135Figure 6-11. Anchorage zone reinforcement for critical 19 strand bottom tendon..................................... 136

7.0 FIXED END ABUTMENT .............................................................................................138Figure 7-1. Bras de la Plaine Bridge, Reunion Island, France (adapted from Tanis, 2003) ....................... 139Figure 7-2. Ballasted abutment of Bras de la Plaine Bridge (adapted from Tanis, 2003)........................... 139Figure 7-3. Counterweight abutment concept (superseded)........................................................................ 141Figure 7-4. Tie-down abutment concept (superseded) ................................................................................ 142Figure 7-5. Proposed conceptual abutment design...................................................................................... 143

8.0 MATERIAL UTILIZATION .........................................................................................145Figure 8-1. Partial elevation view of Windward Viaduct (adapted from Hawaii DOT, 1991) ................... 146Figure 8-2. Typical cross section of Windward Viaduct (adapted from Hawaii DOT, 1991)..................... 146Figure 8-3. Post-tensioning arrangement for typical span of Windward Viaduct (adapted from Hawaii DOT,

1991) ................................................................................................................................... 147Figure 8-4. General arrangement of Islington Avenue Underpass (adapted from MTO, 1990) ................. 148Figure 8-5. Typical cross section of Islington Avenue Underpass (adapted from MTO, 1990) ................. 148

xi

Page 12: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

xii

LIST OF TABLES2.0 BACKGROUND ..................................................................................................................7

Table 2-1. Reference concrete mix quantities (Susetyo, 2007)....................................................................... 9Table 2-2. Simplified constitutive law for the nominal 80Mpa stress-strain behaviour used in analysis ..... 12

3.0 LONGITUDINAL FLEXURE..........................................................................................28Table 3-1. Procedure for calculating the stress in an unbonded tendon due to girder deflections ................ 34Table 3-2. Procedure for calculating uniform applied load corresponding to a desired tendon stress.......... 37Table 3-3. Estimation of dead load for typical segment................................................................................ 40Table 3-4. Estimation of superimposed dead load for typical segment ........................................................ 40Table 3-5. Preliminary prestressing scheme for cantilever tendons based on ultimate capacity .................. 46Table 3-6. Calculation of long-term bending moment redistribution............................................................ 50Table 3-7. Preliminary prestressing scheme for continuity tendons based on ultimate capacity.................. 54Table 3-8. Secondary prestress moment calculation ..................................................................................... 56Table 3-9. Tendon unit sizes (15mm strands) used for cantilever post-tensioning ....................................... 70

6.0 ANCHORAGE ZONE.....................................................................................................125Table 6-1. Area of steel required for local spreading reinforcement for top tendons (in mm2).................. 137Table 6-2. Area of steel required for local spreading reinforcement for bottom tendons (in mm2) ........... 137

8.0 MATERIAL UTILIZATION .........................................................................................145Table 8-1. Concrete consumption for one span of the Windward Viaduct (Hawaii DOT, 1991) ............... 149Table 8-2. Concrete consumption for Islington Avenue Underpass (MTO, 1990) ..................................... 149Table 8-3. Concrete consumption for proposed cantilever box girder ........................................................ 149Table 8-4. Mild reinforcing steel utilization for one span of the Windward Viaduct (Hawaii DOT, 1991)150Table 8-5. Mild reinforcing steel utilization for Islington Ave Underpass (MTO, 1990) ........................... 151Table 8-6. Mild reinforcing steel utilization for proposed box girder......................................................... 151Table 8-7. Post-tensioning utilization for one span of the Windward Viaduct (Hawaii DOT, 1991) ......... 152Table 8-8. Post-tensioning utilization for Islington Ave. Underpass (MTO, 1990) .................................... 152Table 8-9. Post-tensioning utilization for proposed cantilever box girder .................................................. 152Table 8-10. Transverse prestressing steel utilization for one span of the Windward Viaduct (Hawaii DOT,

1991) ................................................................................................................................... 153Table 8-11. Transverse prestressing steel utilization for Islington Avenue Underpass (MTO, 1990) ........ 153Table 8-13. Relative material consumption of proposed box girder ........................................................... 154Table 8-12. Transverse prestressing steel utilization for proposed box girder............................................ 154

Page 13: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

LIST OF SYMBOLS

Ap area of unbonded prestressing steel

Aps area of prestressing strand

As area of bonded steel

b width of compression flange

bbf width of bottom flange

bc concrete width function for idealized cross section

btf width of top flange

c distance from neutral axis to extreme compression fibre

d depth of cross section; centre-to-centre distance between tension and compression piles

db diameter of bonded steel

dp distance from centroid of prestressing steel to extreme compression fibre

dv effective shear depth taken as the greater of 0.9d or 0.72h (CSA 23.3-04)

e eccentricity of prestressing steel

eb continuity tendon eccentricity from girder axis

Ec compressive modulus of elasticity of concrete

Ep elastic modulus of prestressing steel

Et tensile modulus of elasticity of concrete

et cantilever tendon eccentricity from girder axis

fc stress in concrete

Fc force in concrete

f’c cylinder compressive strength of concrete

fci initial concrete strength at transfer

long-term effective prestress

fps stress in bonded prestressing strand

fpu ultimate stress of prestressing steel

fpy yield stress of prestressing steel

fse effective stress in bonded pretensioning steel after losses

fsi stress in bonded prestressing strand just prior to transfer

fsy yield stress of bonded reinforcing steel

f’t dogbone tensile strength of concrete

fp∞

xiii

Page 14: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

H horizontal force in pile caused by deflection

Ic uncracked sectional moment of inertia

K torsional constant

kb distance from girder axis to centroid of compression block in bottom flange

kt distance from girder axis to centroid of compression block in top flange

L span length

la moment development length

Lb length between anchorage blister ties for spreading of forces

Lc length of cantilever

lf flexural bond length for 14mm pretensioning strand

lt transfer length for 15mm pretensioning strand

Lten length of continuity tendon

M bending moment

fully redistributed bending moment

Mclo dead load moment applied to statically determinate cantilever at time of closure

Mf factored total moment at a section

Mfal bending moment obtained assuming entire structure was cast simultaneously on falsework

live load moment applied to statically indeterminate continuous girder

MP primary prestress moment,

Mpb prestress moment due to bottom continuity tendons

Mq moment due to external load

Mr factored moment resistance

superimposed dead load applied to statically indeterminate continuous girder after closure

Mtot total moment due to dead load plus external loads plus prestress loads

N axial force in concrete

P force in prestressing steel

P0 initial force in prestressing steel after jacking

long-term effective prestress force after all losses

Pb bottom prestress force of continuity tendons

pps perimeter of pretensioning strand

Pt top prestress force of cantilever tendons

M∞

MLLindet

P e×

MSDLindet

P∞

xiv

Page 15: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Ptot total prestress force of unbonded tendons for top and bottom prestressing steel

q uniform load

Q applied load

Qu applied load corresponding to the ultimate state

tbf thickness of bottom flange

ttf thickness of top flange

uf average flexural bond stress of pretensioning strand

V shear

Vc shear resistance attributed to concrete

Vf factored shear demand

Vs shear resistance attributed to transverse reinforcement

w width of web; uniform dead load

x coordinate along longitudinal axis of girder

y elevation in cross section (where bottom fibre elevation is y = 0)

yclo cantilever tip deflection at closure

midspan deflection of continuous system due to long-term creep

ratio of the average stress in compression block to the specified concrete strength

live load factor

prestress load factor

ratio of the depth of compression block to the neutral axis depth

girder deflection

centre-to-centre spacing of transverse ribs

change in beam length due to prestress, dead load, and applied load

elongation of unbonded prestressing steel relative to initial prestress

initial change in beam length due to prestress and dead load

increase in tendon force due to girder deformations

loss in prestress due to superimposed dead load

increase in prestress due to superimposed dead load

change in strain in prestressing steel relative to strain due to initial prestress

strain

bottom fibre strain

ultimate strain of concrete

initial longitudinal strain in concrete at level of prestressing steel

α1

αL

αP

β1

∆b

∆lcp

∆lcp ∆lp0–

∆lp0

∆P

∆P∞

∆Pdl

∆εp

ε

εb

ε'cεc0

xv

Page 16: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

average initial strain in concrete due to prestress and dead load

strain caused by concrete stress fc

strain in concrete at level of prestressing steel

ultimate compressive strain

strain in prestressing steel

initial strain in prestressing steel

top fibre strain

cracking strain of fibre-reinforced concrete

strain of concrete at mid depth of member

girder slope; angle of principle compressive stress in the web; angle of force resultant trajectory for arching shear

angle of compression force in slabs due to anchorage bursting forces

angle of compression force in anchorage blister projected onto flange

angle of compression force in anchorage blister projected onto web

coefficient of friction

fully redistributed concrete stress

bottom fibre stress

concrete actual stress at time of closure

concrete stress obtained assuming entire structure is built simultaneously on falsework

stress in prestressing steel

long-term effective prestress

top fibre stress

curvature

resistance factor for concrete

resistance factor for prestressing steel

resistance factor for steel

cantilever tip rotation at closure

εc0 avg,

εcf

εcp

εcu

εP

εP0

εt

ε'tεx

θ

θb

θf

θw

µ

σ∞

σb

σclo

σfal

σp

σp∞

σt

φ

φc

φp

φs

ωclo

xvi

Page 17: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

1

1.0 INTRODUCTION

In modern urban environments, providing increased capacity on existing roadways and bridges

is a significant challenge. Engineers are faced with several demanding factors including tight

construction scheduling, the maintenance of traffic in highly travelled corridors, and the creation

of structures that enhance the urban landscape. Precast segmental construction is one solution that

can provide aesthetically pleasing structures that can be erected on difficult construction sites. By

the method of match-cast precasting, segments can be fabricated and prepared off-site and erected

rapidly with minimal effect on traffic.

Traditional methods of bridge construction over high-volume highways generally require lane

closures, traffic delays, and in some cases total road closures and traffic detours. Faced with a

growing economy and increased need for traffic flow reliability, minimum impact construction and

rapid erection for bridge building is becoming more crucial especially in the dense urban

environment. This necessity is realized not only for commuter traffic but also industries that rely

on just-in-time delivery. In Europe, new technologies have been implemented exclusively for the

purpose of fast erection in order to minimize impact of construction. Recent implementations of

fast construction in Ontario has been the use of full-width precast panels for overpass bridge

construction over Highway 401 in the Municipality of Chatham-Kent (Rapoport, 2006), and the

recent use of a self-propelled modular transporter for replacement of an overpass bridge on

Highway 417 in Ottawa (Tinkess, 2007). The implementation of new rapid bridge building

technology in Ontario is a testament to the increasing need of new construction methods for fast,

low-impact construction.

New opportunities for efficient and aesthetically pleasing designs exist through the

development of high performance materials such as high strength concrete. Due to the low water-

to-cement ratio made possible by the inclusion of water reducing admixtures, high strength

concrete exhibits favourable design properties such as higher ultimate stress and stiffer modulus of

elasticity. The impact on design due to these improved material properties is the ability to create

more elegant structural designs that make better use materials through the use of thinner sections

that result in light weight solutions. The increased stiffness of new materials allows engineers to

design more slender shapes that add to the aesthetic transparency of the structure.

Page 18: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Light weight girders with thin sections for prestressed concrete structures can be made possible

through the use of external unbonded tendons. Flange and web sections with large amounts of

internal bonded steel generally require a significant amount of concrete clear cover solely for the

purpose of durability and corrosion protection. This sacrificial layer of concrete is necessary to

protect the bonded tendons from environmental exposure, but it is not designed as an efficient use

of materials from a strength perspective. By removing the tendons from the concrete section, the

concrete thickness may be reduced to make more efficient use of concrete material, while reducing

the overall weight of the superstructure.

1.1 Statement of Problem

A new minimum-impact solution for rapid highway overpass construction is desired which

makes efficient use of high-performance concrete and external unbonded tendons. The use of

external unbonded tendons in prestressed concrete structures allows for the minimization of

concrete consumption, reduction in dead load due to thinner cross sections, and simplified

construction.

The cantilever method is one type of segmental construction that is self-supporting as it allows

erection of the bridge from the fixed end with no interference with the ground below and no

requirement for falsework. This method is attractive for minimum impact highway overpass

construction since it allows the bridge to be built overhead of the traffic with no requirement for

equipment and machinery below the bridge. For prestressed concrete cantilever-constructed

bridges, cantilever tendons are traditionally designed as internally bonded within the top flange.

Flange thickness is governed by the space required for tendons inside the concrete section, and the

minimum reinforcing steel required for adequate crack control and durability.

Through the development of high-performance materials such as high strength fibre-reinforced

concrete (FRC), improved mechanical properties have been observed to allow engineers greater

possibilities for the design of prestressed concrete structures. Of interest are the benefits to be

gained by the adaptation of high strength FRC and unbonded tendons for use with precast

segmental structures. Quality control of concrete is maximized for precasting in comparison to

cast-in-place concrete, as it is fabricated, cast and cured in a controlled environment. The use of

high strength FRC in combination with external unbonded tendons in concrete structures appears

to be a complimentary pairing of materials as external tendons allow for the minimization of

2

Page 19: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

concrete consumption and high strength concrete allows for large precompression stresses from

external prestressing.

Precast segmental cantilevered-construction using exclusively external unbonded tendons with

no continuous reinforcing steel is traditionally not done in standard practice. One rare example of

the use of total external unbonded tendons for precast segmental cantilever construction is the

Second Severn Bridge across the River Severn between England and Wales. Although the use of

total external unbonded prestressing for this bridge was primarily due to a Government durability

requirement, this design also simplified deck unit production and enabled thinner webs and flanges

(Mizon, 1997). Extending the use of external unbonded tendons to cantilever construction

potentially provides an opportunity for designers to erect medium span overpass bridges faster, off

the right-of-way, with little or no interference with traffic demands.

1.2 Scope and Objective

The purpose of this study is to develop and validate a design concept using a new application

of external unbonded tendons intended for rapid overpass bridge construction requiring a low-

impact construction technique.

The objectives of this study are as follows:

• develop valid design assumptions for consideration of the proposed bridge type

• determine the global member response of the superstructure under serviceability and ultimate limit states

• incorporate the use of high performance (80MPa characteristic compressive strength) fibre-reinforced concrete (FRC) for the design of thin sections to reduce dead load and minimize reinforcing steel

• validate the use of external unbonded tendons as a new application in segmental cantilevered construction

The member response investigated for this study examines explicitly the effects of longitudinal

flexure, longitudinal shear and transverse flexure due to the force effects of dead load, live load,

and prestress.

Several actions have not been considered explicitly within the scope of this thesis. These

effects include: a) thermal action, b) torsional loads, and c) combined effects of moment, shear, and

torsion. These effects are important to the final design and for the full understanding of structural

behaviour, and could possibly have an impact on the proposed design concept. Future work is

3

Page 20: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

necessary for the complete validation of the superstructure concept. In addition, the primary focus

of this thesis has been on the design and development of the superstructure system for a light

weight box girder design using high strength FRC. The design and development of the substructure

system is briefly discussed but a detailed design of the substructure is beyond the scope of this

work.

1.3 Thesis Structure

This document consists of 9 chapters. Chapter 1 introduces the topic and establishes the design

motivation for using high performance concrete and external unbonded tendons for the application

of rapid highway bridge construction. Chapter 2 presents background information on material

properties and the proposed structural system for discussion. Chapter 3 describes the behaviour of

unbonded tendons in a system with curtailed prestressing and the design for longitudinal flexure of

the proposed bridge. Chapter 4 discusses the motivation for minimizing the thickness of the top

flange and provides a feasible solution for transverse bending using a thin top flange and transverse

pretensioning. Chapter 5 presents a variation of design models for the analysis of shear behaviour

in concrete structures prestressed with unbonded tendons, and provides the design for shear for the

proposed bridge. Chapter 6 describes the importance of anchorages for girders prestressed with

external tendons, and provides the design developed for anchorage reinforcement. Chapter 7

briefly discusses a design concept for a substructure design. Chapter 8 reviews the material

consumption for the proposed bridge and compares the material consumption of other

conventionally constructed bridge designs. Conclusions for this work are summarized in Chapter

9.

Chapter 2 begins by describing the benefits of FRC and the improved constitutive laws that are

distinctive of the material. The results of a mini experimental study are presented and used for the

basis of a simplified design model for the behaviour of high strength FRC in compression. Typical

highway overpass bridges are discussed and the proposed alternative solution is described and

compared.

Chapter 3 first describes the flexural response of concrete girders prestressed with unbonded

tendons and convenient analysis techniques for the preliminary design of girders. For concrete

girders that are prestressed under the action of self-weight, a reference state of strain is described

that corresponds to the deformations of the girder after jacking. To begin the design process,

conservative preliminary design assumptions are described for the state of stress of the unbonded

4

Page 21: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

tendons. Following this, the design methodology for cantilever tendons is discussed for a

preliminary estimate of tendon arrangement. The change in structural system due to closure of the

span is discussed and an estimation for long-term redistribution of forces is presented. The design

methodology for continuity tendons is then presented and the effect of the tendon layout on

secondary prestress moments is described. Based on the designed prestress loads, the bottom

flange of the girder is designed for negative bending at the support. The girder behaviour is then

analyzed under factored loads at the ultimate limit state, and unfactored loads for seviceability

performance. Following this, a refinement in tendon stress at the ultimate limit state is investigated

to improve the economy of prestress strand consumption. Finally, the effect of span-to-depth ratio

of the proposed box girder is presented, and a practical range of spans for the proposed box girder

is identified.

Chapter 4 describes the design challenges of satisfying prestress requirements for the thin top

slab in transverse bending due applied vertical loads and prestress forces. The design assumption

of small concrete covers is addressed. The chosen pre-tensioning prestress system is justified based

on the available studies that demonstrate the improved bond strength characteristics between high

strength FRC and 15mm pretensioning strands. The method of analysis to determine the live load

distribution of forces is described. A comparative study between the proposed box girder and an

AASHTO-PCI-ASBI standard box girder (Precast/Prestressed Concrete Institute, 1997) is then

carried out for the validation of the thin top flange design in transverse bending on the basis of

maximum allowable deflections. The deflections and computed stress ranges of the top flange at

service loads is shown for the final design. Finally, a design detail for a proposed PL-2 barrier

defined in the CAN/CSA S6-06 (Canadian Standards Association, 2006) is presented for the

resistance of impact loads.

Chapter 5 begins with a discussion on strut-and-tie models and the description of the funicular

load path in a prestressed girder with curtailed unbonded prestressing tendons. Various approaches

to shear design are discussed based on alternative decompositions of the funicular load path in

concrete structures. A parallel chord model is discussed and the variable angle parallel chord model

used in the CAN/CSA-S6-06 shear provisions is described. Based on the results of a previous

study, an arching model by Noshiravani (2007) is described and compared to the parallel chord

model. For the proposed box girder, a parallel chord model is decided to be most appropriate to

describe the flow of forces in the box girder to transfer vertical loads. The proposed girder is then

5

Page 22: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

analyzed for the shear demands based on CAN/CSA-S6-06 load configurations. Finally, the design

for web reinforcement is presented based on the parallel chord model.

Chapter 6 begins with a discussion of anchorages for prestressed concrete girders with external

unbonded tendons. The fundamentally different flow of forces within the anchorage zone for an

unbonded tendon is then described in comparison to the anchorage zone of an internal tendon. A

three-dimensional strut-and-tie model is developed for the design of reinforcement required to

anchor an external unbonded tendon at a corner blister. Possible reinforcement detailing based on

the developed truss model is proposed to enable the required flow of forces.

Chapter 7 briefly describes a conceptual abutment design for the proposed bridge. The

abutment must: a) adequately provide anchorage of all cantilever prestressing, b) resist large fixed

end moments due to negative flexure of the superstructure, and c) provide longitudinal flexibility

to allow for thermal movements since the proposed bridge is designed to be monolithic at the

midspan with no expansion joints. A detailed analysis of substructure behaviour is beyond the

scope of this work, however.

Chapter 8 first describes two reference bridges which are used to determine the material

utilization of conventional construction in comparison to the proposed box girder bridge. One

reference bridge is representative of traditional cantilever construction, and the other reference is

representative of a typical highway overpass structure in Ontario. Concrete usage, mild steel

reinforcement utilization, longitudinal prestressing and transverse prestressing steel consumption

quantities are compared.

Chapter 9 concludes with a summary of findings for the design components investigated within

this study. Recommendations for the design of cantilever-constructed bridges prestressed with

unbonded tendons are provided. Advantages of using thin concrete sections with unbonded

tendons are described, and the disadvantages are identified. Necessary future work is discussed for

the full validation of the proposed design concept.

6

Page 23: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

7

2.0 BACKGROUND

This chapter begins by discussing the behaviour of fibre-reinforced concrete (FRC) under axial

compression and axial tension. The results of a small experimental program are presented and

briefly discussed for test specimens fabricated with nominal 80MPa strength concrete, similar to

the concrete intended for design. A simple stress-strain model is developed for the FRC specimens

tested which will be used for the behaviour of concrete throughout the work of this study.

Following this, conventional highway overpass bridge construction using precast components is

described and a description of the proposed bridge concept using precast segments is presented. A

visual comparison of the proposed bridge to a conventional overpass is made to evaluate the

aesthetic qualities of the proposed bridge. Finally, benefits and drawbacks of the proposed bridge

concept are identified.

2.1 Constitutive Laws for High Performance FRC

2.1.1 Compressive Stress-Strain Behaviour

The innovation of material technology has introduced a widespread variety of high

performance concrete materials that have improved mechanical properties over traditional 35MPa

concrete mix designs. High performance concretes (HPC) range to approximately 100MPa and

ultra-high-performance fibre-reinforced concrete (UHPFRC) can be as high as 200MPa. Research

programs have been carried out in many countries to study the behaviour high performance

concrete, leading to significant changes to design codes in several countries. (Paultre et al., 2003).

Design codes are continually evolving documents using a consistent philosophy and the latest

research results, often reflecting the prior state of the art and tradition of the country of origin

(Paultre et al., 2003). However, because concrete is a heterogeneous material without standardized

mixture designs, the compressive stress-strain responses of different concrete mixtures exhibit

significant scatter (Graybeal, 2007). As a result, the empirical definitions of compressive stiffness

of HPC is not entirely consistent among design codes. A comparison of elastic modulus code

definitions and recent results of some researchers are displayed in Figure 2-1 below.

Page 24: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 2-1. Variation in modulus of elasticity with the ultimate compressive strength of concrete

While these design values for elastic stiffness are useful for the calculation of structural

behaviour at the serviceability limit state, very little information is available in the literature that

discusses the post-peak descending branch of FRC. The post peak behaviour is important since it

has a considerable effect on the flexural ductility of a member at the ultimate limit state. Due to the

large amount of scatter in the elastic stiffness constitutive laws, and the lack of information

published on the compressive post-peak response of FRC, compression cylinder tests were

performed to develop a full range stress-strain material model for use in the analysis of the

proposed design.

A concrete mix design was adapted from Susetyo (2007) for a nominal 80MPa FRC containing

1.5% hooked steel fibre volume ratio. The reference concrete mix design is outlined in Table 2-1.

Two compression cylinder specimens were tested for the basis of the simplified stress-strain model

shown in Figure 2-2. The average peak stress was observed to be 74.5MPa, and the average

uncracked elastic modulus was observed to be 43680MPa. A more thorough description of the

experimental program is given in Appendix B. The compression cylinder test results were

compared with a previous study by Chao et al. (2006), who performed cylinder tests for a 75MPa

FRC mortar matrix containing no coarse aggregate and 1% steel fibres, and found there was good

agreement in the results. The curves are comprised of a stiff initial ascending branch, followed by

a ductile post-peak descending branch.

0

00001

00002

00003

00004

00005

00006

00007

00008

00009

00108060402

80Mpa concrete strength intended for design

Ec [MPa]

f‘c [MPa]

Graybeal 2007 (f’c ≤ 200MPa)CHBDC 2006 (f’c ≤ 85MPa)ACI 1992 (f’c ≤ 83MPa)ACI 2005 (f’c ≤ 40MPa)

Ma et al. 2004 (f’c ≤ 200MPa)CEB-FIP 1990 (f’c ≤ 88MPa)CSA 2004 (f’c ≤ 80MPa)CSA 2004 (f’c ≤ 40MPa)

8

Page 25: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 2-2. Compressive stress-strain relationship for nominal 80MPa FRC test specimens and simplified model

The presence of fibres in the concrete matrix provides excellent confinement which facilitates

the sustained compressive stresses after concrete crushing and allows for significant ductility in the

post-peak response. Figure 2-3 displays a failed FRC compression cylinder after the test. Even

after significant post-peak straining, the concrete remains intact and very little spalling from the

surface is observed. The resistance to spalling not only contributes to the strength of FRC at the

ultimate limit state, but it also provides an important benefit for durability of structures as it

improves the performance for frost resistance during freeze-thaw cycles (Xu et al., 1998).

Table 2-1. Reference concrete mix quantities (Susetyo, 2007)

Material Measured Unit Quantity Equivalent Volume

HSF Cement kg 600 0.191

Sand (SSD) kg 1133 0.419

10mm limestones (SSD) kg 802 0.292

Water L 162 0.162

Water Reducer mL 4200 -

Superplasticizer mL 9600 -

Entrapped Air 0.02

Total Volume m3 1.083

0

01

02

03

04

05

06

07

08

09

001

800.0700.0600.0500.0400.0300.0200.0100.00

Chao (2006)

UofT testscompression model developed for design

# specimens tested = 2Eavg = 43680 MPaε’c avg = 0.00209 mm/mmf’c avg = 74.6 MPa

fc [MPa]

εc [mm/mm]

9

Page 26: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 2-3. Photo of intensively cracked compression cylinder after failure and significant post-peak straining

2.1.2 Tensile Stress-Strain Behaviour

Since fibre-reinforced concrete exhibits a more ductile tension stress-strain response in

comparison to the brittle nature of normal concrete, the standard ASTM C496 cylinder splitting

test is insufficient to measure the tensile response of FRC. In lieu of the ASTM C496 test,

deformation-controlled dogbone-shaped test specimens loaded in uniaxial tension have been

performed by previous researchers to provide a tensile stress-strain relationship of FRC. Susetyo

(2007) performed a dogbone tension test on the nominal 80MPa FRC similar to the test set-up by

Habel et al. (2006), shown in Figure 2-4 below.

10

Page 27: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 2-4. a) Dogbone-shaped specimen: dimensions and sensors, b) test set-up of the uniaxial tensile test, showing the front side of the specimen (adapted from Habel et al., 2006)

The result of the dogbone tension test performed by Susetyo (2007) is presented in Figure 2-5.

The FRC mix design predominantly exhibits tension softening following the occurrence of the first

crack. In general, the formation and opening of one primary crack dominates the post-cracking

response due to the relatively low fibre ratio of 1.5%. Although tension stiffening behaviour is

possible for some FRC mix designs (which generally requires a larger amount of fibres), an

increased consumption of steel fibres to achieve this behaviour is not necessary or economical for

the application of FRC in precast segmental construction. For either dry joints or epoxy joints

between precast units, there is no embedment length of fibres across the joint; therefore, the

smeared tension behaviour of FRC can not be relied upon at the joint interface undergoing flexural

tension forces. For epoxied joints, a brittle tensile behaviour would be expected that is related to

the cracking strength of the concrete, with no contribution of the fibres since internal equilibrium

of tension stresses in the fibres can not be established at the joint. Thus, the inclusion of steel fibres

in the concrete mix is included predominantly for the benefit of controlling plastic shrinkage

cracking which replaces typical temperature and shrinkage reinforcement.

Sivakumar (2007) has provided a study which indicates that for a high strength 60MPa silica

fume cement, the inclusion of only 0.5% steel fibres reduces the plastic shrinkage cracking by 49%

compared to the same mix with no fibres. Some hybrid mix designs including steel fibres and

polyester fibres provided a better reduction in shrinkage cracking by 95% while maintaining

a) b)

1005050

100

5050

100

100200

LVDT

U4

Front Backspecimendepth = 50mm

U4 - deformation transducersLVDT - linear variable differential transformer

11

Page 28: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

reasonable workability. For the material considered in this report, the fibre ratio of 1.5% steel fibres

should provide reasonable control over plastic shrinkage cracking; however, no tests have been

performed to measure the cracking behaviour of this mix design.

A simplified model for the tension behaviour has been estimated using a tri-linear response

(Figure 2-5). Based on the dogbone test described above, a simplified constitutive material model

has been created for design of the proposed bridge concept. A summary of the response for both

tension and compression is given in Table 2-2 below, where tension stress is positive and

compression stress is negative

Figure 2-5. Tensile stress-strain relationship for nominal 80MPa FRC test specimens and simplified model:

Table 2-2. Simplified constitutive law for the nominal 80Mpa stress-strain behaviour used in analysis

Strain[mm/mm]

Stress [MPa]

-0.016 0

-0.00209 -74.6

-0.0008539 -37.3

0 0

0.00012 5.2

0.0025 2.2

0.01 0

0

1

2

3

4

5

6

10.0800.0600.0400.0200.00

Susetyo (2007)

# specimens tested = 1Et = 63600 MPaε’t = 0.00012 mm/mmf’t = 5.15 MPa

tension model developedfor design

ft [MPa]

εc [mm/mm]

12

Page 29: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

2.2 Typical Modern Highway Overpass Bridge Design Currently in Ontario

Since the 1950’s, prestressed concrete bridges have become increasingly popular, and

comprise about two-thirds of all bridges with spans between 18 and 36m (Lounis et al., 1997). The

precast concrete I-girder system represents about 50% of all prestressed concrete bridges built up

until the early 1990’s (Dunker et al., 1992). Precast I-girder systems have expanded their use even

further in the last thirty years through the concept of spliced girders to extend the span capability

(Rabbat et al., 1999; CPCI, on-line). Although transportation equipment and available cranes

limited the length of precast pretensioned girders to around 34m in the 1960’s, precast I-girders

now can be fabricated and transported in lengths of 40m to 50m and weights up to 75 to 90 tonnes

(CPCI, on-line). The standardization of girder sections has led to simplified designs, speed of

construction, and resulted in economy (Rabbat et al., 1999).

Since slab-on-girder bridges represents such a large proportion of bridges of short to medium

spans, precast slab-on-girder bridges will be regarded as a typical highway overpass design for

comparison to the proposed precast segmental box girder design. Both of these systems are precast

and are comparable in terms of low impact of construction solutions. A visual example of a typical

modern slab-on-girder structure is shown in Figure 2-6. This bridge depicts a typical continuous

two-span overpass structure supported by a centre pier in the median of the highway. The

abutments have parallel wingwalls oriented parallel to the axis of the bridge. A concrete parapet

lines the edges of the deck slab along the length of the bridge and continues to the furthest extents

of the wingwalls.

Figure 2-6. Typical slab-on-girder highway underpass built in Markham, Ontario on Highway 407. Photograph courtesy of Scott Steves, 2006.

13

Page 30: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

To maximize economy, bridge girder sections have been standardized as it provides a basis for

consistency and allows for the multiple reuse of standard forms for several bridge projects (Figg,

1997). The Canadian Precast/Prestressed Concrete Institute (CPCI) provides a range of standard

sections that are available in Canada. A suggested range of spans is provided for the use of each

standard section and a recommended transverse girder spacing is provided in relation to the girder

span length (shown in Figure 2-7). Although this recommendation is given only as a guideline, it

provides engineers with a launching point in which to form a slab-on-girder bridge concept.

Figure 2-7. Recommended girder spacing with respect to span length for standard CPCI precast I-girders (adapted from CPCI, 1996)

In addition to precast concrete girder sections being standardized, the design of slab-on-girder

bridges are also somewhat standardized as they must conform to several geometric constraints for

evaluation using the Simplified Method of Analysis detailed in the CAN/CSA-S6-06 (cl. 5.6.1, cl.

5.7.1). As the Simplified Method outlines an empirical approach to provide safe designs for a wide

range of span lengths and bridge widths, the design of bridges using these methods in general

produces conservative designs.

2.2.1 Standardization of Precast Components

Following the standardization of precast I-girder sections, an evolution for prestressed precast

concrete applications emerged in 1997 for standardized segmental bridge construction (Rabbat et

al., 1999). The AASHTO-PCI-ASBI Segmental Box Girder Standards provided a new product for

grade separations and interchange bridges for span lengths up to 61m (Freyermuth, 1997). Two

families of standard box girder sections exist: one intended for span-by-span construction, and

another intended for cantilever construction. The standard box girder for span-by-span

construction, which has a maximum intended span of 45.7m, exhibits thinner flanges since all post-

14

Page 31: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

tensioning steel is comprised of external unbonded tendons. The standard box girder for cantilever

construction has thicker top and bottom flange components to incorporate the internal bonded

cantilever and continuity post-tensioning tendons and a thicker web since cantilevering can achieve

longer spans.

Segmental spans can be erected very quickly, typically in a few days, minimizing disruption of

traffic (Figg, 1997). For bridge projects intended for the standard precast box girders, it is generally

anticipated that segments are erected by crane (Freyermuth, 1997). The large amount of work that

can be done off site under a controlled environment is favourable as it reduces the amount of on-

site construction time and permits the best curing conditions for the concrete to achieve a higher

standard of durability. Quality control can be maintained as segments fabricated with high-quality

precast concrete can be erected during non-peak traffic hours when disruption of traffic is kept to

a minimum.

The recommended minimum concrete strength for the standard segmental sections is 34MPa.

In some cases, concrete with a greater compressive strength can be used; however, no changes to

the standard AASHTO-PCI-ASBI cross sectional dimensions and cross sectional thicknesses can

be made to exploit the full potential and economy of high-strength concrete mix designs. These

standard sections have been developed for the application of traditionally reinforced normal

concrete mixes. For the application of high-performance FRC and total external unbonded post-

tensioning, a new structural section, respecting the inherent advantages of FRC, needs to be

developed to achieve economical designs and efficient use of the material.

2.3 Description of the Proposed Segmental Box Girder Concept

The proposed bridge system for precast segmental cantilever construction is a constant depth

single cell box girder with a span-to-depth ratio of 25:1. The box girder cross section is designed

to have thin flanges and thin webs to make efficient use of high-strength concrete and to reduce

dead load of the superstructure. The maximum bridge span considered for this study is 90m for

practical and visual considerations. As a practical consideration, the maximum depth of the girder

segments should be limited to accommodate ease of transportation which is constrained by

standard vertical clearances for existing overpass structures. As a visual consideration, the depth

of the girder should be limited to maintain a reasonable level of slenderness for the design. At a

25:1 span-to-depth ratio, a 90m span length corresponds to a girder depth of 3.6m.

15

Page 32: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

For the transportation of tall segments, a “low-boy” trailer is necessary which has a lowered

section of the trailer to reduce the top of the bed to within 610mm of the roadway surface (Precast/

Prestressed Concrete Institute, 1997). The minimum vertical clearance provided under highway

bridges in Ontario is 4.7m for existing structures and 4.8m for new structures (Ontario Ministry of

Transportation, 2002). The barrier walls of the precast segments are intended to be cast-in-place

after the cantilever structure has been closed; therefore, the 3.6m deep girder may be transported

along any major highway in Ontario with a minimum vertical clearance of 490mm for protruding

reinforcing steel for the cast-in-place barrier wall.

A conceptual plan and elevation of the proposed cantilever girder concept for the longest span

considered of 90m is shown in Figure 2-8. The girder is erected segmentally from the abutments,

and each cantilevered segment during construction is post-tensioned to the structure with

permanent external unbonded post-tensioning tendons. The typical precast segment length is 3m

to accommodate standard traffic lane widths for delivery of segments. The minimum vertical

clearance provided over traffic lanes is 5000mm, which is consistent with standard policy for new

structures over roadways for this bridge type (Ontario Ministry of Transportation, 2002) and a

700mm construction tolerance has been provided for safety during construction. The 90m

mainspan is continuous across the abutment face and the short 15m sidespan is ballasted by backfill

material to maintain rotational restraint at the abutment face and to prevent uplift of the backspan.

Figure 2-8. Plan and elevation view of a possible cantilever-constructed overpass bridge

16

Page 33: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Assuming the overpass bridge structure services an arterial undivided urban roadway, the

minimum side clearance from the edge of the travelled way to the face of the barrier is 2.0m

(Ontario Ministry of Transportation, 2002). The arrangement of traffic lanes intended for the box

girder bridge consists of two centrally located lanes with a width of 3.6m each, two exterior

shoulders with a width of 2.0m each, and an allowance for traffic barriers with a width of 400mm

each. Therefore, the total cross sectional width of the top flange considered for this design is 12.0m

(shown in Figure 2-9). As there is no thickness requirement to place internally bonded cantilever

post-tensioning tendons, the top flange has been designed to minimize the usage of high-

performance FRC. The continuous top flange is 85mm thick, which is stiffened with transverse ribs

equally spaced at 1500mm centre-to-centre. The transverse ribs are cast monolithically with the top

slab and have a total depth of 200mm from the top of the flange surface. For each typical precast

segment, there is a 500mm wide transverse rib located in the middle of the segment, and 250mm

wide transverse ribs located at the both edges of the segment. The 250mm wide transverse ribs

provide a larger face for bulkhead detailing of match-cast joints, and it provides a seamless

transition of joined segments since the combination of two adjacent edge ribs are of the same size

as the centre rib.

Figure 2-9. Cross section and details of the constant depth box girder shown in Figure 2-8

The webs of the box girder are 120mm in thickness and are tapered inward towards the bottom

flange. For a 3600mm deep girder, the height-to-thickness slenderness ratio is 30:1. Dilger et al.

(2003) have shown that slender webs for box girder bridges are stable in the uncracked state. Box

girder bridges with webs of comparative slenderness are generally cantilever-constructed variable

depth girders where the web height is greatest in the high shear region near the support. As an

example, the Confederation Bridge in Eastern Canada has a girder depth of 14m at the support and

17

Page 34: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

a web thickness of 400mm, resulting in a slenderness ratio of 35:1 (Dilger et al., 2003). The webs

of the proposed girder are connected to the bottom flange which is 4000mm wide. The bottom

flange has a uniform thickness of 150mm for typical segments, and increases in thickness to

305mm at the support.

Figure 2-10. Typical standard segment components of proposed box girder concept

A three-dimensional representation of a typical segment is shown in Figure 2-10. The main

components of the box girder segment are identified. On the exterior of the box, steel struts are

installed to provide stiffness to the thin top flange. Each segment has a pair of steel struts, one on

each side of the box girder located in the mid-plane of the segment and coincident in plan with the

middle transverse rib of the top flange. When adjacent segments are assembled, the steel struts are

equally spaced at 3000mm centre-to-centre. Steel struts have proved to be an economical and

aesthetically pleasing solution for box girder bridges requiring a light weight design. The

Shibakawa Viaduct, which is part of the Second Tomei Expressway in Japan, incorporated the use

of inclined steel struts to support the cantilever slab. Through the use of these inclined struts, the

dead load of the superstructure was reduced to approximately 80% of the original box girder design

concept that did not use inclined struts (Mutsuyoshi, 2004). In addition, further savings were

realized in the substructure due to the innovation of this strutted box girder section. Figure 2-11

displays the successful design of steel struts used on the cantilever-constructed Shibakawa

steel strut

top corner blistersfor cantilever tendonanchorage

bottom corner blisterfor continuity tendonanchorage

continuous transverse rib through box(500mm width)

continuous transverse ribs through box(250mm width)

18

Page 35: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Viaduct, and the adaptation of steel struts to the proposed box girder design for overpass

construction.

Figure 2-11. Top flange supported by steel struts, a) example shown of Shibakawa Viaduct, Japan (photo courtesy of Takashi Kosaka, 2006), b) rendering of proposed 90m overpass structure

On the interior of the box girder, the external unbonded post-tensioning is anchored at formed

anchorage blisters at the inside corners. During construction, each segment is attached to the

cantilever girder by stressing a pair of external unbonded post-tensioning tendons at the top corners

of the box girder. In a similar fashion, continuity of the box girder is achieved by stressing a pair

of external unbonded post-tensioning tendons at the bottom corners of the box girder. External

unbonded post-tensioning for precast segmental cantilever construction has been successfully

designed for the Second Severn Bridge across the River Severn between England and Wales

(Mizon, 1997). Figure 2-12 displays the adaptation of total external prestressing inside the viaduct

box girder of the Second Severn Bridge, and the bulkhead detailing showing no continuous bonded

reinforcement at the joint. The top cantilever tendons are placed at a minimal distance below the

underside of the top flange in order to maximize the flexural lever arm for negative bending.

Likewise, the bottom continuity tendons are placed at a minimal distance above the bottom flange

to maximize the flexural lever arm for positive bending.

a) Shibakawa Viaduct, Japan b) Proposed overpass structure

steel strut steel

strut

19

Page 36: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 2-12. Second Severn Bridge, a) total external unbonded prestressing tendons, b) cantilever construction showing bulkhead face of segment with no continuous bonded steel across the joint and no internal bonded tendons (adapted from Mizon, 1997)

The tendon arrangement proposed for the constant depth overpass structure involves the use of

curtailed post-tensioning tendons. The curtailment of post-tensioning tendons refers to the

termination of prestressing steel at discrete locations for every precast segment. Thus, unlike box

girders built by the span-by-span method, the area of prestressing steel is not constant at all cross

sections along the length of the girder. The result of this curtailment of tendons is an efficient

arrangement of prestressing steel since the flexural capacity of the box girder closely matches the

flexural demand. This point will be illustrated further in Chapter 3. One possible arrangement of

prestressing tendons is shown in Figure 2-13. The cantilever post-tensioning is anchored at the

inside corner blisters of the box and deviated towards the middle of the box girder at sections closer

to the abutment. At the abutment wall, the cantilever tendons are deviated in the vertical plane and

splay into three distinct trajectories. At an internal deviator wall 5m from the abutment wall, the

tendons are deviated and splayed horizontally to provide adequate space for all tendon anchorages

at the back face of the abutment. The continuity tendons follow a similar arrangement as the

cantilever tendons. Each pair of continuity tendons are anchored at a bottom anchorage corner

blister inside the box girder and deviated towards the middle of the box girder for sections closer

to the midspan. A deviator beam is located near the midspan to provide deviation of tendon units

a) b)

20

Page 37: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

in the horizontal plane for the continuity tendons. Similarly, a deviator beam is also located near

the abutment to provide deviation of tendon units in the horizontal plane for the cantilever tendons.

Figure 2-13. Possible tendon arrangement shown for half the span (symmetrical about span centreline)

As a way to provide reference to specific segments in the structure, the segments have been

numbered in the order in which the segments are erected. Thus, Segment 1 is the segment that is

21

Page 38: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

closest to the abutment wall; the segment which is match-cast against Segment 1 is numbered

Segment 2, and so forth. This numbering scheme is represented in Figure 2-14.

Figure 2-14. Numbering scheme for segment labels

2.4 Visual Comparison of Proposed Box Girder with Typical Slab-on-Girder Overpass

As a basis for visual comparison, we shall contrast the unique aesthetic characteristics between

the proposed box girder design and more traditional slab-on-girder highway overpass bridge using

CPCI precast I-girders. For the purpose of this section, only a visual comparison of aesthetic

qualities will be made; implications on economy will be addressed in Chapter 9 of this thesis. As

shown in Figure 2-7, the recommended transverse girder spacing for a span length of 45m using

standard 2300mm deep CPCI I-girders ranges from 1.61m to 1.85m (CPCI, 1996). Assuming that

the reference bridge has been designed using the simplified method of analysis defined by the

CAN/CSA-S6-06, the deck overhang must not exceed 1.8m or 60% of the typical girder spacing

(cl. 5.6.1, cl. 5.7.1). Therefore 7 girders are required at a typical spacing of 1.67m. For a deck slab

thickness of 225mm and a slab haunch of 75mm, the total structural depth of the slab on girder

superstructure is 2.6m. This corresponds to a span-to-depth ratio of 17.3:1 for two equal 45m spans

comprising the 90m total bridge length.

The single span cantilever-constructed segmental box girder is designed at a more slender

span-to-depth ratio of 25:1. For the 90m span, the structural depth of the box girder is 3.6m. The

box girder maintains a constant span-to-depth ratio for the entire length of the bridge. A minimum

vertical clearance of 5m is provided for each of the overpass bridge concepts.

A 3-dimensional solid model for each conceptual design under comparison has been developed

to illustrate the visual experience of the observer for selected vantage points. These vantage points

portray realistic perspective views of each structure to provide a credible basis for visual

comparison of each type of highway overpass. The views shown in Figure 2-15 are indicative of

the sights experienced on an overcast day where direct sun and shadow casting is not present.

Figure 2-16 provides the cross sections for the superstructures under comparison.

22

Page 39: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 2-15. Rendered perspective views on typical highway overpass construction using precast CPCI girders (left) and proposed cantilever-constructed single cell box girder (right) as seen on an overcast day.

Figure 2-16. Deck cross sections for a) two-span precast CPCI girder overpass and b) proposed cantilever-constructed single cell box girder.

a) elevation view

b) side profile

c) view from abutment

a) Two-span slab on CPCI girders b) Single span segmental box girder

23

Page 40: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The first comparison to be made is shown in Figure 2-15a), where the entire bridge is viewed

from abutment to abutment at an angle which is orthogonal to the axis of the bridge direction. As

the eye of the observer tends to trace the entire continuous length of the structure, the most

dominant shape that is observed is the clearance window below the superstructure from abutment

to abutment. This clearance window is defined by the 90m horizontal opening and the 5m vertical

clearance. For the CPCI girder design, the clearance window is broken visually due to the presence

of the centre pier; however, notwithstanding this discontinuity the overall appearance of the

superstructure remains relatively slender as the presence of the pier appears to have a minor impact

on the overall transparency of the bridge at this vantage point. If the observer pays minimal

attention to the presence of the pier, the perceived span-to-depth ratio of the superstructure

effectively has the appearance of being thinner than the actual slenderness ratio of 17.3:1. The box

girder design provides a clearance window that is completely unbroken from the outermost

supports with no intermediate supports between abutments. From this vantage point, the structure

provides an open unobstructed view throughout the full length of the bridge; however, the cross

section in contrast to traditional designs is noticeably deeper due to the low proximity to the

ground. This absolute depth is most pronounced on an overcast day under ambient lighting;

however, on a bright day when directional light creates shadows on the structure, the perceived

depth is more slender.

Visual slenderness can be achieved by taking advantage of shadow casting that results from

overhead sunlight. The shadow created on the webs of I-girders and box girders reduces the visual

prominence of the structural depth, and therefore portrays a lighter appearance to the viewer.

Larger deck overhangs produces a deeper shadow, thus a larger benefit due to shadow casting is

gained through the incorporation of a large deck overhang. As the overhang is restricted to 60% of

the typical girder spacing for slab-on-girder bridges designed using the simplified method, the

maximum overhang allowed following this restriction is 1000mm from the centre of the web for

the reference bridge. The deck overhang to girder depth ratio for this structure is 0.44, providing

only a small benefit in shadow casting to enhance the visual slenderness of the structure. For the

box girder shown in this example, the deck overhang is approximately 3500mm from the centre of

the web at the elevation of the top flange which provides a deck overhang to girder depth ratio of

0.97. Shadow casting is enhanced even more for a box girder with inclined webs in comparison to

a girder with vertical webs as it provides a deeper shadow projection from the top flange. Figure 2-

17 displays the reduction in the dominance of the girder that can be attained through the

24

Page 41: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

incorporation of a larger deck overhang inherent in box girder superstructures. This reduction in

dominance is clear when comparing to Figure 2-15b which shows the same view under ambient

lighting.

Figure 2-17. Visual slenderness attained through shadow casting for a superstructure with a a) small deck overhang and b) large deck overhang.

Lastly, Figure 2-15c provides a vantage point that is seen when the observer is up close to the

structure, viewing from a pedestrian level at the foot of one abutment and viewing along the axis

of the bridge. From this vantage point, the traditional 2-span slab-on-girder structure exhibits an

array of parallel girder lines along the axis of the structure. These girders are supported by the

centre pier which has an increased visual prominence when seen from a more acute angle. The

centre pier detracts from the overall transparency of the structure, and creates a visual barrier when

the observer is in the vicinity of the bridge. The CPCI girders are pin-supported at the abutments

where bearings and expansion joints are normally present; the abutment walls are prone to staining

and delamination spalling due to drainage runoff and corrosion from de-icing salts. In contrast,

from this vantage point the box girder design provides a unique visual experience as the steel struts

exhibit a cascading effect along the continuous flange lines. The depth of the box girder is not

prominent when viewing the bridge from this angle. The observer also has an unobstructed view

for the full width of the right-of-way with no intermediate visual barriers. The box girder is

designed to be integrally connected to the abutments; therefore, no expansion joints are present at

the abutments which provides protection to the abutment from staining and deterioration.

2.5 Benefits and Drawbacks of the Proposed Design

2.5.1 Benefits

1) The absence of a pier in the median of the mainline highway provides motorists with a

greater level of safety. Piers in the vicinity of the right-of-way present a risk for vehicular collision.

a) small deck overhang b) large deck overhang

25

Page 42: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Catastrophic vehicular impacts with bridge piers have occurred in the past, causing loss of the

structure and loss of life (El-Tawil et al., 2005). As an example, Figure 2-18 displays the

catastrophic failure that was caused to a slab-on-girder structure in Nebraska, USA due to the

impact of a heavy truck. Following the expensive traffic delays and demolition and cleanup of the

failed structure, a new structure was designed with fewer bridge piers to improve safety (ENR,

2003).

Figure 2-18. Collapse of slab-on-girder bridge after being struck by tractor-trailer (adapted from El-Tawil, 2005)

2) Box girder sections for overpass structures in comparison to slab-on-girder bridges also

provide enhanced durability due to a smaller surface of concrete exposed to salt spray. For the cross

sections that were considered in Section 2.4 and shown in Figure 2-16, the slab-on-girder bridge

has a surface area of 76.8m2 per meter of superstructure directly exposed to the atmosphere, while

the box girder bridge only has a surface area of 34.4m2 per meter of superstructure directly exposed

to the atmosphere.

3) As box girder sections are integral units comprising the deck surface itself, the entire

structural section is fabricated with high-quality precast concrete that is cast and cured in a

controlled environment. Typical slab on girder bridges are constructed using precast girders with

a cast-in-place concrete deck. The concrete deck, which is arguably the most damage prone

component of the bridge superstructure, provides enhanced durability due to the improved quality

of the exposed concrete. Improved quality translates to lower life-cycle costs and longer life (Ralls,

2003).

4) Not only does the absence of the centre pier provide improved safety for motorists, but it

also allows for the maximum flexibility in lane arrangement possibilities for the mainline highway.

This may be especially significant in dense urban areas where express and collector lanes are

required.

26

Page 43: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

5) Accelerated bridge construction inherently provides greater service to the public by reducing

the amount of time required to erect a structure. In addition, greater savings in labour can be

realized due to the accelerated scheduling. Ralls (2003) describes that traffic control can run

anywhere from 20 to 40 percent of construction costs and user delays are priced at thousands of

dollars per day in heavy traffic areas.

2.5.2 Drawbacks

1) In comparison to the two-span slab-on-girder alternative, the single-span box girder design

requires a deeper structural section. The larger bridge depth has implications on the approach

design as the road elevation is necessarily higher at the abutment to accommodate the depth of the

section. The top of roadway elevation of the approach must match the top of roadway of the

structure. Therefore, for an increase in structure height, two situations may exist for the approach

road: either the approach grade must be steeper for a fixed length of approach, or the approach must

be elongated to meet the desired elevation for a fixed approach grade (shown in Figure 2-19). In

both cases, an increase in sub-grade material is necessary for the approach road.

Figure 2-19. Elevation (1:2500 scale) showing the result of elevating the overpass road grade

2) The general understanding of the behaviour of external unbonded post-tensioning is not well

understood with respect to the application of cantilever construction with curtailed tendons. To the

authors knowledge, no published material exists that describes the behaviour of curtailed external

unbonded tendons both at serviceability limit states (SLS) and at ultimate limit states (ULS). The

increase in tendon stress at ULS through the application of FRC and the potential benefits of the

ductility of FRC is presently unknown. It is the intent of the following chapter to explore and

determine the potential benefits of using FRC in conjunction with external unbonded tendons for

cantilever girders and to determine the level of efficiency that can be obtained using the proposed

cantilever-constructed bridge as a vehicle to uncover these questions.

box girder elevation slab-on-girder elevation

27

Page 44: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

28

3.0 LONGITUDINAL FLEXURE

In this chapter, the design for longitudinal flexure of a cantilever-constructed girder prestressed

exclusively with external unbonded tendons will be discussed. Sections 3.2 and 3.3 discuss the

calculations involved for determining the response of unbonded tendons. Section 3.4 outlines two

procedures for the calculation of tendon stress due to global deformations of the box girder.

Following this, Section 3.5 describes simplifying design assumptions made to develop a

preliminary design concept for the box girder. Sections 3.6 to 3.8 discuss the strategies for

designing cantilever and continuity external post-tensioning tendons. Section 3.9 describes the

design method for sizing the bottom slab thickness in negative moment regions. Sections 3.10 to

3.12 describe the girder behaviour at SLS and ULS. Section 3.13 investigates the possible

refinement in tendon stress at ULS to achieve greater economy in prestressing steel. Finally,

Section 3.14 provides a discussion on the effect of span-to-depth ratio on material consumption,

and the practical range of spans of the proposed box girder.

3.1 Introduction

The design of tendons for a cantilevered bridge achieves significant economy in steel

consumption since the area of prestressing steel at any section generally corresponds to the flexural

demands of the bending moment diagram. Classical cantilever-constructed bridges are built out

symmetrically from pier tables using cast-in-place construction having segments ranging from 3 to

5 meters in length (Menn, 1990). These classical cantilever bridges are normally haunched, where

the depth of the cross section varies as the bottom flange follows a parabolic profile to increase the

internal lever arm at the support. The cantilever tendons are typically internal and bonded in the

top slab, and the continuity tendons are typically internal and bonded in the bottom slab. In general,

prestressed concrete cantilever type bridges carry a significant amount of dead load compared to

live load, typically in the order of 4:1 (Menn, 1990). This ratio is indicative of the total load

proportions on a box girder, although this ratio may decrease some for shorter spans since the unit

weight of a shallower box girder is decreased, all other things being equal. A common measure

taken in cantilever-constructed concrete bridges is the reduction of weight for segments placed

closer to the midspan in order to reduce the negative bending moment at the support. The result of

this design effort is a haunched girder profile, where the depth of the section is reduced at the

Page 45: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

midspan. The bottom slab of cantilever girders also varies in thickness in some cases, being thicker

at the support for strength requirements and thinner near the midspan to reduce dead load. In light

of this design strategy, the reduction of dead load for a cantilever-constructed bridge presents great

value when designing for the ultimate limit state.

3.2 Moment-Curvature Response

Using a sectional analysis, a typical concrete box girder can be idealized as a simple wide

flanged section when considering the global flexural response. The idealized cross sectional

geometry can be described as:

(3-1)

where bbf is the width of the bottom flange, w is the web thickness of one web, btf is the width

of the top flange, tbf is the thickness of the bottom flange, ttf is the thickness of the top flange, and

d is the depth of the section. This sectional idealization is illustrated in Figure 3-1 below.

Figure 3-1. Idealization of box cross section for analysis of longitudinal flexure

Assuming plane sections remain plane, the strain distribution over the depth of the section and

corresponding curvature are given as

(3-2)

(3-3)

where are the top and bottom fibre strains, respectively. Internal equilibrium of

forces for axial load and internal moment are given by

bc y( )

bbf

2wbtf⎩

⎪⎨⎪⎧

=

if 0 y tbf <≤

if tbf y d ttf–<≤

if d t– tf y d ≤ ≤

wtbf

ttf

btf

d

bbf

tbf

ttf

btf

d

bbf

2w

a) actual cross section b) idealized cross section

y

εcf y εb εt, ,( ) εbεt εb–

d---------------y+=

φ εb εt,( )εb εt–

d---------------=

εb and εt

29

Page 46: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

(3-4)

(3-5)

where P is the axial prestress load of the unbonded tendons at the section under consideration,

fc is the concrete stress, and are the concrete and prestress material resistance factors,

respectively.

For structures prestressed with external unbonded post-tensioning where tendons are attached

only at discrete locations (such as anchorages and deviators), a change in tendon eccentricity

occurs due to deformation of the girder. This concept is illustrated in Figure 3-2, where an example

of a continuously guided tendon and an example where the tendon is attached only at the cantilever

tip and at the support are shown. For the tendon attached only at discrete locations, a decrease in

internal lever arm from the elevation of the tendon to the compression flange results in a decrease

in ultimate resistance. In this thesis, the external unbonded tendons are assumed to behave as

internal tendons (that is, like the continuously guided tendon). This estimation is reasonable

provided that tendon guides are located at sufficient intervals along the length of the girder.

Treating the tendons as internal tendons, total moment can be calculated based only on applied

loads and prestress force when examining the response of a girder in flexure.

Figure 3-2. Internal lever arm from tendon elevation to the compression flange, a) for a continuously guided tendon, b) for a tendon attached only at discrete locations

3.3 Behaviour of Unbonded Tendons

In addition to satisfying equilibrium with external loads, internal forces must correspond to

deformations that satisfy conditions of geometrical compatibility appropriate to the materials and

structural components used. In concrete girders reinforced entirely with bonded steel, internal

forces at a given cross section can be computed using the compatibility conditions at that section

only. For beams of dimensions commonly used in bridge design practice, compatibility is defined

by the assumption that plane sections remain plane after deformation and are orthogonal to the

N εb εt P, ,( ) φcfc εcf y εb εt, ,( )( )bc y( ) yd0

d

∫ φpP+=

M εb εt P, ,( ) φcfc εcf y εb εt, ,( )( )bc y( )y yd0

d

∫ φpPyp+=

φc and φp

a) continuously guided tendon b) tendon attached only at discrete locations

yp(x) = const

yp(x=a)

yp(x=a) < yp(x=0) x x

30

Page 47: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

deformed member axis, and that perfect bond exists between the bonded steel and the concrete. In

girders prestressed with unbonded tendons, however, the plane sections compatibility is no longer

valid at a specific section. This is illustrated in Figure 3-3 where for the unbonded system is

shown to be less than for that of a system with bonded steel. The total strain in the unbonded tendon

is therefore the summation of the effective prestress strain, , and the overstress strain, , due

to deformation of the girder system. The effective prestress strain is shown as a broken line as it is

generally much larger than the initial concrete strain, . This is illustrated more clearly by

comparing material strains, shown in Figure 3-4.

Figure 3-3. Strain Distribution at ultimate limit state (adapted from Naaman, 2006)

Figure 3-4. Material strains due to equilibrium in the reference state and in the ultimate state

3.3.1 Reference State of Strain

When a beam with unbonded tendons is stressed, assuming deflections are free to occur during

the stressing operation (which is true for a cantilevered beam), there will be an initial elastic

deflection due to the axial prestress load, P0, the induced prestress moment, Mp, and the dead load

∆εp

εp0 ∆εp

εc0

εcu

Ultimate State(P0 + ΔP), Mu

P0, Md Reference State

εc0

εp0

CompressionTension

εcp

dp

(∆εp)bonded

(∆εp)unbonded

c

-80

-60

-40

-20

0-0.02-0.015-0.01-0.0050

0

400

800

1200

1600

0 0.005 0.01 0.015 0.02

fc [MPa]

εc [mm/mm]

fp [MPa]

εp [mm/mm]

εp0

εc0

εcu εp0+ Δεp

a) concrete strains

Δεp

b) prestressing steel strains

εc0 corresponds toinitial strain aftertransfer

εp0 corresponds toinitial effectiveprestress strain

31

Page 48: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

moment due to the girder self-weight, Md. The integration of elastic straining in the concrete at the

level of the prestressing tendon when the prestress force is anchored gives the initial change in

length of the beam at the elevation of the tendon, and is given by

(3-6)

The average straining in the concrete at the level of the prestressing tendon is given as

(3-7)

The reference state of deformations is chosen to be the initial length of the tendon after transfer,

where the initial concrete deformation due to the jacking effort under dead load is . This

choice of reference state takes into account structural behaviour, and particularly, the construction

sequence of sequential cantilevering.

For a cantilever-constructed bridge, each tendon set has its own unique reference state of elastic

strain in the cantilever beam that corresponds to the jacking stress after transfer (which is assumed

to be 0.7fpu). The state of elastic strain that defines the reference point for one set of cantilever

tendons depends on the dead load applied to the girder during the stressing operation, and the force

in tendons overlapping the tendon being stressed. The addition of dead load and the addition of

subsequent pairs of unbonded tendons due to the successive placement of segments will affect the

state of deformations of the previously stressed tendons. Since dead load is active during the

stressing operation of tendons, the deformations due to dead load are included in the calculation of

for the tendon pair being stressed and not in the calculation of , which corresponds to a

change in length after the stressing operation. The calculation for is the same as for as

described in Equation 3-6, except all loads applied to the girder are considered. Any changes in

tendon stress after the initial state is determined by the tendon deformations relative to the initial

deformation, .

3.3.2 Long-Term Effective Prestress

The effective initial prestressing force is defined as , which includes the force jacked into

the tendon and the effect of dead load during prestressing. The long-term effective prestress that

includes the added effect of permanent dead load added after prestressing, plus any long term

losses due to creep and shrinkage is given as:

∆lp0

∆lp0 εcp P Md L x,( ) x,,( ) xd0

L

∫=

εc0 avg,∆lp0

L----------=

∆lP0

∆lP0 ∆lcP

∆lcP ∆lP0

∆lcP ∆lP0–

P0

32

Page 49: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

(3-8)

where is the increase in tendon stress due to additional permanent dead load, and is

the change in prestress force due to long-term losses. is required in design for the calculation

of ultimate strength of girders with unbonded prestressing steel. For concrete girders of typical

proportions and reinforcing ratios prestressed with bonded steel, large increases in stress causing

yield of steel in regions of high moment can be attained at the ultimate limit state. However, for

girders prestressed with unbonded tendons, the increase in tendon stress is generally far lower since

the stress increase is averaged out over the tendon length between anchorages (Collins and

Mitchell, 1997). The ultimate force in unbonded tendons can be calculated by the summation of

the long-term effective prestress force plus additional tendon force gained due to girder

deformations from short-term applied loads, .

The ultimate strength of girders in general is proportional to two main parameters: the force in

tension steel, and the lever arm from tension steel to the compression flange. For a box girder with

a predetermined depth having tendons inside the cross section, the maximum flexural lever arm

effectively is also predetermined. Therefore, the ultimate strength of girders prestressed with

unbonded tendons is dependent on the maximum force that can be attained in the prestressing steel.

Since tends to be smaller for unbonded tendons, the effective prestressing force is important

for the ultimate strength of the girder. To achieve efficient and safe use of unbonded tendons in

design, a reasonable estimate of ultimate stress must be determined.

3.4 Tendon Stress Calculation Methods

To calculate the response of concrete girders prestressed with unbonded tendons, a general

procedure is given in Table 3-1 below. The state of stress in an unbonded tendon corresponding to

a known arrangement of applied loads can be determined by first assuming the effective prestress

in the tendon and iterating the tendon stress until an acceptable convergence is achieved. This

procedure returns an ‘exact’ tendon stress for the given arrangement of load and state of

deformations of the girder. When the girder is in a linear elastic state of stress, convergence is

generally achieved within 4 iterations.

P∞ P0 ∆Pdl ∆P∞+ +=

∆Pdl ∆P∞

P∞

P∞ ∆P+

∆P

33

Page 50: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

This method requires the iteration of tendon stress, which consequently requires the

recalculation of moment-curvature response corresponding to the state of stress in the tendon. If

we are interested to determine the associated response due to a known arrangement of applied loads

on a girder, an interpolation scheme between known moment-curvature responses (corresponding

to constant prestress forces) is a convenient method to save time and computational effort for the

iterative calculation of tendon stress. If the prestress force is assumed to be concentric to the girder

(where Mp = 0) then the linear elastic region of all curves will be coincident along

and the post cracking responses will deviate from the linear elastic state toward the ultimate

resistance and ultimate curvature (shown in Figure 3-5). Interpolating between these lines is a fast

calculation and it accelerates the speed for each iteration step since it eliminates the need to

reproduce moment curvature responses for every unique tendon stress. Several chosen responses

may be calculated and stored in matrix form to serve as a database for the automated interpolation

of responses, illustrated by the following matrices below:

Table 3-1. Procedure for calculating the stress in an unbonded tendon due to girder deflections

Step Calculation Comment

1 Guess tendon stress to be the effective prestress

2 Calculate total moment on girder including applied loads, plus primary and secondary prestress loads

3Calculate curvature and tendon level strain distribution over the length of the girder using member response corresponding to lines of equal force

4 Calculate absolute tendon elongation by integrating concrete strain due to girder deformations

5 Calculate difference in tendon elongation from the reference state

6 Calculate difference in tendon strain from the reference state

7 Calculate new tendon strain

8for linear elastic

Calculate new tendon stress with respect to known constitutive material laws

9 Check Check to see if the difference between the new calculated stress and the assumed stress is sufficiently small. If so, finish. If not, return to step 1 with

σp σp∞=

Mq x( )

Mp σp x,( ) P σp( ) e x( )( )=

Mtot σp x,( ) Mq x( ) Mp σp x,( )+=

+ Mps σp x,( )

φ σp x,( )

εcp σp x,( )

∆lcP εcp σp x,( ) xd0

L

∫=

∆ldiff ∆lcP ∆lP0–=

εpdiff∆ldiff

L-------------=

εpnew εp0 εpdiff+=

σpnew εpnew( )

σp Epεp=

σpnew σp–σp σpnew=

M φ EcIc( )⁄=

34

Page 51: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The vector n contains the prestress force information, where n is the number of prestress

loads chosen. The vector m contains the x-ordinate information for the moment-curvature

response and is consistent for all calculated responses, where m is the number of data points in each

curve. The matrix m,n contains the y-ordinate information for all calculated responses, where

the column i corresponds to prestress load Pi. As can be seen in Figure 3-5, the variation

between responses is approximately linear, thus, an linear interpolation should yield sufficient

accuracy in design. A similar interpolation scheme is also necessary for the variation of

with moment.

Figure 3-5. Family of moment-curvature responses for a specific cross section and varying prestress force assuming a concentric prestress

In some cases when the girder response is approaching the ultimate limit state, problems with

convergence of tendon stress may arise due to the high sensitivity of the response in the vicinity of

the ultimate resistance. As illustrated previously in Figure 3-5, the moment-curvature response is

nearly bi-linear, where the slope of the response in the linear elastic range is EI, and the slope of

the response when the girder is undergoing plastic deformation is nearly flat. This drastic change

in response over a very small range of bending moment causes the tendon stress calculation to have

a diverging iterative solution. If a girder undergoes significant plastic behaviour for a given

arrangement of load, the first iteration through the calculation will indicate that a large stress

[ ]

⎥⎥⎥⎥

⎢⎢⎢⎢

⎥⎥⎥⎥

⎢⎢⎢⎢

nm,m,1

2,22,1

n1,1,21,1

m

2

1

n 21

MM

MMMMM

P P P

LL

MOM

M

L

M

L

φ

φφ

P[ ]

φ[ ]

M[ ]

M{ }

εt and εb

-60000

-50000

-40000

-30000

-20000

-10000

0

10000

20000

30000

40000

2MN6MN10MN14MN18MN22MN

d = 3600mm

EI

Linear Elastic Response

M [kNm]

Ф [1/m]-0.002 -0.001 0 0.001 0.002

35

Page 52: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

increase in the tendon has occurred which is indicated by the large integrated strain in the hinge

region. For the second iteration, a higher stress will be guessed to perform the calculation;

however, this new guess causes the response to be primarily linear elastic indicating that the tendon

stress is much smaller than the guessed stress. Regardless of the accuracy of the guess value, the

method still renders divergent behaviour. A typical example of this behaviour is represented in

Figure 3-6 showing both the convergent and divergent solutions.

Figure 3-6. Iterative method behaviour for tendon stress calculation for a) girder impending plastic hinging, b) girder experiencing plastic hinging

A more stable iterative method has been observed to fix a desired target tendon stress, and

iterate the applied load until convergence is achieved. This procedure is given in Table 3-2, in

which the applied load must be approximated by a uniformly distributed load. The advantage of

this method is that only one moment-curvature response needs to be computed (for the assumed

tendon stress), and iterating of applied load is a straightforward process. This is typically not done

in design since the applied loads are known; however, this method provides a stable method to

determine the tendon stress increase for a range of applied loads up until crushing of the

compression flange. This method requires one to choose an increment value of load to iterate

towards a solution; therefore, the speed and accuracy of this solution is dependent on the

coarseness of the load increment and the closeness of the guess value to the actual value. A typical

example of this method is represented in Figure 3-7 showing termination after reaching the target

value.

100010501100115012001250130013501400

0 5 10 15 20 25 30100010501100115012001250130013501400

0 5 10 15 20 25 30

guess fp = 1190MPa

guess fp = 1220MPa

guess fp = 1240MPa

guess fp = 1020MPa

guess fp = 1350MPa

guess fp = 1240MPa

fp [MPa]

iteration

fp [MPa]

iteration

no solution foundconvergence obtained

36

Page 53: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 3-7. Iterative method solution for obtaining tendon stress by incrementing applied load

3.5 Preliminary Design Assumptions

The previous section has described methods which can be used to determine the stress in an

unbonded tendon due to girder deformation. These iterative methods allow for the accurate

Table 3-2. Procedure for calculating uniform applied load corresponding to a desired tendon stress

Step Calculation Comment

1Guess

Set target Set a target tendon stress. Guess an load to correspond to this tar-get stress

2 Calculate target tendon strain corresponding to the set target stress using known constitutive material law

3 Calculate a target tendon elongation measured from the reference state

4 Calculate total moment on girder including assumed applied load, plus primary and secondary prestress loads.

5Calculate curvature and tendon level strain distribution over the length of the girder using member response corresponding to lines of equal force

6 Calculate absolute tendon elongation by integrating concrete strain due to girder deformations

7 Calculate difference in tendon elongation from the reference state

8 Check

Check if the change in tendon elongation due to the assumed load is greater than or less than the change in tendon elongation due to the target stress. Increment the applied load

to get closer to the target stress and return to step 4. When target stress is reached, finish.

0

1.0

2.0

3.0

4.0

5.0

0 5 10 15 20 25

guess wf = 143.1kN/m

guess wf = 139.6kN/m

Δltarget = 1.92mm

Δldiff [mm]

iteration

criteria met whereΔldiff < Δltarget

criteria met whereΔldiff > Δltarget

q q0=

σp σpt etarg=

εpt etarg σpt etarg( )

∆lt etarg εpt etarg εp0–( )L=

Mq q x,( )

Mp σp x,( ) P σp( ) e x( )( )=

Mtot σp x,( ) Mq x( ) Mp σp x,( )+=

+ Mps σp x,( )

φ σp q x, ,( )

εcp σp q x, ,( )

∆l εcp σp x,( ) xd0

L

∫=

∆ldiff ∆l ∆lP0–=

∆ldiff ∆lt etarg or>

∆ldiff ∆lt etarg< q q qincrement+=

37

Page 54: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

calculation of tendon stresses by accounting for the distribution of strain along the girder at the

elevation of the prestressing tendon. The determination of this strain distribution is a

computationally intensive task as it requires the accurate calculation of stresses for all unbonded

tendon sets for a given arrangement of applied load. The stress in one tendon set is not independent

from the stress in another tendon set, rather all overlapping tendon sets that share a union of girder

length are coupled in the calculation of tendon stresses. The effect of increasing stress for a long

tendon that envelops the entire structure will place increased axial prestress, P, and its associated

prestress moment, , on the system. This change in prestress force will directly affect the state

of strain for any shorter tendons that are embodied within the length of the longest tendon.

Conversely, any change in tendon force of short tendons will have an effect on the average

straining in the longest tendon. It is not at all clear what increases in unbonded tendon stress can

be relied upon at the ultimate limit state due to member deformation for a girder with tendons of

varying lengths.

In order to construct a working bridge concept, the design process is simplified initially by

making conservative assumptions for the state of stress in the external unbonded tendons. At

serviceability limit states (before cracking), practice has been to assume the tendon stress is

constant using the long-term effective prestress, . This assumption is always

conservative (Menn, 1990) since the increase in tendon stress that is gained due to any deformation

of the girder is ignored. At the serviceability limit state, this assumption is reasonably accurate

since the average elastic straining of concrete over the length of the tendon due to deformation of

the girder is in general much smaller than the effective prestress strain in the prestressing steel.

For prestressed structures that have permanent unbonded tendons, it is economical to calculate

the ultimate resistance with the actual value of as it will minimize the consumption of

prestressing steel (Menn, 1990). The stress increase in unbonded prestressing due to girder

deformation typically is significant since flexural cracking and opening of joints is expected at the

ultimate limit state. In addition, the nominal compressive stress of 80MPa is relatively high

compared to conventional construction; therefore, the girder strain at the level of the prestressing

tendon at ULS is expected to be larger than a box girder using conventional strength concrete for

a cross section at the critical bending location. A refined treatment of is preferable at ULS for

an accurate prediction of ultimate strength; however, this refined treatment will be dealt with after

a working concept has been established using conservative assumptions. For the proposed box

girder which incorporates a complex arrangement of overlapping unbonded tendons of varying

P e×

σp σp∞=

∆σp

∆σp

38

Page 55: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

lengths, the tendon stress at ultimate limit state will also initially be assumed to be equal to the

long-term effective prestress .

Post-tensioned tendons are limited by the CAN/CSA-S6-06 to a tendon stress of at

jacking to provide some reserve capacity in case of unanticipated strand breakages (cl. 8.7.1). The

stress limit specified at transfer, which accounts for anchorage set loss, friction, initial relaxation

and elastic shortening, reduces the tendon stress to (cl. 8.7.1). Because the straight external

tendons have practically no angle break in the vertical plane, frictional losses will be small.

Regardless, this stress limitation at transfer of 0.7fpu still applies since the intent of this

specification is to reduce the probability of breakage of strand wires at the grips (cl. C8.7.1).

After jacking, long-term losses are realized which lower the effective prestress even further. In

lieu of considering long term losses explicitly, for the purpose of preliminary design and validation,

long term losses in tendon stress are incorporated into a reduced effective tendon stress. For design

of the proposed girder, is conservatively assumed to be 0.6fpu, which includes immediate

losses due to friction and anchorage set, and long term losses due to creep, shrinkage and

relaxation.

3.6 Preliminary Design Methodology for Cantilever Tendons

The following section will discuss recommended design considerations for the initial design of

a precast segmental cantilever bridge prestressed with unbonded tendons and no longitudinal

continuous bonded steel. First, the loads considered for the continuous girder will be described and

computed. Following this, the design for ultimate bending resistance will be discussed.

3.6.1 Loading

For a single span continuous system, it is relatively straightforward to calculate the longitudinal

factored bending moments, ignoring redistribution due to long-term effects. The total factored

load, , is considered as the following summation:

(3-9)

where is the cantilever dead load which is applied to the statically determinate

cantilever system, is the secondary prestress moment due to stressing of the continuity

tendons, and are the superimposed dead load and live load, respectively,

which are applied to the continuous statically indeterminate structure.

σp σp∞=

0.8fpu

0.7fpu

σp∞

Mf x( )

Mf x( ) αDMDdetx( ) αPMps αDMSDLindet

x( ) αLMLLindetx( )+( )+ +=

MDdetx( )

Mps x( )

MSDLindetx( ) MLLindet

x( )

39

Page 56: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

For the typical cross section shown in Figure 3-8, the unfactored dead load smeared over the

length of the segment is 88.2kN/m and the unfactored superimposed dead load of 28.7kN/m which

includes traffic barriers and an 80mm asphalt layer. These dead load derivations are summarized

below in Table 3-3 and Table 3-4.

Figure 3-8. Typical segment cross section

The secondary prestress moment, Mps, is a redundant moment of the statically indeterminate

system applied to the ends of the girder at the fixed end. Mps is constant along the entire length of

Table 3-3. Estimation of dead load for typical segment

Item Quantity[m3]

Density[kN/m3]

Weight[kN]

Continuous concrete 7.29 24.5 179

Transverse ribs 1.35 24.5 33.1

Anchorages 1.64 24.5 40.2

Subtotal [kN] 252

5% Allowance for Tendons and Struts 12.6

Total segment weight [kN] 265

Smeared weight over 3m segment [kN/m] 88.2

Table 3-4. Estimation of superimposed dead load for typical segment

Item Quantity[m3]

Density[kN/m3]

Weight[kN]

Traffic barriers 2.12 24.5 51.9

80mm asphalt layer 2.69 12.7 34.2

Total weight per segment [kN] 86.1

Smeared weight over 3m segment [kN/m] 28.7

40

Page 57: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the girder. It is necessary to begin the cantilever design with an initial estimate of Mps since it can

not be determined until the continuity prestressing is designed. Mps is assumed to be equal to 56%

of the total factored positive moment at midspan due to and applied to the

continuous system plus factored positive moment due to long-term redistribution of dead load. The

validity of this assumption for the proposed box girder will be demonstrated later on in the design

process when the continuity prestressing steel is designed.

Following the live load model defined in the CAN/CSA-S6-06, for a 12m wide bridge deck, 3

design lanes must be considered for the maximum effect (cl. 3.8.2). For the span of 90m, the lane

load model (cl. 3.8.3.2) tends to govern the design at the maximum and minimum moment

locations since the uniform applied load becomes more critical for midspan moment and end

moment than the CL-625-ONT truck model (cl. 3.8.3.3). However, the CL-625-ONT truck model

produces significant moment demands in other regions of the girder since it is a moving load

amplified by the Dynamic Load Allowance (DLA) (cl. 3.8.4.5.3). It is observed that the bending

moment diagram due to loads from the truck model produces critical locations for the positive

moment envelope from the abutment to approximately the quarter points of the span, and the truck

model is also critical for the negative moment envelope near the third points of the span. The

factored live load bending moment envelopes for the CL-625-ONT truck model and the lane load

model are shown on the tension side in Figure 3-9.

Figure 3-9. Comparison of live load models defined in CAN/CSA-S6-06

MSDLindetMLLindet

-60000

-50000

-40000

-30000

-20000

-10000

0

10000

20000

30000

40000

0 10 20 30 40 50 60 70 80 90

truck load governsnegative envelope

truck load governspositive envelope

truck load governspositive envelope

lane load governspositive envelope

M [kNm]

distance [m]

lane load governsnegative envelope

truck load governsnegative envelope

lane load governsnegative envelope

Max Lane Load (cl. 3.8.3.3)Min Lane Load

Max Truck Load (cl. 3.8.3.2)Min Truck Load

CL-625-ONT TruckNo. Lanes = 3MLF = 0.8

41

Page 58: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Considering dead load, live load, and secondary prestress moment for the ultimate limit state,

the ultimate demand is constructed using the load factors specified in ULS Combination 1 (cl.

3.5.1). The critical factored live load for the single span girder is taken as the maximum and

minimum demand envelope that encompasses the demands for both the truck load case and the lane

load case. The dead load and superimposed dead load demands are based on the estimates

previously outlined in Table 3-3 and Table 3-4. Secondary prestress moment is temporarily

assumed, and it will be calculated directly during the preliminary design of continuity tendons. If

the actual value Mps is found to be significantly different than the assumed value, the tendon design

must be refined with a subsequent iteration. The total loads following ULS Combination 1 to

produce the largest negative moment are illustrated in Figure 3-10 below.

Figure 3-10. ULS bending moments on continuous girder

3.6.2 Ultimate Moment Resistance

In preliminary design, an initial estimate for the sizing of flexural steel can be based on the

flexural resistance of the girder at the ultimate limit state, given by the equation below,

(3-10)

where is the long-term effective prestress of the unbonded tendons at the section, and z is

the internal flexural lever arm which can be assumed to be the height taken at the centroid of the

Mps

DL+Mps

DL+SDL+Mps

DL+SDL+LL+Mps

-200000

-150000

-100000

-50000

0

50000

100000

150000

0 10 20 30 40 50 60 70 80 90

LL envelope

Assumed Mps

M [kNm]

distance [m]

Mr φsAsσsy φpApσp∞+( )z=

σp∞

42

Page 59: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

thin compression slab to the level of the flexural steel. Since there is no continuous bonded

reinforcing steel crossing the joints for the proposed girder, the term goes to zero, and the

sectional moment resistance is based purely on the unbonded tendons.

Unbonded cantilever tendons are anchored to top corner blisters of the box girder segments.

Using standard 3m length segments, the 90m span continuous girder will be erected by two

adjacent 45m cantilevers consisting of 15 segments each. Therefore, with the cantilever length

discretized over 15 segments, the curtailment of cantilever tendons may be chosen in several ways.

Cantilever tendons are typically 12, 15, or 19 strand tendons composed of 0.6” diameter strands for

traditional cantilever-constructed bridges of typical weights (LoBuono, 1997). These tendon sizes

require the use of two anchorage unit sizes. In general, it is advantageous to minimize the variety

of anchorage units when designing the tendon arrangement so that the number of required stressing

jacks is also minimized.

For traditional cantilever construction, it is sometimes possible to design the cantilever

prestressing with only one size anchorage for all tendon units. This can be attributed to the classical

haunched profile which has an increase in internal lever arm as the section becomes deeper near

the support. For a parabolically haunched girder, the elevation of the girder axis also approximately

follows a parabolic profile, which allows the linearly increasing prestress force to provide adequate

resistance for the bending moment demands at serviceability. The parabolically haunched girder

axis is significant for the behaviour of the girder at serviceability, and the parabolically haunched

bottom flange is significant for the lever arm at the ultimate limit state. This concept is illustrated

in Figure 3-11.

For a constant depth girder, the girder axis is horizontal with a constant eccentricity to the

flexural steel. Thus, to satisfy for all x and if , an economical tendon

arrangement which minimizes steel consumption varies parabolically from the cantilever tip to the

φsAsσsy

Mr x( ) Mf x( )> Mf x( ) x2∝

43

Page 60: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

support since the flexural lever arm is constant. This arrangment, however, requires a larger variety

of tendon anchorages.

Figure 3-11. Cantilever prestress concepts for haunched and constant depth girders

3.6.3 Sizing of Cantilever Tendons

The maximum bending moment at the support will govern the design of the cantilever tendons.

The tendon arrangement should be chosen to satisfy a reasonable trade-off between the variation

of anchorage units used and the amount of prestressing steel consumed in the design. As discussed

previously, a varied arrangement of tendon units increasing in size from the midspan to the support

will provide the greatest economy in strand consumption. However, for the preliminary design, the

emphasis has been placed on the minimization of anchorage unit variation at the expense of

consuming more prestressing strand. With the motivation for fast construction and repetition of

standard details, only one anchorage unit (VSL ES6-12) has been used for the cantilever post-

tensioning which requires only one prestressing jack for all tendons. The VSL ES series multi-

strand post-tensioning anchorage has been chosen as it is more compact than the E series

anchorage. This design provides a reasonable compromise between required equipment and tendon

utilization. In comparison to an alternative tendon arrangement designed for minimum steel

consumption (shown in Figure 3-12), this design requires approximately 24% more cantilever

P(x)

Mr(x) > Mf(x)

z(x) z = const

Mf(x) Mf(x)

Pa

PbPa >Pb

Pi = const

haunched girder constant depth girder

Mr(x) = P(x)z(x) Mr(x) = P(x)z

44

Page 61: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

post-tensioning steel. The design for minimum steel, however, comes at the expense of the need

for several prestressing jacks and is a less than perfect rationalization of precasting.

Figure 3-12. Preliminary ULS demand and capacity for cantilever PT

The range in number of strands for a 12-strand anchorage is between 8 and 12 strands (Collins

et al., 1997). Using only one anchorage size for cantilever prestressing, the strand consumption can

still be optimized effectively by reducing the number of strands in the tendon and provide spares

in the anchorages. Therefore, to provide economy in prestressing strand consumption, the

minimum number of strands will be included for the longer tendon sets.

Mps

DL+Mps

DL+SDL+Mps

DL+SDL+LL+Mps

Capacity (minimum anchorage units)

Capacity (minimum steel consumption)critical location fornegative bending(governs design)

LL

DL+SDL+Mps

LL/(DL+SDL+Mps) = 0.55

-200000

-150000

-100000

-50000

0

50000

100000

150000

0 10 20 30 40 50 60 70 80 90

LL envelope

Assumed Mps

M [kNm]

distance [m]

45

Page 62: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Segments are numbered according to the order of placement. (i.e. segment 1 is placed closest

to the support, and segment 15 represents the cantilever tip). The initial prestressing scheme is

illustrated below.in Figure 3-13 and given in Table 3-5.

Figure 3-13. Layout of cantilever tendons

One important point to consider is the impact on prestressing steel consumption when

designing light weight cantilever bridges having thin slabs and thin webs. In general, the dead load

moment of cantilever-constructed bridges consumes a large majority of the steel capacity. Menn

(1990) states that the ratio of prestress moment to dead load moment near the supports for

cantilever-constructed bridges is typically lower than for conventionally built bridges. This is a

consequence of a low ratio of live load moment to dead load moment, which is typically equal to

0.2 for cantilever constructed bridges. In this design example, under the initial approximation of

, the ratio of factored live load moment to dead load moment is 0.55 (which includes

superimposed dead load). The unfactored live load moment to dead load moment ratio is computed

as . Since the weight of the cantilever during construction is a large

contributor to the consumption of prestressing steel, the overall consumption of materials is

expected to be less than a classical cantilever-constructed bridge. Assuming that live load moment

is constant between the traditional girder and the proposed box girder, for the ratios given above,

Table 3-5. Preliminary prestressing scheme for cantilever tendons based on ultimate capacity

Segment 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Strands 8 12 12 12 12 12 12 12 12 12 12 8 8 8 8

Mps

0.55 1.2 1.7⁄( )× 0.39=

46

Page 63: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

a traditional cantilever-constructed bridge consumes 1.95 times more concrete material. In light of

this difference, it is clear that a relative design ratio of tendon utilization in terms of kilograms of

steel per cubic meter of concrete becomes less meaningful when comparing the proposed design

concept to existing structures. An explicit quantitative comparison of material consumption of the

proposed girder and other reference bridges is given later in Chapter 8.

3.7 Change in Structural System

A change in structural system occurs after forming a monolithic closure at the centre segment

and stressing the continuity tendons. This change in structural system modifies the erected

structure from two statically determinate cantilevers to a statically indeterminate continuous

girder. The force effects which undergo a redistribution are the permanent loads which are initially

applied to the statically determinate structure. For the proposed bridge under consideration, these

forces include the self-weight of the girder at the time of erection, and the effects of cantilever

prestressing tendons. The permanent load on the final structure due to superimposed dead load

does not undergo any long-term redistribution since it is placed on the continuous structure after

closure of the midspan. Figure 3-14 describes the change in structural system and the associated

force effects. The self-weight of the cantilevers produces a moment:

(3-11)

at both ends with an initial deflection and rotation due to the combined effect of dead load and

prestressing, , respectively. Since the prestress moment generally offsets the dead

load moments, this girder deflection tends to be small. In some cases, the designed cantilever

tendons cause prestress moments that exceed the dead load moments so a positive cantilever tip

deflection is possible. Through the match-casting of segments and adequate geometry control, the

initial deflection of the cantilever tips at the time of closure can be controlled within a reasonable

tolerance (Post-Tensioning Institute, 2006). Proper geometry control can also minimize the angular

rotation at the tip, , however a small discrepancy in tip angle can be easily accommodated by

the cast-in-place closure segment. Through long-term concrete creep, additional deformations

MclowLc

2

2---------- wL2

8----------= =

yclo and ωclo

ωclo

47

Page 64: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

occur in the structural system due to self-weight, , as creep has a softening effect on the long-

term elastic modulus of concrete..

Figure 3-14. Redistribution of forces through concrete creep due to change in structural system (adapted from Podolny et al., 1982)

Considering the concrete strain at any point of the structure, Podolny et al. (1982) describes the

total strain to be the sum of two terms: the strain before continuity is achieved, and the strain after

continuity is achieved. Menn (1990) describes this sum more explicitly as:

(3-12)

where is the initial concrete stress due to permanent load, is the elastic modulus of

concrete, is the time-dependent creep function, and is the aging function for which a typical

value is 0.8. Based on this relationship, Menn (1990) provides a useful engineering approximation

for the calculation of redistributed stress based on an average creep coefficient for dead load plus

prestressing:

(3-13)

where is the fully redistributed stress, is the actual stress at the time of closure, and

is the stress obtained assuming the entire structure was cast simultaneously on conventional

falsework. The distribution factor 0.8 is a typical value for cast-in-place concrete structures.

Mclo

Lc Lc

L = 2Lcyclo

yφ2ωclo

MφMφ

Mclo - Mφ

(a)

(b)

(c)

(d)

ε t( )σ0Ec------ 1 φ t( )+( )

σ t( ) σ0–Ec

---------------------- 1 µφ t( )+( )+=

σ0 Ec

φ t( ) µ

σ∞ σclo 0.8 σfal σclo–( )+=

σ∞ σclo

σfal

48

Page 65: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

As a general rule of thumb, the long term creep coefficient is approximated to be for

precast structures. In comparison to creep expressions given in the CAN/CSA-S6-06 (cl. 8.4.1.6.3),

this approximation appears to be reasonable and perhaps even conservative. Assuming that the

precast segments are 28 days old before erection, a creep coefficient of 1.5 corresponds to an

average relative humidity (RH) of 32% over 75 years. For higher relative humidities, the creep is

even less. The expression for creep defined in the Code is illustrated in Figure 3-15.

Figure 3-15. Time-dependent creep coefficient assuming t0 = 28 days, a) for 75 year period, b) for 5 year period

A modified relationship which is similar to Equation 3-13 has been given that is acceptable for

the determination of redistributed stress for precast structures (Post-Tensioning Institute, 2006).

For serviceability, bending moment is linearly proportional to the stress in a girder, and so a similar

relationship can be formulated for the calculation of fully redistributed bending moment. Thus, the

distribution equation is given as:

(3-14)

where is the fully redistributed moment, is the actual bending moment at the time of

closure, and is the bending moment obtained assuming the entire structure was cast

simultaneously on conventional falsework. The distribution factor 0.68 is typical for precast

concrete structures. Using this relationship, it is possible to determine the moment redistribution

due to creep, equal to the distribution term in Equation 3-14 as .

φ∞ 1.5=

00.20.40.60.8

11.21.41.61.8

2

0 500 1000 1500 20000

0.20.40.60.8

11.21.41.61.8

2

0 5000 10000 15000 20000 25000daysdays

φφ

5 years1 year1 month

RH = 32% a = 10 MPa rv = 82mmf’c = 80MPa t0 = 28 days

φ∞ = 1.5φ∞ = 1.5

1 yr

5 yrs

75 yrs

a = 10 MPa rv = 82mmf’c = 80MPa t0 = 28 days

RH = 70%RH = 50%RH = 32%

a) b)

M∞ Mcloφ∞

1 µφ∞+------------------- Mfal Mclo+( )+=

Mclo1.5

1 0.8 1.5⋅+---------------------------- Mfal Mclo–( )+=

Mclo 0.68 Mfal Mclo–( )+=

M∞ Mclo

Mfal

Mφ 0.68 Mfal Mclo–( )

49

Page 66: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

For the single span continuous structure we are considering, the redistribution of bending

moment is a constant value for all locations on the girder since Mclo and Mfal are offset only by the

redundant moment that provides continuity at midspan. Therefore, the redistribution can be easily

computed by considering the end moments of Mclo and Mfal. Using the uniform dead load of wD

= 88.2kN/m derived previously in Table 3-3, the redistributed moment is shifted towards positive

bending by 24300kNm. This simple calculation is demonstrated in Table 3-6 below.

3.8 Preliminary Design Methodology for Continuity Tendons

The following section will discuss recommended design considerations for the initial estimate

of continuity tendons for a monolithic closure at the midspan.

3.8.1 Ultimate Moment Resistance

Similar to the preliminary design of cantilever tendons, an initial estimate for the sizing of

flexural steel can be based on the behaviour of the system at the ultimate limit state. For sections

which have both cantilever and continuity prestressing tendons, a simple calculation to account for

both top and bottom prestressing is:

for positive flexure (3-15)

for negative flexure (3-16)

where Ptot is the total prestressing force on the section, kt is the distance from the girder axis to

the centroid of compression in the top flange, kb is the distance from the girder axis to the centroid

of compression in the bottom flange, and Mpt and Mpb are the prestress moments due to top and

bottom prestress tendons, respectively. These parameters are illustrated in Figure 3-16 below. In

general, the prestressing tendons are located as close as possible to the flanges to maximize the

Table 3-6. Calculation of long-term bending moment redistribution

Item αD wD [kN/m]

L [m]

End Moment Equation Moment [kN/m]

Mfal 1.2 88.2 90 107000

Mclo 1.2 88.2 90 71400

0.68(Mfal - Mclo) - - - - 24300

wDLc2

2--------------wDL2

8--------------=

wDL2

12--------------

Mr φp Ptotkt M+ pt Mpb+( )=

Mr φp Ptotkb M+ pt Mpb+( )=

50

Page 67: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

internal lever arm. The ultimate resistance of the section is determined by moving the eccentrically

applied loads to the centroid of the cross section and considering the equivalent concentric

sectional forces.

Figure 3-16. Sectional forces due to prestressing with top and bottom tendons

3.8.2 Spreading of Forces

For structures prestressed with external unbonded tendons, moment resistance is essentially a

product of the tension steel force and the internal lever arm between the tension steel and the

centroid of compression stress in the concrete. Since the increase in steel stress due to the formation

of cracks for structures with unbonded tendons is in general less sensitive than those with bonded

tendons, it is especially important to ensure that the internal lever arm may be developed at a

section where the resistance is required. The development of internal lever arm for the box girder

is dependent on an induction length which takes into account the spreading of forces in the concrete

from the point of origin of the applied prestress force (Canadian Standards Association, 2006).

Menn (1990) illustrates this concept clearly using a three dimensional truss model to describe the

flow of forces due to an anchorage force located at the bottom corner of a box section (Figure 3-

17). This model depicts an extruded cross section of a concrete box girder by considering only the

closed box shape. On the left side, the diagram illustrates the state of stress in the concrete, fc, at

various distances from the applied prestress force. Menn (1990) shows that the plane sections

hypothesis is not valid in the vicinity of the applied load where there is a significant peak stress at

the the corner. On the right side, the diagram depicts a simple truss model that represents the

spreading of force in the cross section. Menn (1990) assumes a spreading angle of which

et

eb

Applied Forces Equivalent Sectional Forces

Ptot = Pt + PbPt

Pb

Mpt = Ptet

Mpb = Pbeb

Ptot

kt

kb

Cross Section

45°

51

Page 68: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

results in a required development length of la = h+b/4 from the bottom corner where the prestress

force is assumed to originate.

Figure 3-17. Moment development length, la, due to anchorage of continuity tendon (adapted from Menn, 1990)

The CAN-CSA-S6-06 requires for segmental cantilevered construction that continuity tendons

be anchored at least one segment beyond where they are theoretically required since the prestress

force requires an induction length before it may be assumed to be effective over the whole section.

This requirement is consistent with the model by Menn (1990) for girders that have a web height

less than the segment length minus b/4. For precast segmental girders that have a limited segment

length of only 3m and have a web height that is deeper than the segment length, providing only one

segment length for moment development is unconservative. The moment development length

specified by the Code also does not distinguish between the required development length for

girders post-tensioned with internal bonded or external unbonded tendons.

For the segmental design we are considering, the external unbonded tendons applies a force to

the cross section at external anchorage blisters located on the inside of the box near the junction of

the web and flange. However, since this load is applied from within the box itself, the origin of this

point load is eccentric to the centroid of the web and flange intersection. It is necessary for the

prestress force to spread towards the corner before it can begin to develop within the section. This

additional spreading of forces due to the eccentric anchorage of prestress force is an additional

consideration that must be taken into account for the spreading of forces in the box girder. For the

purpose of this chapter, only the longitudinal spreading length of the anchorage force is important

when designing the post tensioning arrangement for global bending. The spreading length is

approximately equal to twice the length of the anchorage blister (shown in Figure 3-18). This

spreading length has been determined based on a local spreading truss model which is discussed in

V

MP = const

V

P

PFcFc

fc

hb/4

P

MPM

la = h +b/4

52

Page 69: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

detail in Chapter 6. The total required moment development length is illustrated below in Figure 3-

18.

For the girder considered in this study, the anchorage length, girder depth, and top flange span

between webs is 1495mm, 3600mm, 5000mm, respectively. Therefore, the required moment

development length for the local spreading of forces to the corner of the box is approximately

7840mm. This corresponds to over two segment lengths required for the spreading of forces to

develop the compressive stresses in the top flange of the girder.

Figure 3-18. Required moment development length for a girder post-tensioned with unbonded tendons

3.8.3 Sizing of Continuity Tendons

The maximum effect for positive bending is greatest following long-term redistribution of

forces; therefore, the full redistribution of bending moment previously calculated in Table 3-6 will

be assumed to consider the worst case for positive bending. Loads which are applied to the girder

immediately after construction and before much redistribution has taken place will result in smaller

positive bending demands; therefore, it does not present a critical load case for the design of

positive bending resistance. Up until now, an initial assumption of Mps has been used for the design

of cantilever tendons since it is undefined without a continuity post-tensioning arrangement.

Following the preliminary design of continuity tendons, this initial assumption of Mps must be

checked by calculating Mps directly. Afterward, the tendon design must undergo a subsequent

design iteration based on the Mps calculated with the preliminary continuity tendon design if the

assumed value is unconservative or uneconomical.

For consistency with the tendon selection of the cantilever post-tensioning, the continuity

tendons have been designed to minimize the variety of anchorage sizes and the associated

MP = const

2Lb h b/4

M

V V

P

53

Page 70: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

prestressing jacks required for construction while satisfying a reasonable utilization of prestressing

steel. Using the same VSL ES6-12 and VSL ES6-19 size anchorages, the prestressing schedule is

shown in below in Table 3-7.

Figure 3-19. Layout of continuity tendons

This arrangement of tendons provides reasonable economy in steel consumption for the design

of continuity tendons while maintaining minimum anchorage variety. The adjusted factored

demands including long-term creep and the factored capacity envelope is shown in Figure 3-20.

When considering the long-term effects of the structure, the live load moment to dead load moment

at the support is further increased from 0.55 to 0.74 due to the redistribution of forces. At midspan

the live load moment to dead load moment is 0.69. For long term redistribution of forces, a savings

in continuity tendons can also be realized due to the design of light weight box girders since there

is a lesser proportion of the total load on the bridge that has been applied to the statically

determinate system during erection.

To satisfy the safety requirements for longitudinal flexure, the factored demands must be

smaller than the factored resistance at all locations along the girder. Factored demands have been

calculated according to ULS Combination 1 (cl. 3.5.1). It is seen in Figure 3-20 that there is a

significant amount of reserve capacity in negative bending for most locations along the girder, and

Table 3-7. Preliminary prestressing scheme for continuity tendons based on ultimate capacity

Segment 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Strands - - - 19 19 19 12 12 12 8 8 - - - -

54

Page 71: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

only one critical location in negative bending at the support. The demand for positive flexure shows

a smaller margin for reserve capacity with respect to the positive bending capacity envelope. The

moment development length due to spreading for positive bending has been accounted for by

shifting the demand envelope to provide the required development length determined previously

as 7840mm (Figure 3-20).

Figure 3-20. Preliminary ULS demand and capacity for continuity tendons

3.8.4 Secondary Prestress Moment due to Continuity Tendons

Once a preliminary prestressing scheme for continuity tendons is determined, the initial

estimate made for must be adjusted to reflect the preliminary continuity tendon arrangement.

Following this adjustment, the system must be checked so that the initial estimate did not result in

an unconservative design, or alternatively, an overly conservative design. For a constant depth

girder with equal stiffness along its length, the secondary prestress moment can be calculated

directly due to the influence of each tendon using a frame section analysis. This analysis assumes

that the abutments are longitudinally flexible, meaning that there is no axial restraint provided by

the abutments due to the stressing of continuity tendons.

A frame analysis consisting of a single beam element with constant stiffness having moment

fixity at the supports is considered. As shown previously in Figure 3-16, the application of

prestress forces on a section can be decomposed into an equivalent prestress force acting at the

-200000

-150000

-100000

-50000

0

50000

100000

150000

0 10 20 30 40 50 60 70 80 90

Mps DL+Mps+Mφ

DL+SDL+Mps+Mφ

DL+SDL+LL+Mps+Mφ

ULS Capacity

DL+SDL+LL+Mps+Mφ (shifted)Mps+Mφ

LL

DL+SDL+Mps+Mφ

LL/(DL+SDL+Mps+Mφ) = 0.74

LL envelope

Assumed Mps

M [kNm]

distance [m]

shifted demandenvelope to accountfor spreading

Mps

55

Page 72: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

girder axis and a corresponding prestress moment equal to . For a continuity tendon that is

stressed symmetrically about the midspan of the structure, the secondary prestress moment due to

the deformations of the indeterminate system are proportional to Lten/L where Lten is the length of

the tendon and L is the span length (Figure 3-21). Since this design includes only straight tendons,

no deviation forces are considered for Mps since deflections are assumed to be small.

The secondary prestress moment breakdown that results from the preliminary continuity

tendon design is summarized in Table 3-8 below.

Table 3-8. Secondary prestress moment calculation

Segment Tendon Unit Pba

[kN]

a. Prestress force assumes a constant tendon stress of 0.6fpu.

Mp [kNm]

Lten [m]

Lten/L Mps [kNm]

4 ES-19 5640 10600 72 0.800 8450

5 ES-19 5640 10600 66 0.733 7750

6 ES-19 5640 10600 60 0.667 7040

7 ES-12 3560 6670 54 0.600 4000

8 ES-12 3560 6670 48 0.533 3560

9 ES-12 3560 6670 42 0.467 3110

10 ES-8 2370 4450 36 0.400 1780

11 ES-8 2370 4450 30 0.333 1480

Σ = 37200

P e×

56

Page 73: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 3-21. Secondary prestress moment due to prestressed continuity tendon

The secondary prestress moment due to the summation of all continuity prestressing tendons

described in Table 3-8 is 37200kNm. This corresponds to 56% of the factored positive moment due

Lten

tendon

eP P

a) elevation of actual structureLLside Lside

Mp Mp

Mps Mps

b) assumed fixity at abutment

c) simplified beam model for analysis of moment only

d) primary system and applied prestress moment

Mp Mp

e) primary system bending moment Mp

f) primary system deflection

g) redundant moments to restore continuity

Mps Mps

Mps

Mp(1-Lten/L)

h) redundant bending moment, Mps

i) total prestress moment, Mptot = Mp + Mps

Mps = Mp(Lten/L)

57

Page 74: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

to MSDL, MLL, and . Therefore, the initial assumption used for the proposed box girder is valid

for the designed prestress arrangement. There is no need to adjust the loads on the system for a

second iteration of tendon arrangement.

Figure 3-21 shows that the rigid abutment attracts more positive bending moment for tendons

that are anchored further from the midspan. Thus, the longest continuity tendon lengths have a

greater influence on the secondary prestress moments than the shorter tendon lengths anchored

closer to the midspan. Since Mf varies directly with Mr due to Mps, it is often necessary to compute

a few iterations of tendon arrangement designs before a final design is selected where Mr > Mf for

all sections.

3.9 Bottom Flange Sizing

In preliminary design, the box girder cross section is assumed to have a typical geometry with

a constant bottom flange thickness of 150mm (shown in Figure 3-22 below). For cantilever-

constructed bridges, large negative moments exist at the supports for both dead load moments and

live load moments. In Section 3.6.3 the cantilever tendons were designed based on the ultimate

limit state capacity which assumes an internal lever arm, z. Now that the tendon arrangement is

known and the demands on the structure have been determined, the capacity must be checked for

negative flexure at the critical bending location.

The internal lever arm at the ultimate limit state is directly related to the axial prestress force,

P, applied by the unbonded cantilever tendons. At the support, the axial prestress force in the girder

is due to the accumulation of all cantilever prestressing tendons. At the ultimate limit state, P =

47.5MN (under the initial assumption that all tendon stresses are 60% of ultimate strength). A

bottom flange thickness of 150mm is seen to be insufficient for the large prestress force since

premature crushing of the bottom flange occurs at a very small curvature (Figure 3-22).

58

Page 75: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 3-22. Ultimate state of stress for section with undersized bottom flange

For design, the bottom flange thickness should be sized based on an idealized stress block at

ultimate limit state. Sizing the bottom flange so that at ultimate will maximize the internal

lever arm and the flexural ductility of the girder in negative bending. Internal equilibrium of forces

provides the design equation for sizing of the bottom flange due to negative bending:

(3-17)

where tbf is the bottom flange thickness, P is the prestress force, is the concrete material

factor, are the stress block factors, is the ultimate material stress, and b is the bottom

flange width.

Figure 3-23. Appropriately sized bottom flange thickness for negative bending at ULS

Using the FRC constitutive material law model developed for analysis, the ultimate moment

resistance has been observed to occur when . The associated stress block factors

derived for this state of stress are and (Figure 3-23). The stress block

y c

P

Stress Distribution

fc(y)

UltimatePlane ofStrain

Cross Section

P

Force Distribution

fc(y)ּb(y)

z < zassumedΔy

tbf c=

tbfP

φcα1f'cβ1b---------------------------=

φc

α1 and β1 f'c

y c β1c

α1f’c

P P

IdealizedStress Block

Stress Distribution

fc(y)

Ultimate Plane ofStrain

Cross Section

P

Force Distribution

fc(y)ּb(y)

z

εb 1.6ε'c=

α1 0.903= β1 0.774=

59

Page 76: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

factor is a value which is used to match the elevation of the actual compression stress centroid

to the centroid of an idealized equivalent rectangular stress block. Knowing that the centroid of a

rectangle is the half depth, the value is a multiplier to give the height of the rectangular stress

block expressed as a scalar multiple of the height of compression, c. The stress block factor is

determined by equating the centroids of the actual stress distribution to the rectangular distribution:

Once the centroids of the compression stress distributions have been matched, the stress block

factor , which is a scalar value to determine the equivalent stress of the idealized stress block,

is determined by equating the compression force in the concrete:

Using Equation 3-17, the bottom flange at the support should be 305mm thick. It is

recommended for design that the bottom flange thickness varies linearly from 305mm at the

abutment to the typical thickness of 150mm at the point of inflection.

3.10 Prestress Distribution

Axial prestress force is introduced at every anchorage location along the length of the girder.

Cantilever tendons introduce an axial force on girder segments that are located between the

cantilever tendon anchorage and the abutment. The prestress force on the girder due to cantilever

tendons is accumulative from the midspan region towards the support. Continuity tendons

introduce an axial force on the girder between symmetrical anchorages across the midspan region.

Similarly, the prestress force on the girder due to continuity tendons is accumulative from the

continuity tendon anchorages towards the midspan region. Considering the total prestress effects

which combines the cantilever and continuity tendons, the resultant prestress force distribution

over the length of the girder obtained from the post-tensioning design shows that the axial force is

β1

β1

β1

yφc fc y( )y yd

0

c

φc fc y( ) yd0

c

∫---------------------------------

β1c2--------==

β1

2 fc y( )y yd0

c

c fc y( ) yd0

c

∫-------------------------------=

α1

y φc fc y( ) yd0

c

∫ b y( ) φcα1f'cβ1cb=⋅=

α1

fc y( ) yd0

c

∫f'cβ1c------------------------=

60

Page 77: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

relatively constant over the length of the girder. The primary prestress moment that is induced due

prestress forces applied eccentrically to the concrete centroid varies along the length of the girder

in a shape that corresponds to the inverse of the applied bending moments.

Figure 3-24. Primary prestress force distribution

a) girder elevation

-80000

-60000

-40000

-20000

0

20000

40000

60000

80000

0 5 10 15 20 25 30 35 40 45

0

10000

20000

30000

40000

50000

60000

70000

80000

0 5 10 15 20 25 30 35 40 45

P [kN]

Ptot

Pt

Pb

cantilever tendonprestress, Pt

continuity tendonprestress, Pb

total prestressforce, Ptot = Pt + Pb

b) girder prestress force, P

c) girder prestress moment, Mp = Pּe

Mp [kNm]

continuity tendonprestress moment,Mpb = Pbeb

cantilever tendonprestress moment,Mpt = Ptet

total primary prestress moment, Mp = Mpt + Mpb

Mp

Mpt

Mpb

61

Page 78: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The deformation response and stress distribution along the length of the girder due to applied

loads and prestress forces is determined through the unique moment-curvature relationships that

pertain to each segment. As shown in Figure 3-24, the sectional prestress forces vary along the

length of the girder where each segment corresponds to a specific prestress force, P, and a total

primary prestress moment, Mp. For the typical box girder cross section, these two values determine

the member response for the segment under consideration which corresponds to the constant

prestress force, P. Using the interpolation method that was described previously in Section 3.4, the

distribution of sectional effects along the length of the girder can be determined. This procedure is

illustrated below in Figure 3-25 for the curvature distribution along the girder length. Only two

examples are shown for the girder at segments 3 and 14. A similar procedure is performed to

determine the distribution of stresses and strains along the length of the girder.

62

Page 79: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 3-25. Girder response for varying prestress force

3.11 Ultimate Limit State

Figure 3-26 displays the girder response at ultimate limit state due to permanent factored load

(dead load plus prestressing) and maximum factored live load. Flexural cracking is observed to

-150000

-100000

-50000

0

50000

100000

0 10 20 30 40

-0.0005-0.0004-0.0003-0.0002-0.0001

00.00010.00020.00030.00040.0005

0 10 20 30 40

-150000

-100000

-50000

0

50000

100000

-0.002 -0.001 0 0.001 0.002

-150000

-100000

-50000

0

50000

100000

-0.002 -0.001 0 0.001 0.002

M [kNm]

Φ [1/m]

interpolated curvesconstant force datums

P14

Φ [1/m]

M [kNm]

Segment 14 P14 = 37.1MN

Mp14 = -39.8MNm

Mp3

Mp14

P3

M [kNm]

Segment 3 P3 = 41.6MN

Mp3 = -66.0MNm

Φ [1/m]

a) girder elevation b) interpolation between known curves of constant force

P3,Mp3

P14,Mp14

c) ULS bending moments d) primary prestress moment as internal resistance

e) cuvature distribution along girder length

interpolated curvesMp internal resistance

63

Page 80: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

occur at the support due to negative bending. Localized tension strains on the top surface are

observed to occur in the vicinity of the support.

Figure 3-26. Girder response at ULS (half span shown)

3.12 Serviceability Limit States

The following section will describe some of the key components of serviceability required for

adequate durability, crack control, and deflection limit states.

-200000

-150000

-100000

-50000

0

50000

100000

150000

0 10 20 30 40

-0.0005-0.0004-0.0003-0.0002-0.0001

00.00010.00020.00030.00040.0005

0 10 20 30 40

-0.002

-0.0015

-0.001

-0.0005

0

0.0005

0 10 20 30 40

-0.002

-0.0015

-0.001

-0.0005

0

0.0005

0 10 20 30 40

-80

-70

-60

-50

-40

-30

-20

-10

00 10 20 30 40

-80

-70

-60

-50

-40

-30

-20

-10

00 10 20 30 40

M [kNm] Φ [1/m]

εt [mm/mm] εb [mm/mm]

σt [MPa] σt [MPa]

CapacityDL+SDL+LL+Mps

DL+SDL+Mps

critical location for negative bending

concrete model assumes zero tensile capacityaccounting for no tensile strength across the joint

64

Page 81: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

3.12.1 Construction

For the erection of segmental cantilever girders with unbonded tendons, large prestressing

forces are applied to the top corner of the box girder at the location of anchorage blisters. The

prestress moment, , can be significant and often exceeds the dead load bending moment due

to the light weight girder design, especially in the vicinity of the cantilever tip where the dead load

moment is zero. For segmental bridges, epoxied joints are necessary to prevent corrosive de-icing

salts from leaking into the box girder. At the time of erection, a minimum compressive stress of

350kPa must be applied to the entire cross section for adequate adhesion of joints (cl. 8.22.6.1.2).

Temporary bottom post-tensioning is necessary for full compression of the joint face in the vicinity

of the cantilever tip. The critical loading condition for each joint face corresponds to the jacking

force of the cantilever tendon immediately ahead of the joint.

The largest cantilever tendon unit considered in this analysis consists of 12 strands of 0.6”

diameter. As the jacking force corresponds to 80% ultimate stress of the tendon, the temporary post

tensioning must be designed for the associated jacking moment. The DSI Threadbar system

(Dywidag-Systems International, 2006) can be easily adopted for the temporary post tensioning of

the cantilever during construction. Typically, only the first three to four segments at the cantilever

tip are critical - past this point, the dead load of the cantilever applies sufficient dead load moment

to provide full compression on the joint face. Temporary post-tensioning is assumed to be reusable,

so the maximum temporary jacking stress is only 0.6fpu. The temporary post-tensioning steel can

be optimized by using couplers and leapfrogging bars for the progressive erection of the cantilever.

Assuming a linear elastic plane section at the joint, and neglecting the benefit of dead load, the

temporary post-tensioning to provide the required compressive concrete stress at the critical joint

is two 36mm diameter Threadbar post-tensioning rods stressed to 60% of ultimate stress.

3.12.2 Cracking

The SLS girder response due to unfactored loads is determined based on SLS Combination 1

(cl. 3.5.1). It is observed that the girder is fully compressed for serviceability using the tendon

arrangement designed to satisfy strength requirements at the ultimate limit state. The girder

response is shown in Figure 3-27 below.

P et×

65

Page 82: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 3-27. Girder response at SLS (half span shown)

-0.0005-0.0004-0.0003-0.0002-0.0001

00.00010.00020.00030.00040.0005

0 10 20 30 40

-0.002

-0.0015

-0.001

-0.0005

0

0.0005

0 10 20 30 40

-0.002

-0.0015

-0.001

-0.0005

0

0.0005

0 10 20 30 40

-80

-70

-60

-50

-40

-30

-20

-10

00 10 20 30 40

-80

-70

-60

-50

-40

-30

-20

-10

00 10 20 30 40

-0.001

-0.0005

0

0.0005

0.001

0.0015

0.002

0 10 20 30 40

-15-10-505

1015202530

0 10 20 30 40

-200000

-150000

-100000

-50000

0

50000

100000

150000

0 10 20 30 40

M [kNm] Φ [1/m]

εt [mm/mm] εb [mm/mm]

σt [MPa] σt [MPa]

Θ [rad] Δ [mm]

DL+SDL+LL+Mps

DL+SDL+Mps

66

Page 83: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

3.12.3 Global Deflection

The vibration limit state defined in CAN/CSA-S6-06 for SLS Combination 2 does not govern

the design of the box girder. For the 90m span with a 25:1 span-to-depth ratio, the vertical

deflection due to longitudinal bending for one passing truck (cl. 3.8.4.1c) is 11.4mm. The

additional sidewalk deflection as a result of transverse slab bending at the barrier face has been

computed as 2.8mm. This result is obtained from a grillage analysis which is described later in

Chapter 4. The girder twist caused by torsion in the box girder due to a truck placed eccentrically

from the bridge centreline has been computed as 3.1mm from a linear elastic analysis. The

combined deflection of 17.3mm is well within the deflection limit of 42mm for bridges with

occasional pedestrian use corresponding to the first-order vibration computed as 1.9Hz. For typical

highway overpass bridges, this deflection is acceptable.

3.13 Refinement in Tendon Stress

The proposed box girder bridge concept has been developed under the assumption that

at the ultimate limit state. The consequence of this assumption is an increase in

consumption of required prestressing steel. It is of interest to investigate the stress increase in the

unbonded prestressing steel that is possible at the ultimate limit state due to girder deformation for

the purpose of achieving greater economy prestressing steel consumption.

3.13.1 Ductility of Fibre-Reinforced Concrete

With reference to the demand and capacity envelopes shown previously in Figure 3-12 and

Figure 3-20, the critical flexural regions of the girder are observed to be at the midspan and at the

supports with the least reserve capacity. Due to the large concrete area provided in the 12m wide

top flange and the relatively small factored axial prestress of 34.7MN, there is a substantial

capacity for rotational ductility for positive bending at the midspan. At the support, however, the

bottom flange has a considerably smaller width and a higher factored axial prestress force of

47.5MN. At the ultimate limit state, the rotational capacity is limited by the crushing of concrete.

In comparison to a typical compressive stress-strain curve for concrete with a nominally similar

compressive strength, FRC exhibits far greater ductility in compression for the post-peak response

(Figure 3-28). This ductility in compression becomes abundantly important when considering the

ultimate flexural response of girders prestressed with unbonded tendons. The increment of strain

σp σp∞=

67

Page 84: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

in unbonded tendons is generally lower than bonded tendons (Ramos et al, 1996); therefore,

yielding of steel at the ultimate limit state is less likely to take precedence over crushing of the

compression flange. Typically, the failure of normal concrete in crushing is a brittle mechanism,

allowing little flexural ductility and plastic redistribution of forces. The capability of FRC to

sustain significant compressive stress for large strains following the peak stress allows structural

members to undergo larger curvatures before failure. For this reason, it is expected that girders cast

with high strength FRC can maintain equilibrium under large curvatures due to a ductile crushing

of the compression flange.

Figure 3-28. Stress-strain response of FRC and typical normal-weight concrete

The use of FRC in combination with structures prestressed with unbonded tendons appears to

be complementary since FRC allows for increased girder deformation to produce higher strains in

the tendon at ultimate limit state. The moment-curvature response shown in Figure 3-29 compares

the behaviour of identical box girder sections with normal and fibre reinforced concrete for

negative bending. For highly prestressed sections, the rapid degradation in compressive strength of

normal concrete when causes crushing failures to occur at small curvatures. For

geometrically identical box girders, the bottom flange of the girder fabricated with regular concrete

fails when whereas the bottom flange of the girder fabricated with FRC fails when

. Since negative bending at the abutment is expected to control the failure mechanism

at the ultimate limit state, it is prudent to maximize the ductility of this failure mode for safety. For

maximum curvature when , the idealized stress block factors have been computed to

be and . The calculation method for stress block factors has been

0

01

02

03

04

05

06

07

08

210.010.0800.0600.0400.0200.00

FRC model (from tests)

normal-weight concrete (Popovics, 1970)ε’c = 0.00271Ec = 37100

post-peak ductility of FRCdue to fibre confinement

fc [MPa]

εc [mm/mm]

ε ε'c>

εb 1.4ε'c=

εb 3.07ε'c=

εb 3.07ε'c=

α1 0.812= β1 0.921=

68

Page 85: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

described previously in Section 3.9. Maximum ductility, however, can be achieved using the

bottom flange concrete thickness designed for maximum strength.

Figure 3-29. FRC comparison for total moment-curvature response in negative bending

3.13.2 Deformation Capacity

The rotational ductility of the concrete box girder allows the structure to undergo plastic

deformation to produce large cracks at the ultimate limit state. The available rotation capacity can

be derived from the curvature distribution over the length of the girder. Large curvatures result

when the bending moment demands approach the ultimate capacity of the box girder. Under this

condition, the centroid of compression stress block is entirely in the compression flange and large

tensile strains result at the level of the prestressing tendon (shown previously in Figure 3-23). If the

girder is loaded to failure, the bottom flange crushing governs the failure mode at the support due

to negative bending.

It was found that for cantilever girders with straight unbonded cantilever tendons are not well

adapted to getting large tendon stress increases at the ultimate limit state. Although higher

curvatures are possible using high strength FRC, the rotational capacity for girders in negative

bending regions is still somewhat small. This is because large curvatures only exist over a small

plastic hinge length since the bending moment demand in negative moment regions approaches the

member capacity at a fast rate (shown in Figure 3-30). The integration of curvature over the small

plastic hinge length results in a small rotational capacity before failure. Therefore, the tendon stress

increase due to girder deformation is not significant at the ultimate limit state due to the limited

girder deformation capacity. Since the abutment location is the only critical negative bending

region, only one predominant crack (or joint opening) occurs at the abutment face to provide strain

00008-

00006-

00004-

00002-

00500.0-10.0-510.0-20.0-

normal concrete girder failure

FRC girder failure

M [kNm]

φ [1/m]

responses shown are for a constant prestress force, P

69

Page 86: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

increases for the unbonded tendons. This single crack has the largest influence on tendon stress

increase for the shortest tendon, and a diminishing influence for longer tendons.

Figure 3-30. Plastic hinging in negative moment regions

3.13.3 Example of Tendon Stress Increase for Negative Bending

An investigation was carried out to consider an alternative 90m span bridge design which is

hinged at the midspan and uses a tendon design which more closely matches the moment demand

envelope for cantilever bending. The intent of this design is to promote several joint openings at

the ultimate limit state to provide larger tendon stress increases at the ultimate limit state. The

tendon arrangement given in Table 3-9 is assumed for the cantilever system.

For simplicity, only a uniform distributed load is applied to the hinged cantilever system to

investigate the structural response at failure. As can be seen in Figure 3-31, the capacity and

demand envelopes are closely matched which causes several locations along the girder to approach

Table 3-9. Tendon unit sizes (15mm strands) used for cantilever post-tensioning

Segment No. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Strands per tendon 19 19 19 12 12 12 12 7 7 7 7 4 4 4 4

Φ [1/m]

small plastichinge length

-200000

-150000

-100000

-50000

0

50000

100000

150000

0 10 20 30 40

M [kNm]

CapacityDemand

rapidly approaching demand to the ultimate resistance in negative bending regions

Segment 1Segment 1

70

Page 87: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the ultimate resistance of the cross section. Therefore, it is observed that plastic deformation occurs

at multiple segments at the ultimate limit state to increase the stress in the unbonded tendon.

Figure 3-31. 90m span with midspan hinge when load is close to failure

Cantilever tendon stresses have been computed for each construction stage, and for short-term

loading (shown in Figure 3-32). During the construction process, it can be seen that there is an

initial decrease in cantilever stresses due to the addition of a new segment and corresponding

prestressing force. The tendon stresses reach a minimum as the effect of additional prestressing is

offset by the additional dead load moment due to the weight of the cantilever. In the last stage of

construction, the cantilever is assumed to be closed with a hinge at the midspan.

The change in tendon stresses of unbonded tendons due to girder deformations was then

examined by applying additional load to the girder after closure. It is seen that the increase in

tendon stress for curtailed unbonded prestressing is linear for the uncracked state. Following the

first cracking, the rate of change in tendon stress increases due to large local strains in the vicinity

of the crack at the level of the prestressing. Even though this system provides the most optimal

conditions for increases in tendon stress at the ultimate limit state, the average increase in tendon

stress before failure was only 0.07fpu. For the proposed box girder designed with cantilever and

-200000

-150000

-100000

-50000

00 10 20 30 40

-0.0025

-0.002

-0.0015

-0.001

-0.0005

0

0.0005

0 10 20 30 40

M [kNm]

Φ [1/m]

capacity envelope

uniform distributed load

plastic rotation at several locations

71

Page 88: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

continuity post-tensioning which allows only one segment joint to open, the average increase in

tendon stress at the ultimate limit state is negligable.

Figure 3-32. Stresses in unbonded tendons for curtailed prestressing. a) change in tendon stresses during construction due to the addition of segments and post-tensioning, b) change in tendon stresses due to the application of increasing uniform load until failure

3.13.4 Discussion

While it has been observed that the use of FRC provides a higher capacity for curvature at the

ultimate limit state, it appears that there is no great benefit in the increase in tendon stress at the

ultimate limit state. For cantilever type systems with curtailed prestressing tendons, the critical

moment location is seen to occur at the support where the bending moment is greatest. Since the

bending moment approaches the critical section at a fast rate, the region in which the plastic

rotation occurs is very small. Although the absolute curvatures that can be realized for structures

with fibre reinforced concrete is large, the total crack opening determined by integrating the strain

distribution over the small plastic region is still quite small. It appears that a cantilever system

using straight external unbonded tendons for the cantilever and continuity post-tensioning is not

well adapted to provide multiple crack opening to give larger stress increases at the ultimate limit

state. Tuning the capacity envelope provides a marginal benefit to the increase in tendon stress at

ultimate limit state; however, this is neither a practical nor a desirable approach to realistic designs

for constant depth girders.

Unbonded tendons are more efficiently and appropriately used for beams in positive moment

regions where the capacity envelope is more closely matched to the demand which allows for

multiple discrete cracks to open, providing larger strain increases in the tendon. For the proposed

box girder with straight tendons, the resulting prestress force distribution for the combination of all

prestressing steel is relatively constant and the resulting prestress moment is inversely related to

12001250130013501400145015001550160016501700

0 5 10 15 0 20 40 60 80 100w [kN/m]Stage of Construction

girder failure at w = 99.5kN/m

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

yield stress yield stress

first cracking

a) b)fp [MPa]

72

Page 89: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the applied loads to the structural system. This resulting distribution resembles that of a girder

designed using draped tendons that are anchored at the supports. An example of a draped tendon

arrangement is shown in Figure 3-33. For the draped system described by Muller et al. (1990), it

can be seen visually that tendon stress increases for this arrangement is largely due to the multiple

crack openings that occur in the positive bending region, shown by the large opened cracks at the

midspan.

Figure 3-33. Ultimate behaviour for a continuous girder using unbonded draped tendons (adapted from Muller et al., 1990)

It can be concluded that the shape of the bending moment diagram for negative bending

produces a small plastic hinge length which greatly reduces the rotational capacity of the girder.

The straight cantilever tendons which are at a constant elevation below the top flange do not

experience an appreciable tendon stress increase since the plastic rotation capacity in negative

bending is small. To this end, the proposed box girder design is not revised further to optimize the

tendon steel consumption.

3.14 Global System Considerations

3.14.1 Span-to-Depth Ratio for Constant Depth Box Girder

A range of span-to-depth ratios were considered for the design of the proposed constant depth

box girder to investigate the effect of span-to-depth ratio on material economy. The general trend

for material consumption showed that the concrete usage varied slightly for a wide range of span-

to-depth ratios but overall the concrete consumption is relatively constant for regardless of girder

depth. The reason for this is because the top flange design, which accounts for a large portion of

the total concrete usage, is governed by the design for transverse bending. The bottom slab is also

chosen to have a typical constant thickness for varying girder depths. Internal corner blisters for

73

Page 90: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

cantilever and continuity tendon anchorages must remain the same size since minimum clearance

requirements for stressing jacks must be satisfied for construction and the anchorage details are

intended to be standardized for precast construction. Therefore, only the height of the web is the

predominant factor which affects the overall concrete consumption due to varying girder depth.

Since the webs are slender, this change is relatively small (shown in Figure 3-34).

The tendon consumption varies approximately linearly with span-to-depth ratio. Since the

ultimate resistance of the cross section is essentially a the product of tendon force and internal lever

arm, the depth of the girder has a significant effect on the prestressing steel demand. As the internal

lever arm is decreased due to a shallower box girder, the tendon consumption increases.

Figure 3-34. Material consumption for varying span/depth ratios

For the 90m span under consideration, a 25:1 span-to-depth ratio was chosen as a reasonable

compromise of economy and visual girder slenderness while using efficient use of high strength

fibre reinforced concrete. Deeper box girders have increased visual prominence which detracts

from the transparency of the structure. Thinner box girders require large prestress forces which

decrease the tendon utilization economy. The need for large prestress forces also requires larger

tendon units. Due to anchorage detailing (which is discussed in Chapter 6), the maximum tendon

unit size presents a limitation for girder slenderness.

3.14.2 Range of Spans

Assuming that alternative span lengths would also be designed at a span-to-depth ratio of 25:1,

the change in total concrete consumption for varying girder depths is small since the only changing

variable for the typical section is the web depth (as shown for shallower sections in Figure 3-34).

Due to this, the uniform dead load is similar for any span length. For shorter spans, the

0.0

20.0

40.0

60.0

80.0

100.0

120.0

16 18 20 22 24 26 28 30 32 34 360.00

1.00

2.00

3.00

4.00

5.00

6.00Tendon consumption

Concrete consumption

span/depth ratio

Concrete Consumption[m3/m]

TendonConsumption[kg/m3]

74

Page 91: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

superimposed dead load is unchanged for equal bridge widths. Therefore, the effects of dead load,

superimposed dead load and long-term redistribution of dead load are proportional to L2.

The live load model applied to the box girder defined by the CAN/CSA-S6-06 is also the same

for all span lengths for equal bridge widths. Since the top flange is governed by transverse bending

due to live load, the top flange design for all span lengths does not change. The continuous slab

thickness is governed by the chosen concrete clear cover to the transverse post tensioning. Since

the proposed bridge design concept corresponds to only one single continuous span, there are no

influences due to multiple span effects. The prestressing steel arrangement may be kept similar for

alternative span lengths, and all other things being equal, the quantity of steel consumed for shorter

span lengths is expected to be scaled proportionally to the span length since:

The 90m span considered for this concept is an upper boundary for the range of spans since

deeper cross sections required for longer spans is not recommended due to the visual prominence

of the girder. The maximum depth of segments is also governed by ease of transportation as the

minimum vertical clearance for existing bridges on Ontario highways is 4.7m (Ontario Ministry of

Transportation, 2002). With the consideration for trailer bed height and protruding reinforcing steel

for the cast-in-place barrier, the maximum girder depth for segments is 3.6m.

The smallest span for this concept is anticipated to be 45m due to practical and economical

reasons. Maintaining a 25:1 span-to-depth ratio results in a 1.8m deep box girder. To allow

sufficient working space for continuity tendon stressing inside the box girder and to provide ease

of inspection, a minimum clearance must be provided for workers and inspectors to walk inside the

box girder. Also, at a span length of 45m, precast I-girders can also offer a single span solution

since I-girders can be transported in lengths up to 50m (CPCI, on-line) and may provide a more

economical solution.

Mr Mf>

PdvwL2

const-------------->

P 0.9L25-----------⎝ ⎠

⎛ ⎞ wL2

const-------------->

P L const×>

75

Page 92: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

3.14.3 Alternative Girder Designs

The choice to use a constant depth girder over a haunched girder for a light weight design

prestressed with unbonded tendons has significant advantages: a) the extra premium in dead load

due to the full depth of the web is a small increase from the haunched profile, b) a significant

increase in internal lever arm is gained at the midspan (typical 50:1 span-to-depth for haunched

girders (Menn, 1990), 1:25 span-to depth for constant depth) which provides a considerable

increase in flexural resistance for positive bending, and c) the constant girder depth facilitates

repetitive and simpler detailing for the strutted connections. Figure 3-35 illustrates the alternative

haunched girder concept, using span-to-depth proportions suggested by Menn (1990).

Figure 3-35. Concept comparison of haunched girder and constant depth girder elevation

Several girder profiles were investigated for the proposed cantilever-constructed bridge. A

constant depth girder was found to be the most structurally efficient girder shape, unlike traditional

haunched girders designed for cantilever construction. Since dead load typically contributes a

substantial bending moment demand at the pier for traditional cantilever-constructed bridges,

reducing the dead load of segments nearing the midspan relieves the flexural demands at the

support. However, for light weight girders with slender webs, reducing the depth of the girder has

an insignificant reduction of dead weight since most of the concrete volume is in the flanges. The

savings in cantilever post-tensioning due to the reduction in dead load of the cantilever do not

outweigh the additional continuity prestressing required to compensate for the loss in lever arm at

midspan. Since unbonded tendons do not experience any appreciable increase in tendon stress at

76

Page 93: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the ultimate state, the midspan flexural resistance is reduced significantly by the loss in lever arm

to the compression flange. For a haunched girder with span-to-depth ratio of 17:1 at the support

and 50:1 at the midspan, the required demand for prestressing steel was found to be approximately

15% more than for the 25:1 constant depth girder.

Unlike typical balanced cantilever construction, the supports of the proposed bridge must resist

a significant overturning moment in the final stage due to factored dead load and live load. Large

flexural demands at the support introduce large forces in the substructure due to the overturning

effect; therefore, it is advantageous to maximize the redistribution of moments to the midspan by

making a constant depth girder monolithic at the closure segment. The installation of straight

continuity tendons in the superstructure is uncomplicated and provides a straightforward and

efficient solution to resist flexural demands. It is more economical to provide a monolithic

connection at midspan to relieve negative bending moment at the abutment. The elimination of a

joint at midspan also simplifies the detailing at closure and provides added protection against

corrosion due to de-icing salts.

Since the top slab has been designed for minimum weight and minimum thickness, a strutted

connection is necessary at the cantilever tip to provide adequate strength. The strutted connection

is standardized for a constant depth girder since the length of the strut and the angle in which it

connects to the web and top flange is consistent. This repetition of details provides further economy

in the design concept. For a haunched girder, the strutted connection becomes more complicated.

Either the connection of the strut to the box frames into the bottom flange which varies in elevation

at every segment, or the strut is connected to the box at a constant elevation below the top flange.

In the latter configuration, a compression strut inside the box must be installed to resist the

horizontal forces of the external diagonal struts. In both cases, the detailing required for the strutted

connection becomes more complex and is less economical from a standardized precasting

perspective.

77

Page 94: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

78

4.0 TRANSVERSE FLEXURE

This chapter begins by describing the main objective for the transverse design, and discusses

typical transverse prestressing in existing structures. The motivation for the design of a thin top

slab is described, and the feasibility of post-tensioned or pre-tensioned transverse prestressing

alternatives are identified. Following the selection of the pretensioning concept, a grillage analysis

is selected as the method of analysis for determining the live load distribution of transverse bending

moments on the top slab. The sizing of concrete slab dimensions is designed with respect to

maximum allowable transverse bending deflections which is determined by comparing the linear

elastic deflections of a standard AASHTO-PCI-ASBI girder (Precast/Prestressed Concrete

Institute, 1997). The behaviour of the structure is described for service loads and for the ultimate

limit state. A traffic barrier design is presented which satisfies the CAN/CSA-S6-06 Performance

Level 2 (PL-2) requirements (cl. 12.4.3.2) for factored impact loading.

4.1 Introduction

The structural components of the box girder cross section that provide vertical support

reactions for the top slab are the webs which support the top flange continuously along the web

lines and the diagonal struts which support the top flange at discrete points along the cantilever

wing. Live load is applied vertically to the top slab which must carry the forces in transverse

bending to the support reactions. The bending moments in the top slab that act in the transverse

direction vary depending on the location of the vertically applied live load forces. In general, the

maximum negative moments occur above the web supports since the top flange is continuous over

the webs. For girders with a free cantilever wing and no struts, the maximum positive moments

occur between the webs where there is a large influence for positive bending. However, for the

proposed box girder with struts, the boundary conditions for the top slab are modified which

produce a significant positive bending moment between the webs and the strut connections. The

critical transverse bending effects can be determined by accounting for all possible load

configurations on the bridge deck and by taking the maximum and minimum live load distribution

envelopes for all cases.

For the given number of traffic lanes to design for, the distribution of live load in the transverse

direction can be modified by altering the boundary conditions which support the top slab.

Page 95: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Maximum positive bending moment between the webs can be reduced by moving the webs closer

to the centreline of the bridge; however, the maximum negative bending moment at the web is

increased due to this repositioning. The bending moment distribution is also affected by the relative

stiffnesses of the box girder structural components, such as the top flange stiffness and the web

stiffness. Since the top flange is rigidly supported at the web, a distribution of moments occurs due

to frame action. However, the bending stiffness of the slender webs is much smaller than the

required bending stiffness of the top flange, so very little moment is distributed into the webs. This

result is presented later using a frame analysis of the cross section. A favourable distribution of

bending moment demands is found to design the top slab for transverse bending which satisfies

strength requirements and accounts for geometrical constraints.

The design intent of the transverse prestressing is to make the top slab intensively prestressed

transversely for adequate crack control under service loads. Since it is expected that the girder will

behave elastically for SLS conditions, a linear elastic analysis of forces is appropriate for analysing

service loads and structural response.

4.2 Typical Transverse Prestressing for Segmental Bridges

Typical box girder bridges are transversely post tensioned using multi-strand flat post-

tensioning ducts. The VSL Type SO anchorage is one type that allows for the bundling of four

15mm strands placed in a flat duct for slab prestressing (VSL, 2007). The east approach girders of

the Confederation Bridge has a deck width of 11.64m that is post-tensioned using transverse

tendons containing three 15mm strands, with an even tendon spacing at 1000mm and a minimum

concrete cover to the top surface of 85mm (Straight Crossing, 1994). The advantage of flat post-

tensioning ducts is that they can be deviated in the slab vertically to provide deviation forces

opposite in direction to the applied live loads in addition to maximizing the internal lever arm for

flexural resistance. For post-tensioning, the casting operation and stressing of the tendons are

uncoupled processes.

4.3 Prestressing Concept for Light Weight Slab Design

The main objective of the deck design is to reduce the dead weight. In fulfilling this design

motivation, smaller concrete covers are assumed to be valid for the application of high-

performance FRC. The CAN/CSA-S6-06 requires that the top concrete cover for transverse post-

tensioning be mm and the soffit concrete cover for pretensioning strands of deck slabs less 80 10±

79

Page 96: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

than 300mm thick to be mm (cl. 8.11.2.2). Following these strict provisions for the use of

FRC in bridge decks results in overdesigned strength and poor efficiency of materials. For the

efficient utilization of high-performance concrete and the minimization of dead load, the transverse

design has been based on strength requirements for transverse flexure and serviceability

requirements for acceptable crack control and deflection. The concrete cover recommendations

given in the CAN/CSA-S6-06 have not been used. A top clear cover of 40mm to the pretensioning

strand and a bottom clear cover of 30mm to the pretensioning strand has been chosen for the thin

slab design using high strength concrete. High-strength concretes with low water-to-cement ratios

generally are less porous and have increased durability properties due to low permeability (Neville,

1995). The concrete mix tested for this study contained 8% silica fume of the total cementitious

material. The influence of silica fume upon the permeability of concrete is very large - a 5% content

of silica fume was reported by Khayat and Aïtcin (1993) to reduce the coefficient of permeability

by 3 orders of magnitude. In addition, the low water-to-cement ratio of silica fume concretes, using

water reducing and superplasticizing admixtures, can result in total elimination of capillary

porosity in the concrete mixture provided sufficient curing had been ensured (Mehta, 1986). With

the combination of a high-performance waterproofing membrane, smaller concrete covers may be

acceptable for design; however, a definitive proof for this assumption is not investigated in this

thesis.

4.3.1 Post-Tensioning

A typical transverse post-tensioning arrangement, using flat multi-strand slab tendons, was

originally investigated for the transverse prestressing concept. It was found that using the largest

flat slab tendon of 4-15mm strands did not provide sufficient resistance for the design of a thin

concrete section under consideration. The closest specified VSL SO6-4 tendon spacing is 352mm

centre-to-centre which is governed by the wide anchorage unit that must be cast at the cantilever

wing tip (VSL, 2007). However, the practical limitation for minimum tendon spacing is larger

because the tendons can only be placed within transverse ribs since the tendon duct diameter is

larger than what can be reasonably placed within the thin slab section.

In addition, the prestress force applied by the anchorages requires a development length for the

prestress force to spread over the width of the deck slab. Within the spreading length and between

the anchorages, there are dead zones of deck slab which are not prestressed. For a thin slab that

relies on the precompression force due to prestressing for good performance at serviceability, it is

60 5±

80

Page 97: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

necessary to provide a more uniform application of transverse prestress force along the length of

the girder. Therefore, to provide sufficient transverse bending strength and uniform prestressing at

serviceability and to avoid the use of large diameter tendons, post-tensioning is not a feasible

solution for the design of transverse prestressing in thin slabs. A pretensioning design will be

investigated for the transverse prestressing concept.

4.3.2 Pretensioning

The use of regularly spaced pretensioning strands for the design of transverse flexure in precast

segmental construction appears to be well-adapted for the design of standardized segments with

thin slabs. For the design of a thin top slab strengthened by transverse ribs, only straight

pretensioning strands are considered. Since prestressing strands are not deviated in the slab to

accommodate negative and positive flexural regions, a prestressing arrangement that results close

to a concentric prestress has been decided to be the most effective arrangement for constructability

and structural behaviour. For a prestress arrangement that is not concentric to the top slab,

secondary prestress moments will be produced due to the indeterminate structural system of the

strutted box girder. Prestressing strands are assumed to be evenly spaced in the slab, and in the

bottom of the transverse ribs to maximize the ultimate strength. The advantage of evenly spaced

strands is the uniform introduction of prestress force on the section. The top concrete cover has

been chosen to be 40mm and the bottom concrete cover is chosen to be 30mm. These covers

provided are only half of what the CAN/CSA-S6-06 requires, however this assumption is

maintained in design for high performance concrete notwithstanding the Code recommendations.

Using these assumed concrete covers, the required slab thickness including the diameter of a 15mm

strand is 85mm. A possible arrangement of pretensioning strands and the clear covers to the strands

is shown in Figure 4-1. Since pretensioning does not use any mechanical anchorage at the face of

the slab, an investigation of bond strength is required to determine the required transfer and

development length which is necessary for the prestress to be effective.

81

Page 98: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-1. Possible pretensioning arrangement

4.4 Bond Strength of 15mm Pretensioning Strands in FRC

4.4.1 Mechanisms of Bond

The bond resistance of prestressing strands embedded in concrete depends primarily on

frictional resistance and mechanical interlock. The chemical adhesion bond, if any, fails at very

small slips (Chao, 2006). A slip is the movement of the bonded steel relative to the surrounding

concrete. Due to the differential slip between steel and concrete, a shear stress due to frictional

bond is activated. Frictional bond provides initial resistance against loading and further loading

mobilizes the mechanical interlock between the concrete and strands. The helical outer wires

around the straight centre wire of a seven-wire strand are responsible for the mechanical action in

a pretensioning strand (den Uijl, 1998).

In seven-wire strands embedded in a FRC matrix, the confining effect of the fibres enhances

the frictional and mechanical interlocking significantly (Figure 4-2). As strands slip in a FRC

matrix, the pullout mechanism is formed by the helical path of the outer wires. Crack widths due

to the radial tension field for seven-wire strands are generally smaller than crack widths for

deformed bars, due to both the fibre confinement and due to the helical pullout path (since

deformed bars pull out directly). Thus, the strand exhibits higher bond strength and maintains bond

stresses for longer slip distances (Chao, 2006). A ductile stress-slip response is exhibited for

82

Page 99: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

strands embedded in FRC, which is characteristically different from the response of reinforcing

bars (Chao, 2006).

Figure 4-2. Bond mechanisms between FRC and 15mm pretensioning strand (adapted from Chao, 2006)

4.4.2 Development of Stress in a 15mm Pretensioning Strand

The initial function of bond is to transfer the prestress force from the pretensioned strand to the

concrete member. The bonded length of strand needed for such transfer is known as the transfer

length, lt. Since the concrete is cast around the strand when it is pretensioned, the strand is slightly

thinner due to the Poisson's ratio. When the strand is released after the concrete has cured, there is

a swelling of the strand along the transfer zone which gives rise to considerable radial pressure and

thus larger friction forces against slip. This increase in frictional bond is termed the Hoyer's effect.

When a member is subjected to loads in the range from SLS to the ULS, the stress in the

prestressed strand increases substantially beyond the prestress level and a bonded embedment

length beyond the transfer length is required to develop the increased steel stress. This length is

referred to as the flexural bond length, lf. The sum of the transfer and the flexural bond length is

known as the development length (Figure 4-3).

Figure 4-3. Idealized strand stress profile in a pretensioned strand under applied load (adapted from Kahn, 2002)

BondStress

Slip

without fibers

with fibers

strand pulloutdirection

twistdirection

mechanical interlocking

mechanical interlocking

friction

narrowlongitudinal

concrete crack

SteelStress

Distance from free endlt ld

transferlength

flexural bond lengthdevelopment length

increase in strand stressdue to applied loads

fps

fse

83

Page 100: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The CAN/CSA-S6-06 (cl. 8.15.4) defines the transfer length and the flexural bond length based

on the work done by Zia and Mostafa (1977):

(4-1)

(4-2)

where fsi is the stress in the prestressing strand just prior to transfer (which is typically 0.78fpu),

and fse is the effective stress in prestressing steel after losses (which is typically 0.74fpu). A

conservative estimate of development length is obtained by adopting these empirical equations

from the Code where lt = 436mm and lf = 803mm assuming fci = 59Mpa (based on 7-day cylinder

strength) and fps = fpy.

4.4.3 Improvement of Bond Strength due to High Strength FRC

Previous research (Hota and Naaman, 1995) has shown that the addition of hooked steel fibres

into a cementitous matrix can improve the bond properties of bonded steel. The presence of fibres

increases the confinement to resist the radial forces which increases the frictional bond on the

strand.

Recent studies have been carried out to investigate the increased bond properties on 15mm

prestressing strands due to the use of high-strength concrete and the influence of fibre

reinforcement. Kahn (2002) has investigated the influence of high-strength concrete ranging from

70 MPa to 100 MPa on the transfer and development lengths of 15mm pretensioning strands, and

showed that they are less than calculated by the AASHTO (1996) and ACI (1999) code provisions.

Chao (2006) has studied, among many other variables, the influence of hooked steel fibres on the

flexural bond of 15mm pre-tensioning strands for 76 MPa (11 ksi) concrete, but did not perform

tests to measure the transfer length.

It has been shown by Russell et al. (1993) that the average flexural bond stress based on ACI

(AASHTO) development equations can be calculated by solving equilibrium on a strand:

(4-3)

lt 1.5fsif'ci------db 117–=

lf 0.18 fps fse–( )db=

lf uf pps⋅ ⋅ fps fse–( )Aps=

uffps fse–( )Aps

lf pps⋅--------------------------------=

84

Page 101: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

where is the perimeter of a seven-wire strand, and is the cross

sectional area of a seven-wire strand.

Chao (2006) performed several monolithic pull-out tests for specimens containing a concrete

matrix strength of 76 MPa (11 ksi) and 1% fibre volume fraction. The composite material

containing hooked steel fibres exhibited a peak bond strength of approximately 6.89 MPa (1000

psi). Assuming that the hooked steel fibre specimen responds similarly to the concrete being

considered for design, the flexural bond length may be determined on the basis of these tests. Since

the work done by Kahn (2002) investigated only the use of high-strength concrete with no fibres,

the transfer length determined in his study is too conservative for FRC. Thus, for analysis, the

benefit due to the Hoyer’s effect will be conservatively ignored, and the flexural bond strength

determined by Chao (2006) for FRC will be used for the transfer length. Assuming a bond strength

of 6.89MPa, the strand length to develop the effective stress of 0.74fpu and the yield stress of 0.9fpu

is 440mm and 535mm, respectively (Chao, 2006). Assuming a linear increase in the effective

prestress over the transfer length, the top slab prestressing is sufficient on the inside face of the

barrier for service loads with truck loading next to the barrier. It is now reasonable to proceed with

a pretensioned design with a detailed analysis of transverse loads.

4.5 Methods of Analysis

The lateral distribution of live load forces in the girder cross section must first be determined

in order to obtain the governing transverse loads. The maximum load effect can be obtained by

applying various load arrangements on the deck and then taking the maximum and minimum

envelope for all load arrangements. The distribution of load can be obtained in several ways:

conservative two-dimensional hand calculations using Pucher charts (Pucher, 1977) or Homberg

charts (Homberg, 1968), two-dimensional frame analyses of a cross section, three-dimensional

folded plate analyses, or three-dimensional finite element analyses (LoBuono, 1997). Pucher and

Homberg charts provide elastic plate influence surfaces for a variety of load and support cases.

Most support and load combinations are available that are typical for bridge design; however, the

boundary conditions which correspond to the cantilever overhang for a strutted box girder is not

available in these charts. This case pertains to a continuous plate that is supported continuously at

the web and supported only at discrete locations at the struts. In this design, a three-dimensional

finite element model using a grillage of beam elements is used for analysis to accurately model the

boundary conditions of the top flange.

pps43---πdb= Aps

736------πdb

2=

85

Page 102: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

4.5.1 Grillage Analysis

In bridge engineering, grillage analyses have primarily been used to determine overall bridge

behaviour (Sotelino et al., 2004). The grillage analysis method is inexpensive and simple to

implement and comprehend, thus it has been favoured over finite element analyses in the field of

bridge engineering (Sotelino et al., 2004). Menn (1990) states that grillage models will yield

reasonably accurate sectional forces and deformations provided the member stiffnesses in flexure,

shear, and torsion have been properly chosen. Results of acceptable accuracy are always obtained

when the member stiffnesses are calculated from properties of the homogeneous uncracked

section.

The anticipated design of the top flange includes a thin continuous top slab that is stiffened by

discrete monolithic transverse ribs. The composite action of the slabs and ribs acting together is

similar to that of a multiple web T-girder. For the developed grillage model, the transverse beam

elements are assigned smeared sectional properties that are obtained directly from the cross

sectional geometry of a transverse strip of slab. The longitudinal beam elements are assigned

sectional properties due to a similar procedure, assuming a fixed number of longitudinal beams.

Menn (1990) provides a guideline to accurately model beam stiffnesses within a grillage model,

given in Figure 4-4.

86

Page 103: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-4. Generalized frame element properties for grillage analysis (adapted from Menn, 1990)

For the table provided in Figure 4-4, G is the shear modulus of concrete, E is the elastic

modulus of concrete, A is the shear area of the beam, K is the torsional constant of the beam, ttf is

the thickness of the top slab, and EItot is the total flexural stiffness of the whole cross section before

dividing into discrete beam elements.

A three-dimensional finite element model was developed using the SAP2000 software to

determine the transverse bending moment distribution due to lane loading on the bridge. It was

found that the slab acts primarily as a one-way slab since the ribs provide a much higher stiffness

in the transverse direction. The presence of equally spaced strut supports at the cantilever tip also

diminishes the transfer of longitudinal bending in the slab with respect to a cantilever wing without

struts. Therefore, it is not necessary to model the full length of the bridge to determine the

Stiffness Transverse Beam (TB) Longitudinal Beam (LB)

Flexural where

Shear

Torsion

Δb Δb Δb Δb Δb Δb

beam width

longitudinalcross section

transversecross section

m transverse beams

n longitudinal beams

EITB EItotm-----------≅ EILB 1

12------E ∆l ttf( )3⋅ ⋅≅ ∆l ln 1+------------=

αGATB ∞≅ αGALB ∞≅

GKTB GKTB

m---------------≅ GKLB 0≅

87

Page 104: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

transverse bending effects in the transverse direction. It was observed that modelling a short bridge

span with only 5 segments is sufficient to obtain accurate results in transverse bending for the top

slab. To provide sufficient resolution for the transverse bending moment demands due to moving

load cases, longitudinal beams were chosen to be 500mm centre-to-centre. The SAP2000 model is

illustrated below in Figure 4-5.

Figure 4-5. 3-dimensional grillage model

4.5.2 Frame Analysis

For static loads that are evenly distributed over the section such as dead load and prestress load,

a simple two-dimensional frame analysis of the cross section can be performed. By assuming a

transverse strip of box girder, sectional properties can be calculated and assigned to the appropriate

cross section components. Anchorage and deviation forces acting on the box girder due to the

prestress can be applied to the frame model as external applied loads to determine the distribution

of forces in the box girder. A frame model showing the transverse moment distribution due to a

non-concentric prestress is discussed later.

4.6 Comparative Study for Transverse Stiffness

Using the method of analysis described above, a grillage model is developed for the design of

the top slab in transverse flexure due to the application of live load. The CAN/CSA-S6-06 provides

section at strut

section between struts

longitudinal beamsspaced at 500mm

edge stiffening member along cantilever tip

transverse beamsspaced at 750mm

pin-ended strutsspaced at 3000mm

shaded area indicates web meshed withshell elements

88

Page 105: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

no specific guidelines pertaining to maximum acceptable deflections for deck slabs in transverse

bending. To provide a basis for validation, the design for transverse stiffness is carried out by

allowing similar maximum deflections as that of an existing reference structure. The reference

structure chosen for comparison is the standardized AASHTO-PCI-ASBI precast box girder

(Prestressed/Precast Concrete Institute, 1997). This reference structure allows for a direct

comparison of live load deflections for the same arrangement of design lanes; however, it is

possible that this chosen standard is conservative. A similar grillage model, making consistent

modelling assumptions, is created for the AASHTO-PCI-ASBI box girder for direct comparison

of transverse stiffness to the proposed segmental strutted box girder.

4.6.1 Cross Sections

There are two types of standard AASHTO-PCI-ASBI box girder sections - one type is intended

for span-by-span construction, and the other is intended for cantilever construction (shown in

Figure 4-6). With regard to the top slab proportions, the box girder intended for cantilever

construction has a much deeper haunch at the junction of the webs. This haunch tapers to the

typical thickness of 225mm near the quarter point of the interior slab span and approximately the

midpoint of the cantilever wing. The depth of the top slab for this design is not governed by the

required transverse stiffness of the deck slab to satisfy minimum slab deflections due to transverse

bending, rather it is governed by the spatial requirement for internal bonded tendons.

The cross section chosen for comparison is the 2400mm deep box girder intended for span-by-

span construction (Prestressed/Precast Concrete Institute, 1997). This section is chosen since the

top flange thickness for span-by-span constructed girders is not governed by the inclusion of

bonded tendons as it is with cantilever girders with internal tendons. Since the span-by-span girder

top flange is thinner, it provides a better comparison for minimum transverse stiffness. The 12m

wide standard section has an interior span of 6.54m and a cantilever span of 2.73m.

89

Page 106: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-6. Comparison of top flange thicknesses of AASHTO-PCI-ASBI standard box girder examples (adapted from Prestress/Precast Concrete Institute, 1997)

The proposed box girder has an interior span of approximately 5m from centre-to-centre of web

at the centroid elevation of the top flange. The cantilever span is approximately 3.5m from the

centre of the web to the cantilever tip. The cantilever span is continuously supported at the web and

supported by discrete struts equally spaced at 3m centre-to-centre at the cantilever tip. The slab

thickness is chosen to be 85mm and the depth of the ribs is chosen to be 200mm from the top

surface. Transverse ribs are 500mm wide and are spaced at 1500mm centre-to-centre. Figure 4-7

displays the top flange dimensions for a typical segment.

b) AASHTO-PCI-ASBI standard section for span-by-span construction

a) AASHTO-PCI-ASBI standard section for cantilever construction

90

Page 107: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-7. Cross section dimensions for typical segment

4.6.2 Material Definitions used in Analysis

The recommended minimum strength concrete for standardized AASHTO-PCI-ASBI

segmental box girders is 5000psi (34.5MPa) (Precast/Prestressed Concrete Institute, 1997).

Normal concrete of this strength corresponds to a modulus of elasticity of approximately

26400MPa (Collins et al., 1997). The standard PCI section is assumed to use the recommended

minimum strength concrete for the determination of minimum transverse stiffness. The proposed

girder is expected to be fabricated with high-strength FRC similar to the concrete cylinders tested

at the University of Toronto which were described previously in Chapter 2. The modulus of

elasticity observed in the laboratory tests for the FRC is approximately 43680MPa. Figure 4-8

displays the concrete models considered for the comparative study. A linear elastic analysis

considering only the uncracked stiffness is performed for the SLS investigation.

Figure 4-8. Assumed stress-strain material properties for investigation of transverse stiffness

08-

07-

06-

05-

04-

03-

02-

01-

0400.0-300.0-200.0-100.0-0

FRC compression cylindersf’c = 74.6 MPaEc = 43680 MPa

PCI recommended minimumstrength concretef’c = 34.5 MPaEc = 26400 MPa

fc [MPa]

εc [mm/mm]

91

Page 108: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

4.6.3 Loading

The girder considered for the proposed design has a total top slab width of 12m. The CAN/

CSA-S6-06 requires that a deck width of 10m to13m must be checked for 2 or 3 design lanes,

whichever causes the worst effect (cl. 3.8.2). This distribution of design lanes is displayed in

Figure 4-9. The 2 lane model developed comprises two design lanes side-by-side centrally located

on the bridge to maximize the positive bending moment and deflection in the interior span of the

box girder. The 3 lane model comprises a centre lane coincident with the bridge axis, and two

exterior lanes. The exterior lanes are located so that the centre of the truck tire load is 0.5m from

the cantilever tip. The CAN/CSA-S6-06 requires that the centre of the truck tire load be positioned

0.3m from the inside face of the barrier. Thus, the live load analysis assumes a barrier width is

0.2m, which is conservative.

The truck load used for design considers a CL-625-ONT truck defined in the CAN/CSA-S6-

06 (Annex A3.4). A Dynamic Load Allowance (DLA) of 0.25 was used in the analysis which

corresponds to multiple-axle loading. The CAN/CSA-S6-06 Lane Load (cl. 3.8.3.3) comprises a

3m-wide uniform load of 9kN/m superimposed with 80% of the CL-625-ONT truck load (cl.

3.8.3.2) without the DLA multiplier. Figure 4-9 displays the transverse load arrangements

considered in the analysis. The load arrangement shown is applied to both the proposed box girder

and the standard PCI girder.

Figure 4-9. Live load models for transverse load analysis

4.6.4 Deflections

Maximum live load deflections were determined through the use of a three-dimensional

grillage analysis and moving load cases corresponding to the live load models described

previously. The maximum downward deflection due to vertical applied load for the AASHTO-

PCI-ASBI was observed to be approximately 7.0mm cantilever tip and 5.9mm in the middle of the

a) 3 design lane model b) 2 design lane model

92

Page 109: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

interior span. The maximum deflection in the interior span corresponds to the two-lane model

where both traffic lanes are placed as close as possible to the centre of the bridge within the design

lanes. The maximum tip deflection of the cantilever also corresponds to the two-lane model the

traffic lanes are acting on the furthest extents of the bridge deck. The centres of the traffic wheels

should be placed 0.3m from the face of the barrier (cl. 3.8.4.3).

Figure 4-10. Vertical slab deflection envelope due to live load, AASHTO-PCI-ASBI box girder

Strictly using the transverse and longitudinal beam properties for the typical ribbed cross

section yields large deflections at the cantilever tip between the struts. As the thin top slab is

flexible in the longitudinal direction, the slab transfers load primarily in the transverse direction

but large deflections occur for loads applied near the cantilever tip between the strut supports.

Therefore, the inclusion of a stiff longitudinal member at the cantilever tip improves the flexural

stiffness in the longitudinal direction significantly. The edge stiffening beam is given dimensions

250mm thick by 400mm wide (shown in Figure 4-11).

Figure 4-11. Edge beam stiffening

For a girder with no continuous edge stiffening beam, the slab deflection due to live load at the

cantilever edge between struts is observed to be approximately 8mm (shown in Figure 4-12).

However, with the edge stiffening beam incorporated into the cantilever tip, the live load

deflections are reduced markedly between struts from 8mm to approximately 1.5mm. The

deflection envelope for the improved system is shown in Figure 4-12 below. For the proposed box

Max/Min deflection

8-

4-0

4Δ [mm]

93

Page 110: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

girder, the maximum vertical deflection at the middle of the interior span is 7.2mm and the

maximum deflection on the cantilever wings is 3.5mm.

Figure 4-12. Vertical slab deflection envelope due to live load of proposed box girder: a) thin slab stiffened with transverse ribs only, b) thin slab stiffened with transverse ribs and longitudinal edge beam

The maximum deflection of the proposed box girder is comparable to that of the standard

AASHTO-PCI-ASBI box girder. Therefore, based on this reference structure, the transverse

bending deflections of the proposed box girder is acceptable.

4.7 Prestress Moments

The prestressing arrangement for the transverse design includes equally spaced 15mm strands

placed in the slab with 40mm top clear cover, and additional 15mm strands equally spaced in the

rib with 30mm bottom clear cover. This arrangement of steel was shown previously in Figure 4-1.

The resultant of the prestress force applied to the section is not exactly concentric, so a prestress

moment is induced in the transfer length of the strands. The prestress moment causes negative

curvature at cantilever tip and induces a compressive reaction on the struts since the struts provide

a redundant reaction to the indeterminate system. This arrangement of steel counteracts the live

load bending moments in the cantilever wing when load is placed on the cantilever, and it provides

a clamping force on the strut in the unloaded condition after fabrication. Since the strutted box

girder is a statically indeterminate frame system, the secondary prestress moments due to slab

prestressing must be determined to evaluate the total moments on the slab section. Assuming a

large deflections between strutsMax/Min between strutsMax/Min at strut

8-

4-0

4Δ [mm]

8-

4-0

4Δ [mm] improved stiffness

at the cantilever tipMax/Min between strutsMax/Min at strut

a) no edge beam

b) with edge beam

94

Page 111: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

linear elastic response, a frame analysis is conducted to determine the secondary prestress moments

due to a unit load. The prestress moment is assumed to vary linearly over the transfer length.

Figure 4-13. Moment distribution due to non-concentric prestress applied at the transfer length

4.8 Service Loads

Using the live load demands from the grillage analysis, and the applied prestress moments from

the frame section analysis, the prestress design can be determined to resist service loads. By

placing 25-15mm strands equally spaced at 120mm c/c in the slab, and a total of 12-15mm strands

equally spaced at 60mm c/c in the ribs, adequate capacity is obtained for transverse flexure. The

transverse design is not fully prestressed for service loads. For the interior span, the maximum

moment exceeds the cracking stress at the extreme fibre of the cross section which is at the

elevation of the bottom fibre of the transverse ribs. Cracking will occur within the inside of the box

girder where the concrete is protected from atmospheric conditions, and the crack depth is

superficial since the cracking stress, fcr = 5.15MPa, is only exceeded within the bottom 9.3mm of

the transverse rib. The CAN/CSA-S6-06 limits the prestressing steel stress range to 125MPa to

satisfy fatigue requirements (cl. 8.5.3.2). The maximum and minimum stresses experienced by the

bottom prestressing strands at the critical bending location causes a stress range of only 83.7MPa;

therefore, fatigue does not present a problem for the tendon design. The total moments on the

transverse section for serviceability are shown in Figure 4-14. The transverse design based on

serviceability performance has been found to be adequate for ultimate limit states.

-0.120Mp

-0.265Mp

-0.385Mp

Mp Mp-0.265Mp

-0.385Mp

-0.024Mp -0.024Mp

lt lt

MpMp

a) applied prestress moment b) bending moment distribution

95

Page 112: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-14. Total transverse bending moments for transverse flexure

The linear elastic stress ranges in the thin slab due to the total moments applied to the cross

section are displayed in Figure 4-15. The maximum stress experienced by the slab occurs at the

bottom face of the slab at the web support, where the compression stress is 38.4Mpa. The average

transverse stress in the top slab due to transverse prestressing is 15.6Mpa, assuming a tendon stress

of 0.6fpu to account for long-term losses.

dead load

bottom fibre cracking

live load

prestress

250

200

150

100

50

0

-50

-100

-150

-200

-250

M [kNm]

96

Page 113: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 4-15. SLS top slab stress ranges at the extreme fibre due to service loading

4.9 Cross Section Design and Layout

4.9.1 Transverse Rib Proportions

The typical transverse cross section for one segment is shown in Figure 4-16. The layout of ribs

has been chosen to include one central rib of 500mm in width, and two exterior ribs at the segment

faces where are 250mm in width. This geometry facilitates a local thickening at the joint where

shear keys may be incorporated in the bulkhead detailing. Also, the layout allows for a seamless

transition between segments when constructed in the final state since the two exterior ribs placed

-40

-30

-20

-10

0

10

-40

-30

-20

-10

0

10

a) top fibre transverse stressesfc [MPa]

fc [MPa]

c) cross section

-22.6MPa -15.1MPa

-0.32MPa

-14.4MPa

-38.4MPa

b) bottom fibre transverse stresses

-5.15MPa

-29.6MPa

-0.63MPa

97

Page 114: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

together at the joint form an equivalent 500mm rib. The visual appearance in the final state is

evenly spaced ribs of 500mm in width, spaced at 1500mm centre-to-centre.

Figure 4-16. Longitudinal section for a typical segment

4.9.2 Web Spacing

The maximum and minimum transverse bending moments that have been shown in Figure 4-

14 indicate that the web spacing of the box girder can be further optimized since there is a

substantial reserve capacity before cracking in negative bending, and a smaller reserve capacity in

positive bending. By shifting the location of the webs towards the centreline of the bridge, the

maximum positive moment at middle would be decreased and the maximum negative moment at

the webs would be increased. This alternative was considered; however, for the span of 90m a

practical limitation governs the design of the web spacing. As the girder is constructed by the

cantilever method with 3m length segments, 15 pairs of tendons are required to be placed inside

the box girder at the support. All cantilever tendons are desired to be placed in one layer below the

top flange to maximize the internal lever arm. To sufficiently guide the cantilever tendons inside

the box girder, a minimum width of the box is required at the elevation of the cantilever tendons.

For internal bonded post-tensioned tendons, the minimum clear distance between tendons is

40mm (cl. 8.14.2.2.2). For external unbonded tendons, the minimum clear distance has been

chosen to be 60mm since high-density polyethylene (HDPE) sheathing is required to pass through

rigid steel pipe at tendon guide locations. The web spacing requirement to fit all cantilever tendons

in the box is approximately 5m (Figure 4-17). For shorter spans with less cantilever tendons, the

98

Page 115: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

interior span and cantilever span of the top flange can be optimized better with respect to the

transverse bending moment distribution due to live load.

Figure 4-17. Web spacing requirements

4.10 Design for Barrier Impact Loads

The barrier design provided satisfies the ultimate limit state barrier impact for a Performance

Level 2 (PL-2) barrier defined in the CAN/CSA-S6-06 (cl. 12.4.3.2). The detail adequately ties the

barrier into the structural system with sufficient ultimate capacity. The barrier detailing provides a

feasible flow of forces that is described by the truss model in Figure 4-18.

Figure 4-18. Truss model for basis of barrier reinforcement

The moment produced by the horizontal impact load on the barrier is assumed to be taken

primarily by flexure of the transverse ribs. Between the ribs, the concentrated moment due to

impact is transferred to the transverse ribs by torsion in the edge beam. The torsional resistance of

the edge beam is provided by the closed stirrup geometry of the barrier reinforcement (Figure 4-

19). The barrier should be cast to encapsulate the edge of the deck slab, forming a drip nose at the

edge of the deck. A similar detail for barrier design was used for the Confederation bridge. This

detail provides concrete cover to protect the pre-tensioning strands against corrosion and exposure

a) inclined webs to increase interior span b) smaller interior span insufficient

F

99

Page 116: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

to de-icing salts. The drip nose also offers a better solution compared to the typical drip groove for

proper water management to protect the superstructure.

Figure 4-19. PL-2 barrier detail for impact loading

15M @ 250required for impact resistance

Cross Section

drip nose detail for properwater runoff and end protectionfor pretensioning strands

100

Page 117: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

101

5.0 SHEAR

This chapter begins by briefly discussing a background on strut-and-tie models. Section 5.2

describes funicular load path of compression stresses in a traditional prestressed concrete beam and

describes the funicular load path for the proposed segmental girder with curtailed prestressing

tendons. Section 5.3 describes one possible method of decomposing the resultant compression

stress in the girder into a parallel chord truss model. Section 5.4 describes an alternative method of

decomposing the resultant compression stress in a girder to account for the inclined spreading of

forces which contribute to the arching effect of carrying shear forces. Section 5.5 describes the

preliminary shear model chosen for design of the web reinforcement in the proposed box girder.

Finally, the designed web reinforcement is presented. For the design of web reinforcement, there

are two components to be considered: a local spreading of forces due to the intermediate anchorage

of external unbonded tendons, and the global spreading of forces to safely transfer the applied

vertical forces to the supports. Within this chapter, only the global spreading of forces within the

web are discussed. The design for web reinforcement due to local spreading at anchorages is

discussed later in Chapter 6.

5.1 Introduction to Strut-and-Tie Models

The strut-and-tie model, originally introduced by Ritter (1899), is a powerful engineering tool

developed to describe the complex flow of forces in a simple way for cracked reinforced concrete

beams (Figure 5-1). Strut-and-tie models comprise an arrangement of discrete compression struts

and tension ties connected at pin-ended nodes to describe a state of internal force equilibrium.

Compression struts and tension ties are representative of physical concrete elements and

reinforcing steel, respectively. This simple engineering model provides a great benefit to the

designer as it gives physical significance to the calculation of forces in concrete structures. Truss

models provide a visual flow of forces and can be applied to a wide range of multi-dimensional

geometries for varying load conditions (Marti, 1985).

Page 118: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-1. 45 degree truss model (adapted from Ritter, 1899)

Schlaich et al. (1989) provided a comprehensive treatise to the application of truss models in

reinforced concrete and prestressed concrete structures. In the uncracked state, when elastic

concrete strains are small, a concrete girder reinforced with bonded prestressing steel behaves like

the system with unbonded tendons since there is no significant local increases in tendon stress due

to bond action (shown in Figure 5-2). For any cross sectional shape, the resultant compressive

force must assume the funicular shape of the net load to satisfy internal equilibrium (Gauvreau,

1993). After the formation of a crack, bonded reinforcing steel bridging the crack increases in

stress as the mechanical bond is activated. In truss models, this increase in steel stress is

represented by a tension tie at the level of the bonded steel which is equilibrated by diagonal

compression struts and vertical tension ties that connect to the compression chord (shown in

Figure 5-2). The effect of these diagonal and vertical forces on the compression chord is a

modification to the funicular shape to equilibrate the internal forces.

Figure 5-2. Possible truss models for girders with unbonded and bonded prestressing steel (adapted from Gauvreau, 1993)

For structural systems with unbonded steel and no bonded longitudinal reinforcement, in which

tendons are attached only at discrete anchorage locations, there is no mechanical bond present to

provide a local increase in tendon stress at the crack. Since the sectional compatibility of strain in

Q Q

P P

Q Q

Q Q

P P

Q Q

a) beam reinforced with unbonded prestressing only

b) beam reinforced with longitudinal bonded steel

cracked regionuncracked uncracked

tension tie due to bond

resultant funicular shape

102

Page 119: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the concrete and strain in the steel are uncoupled for structures with unbonded tendons, there is no

mechanism for a local transfer of axial force at any location away from the anchorages of an

unbonded tendon. Given this, the classical two chord truss model that depicts an increasing tension

force in the tension chord due to bond action can not reasonably explain the force equilibrium in a

girder prestressed only with unbonded tendons.

5.2 Funicular Load Path for Girders Prestressed with Curtailed Tendons

One unique characteristic of prestressed segmental concrete bridges built by the cantilever

method is the curtailed arrangement of prestressing tendons. As the intrinsic nature of the

construction scheme demands, prestressed tendons must be anchored at each stage of construction

when a new segment has been placed. The implication of this tendon arrangement is that the axial

prestress force in the girder is not constant along the length. To the authors knowledge, this type

of structural system is not extensively described in the literature as it is with prestressed girders

with a constant prestress force.

As a system with curtailed external unbonded prestressing is uncommon, it is easiest to first

envision the funicular load path through the simple familiar example of a simply supported beam

(Figure 5-3). The funicular shape of the compression chord illustrates that the horizontally applied

prestress force is inclined due to the deviation of this force at the support. The inclined compression

force in the concrete provides capacity for the girder to transfer shear from the applied loads to the

support reaction. The implication of the inclined force displayed by this simple model is that there

is no requirement for vertical reinforcing steel in the girder.

To compare the effect of a varying prestress force along the length of the girder, the simply

supported girder is examined for two cases: full-length unbonded tendons, and curtailed unbonded

tendons. It can be seen that the compression force that is applied by the intermediate anchorages at

the level of the prestressing steel must spread through the web and resolve at the compression chord

of the funicular load path. This flow of forces presents a load path that is fundamentally different

from the case with full-length tendons as it requires vertical steel reinforcement to satisfy

equilibrium. This vertical reinforcement is not required for the action of shear, rather it is required

to provide adequate spreading of forces to move the concentrated load applied at the anchorage to

the funicular compression chord. The vertical steel reinforcement required for the anchorage of

curtailed unbonded tendons is analogous to corbel reinforcement that is required to tie a heavy

point load into a support. The accumulation of prestress forces applied at the anchorages increase

103

Page 120: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the compression force in the compression chord at node locations where the diagonal compression

members meet the compression chord. Since all girders shown in Figure 5-3 have the same

dimensions and Q is the same for all girders, the force in the compression chord at midspan must

be equal to P to provide necessary member resistance.

When comparing the curtailed unbonded prestressing arrangement to the case with bonded

prestressing steel (shown in Figure 5-3b and Figure 5-3c), the models display a similar flow of

forces to satisfy internal equilibrium. In both cases, diagonal compression struts form due to the

increase in steel force at the level of the prestressing tendons which must be equilibrated with

vertical reinforcing steel in the web. One major difference between these models is that there is no

tension chord at the level of the prestressing steel for the case with unbonded tendons since there

is no local bond action between steel and concrete.

Figure 5-3. Alternative girder designs to resist applied load Q

Another major difference between the case with bonded steel and unbonded tendons is that the

funicular shape of the compression chord for the case with unbonded tendons is dependent on the

arrangement of horizontal prestress forces. For the case with bonded steel, the elevation of the

compression chord is chosen for the cracked region of the beam; however, for the case with

unbonded tendons the elevation of the compression chord is defined by the resultant location of

compressive force in the section. This concept is illustrated in Figure 5-4 for two different girders

with a different arrangement of prestressing forces. It can be seen that the shape of the compression

Q Q

0.66P 0.66P

Q Q

c) beam reinforced with longitudinal bonded steelincrease in member resistance due tolocal increase in steel stress

Q Q

0.66PQ Q

b) beam reinforced with curtailed unbonded tendons

0.21P 0.13P 0.66P0.21P0.13P

increase in member resistance due to intermediate anchorage force

Q Q

P P

Q Q

a) beam reinforced with full-length unbonded prestressing tendons

resultant funicular shape

104

Page 121: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

chord, or ‘spine’, is modified for a girder with different applied anchorage forces for the same

vertical applied load Q. Note that for both girders to provide the same member resistance at

midspan, the compression chord at the midspan has a compressive force of P. The truss model for

this girder is statically determinate, and the geometry of the compression chord can be computed

by solving for required forces and strut angles through the use of method of joints for trusses or

through the use of graphic statics.

Figure 5-4. Effect of prestress arrangement on funicular compression chord

The extension of this concept to the proposed 90m cantilevered girder with curtailed cantilever

and continuity prestressing tendons is given in Figure 5-5. During the cantilevering process, the

compression spine is developed through the progressive building of segments and anchorage of

tendons. As it was with the simply supported case, the geometry of the truss model is not constant,

rather the geometry is variable depending on the prestress and applied forces. Thus, the shape of

the truss model is unique for every stage of construction and for all combinations of applied load.

Uniformly distributed dead load is assumed to be lumped into discrete point loads based on the

tributary width and positioned at anchorage locations.

For the continuous girder in the final stage (after stressing continuity tendons), the structure is

statically indeterminate due to continuity at midspan. To solve for the truss geometry, the total

bending moment at midspan must first be computed using a frame analysis to provide the boundary

condition for the truss model at midspan. The elevation of the compression spine at midspan can

be calculated by determining the eccentricity of the resultant compression force from the neutral

axis of the girder. This is calculated by dividing the total bending moment at midspan by the

effective prestress force in the girder at midspan. Once this boundary condition is solved, the truss

geometry can be computed using the same methods as previously performed for the simply

Q Q

0.81PQ Q

b) beam reinforced with alternate arrangement of curtailed unbonded tendons

0.10P 0.09P 0.81P0.10P0.09P

spine

Q Q

0.66PQ Q

a) beam reinforced with one arrangement of curtailed unbonded tendons

0.21P 0.13P 0.66P0.21P0.13P

spine

105

Page 122: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

supported girder. A half-span elevation of the funicular truss model immediately after installation

of the continuity tendons is shown in Figure 5-5b.

Figure 5-5. Funicular shape of compression spine for continuous cantilever girder with curtailed tendons

Assuming a symmetrical uniform distributed load is applied to the continuous girder system,

the ultimate limit state for the chosen prestress arrangement is shown in Figure 5-5c. A rotation of

the compressive ‘spine’ about the inflection point is observed. The limiting mode of failure is

crushing of the bottom flange at the abutment due to negative flexure (assuming sufficient web

AbutmentQ/2 Q Q Q Q Q/2

Cross Section

a) Construction stage - 5 segments cantilevered from the abutment

b) Half span elevation with all cantilever and continuity tendons installed

Q/2 Q Q Q Q Q Q Q Q Q Q Q Q Q Q Q/2Abutment CL Midspan

Pspine

Pspine

P1t P2t P3t P4t P5t P6t P7t P8t P9t P10t P11t P12t P13t P14t P15t

P6b P7b P8b P9b P10b P11b P12b

P1t P2t P3t P4t P5t

V

Pspine

V

V

1.0

c) Half span elevation at ultimate load

Qu/2 Qu Qu Qu Qu Qu Qu Qu Qu Qu Qu Qu Qu Qu Qu Qu/2Abutment CL Midspan

Pspine

Pspine

P1t P2t P3t P4t P5t P6t P7t P8t P9t P10t P11t P12t P13t P14t P15t

P6b P7b P8b P9b P10b P11b P12b

(Vs/Vu)

d) Transverse demand envelope for reinforcing steel at the ultimate limit state shown in (c) { { Cantilever tendons govern Continutity tendons govern

Jump caused by change in tendon unit at P2t Change in tendon

unit at P9b

Crushing of bottom flange

106

Page 123: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

reinforcement has been provided to deviate the prestress forces). The transverse demands given by

the model combine an influence of the applied loads and the prestress forces. For the tendon

arrangement assumed in this shear example, the prestress tendon units increase in size for segments

that are nearer to the abutment. This change in prestress force is reflected in the transverse demand,

since a larger tension tie is required to pull the applied prestress force into the section at the location

of the resultant force (Figure 5-5d).

One major deficiency in simple truss models that consider only the resultant funicular

compression force in the concrete is that there is total disregard for the cross sectional shape of the

girder. For the examples discussed above, all truss models have been determined purely on the

basis of graphic statics and equilibrium, and these models give no consideration for the spreading

of forces into predominant flanges of the box girder cross section. To achieve a more realistic

representation of the location of forces within the cross section, the resultant funicular compression

chord should be decomposed into a more refined model having two or more chords that provide a

better representation of the location of forces within the concrete section.

5.3 Parallel Chord Truss Models

One method of decomposing the resultant force in a concrete section with known flexural

compression and tension regions is to divide the resultant compression force into two parallel

chords forming a classical truss with a horizontal upper and lower chord with diagonal struts and

vertical ties in the web. For box girders and other members with predominant flanges, the chords

are typically chosen to be located at the flange elevations. A shortcoming of this chosen geometry

is that the benefit of arching action due to the continuous inclined forces following the funicular

shape is dismissed since the horizontal chords do not provide any component of force in the vertical

direction to resist shear. Rather, shear in the girder is resisted through the discontinuous load path

of diagonal compression struts and vertical tension ties in the web. This model provides one valid

explanation for the possible flow of forces in a girder to resist shear, but it does not account for the

benefit of arching action in girders to resist shear.

Schlaich et al. (1989) have presented a possible flow of forces for flanged sections prestressed

with unbonded tendons where the top and bottom chord are fixed in position at the level of the

flanges (Figure 5-6). According to Schlaich et al. (1989), the truss model consisting of a top and

bottom compression chord already develops when the resultant force is within the kern zone of the

girder section due to the large concentrated forces in the flanges. For the fully prestressed section

107

Page 124: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

shown in Figure 5-6, both the top and bottom chord are in compression for regions where there is

no flexural cracking. As the forces in the truss must resolve to the funicular load path shown, a

change in force must take place in the top and bottom chord. Unlike a classical truss model which

depicts a changing tension force in the bottom flange due to bond action, Schlaich et al. (1989)

demonstrates that a change in force of the bottom chord is also possible through the exchanging of

compression forces between flanges.

Figure 5-6. Strut-and-tie model for a fully prestressed flanged section. a) simplified model; b) through d) detailed model of web, top flange and bottom flange, respectively (adapted from Schlaich et al., 1989)

The required web steel that is determined for parallel chord models depends on the angle of

diagonal compression struts in the web. Steeper angles of compression in the web result in larger

vertical forces in the tension ties and shallower angles of compression result in smaller vertical

forces in the tension ties. For non-prestressed concrete members, Menn (1990) recommends that

the constant angle of compression in the web can be chosen by the designer to be between and

, which is within of the principal stresses for the elastic stress state. This choice of angle

for web compression stress is consistent with current Swiss design practice which is based on

plasticity (SIA 262, 2003). Since the plastic redistribution of forces may require large crack

openings to occur at the ultimate limit state, the shear capacity is determined purely on the shear

resistance of the steel reinforcement.

Centroid of forces coming from the web

Q

P P

Q

Q

P P

Q

P P

a)

b) web

c) top flange

d) bottom flange

30°

60° 15°

108

Page 125: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Alternatively, the Modified Compression Field Theory (MCFT) provides a rational method to

predict the angle of diagonal compressive stresses, θ, from the strain conditions in the web (Bentz,

2006). Through the use of this method to determine θ, a variable-angle truss model can be

developed to more accurately describe the state of stress in the web (shown in Figure 5-7). This

model displays a varying angle of compression stress in the web which is equilibrated by vertical

tension ties. The pattern of diagonal forces in the web is regular everywhere except near the

supports and at concentrated loads (Gauvreau, 1993). In these regions, this local flow of forces

which have a radial spreading of compression stress are called compression fans (shown in

Figure 5-7). The diagonal compression struts require a change in force in both the top and bottom

flange. For the case with internal bonded steel, the bottom flange can take a tension force due to

the action of bond. For the case with only external unbonded tendons, the bottom flange can only

take compression stresses; therefore, this model is only valid in regions where the full depth of the

section is compressed and an exchange of compression forces between flanges can take place, as

shown in the previous model by Schlaich et al. (1989) in Figure 5-6.

Figure 5-7. Variable angle truss model (adapted from Collins et al., 1997)

Shear design using the MCFT is not based on plasticity, but rather it predicts an ultimate

resistance that reflects the actual state of deformation in the web. In this way, the shear resistance

can be determined due to the reistance provided by web reinforcement and aggregate interlock of

the concrete at a crack. The shear resistance of aggregate interlock is significant when the crack

widths in the web are small, and it diminishes as the crack widths in the web increase. Therefore,

the design of web reinforcement must correspond to the diagonal stress state that is related to the

shear resistance of aggregate interlock in the concrete. This requirement provides a method to

determine the shear capacity of reinforced concrete sections, but prevents the designer from

choosing the angle of web stress for the ultimate limit state. The CAN/CSA-S6-06 shear provisions

follow a design approach using the MCFT which is presented in Appendix C.

The parallel chord model used for the CAN/CSA-S6-06 shear provisions does not account for

an arching contribution due to the flow of inclined compression forces following the funicular

shape, but rather it accounts for the presence of prestressing by simply offsetting the force in the

θ compression fan

109

Page 126: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

tension flange by including the compression force due to all prestressing tendons that are on the

flexural tension side of the girder. Alternative models exist that do consider the arching effect of

compressive stress through a modified decomposition of the funicular compression chord. These

model types will be described next.

5.4 Arching Models

The decomposition of the resultant funicular compression chord into the parallel chord model

involved predetermining the distance between the chords which means that the magnitude of force

in the chords must change to satisfy equilibrium. Another possible way of decomposing the

resultant compression chord is to split the resultant force into two compression chords with

predetermined values of compression force for each chord. By setting the force in each of the

chords, the elevation and slope of the compression chords must not be fixed at the flanges to satisfy

equilibrium, rather the geometry of the compression chords must be positioned at elevations that

are statically equivalent to the resultant force.

It is possible to show this spreading of forces through the use of a two-chord model that

demonstrates the bottle shaped flow of forces due to applied point loads. Schlaich et al. (1989) have

presented this possible flow of forces for a rectangular shaped cross section (Figure 5-8). This

model gives account for the benefit of arching action in concrete due to the continuous inclined

compression stresses in the chords. The vertical applied loads to the girder are transferred to the

reactions by the inclined component of force in the compression chords. The model also

demonstrates that vertical web reinforcement is necessary to provide tie forces for the compression

chords to deviate from the line of resultant force.

Figure 5-8. Arch spreading of forces for a rectangular section (adapted from Schlaich et al., 1989)

F

P P

Fa)

F

P P

Fb)

110

Page 127: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Although this model illustrates the arching effect of inclined compression stresses in the

concrete due to prestress forces applied at the girder ends, Schlaich et al (1989) give no explicit

discussion on how to choose the chord geometry of this spreading model.

5.4.1 Tied Arch Model (Noshiravani, 2007)

Noshiravani (2007) has recently performed a study that describes a rational approach to

defining the geometry of a similar spreading model for concrete girders prestressed with external

unbonded tendons without any bonded continuous longitudinal reinforcement. By definition, the

external unbonded tendons must have a constant force between the anchorage locations. Thus,

since there is no mechanical bond between steel and concrete, no local increase in steel stress can

occur in the prestressing tendons. Therefore, in her analysis, the flow of forces in the girder does

not include a bottom chord at the elevation of the tension flange.

The sectional model proposed by Noshiravani (2007) involves splitting the resultant

compression force into two chords, giving each chord an equal magnitude of force. The geometry

of the tied arch model is determined on the basis of linear elastic stress diagrams at several

locations along the length of the girder. Assuming the plane section hypothesis, Noshiravani

(2007) locates the elevation of the compression chords by analyzing several planes of strain at

various sections of the girder and calculates the centroid of force for each chord. Schlaich et al.

(1989) originally proposed a two-chord arching model for the use with rectangular sections only;

however, Noshiravani (2007) extends the application of this concept of a two-chord arching model

to other sectional shapes including predominant flanges. Since sectional force is determined as a

product of the cross sectional stresses and the concrete area of the girder, the shape of the cross

section can be accounted for using this method of plane sections. An illustration of the cross

sectional influence on the chord geometry and the resulting transverse demands are shown below

in Figure 5-9.

Noshiravani (2007) applies the tied arch model between the supports within the B-region of the

simply supported girder. Outside of the support within the D-region of the anchorage zone,

Noshiravani (2007) provides a simple strut-and-tie model which satisfies St. Venant’s principle.

At the support, the elevation of the compression chords are determined on the basis of the elastic

stress diagram which provides the boundary conditions for the truss model within the anchorage

zone. Providing boundary conditions for D-regions based on elastic stress diagrams is typical for

the general design of truss modelling (Schlaich et al., 1989).

111

Page 128: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

.

Figure 5-9. Comparison of tied arch model for a rectangular section and a flanged section (adapted from Noshiravani, 2006)

This new application of a two-chord arching model based on the elastic stress diagram using

plane sections for the B-region provides a straightforward method to determine the shape of the

flow of forces to describe spreading in a girder. However, to provide a complete consideration for

the possible flow of forces in a prestressed section with wide flanges, it is of interest to compare

the results of Noshiravani (2007) to an alternative parallel chord model. On this basis of

comparison, a design model for the proposed box girder may be decided.

For the experimental study performed by Noshiravani (2007), the girder was loaded in static

four-point bending by means of a force-controlled machine head. The load on the girder was

increased incrementally at several load stages to assess the condition of the girder. Strain gauge

measurements in the vertical stirrups along the length of the girder and crack formations in the

webs of the girder were recorded at each load stage. The two load stages under consideration

correspond to: a) the load corresponding to the occurrence of flexural cracking in the bottom flange

of the girder (Load Stage 4), and b) the load corresponding to a load close to the ultimate strength

of the girder (Load Stage 7). The dimensions of the girder tested are shown in Figure 5-10 and the

typical cross section is shown in Figure 5-11. For a full description of the experimental program,

the reader is referred to Noshiravani (2007). The strain gauge measurements of the transverse steel

(Vs / Vu)

Flanged sectionRectangular section

1.0

P P

Q

P P

Q

CompressionTension

Model applied only in B-regionD-region D-region

boundary conditions for D-region provided by model

112

Page 129: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

and the predictions of this two-chord model for these two load stages are presented in Figure 5-12

and Figure 5-13.

Figure 5-10. Girder dimensions (adapted from Noshiravani, 2007)

Figure 5-11. Girder cross section and web reinforcement (adapted from Noshiravani, 2007)

Figure 5-12. Tied arch model for load stage corresponding to flexural cracking (adapted from Noshiravani, 2007)

D4 Cold Drawn Wire Propertiesfy = 530MPaAs = 24.2mm2

05

1015

0200400600800

1000

εv [mm/m]

fv [MPa]

tied arch model predictionactual transverse stress

unaccounted demand

c) Calculated transverse stress and tied arch model prediction

b) Measured transverse strains in vertical stirrups

a) Noshiravani tied arch model

320kN 320kN

2835kN 2835kN

320kN 320kNindicated stirrup yield

yield strain

yield stress

113

Page 130: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-13. Tied arch model for load stage corresponding to ULS (adapted from Noshiravani, 2007)

The observed relationship between the predicted transverse stress of the tied arch model and

the actual transverse stress measured during the test for the load stages shown in Figure 5-12 and

Figure 5-13 show that the tied arch model underestimates the transverse demand for steel. This is

especially evident at ULS where the model predicts virtually no transverse stress at a location

where the stirrups are yielding. One possible explanation for this discrepancy is provided by

Schlaich et al. (1989) where top and bottom parallel compression chords develop for flange

sections when the resultant force is within the kern zone of the girder section with predominant

flanges. To investigate the possibility of this alternative flow of forces, a comparison of

Noshiravani’s (2007) test results will be performed in relation to the parallel chord model described

by Schlaich et al. (1989).

Since the girder has been already designed by Noshiravani (2007), it is possible to compare the

crack patterns of the test to the crack direction that would be predicted by a parallel chord model

assuming yielding of the stirrups at ULS (which was observed in the test). Based on the CAN/CSA-

S6-06 equation for unfactored shear capacity based only on the capacity of the stirrups (Equation

C-3 in Appendix C), the expected web crack angle can be determined by:

tied arch model predictionactual transverse stress

unaccounted demand

05

1015

0200400600800

1000

εv [mm/m]

fv [MPa]

c) Calculated transverse stress and tied arch model prediction

b) Measured transverse strains in vertical stirrups

a) Noshiravani tied arch model428kN 428kN

3585kN 3585kN

428kN 428kNindicated stirrup yield

no strain gauge reading

yield strain

yield stress

114

Page 131: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

(5-1)

For the 2-leg D4 deformed wire stirrups equally spaced in the webs at 100mm centre-to-centre,

the angle of web cracks expected for the load stage at flexural cracking and the load stage near

failure is and , respectively. An alternative parallel chord model and the corresponding

transverse stress demand for the two load stages are presented in Figure 5-14 and Figure 5-15.

Comparing the crack patterns observed during the test to the parallel chord predictions for the

diagonal compressive stresses in the webs shows a remarkable similarity. Since the shear capacity

of concrete has been neglected for the prediction of web crack angle, it is expected that the

predicted angle is slightly shallower than the observed crack angle. For the two load stages

considered, the observed web crack angle is within of the predicted angle at flexural cracking,

and within of the predicted angle near ULS. Based on these observed crack patterns and the

measured stress readings in the vertical steel, the girder tested by Noshiravani (2007) appears to be

symptomatic of the flow of forces described by the parallel chord model proposed by Schlaich et

al. (1989) within the central region of the shear span. However, fan cracking in the vicinity of

concentrated vertical loads was not observed in this test. In these regions where there is flexural

cracking, it seems likely that arching action exists which decreases the demand for transverse steel.

There remains some uncertainty to explain the flow of forces in girders with unbonded tendons;

however, further truss model development is beyond the scope of this work.

Figure 5-14. Crack pattern and alternative parallel chord model for load stage corresponding to flexural cracking

θAvfydv

Vs-----------------⎝ ⎠⎛ ⎞atan=

20° 15°

a) Crack pattern and alternative parallel chord model

θ = 20˚

web crack 21˚320kN 320kN

2835kN 2835kN

320kN 320kN

compression fan

compression fandetailed trusssimplified truss

0200400600800

1000fv [MPa]

parallel chord model designactual transverse stress

conservative capacityin compression fan regions

b) Calculated transverse stress and alternative parallel chord design

yield stress

115

Page 132: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-15. Crack pattern and alternative parallel chord model for load stage corresponding to ULS

It appears that a parallel chord model consisting of two compression chords for fully

compressed girder regions provides a lower bound solution for the design of transverse steel in the

web to resist shear forces for a girder prestressed with unbonded tendons. Due to the intensive

prestressing arrangement provided by the curtailed prestressing tendons of the proposed cantilever

box girder, this fully-compressed state exists for a large portion of the girder which was previously

shown in Chapter 3. It seems reasonable to proceed with a parallel chord model to describe the flow

of forces in the web for the proposed girder and to design the vertical web reinforcement based on

this approach.

5.5 Preliminary Design Model

The truss model developed for the design of the 90m cantilevered segmental girder is shown in

Figure 5-16. This truss has two parallel compression chords that are assumed to be located at the

centroid of the flanges, where there are large force concentrations. In the uncracked state, the truss

is continuous at the supports, since the cantilever prestressing causes the top flange to be fully

compressed under service loads. For the ultimate limit state, flexural decompression of the top

flange occurs at the supports, which modifies the model as the top compression chord at the support

vanishes since it can not carry tension. Although this change in the truss model impacts the flexural

behaviour of the girder, in general it does not change the way the truss model carries shear. The

forces in the chords of the model can be determined by analyzing the truss due to the vertical

applied loads, and including the effects of prestressing in the flanges shown in Figure 5-17.

a) Crack pattern and alternative parallel chord model

WestWest

428kN 428kN

3585kN 3585kN

428kN 428kNθ = 15˚

compression fan

compression fandetailed trusssimplified truss

extended web crack 19˚

parallel chord model design

actual transverse stress

conservative capacityin compression fan regions

0200400600800

1000fv [MPa]

b) Calculated transverse stress and parallel chord model design

yield stress

116

Page 133: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-16. Truss model for shear design. a) fully prestressed state for continuous beam; b) modified model to account for flange decompression (formation of plastic hinge)

The truss model chosen for design is based on plasticity, where the redistribution of web stress

is assumed to occur at the ultimate limit state to satisfy equilibrium based on a chosen angle of web

stress. To ensure a ductile failure mechanism at the ultimate limit state, Menn (1990) recommends

that the shear resistance of concrete members should be taken based only on the shear capacity of

the vertical stirrups. In this regard, yielding of the stirrups will govern the design instead of

compression strut failures due to factored transverse loading. For prestressed members the angle

of shear stress in the web is generally smaller than that of reinforced concrete members (Canadian

Standards Association, 2006). The typical angle of compression stress in the web that is observed

for fully compressed concrete members is (demonstrated by Equation C-5 in Appendix C).

For the purpose of designing the preliminary web reinforcement for the proposed box girder, a

typical compression stress angle will be assumed to be . This angle of stress in the web is

supported by Menn (1990).

Figure 5-17. Prestress forces applied to truss model in Figure 5-16 for each applied anchorage force

θ

θ

a)

b)

Q Q Q Q Q Q Q Q Q Q Q Q Q

Q Q Q Q Q Q Q Q Q Q Q Q Q

V V

V VFlexural decompression of top flange causes chord to vanish from model (no tension capacity across the joint due to no continuous bonded steel)

29°

30°

= = =

QPt

Pb

Pe

Mp=Pe

P

ete

eb

Peb/d Mp/d

Pet/d Mp/d

det

eb

Pt

Pb

WherePt = P (eb/d) + Mp/dPb = P (et/d) - Mp/d

θ

VAppliedprestress

Equivalentsectional forces

d

Model geometry

Distributedchord forces

117

Page 134: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

5.6 Opening of Joints in Segmental Bridges

For critical flexure regions of the girder at the ultimate limit state, the total decompression of a

chord on the flexural tension side occurs. Following this decompression of the flange, an opening

of segmental joints in precast concrete bridges can form since there is no continuous bonded

reinforcing steel across the joint. This has been a concern for designers as the mechanism for

carrying shear stresses across an open dry joint in the vicinity of a support is uncertain since the

assumed diagonal compression strut in the web cannot be established across the open joint.

Virlogeux (1993) has claimed that the opening of joints near the support must not be allowed

for the ultimate limit state design since there is no clear load path for the web shear stresses to be

reacted by the support. In order to bring the shear forces from the superstructure to the support

across an open joint, he states that vertical reinforcing steel or prestressing must be installed in the

vicinity of the open joint to lift the shear forces to an elevation where the shear force can pass

through the support. This additional vertical reinforcing steel, also termed as ‘hanger

reinforcement’, is displayed in Figure 5-18. This explanation of the flow of forces; however, is

fundamentally flawed since horizontal equilibrium is not satisfied at the top node next to the open

joint. Turmo (2005) has conducted a finite element study to show that, indeed, inclined

compressive arching stresses crossing an open joint occurs to transfer the force across an opened

joint. As the girder experiences distress due to joint opening, vertical equilibrium in the vicinity of

the open joint is maintained due to the large inclined forces within the slab and the lower region of

the web in compression. Therefore, it is not necessary to provide additional vertical web

reinforcement in regions where joint opening is possible since the compressive arching of forces

for rotated joints will satisfy vertical equilibrium.

Figure 5-18. Incorrect hanger reinforcement model for shear stresses crossing an open joint (adapted from Virlogeux, 1993)

118

Page 135: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

5.7 Design of Web Reinforcement

To determine the maximum shear demands on the bridge, ULS Combination 1 defined in the

CAN/CSA-S6-06 was considered. For the span of 90m, the lane load model defined in cl. 3.8.3.3

governed the live load demands. For a deck width of 12m, three design lanes are considered to

produce the worst effect, shown in Figure 5-19 below.

Figure 5-19. ULS Combination 1 maximum shear demands

Based on these demands, the arrangement of vertical reinforcing steel in the web can be

quantified that is required to provide the necessary spreading of compression stresses in the web.

The critical sectional force for the design of vertical reinforcing steel has been taken at a distance

of dv from the support (cl. 8.9.3.1). To provide economy in web steel consumption, the bar spacing

has been varied at several regions on the girder. The required vertical steel demand and the chosen

arrangement is shown below in Figure 5-20. At locations where the bar spacing changes, the

capacity envelope can be transitioned over a length equal to the height of the girder, where the

transition of capacity is centred at the location where the bar spacing changes.

V [kN]

distance [m]

-12000

-10000

-8000

-6000

-4000

-2000

0

2000

4000

6000

8000

10000

12000

0 10 20 30 40 50 60 70 80 90

DL

DL+SDL

DL+SDL+LL

Uniform distributedlane load (9kN/m per lane)

Shaded area:Envelope due to CL-625-ONT Truck

119

Page 136: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-20. ULS transverse steel demand and chosen capacity

Due to the slender web design, only one leg of reinforcement is placed at the centre plane of

the web. Equally spaced 20M headed reinforcing bars are placed in the web. The use of headed

reinforcing bars allow for reduced bar congestion in the vicinity of the flange junction.

Figure 5-21. Web reinforcement

5.7.1 Comparison to CAN/CSA-S6-06

To compare the reinforcement design provided by the simple design model to the CAN/CSA-

S6-06, the angle of compression stress in the web needs to be determined based on the code

equations for girders in shear. In order to calculate the angle of compression stress, θ, both the

0

1

2

3

4

5

0 5 10 15 20 25 30 35 40 45

dv

h

h

Av/s

calculated transversesteel required so that Vs = Vf

transverse steelprovided

120

Page 137: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

sectional strain at the mid-height of the web, , and the equivalent crack spacing parameter, sze,

must be known (given in Equation C-5 of Appendix C). The mid-height web strain can be

computed directly based on sectional forces and cross section geometry. The equivalent crack

spacing parameter, however, is an empirical equation which is suitable for use in reinforced

concrete beams designed using standard practice methods. Strictly following the Code, a deep box

girder reinforced only with one single direction of transverse steel will produce a very large

equivalent crack spacing parameter which consequently results in steep predicted angles of

diagonal web compression. This is because the code does not take into account crack control

provided by fibre reinforced concrete. To illustrate this point, consider the case for the proposed

3600mm deep box girder with nominal 20mm aggregate and only vertical steel in the webs:

If the box girder is fully compressed, the strain in the web will be a small compressive value.

Conservatively assuming , the predicted angle of web compression is:

The use of fibre-reinforced concrete in girder webs as a replacement of typical minimum crack

control reinforcement is not current practice; however, current studies are in progress to investigate

the performance of fibre-reinforced concrete panels in shear with only one direction of bonded

reinforcing steel (Susetyo, 2007). Below in Figure 5-22 is a visual comparison of two shear panel

tests performed by Susetyo (2007) which have similar compressive concrete strength, but are

reinforced with different steel arrangements. A high strength 80MPa normal concrete panel is

reinforced with 3.3% reinforcement in one direction (representative of vertical web steel), and

0.4% reinforcement in the orthogonal direction (representative of crack control horizontal steel).

εx

sze 35sz 15 ag+( )⁄=

35(0.9dv ) 15 20+( )⁄=

0.9 3600( )=3240mm=

εx 0=

θ 29 7000εx+( ) 0.88sze

2500------------+⎝ ⎠⎛ ⎞=

29 7000 0( )+( ) 0.88 32402500------------+⎝ ⎠

⎛ ⎞=

63°=

121

Page 138: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

A high strength 80MPa FRC panel containing 1.5% hooked steel fibres has also been tested with

3.3% reinforcement in one direction only.

Figure 5-22. Shear panel tests for normal reinforced concrete and FRC concrete (adapted from Susetyo, 2007)

It is seen that the FRC test specimen has a more distributed crack pattern than the normal

reinforced concrete test specimen prior to failure. The FRC test specimen cracks are more closely

spaced and the average crack width is smaller. These preliminary results do not conclude that this

behvioiur is reliable for every case, but they provide some evidence to speculate the cracking

behaviour of webs in shear made of high strength FRC. For the purpose of calculation using the

Code provisions, the web of the box girder is treated as a typical minimum reinforced web with

adequate crack control reinforcement and the crack spacing parameter, sz, is assumed to be equal

to 300mm (cl. 8.9.3.7). Under this assumption, for the proposed box girder design is evaluated

under the load condition corresponding to ULS Combination 1. The following distribution of

diagonal compression stress and the corresponding parameters can be seen below in Figure 5-23.

a) 80MPa concrete; two orthogonal steel reinforcement directions

b) 80MPa FRC; one reinforcement direction

122

Page 139: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 5-23. Shear parameter predictions according to CAN/CSA-S6-06

The prediction of the CAN/CSA-S6-06 provisions show that the angle of web compression for

the vast majority of the beam is close to for the entire length of the girder since the girder is

fully compressed at the ultimate limit state for all locations at a distance further than dv away from

the support. Notwithstanding the fact that resistance provided by aggregate interlock has been

neglected for the design of web reinforcement, there appears to be good agreement between the

chosen truss model for design and the variable angle truss model predicted by the Code. For the

global spreading of forces, the designed web reinforcement is sufficient to transfer the compressive

-200000

-150000

-100000

-50000

0

50000

100000

0 10 20 30 40

0

2500

5000

7500

10000

12500

15000

0 10 20 30 40

-0.0002

-0.0001

0

0.0001

0.0002

0 10 20 30 40

05

10152025303540

0 10 20 30 40

0

0.1

0.2

0.3

0.4

0.5

0.6

0 10 20 30 400

10000

20000

30000

40000

50000

60000

0 10 20 30 40

Mf [kNm]

Vf [kN]

fpeAps [kN]

εx [mm/mm]

θ [deg]

βcantilever post-tensioning

continuitypost-tensioning

minimum envelope

maximumenvelope

critical point ofinflection

fpoAps for Mf < 0fpoAps for Mf > 0

constant θ = 30˚assumed for designtruss model

εx > -0.0002

dv

all εx < 0 at sections> dv from support

DL+SDL+LL

dv

bottom tensionflange

top tensionflange

CAN/CSA-S6-06

30°

123

Page 140: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

stresses in the web. Web reinforcement corresponding to local spreading at the anchorage will now

be considered in the following chapter.

124

Page 141: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

125

6.0 ANCHORAGE ZONE

This chapter begins by introducing the importance of anchorages for concrete structures

prestressed with external unbonded tendons in Section 6.1. In Section 6.2, the fundamentally

different flow of forces is described within the anchorage zone for an external unbonded tendon

anchorage in comparison to the flow of forces for an anchorage for an internal tendon. A three-

dimensional truss model is developed in Section 6.3 to describe the flow of forces in the anchorage

zone for the corner blisters which is consistent with the previous model used for global transfer of

shear force in the box girder. Section 6.4 discusses the detailing requirements for the anchorage

steel and Section 6.5 presents a possible arrangement of reinforcing steel for a typical anchorage

zone.

6.1 Introduction

The proper anchorage of external unbonded tendons to anchorage blisters is of paramount

importance for the structural integrity of the proposed system during the whole lifetime of the

structure. Unlike internal bonded systems, the transfer of longitudinal force for external unbonded

tendons into a box girder is only facilitated at discrete anchorages located at the ends of the tendon.

For structures with bonded prestressing, there is an inherent redundancy in the system for the

transfer of prestress force. In the case of damage to the original anchorage, the prestressing force

in a bonded tendon may be partly taken over by the activation of mechanical bond of the grouted

tendon to the structure. However, damage and loss of an anchorage for an external unbonded

tendon results in the loss of the entire tendon. Since prestressing of a cantilever girder with no

continuous bonded steel relies exclusively on the anchorage blister, special attention must be paid

to ensure adequate anchorage can be provided.

Anchorages should be located at the corner of the box section for minimum development length

(Menn, 1990). The temporary external unbonded post-tensioning required during the cantilever

construction stage should also be provided near the corners of the cross section for adequate crack

control. Post-tensioning forces applied to the section at the middle of slab components can cause

significant longitudinal cracks due to strain compatibility of local deformations (Moon, 2004).

Also, anchorages located away from the corner can cause significant out-of-plane bending in the

slab element due to the eccentricity of the external anchorage blister (Eibl, 1990). For the proposed

Page 142: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

precast segmental girder, all anchorages have been located at corner blisters (shown below in

Figure 6-1).

Figure 6-1. Top and bottom corner anchorage blisters

6.2 Flow of Forces

The flow of forces for an external unbonded tendon anchored at an intermediate corner blister

is fundamentally different than the flow of forces for an internal tendon that is guided into the slab.

For an internal tendon, the curvature of the tendon profile in elevation creates a vertical deviation

force due to the tendon bearing against the surrounding concrete. The self-equilibrating state of

forces for a girder with internal tendons shows that the resultant prestress force follows the profile

of the tendon. As such, the resultant prestress force is deviated sufficiently into the bottom slab for

the overall equilibrium of the box girder. Figure 6-2 shows the anchorage forces, deviation forces,

and boundary forces for a girder segment prestressed with an internal tendon and a girder segment

prestressed with an external tendon at intermediate anchorages located on the bottom slab. It can

be seen that for the internal tendon, the forces on the tendon and the forces on the concrete are equal

and opposite and act in the same plane.

For the case with the external unbonded tendon, the boundary forces on the segment which

satisfy a statically equivalent set of loads are not coincident with the tendon force. Since the tendon

exits the concrete section, the compression forces in the concrete are not guided into the bottom

flange the same way as it is with the guided tendon. Therefore, a different state of internal

equilibrium must exist to deviate the applied anchorage force into the concrete box girder. The

boundary forces shown in Figure 6-2 follows a consistent method of analysis with regard to the

parallel chord model for the global spreading of forces in the web. Large concentrated forces are

assumed to be carried in the flanges; therefore, the statically equivalent set of loads to equilibrate

the applied anchorage force are shown to be distributed in both the top and bottom flange. For the

126

Page 143: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

anchorage force to be distributed in this way, a local spreading of forces must take place within the

box girder to provide this state of equilibrium.

Figure 6-2. Flow of forces at an intermediate anchorage

6.3 Strut-and-Tie Model for Local Spreading of Forces

To visualize the local spreading of forces in a box girder that is prestressed with intermediate

external unbonded post-tensioning at corner blisters, a simple strut-and-tie model desired. Since

the centroid of the anchorage force and the centroid of the web are not in coincident planes, the

strut-and-tie model required to sufficiently describe the spreading of forces must be in three

dimensions.

6.3.1 Anchorage of External Unbonded Tendons in Slab Blisters (Wollman, 1993)

Wollman (1993) performed a laboratory study to investigate the behaviour of external tendon

anchorages in box girder diaphragms and intermediate slab blisters. His study included eight blister

and rib specimens, one of which was an external tendon corner blister. The load path for a corner

blister is complex as it requires consideration in three dimensions to correctly represent the flow

of forces. He offers a strut-and-tie model to describe the flow of forces for a box girder due to the

anchorage of external unbonded tendon at a corner blister, shown in Figure 6-3. Strong corbel

reinforcement (shown as force T1) is required to tie the blister to the slab, and diagonal tension

reinforcement (shown as force T2) is required at the end of the blister to pull the forces into the

plane of the web and the flange. Web and flange bursting reinforcement (shown as force T3) is also

a) internal bonded tendon

b) external unbonded tendon

forces on concrete forces on tendon

forces on concrete forces on tendon

local spreading of forces into both flangesis required to satisfyequilibrium

deviation forces due to tendon curvature bring prestress force into the flange

127

Page 144: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

provided to take into account the spreading of forces. Wollman (1993) also provides longitudinal

reinforcement in the vicinity of the corner for the truss model to satisfy equilibrium due to the

longitudinal component of force of the diagonal compression struts in the slabs. According to his

additional finite element analysis performed for the corner blister, the highest tension forces occur

behind the anchor at the re-entrant corner, while the slab bursting stresses are much smaller. The

test specimen was reinforced according to the strut-and-tie model, and failure was observed due to

concrete crushing in the local zone, which was typical for all blisters. Wollman (1993) states that

the most consistent predictor of blister failure is the calculation of capacity of local zone based on

the ratio of the area of supporting concrete to plate area (A/Ab) and the confinement provided by

the spiral.

Although the strut-and-tie model given by Wollman (1993) represents a reasonable flow of

forces due to the anchorage force at a corner blister, it appears to be incomplete as it does not satisfy

equilibrium at all nodes. The strong corbel reinforcement, T1, provides a tension force to deviate

the longitudinal anchorage force, P, but the inclined tensile force is not in transverse equilibrium

at node A. In addition, the spreading of compressive forces at the end of the anchorage blister can

not take on the trajectories shown in Figure 6-3 since the force needed in tension tie T2 to pull the

compressive forces into the plane of the slabs requires the resultant of the compressive forces to

converge at the corner. By moment equilibrium of global forces, the equal distribution of the

resultant P/2 forces in each slab shown in this strut-and-tie model can only be true if the application

of the prestress force, P, is equidistant from the web and the bottom flange.

The boundary forces provided on this model may be representative of the specimen tested by

Wollman (1993). The equal boundary forces provided at the mid-height of the web and at the mid-

width of the flange are reasonable assumptions for a specimen with only a vertical web slab and a

horizontal flange slab. However, for the application of this model into a continuous box girder,

some modification is necessary to adapt the idea of this spreading of forces into the full cross

section.

128

Page 145: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 6-3. Strut-and-tie model for corner blister (Adapted from Wollman, 1993)

6.3.2 Design Model for External Tendon Blisters

Using model described previously by Wollman (1993) as a reference, a revised strut-and-tie

model was developed that resolves equilibrium at all nodes, shown in Figure 6-4. The isometric

profile, shown in Figure 6-4a, outlines the centreline of the web in the vertical plane, the centreline

of the bottom flange in the bottom plane, and the anchorage blister at the level of applied force in

the corner. The equilibrium of the tension tie, T1, is equilibrated vertically and horizontally by

inclined compressive struts in the web and in the flange, respectively. The longitudinal component

of force associated with the inclined web and flange compressive struts is tied forward with the

tensile force T3. Compressive struts at the end of the anchorage blister follow the trajectory path

that is required by equilibrium, which resolves in the corner to meet the longitudinal tension tie T3.

The prestressing force, P, resolves into three discretized boundary forces, F1, F2, and F3 at a

distance 2Lb past the face of the anchorage blister, where Lb.is the longitudinal length between ties

T1,and T2. Figure 6-4b displays the front view of the three dimensional model, showing the

eccentricity of the anchorage force, P, to the web and the flange, ew and ef, respectively. Figure 6-

4c displays the angle of the compression struts in elevation. The angle of compression in the slabs

is labelled θs. The angle in the anchorage blister measured from the flange, θf, is defined by the

eccentricity of the force to the slab, ew, and the distance between the centroids of tension ties, Lb.

The top view to define θf is geometrically similar. For longer anchorage lengths, the tensile

demand in transverse ties is decreased since the angle of compression is flatter. However, from a

practical design perspective, the maximum length of the anchorage block limits the maximum

length of Lb for bottom corner blisters due to the required clearance to place a jack between

anchorages.

T1

T2

P

T3

T3

P/2

P/2

ACompressionTension

129

Page 146: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 6-4. Strut-and-tie model for corner blister satisfying static equilibrium; a) 3d isometric view, b) front view, c) side elevation

For the strut-and-tie model shown in Figure 6-4, equilibrium of joints gives the following

tension demands:

(6-1)

(6-2)

P

F3

F2

F1

T4

T5T1

T2

T3

CompressionTension

LbLbew

ef

θf

θfθs

P

F2+F3

F1

Front View Side Elevation

dv

T1P2--- θwtan2 θftan2+ P

2---ewLb------⎝ ⎠

⎛ ⎞2 ef

Lb------⎝ ⎠

⎛ ⎞2

+= =

T2 T1P2---

ewLb------⎝ ⎠

⎛ ⎞2 ef

Lb------⎝ ⎠

⎛ ⎞2

+= =

130

Page 147: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

(6-3)

(6-4)

(6-5)

The reaction forces satisfying global equilibrium of the free body are given as:

(6-6)

(6-7)

(6-8)

It is observed from the demand equations that the demand for steel is largely influenced by the

eccentricity of the anchorage force to the corner and the length of the anchorage blister. To

minimize the demand for steel, the tendon should be anchored as close as possible to the corner of

the box girder.

For the top anchorages, all cantilever post-tensioning is installed during the construction

process at the cantilever tip where there is an open face. Therefore, ample clearance is available for

the stressing of tendons. The largest tendon unit used for the cantilever post-tensioning is a VSL

ES6-12. The spiral that is required for these tendon units is 350mm in diameter. The vertical

limiting factor for placing the anchorage close to the corner is the to ensure that the anchorage

spiral and the transverse pretensioning strands do not interfere with each other. The horizontal

limiting factor is to ensure that the anchorage head clears the inside face of the web.

For the bottom tendons, the construction clearance governs the placement of the anchorages.

Minimum jack clearances for stressing tendons must be provided to allow installation of tendons.

The necessary jack clearances are illustrated in Figure 6-5. The largest tendon units that are used

for the design of longitudinal post-tensioning are VSL ES6-19. The stressing jack type required to

stress these tendons is a ZPE-460/31. For this jack type, the edge clearance required (shown as

T3 P θs θwtan θftan+( )cos P2Lb

4Lb2 dv

2+----------------------------

⎝ ⎠⎜ ⎟⎛ ⎞ ew ef+

Lb----------------⎝ ⎠

⎛ ⎞ 2Pew ef+

4Lb2 dv

2+----------------------------

⎝ ⎠⎜ ⎟⎛ ⎞

= = =

T4P2--- θftan

Pef2Lb---------= =

T5P2--- θwtan

Pew2Lb---------= =

F1P2---

θftanθssin------------- P

2-------

efLb------= =

F2 P T3– P P2

-------ew ef+

Lb----------------⎝ ⎠

⎛ ⎞–= =

F3P2---

θwtanθssin-------------- P

2-------

ewLb------= =

131

Page 148: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

dimension ‘E’) is 305mm and the minimum longitudinal distance required for installation (shown

as dimension ‘C’) is 1505mm.

Figure 6-5. Jack clearances necessary for installation (adapted from VSL, 2007)

Satisfying these clearance criteria, the top anchorages result in a vertical eccentricity of 270mm

and a horizontal eccentricity of 278mm. The bottom anchorages result in a vertical eccentricity of

380mm and a horizontal eccentricity of 305mm (shown below in Figure 6-6).

Figure 6-6. Minimum anchorage eccentricities to box girder corner

According to the truss model shown previously in Figure 6-4, it is observed that the demand

for steel is also minimized by maximizing the longitudinal length of the anchorage blister. For the

bottom tendon anchorages, the maximum length of the anchorage is governed by the longitudinal

jack clearance requirement of 1505mm shown by Dimension ‘C’ in Figure 6-5 above. For a

segment length of 3000mm, the maximum anchorage length for the bottom tendons is limited to

1495mm. For the top tendons, a similar anchorage length has been chosen which also takes into

account the integration of anchorage blister and transverse ribs formed in the top slab. To minimize

the number of surfaces to form, the anchorage blister has been chosen a length of 1250mm which

corresponds to the vertical face of the transverse rib (as shown in Figure 6-7).

B

C

D

E

E

132

Page 149: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 6-7. Anchorage blister lengths

Therefore, adapting the anchorage model to the minimum anchorage eccentricity and the

3600mm deep proposed box girder, the local spreading of forces to provide the statically

equivalent boundary conditions at the flange is shown below in Figure 6-8 for a bottom tendon.

Figure 6-8. Strut-and-tie anchorage model for a bottom tendon

6.4 Detailing

Beaupre et al. (1990) investigated the design of deviator saddles for externally post-tensioned

structures. The saddles studied are generally used to deviate external unbonded tendons vertically

and horizontally for typical span-by-span type structures. Unlike an anchorage blister which must

transfer a large longitudinal force into the structure, these deviation saddles need only provide

resistance in the vertical plane. Notwithstanding this difference, it is useful to note the various

arrangements of reinforcement possibilities to tie a horizontal or vertical load into a box girder due

to deviation forces applied by tendons. Beaupre et al. (1990) show that a typical saddle detail used

in straight-span bridges consists of a direct tension link bar to efficiently tie the deviation force into

the corner node of a box girder section, shown in Figure 6-9. He also proposes a modified

reinforcement scheme to simplify and standardize reinforcement patterns and saddle geometry.

P

F2+F3

F1

133

Page 150: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The closed loop stirrup arrangement relieves reinforcement congestion at the corner node by

placing stirrups around the tendons being deviated.

Figure 6-9. Tension ties linking deviation force to box reinforcement (adapted from Beaupre et al., 1990)

Deviation forces are resisted by combined shear action in the plane of the web and in the plane

of the flange. Since the increase in tendon force is small for the proposed structure at the ultimate

limit state, the critical design force for the deviator reinforcement occurs due to initial jacking of

the tendon. Under CAN/CSA-S6-06 specifications, the maximum allowable initial jacking force is

limited to 80% of the ultimate strength of the tendon (0.8fpuAps).

The truss model shown previously in Figure 6-8 displays diagonal tension tie members at T1

and T2 which is similar to the efficiently oriented steel links shown by Beaupre et al. (1990). A

direct tension link is the most efficient arrangement of steel since it provides two legs of

reinforcement per stirrup that are oriented in the direction of the optimum load path. However, as

an alternative arrangement for simplified and standardized detailing, the closed loop reinforcement

will be used for design instead of the direct tension links. The closed loop arrangement does not

effect the forces on the global system, however, the local load path with the anchorage blister is

somewhat modified (Figure 6-10). The direct tension link ties the force to the corner of the box

directly; however, the closed hoop reinforcement provides a load path where web and flange forces

are met with a diagonal compression strut at the corner.

The quantity of steel required for each detail is different. For the direct tension link, area of steel

which resists the direct load is 2Ab for each link. For the closed loop, the area of steel which resists

an orthogonal component of the direct load is 1Ab in both directions. Therefore, since legs in both

directions have equal area, the number of required closed loops is governed by the leg which

attracts more force. The proportion of forces in the legs can be determined using the geometrically

similar eccentricities shown above in Figure 6-6. The size and shape of hoop is consistent along

the length of the anchorage blister, with concentrated reinforcement at the ends to satisfy the

demand of the truss model given by T1 and T2.

FLinkbar

Direct tension link stirrups

inner loopbarclosed stirrup

Loop stirrups

134

Page 151: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 6-10. Flow of forces for different reinforcement details

Haroon (2006) investigated the benefits of using FRC for anchorage zones in post-tensioned

structures and found that there are significant enhancements in performance. The best performing

anchorage specimen configuration satisfying AASHTO acceptance criteria included 1% hooked-

end fibres and no spiral and no skin reinforcement. The minimum compressive strength of concrete

was 40.7MPa. With the use of high-strength concrete and 1% steel fibres, this allows for 100%

total elimination of the secondary reinforcement. The test specimens in this study were for VSL

type EC5-7 anchorage units; however, Haroon (2006) concludes that the results are expected to be

valid for larger anchorage units. The anchorage blister reinforcement for the proposed girder

design will include a spiral at the anchorage zone to provide confinement at the bearing plate and

the necessary steel required to tie the anchorage force into the box girder, but no secondary skin

reinforcement will be designed since adequate crack control has been observed in tests due to the

benefit of FRC.

6.5 Reinforcement

The anchorage blister reinforcement and all necessary tension steel for local spreading

described by the truss model in Figure 6-4 was used as a basis for the design of anchorage

reinforcement in the box girder. All reinforcing steel has been designed to a limited stress of

240MPa (cl. 8.6.2.7.4). It was found that the required tension forces to deviate the large anchorage

b) closed hoop

a) 2-leg link

truss model reinforcement detail

H

H

direction of forcesat H

magnitude of forces

As = 2Ab

As = 1Ab

135

Page 152: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

forces is significant within the anchorage blister and in the web. The closed loop reinforcement

within the blister has been chosen to be not larger than 20M reinforcing bars to accommodate the

minimum inside bar bend radius of 4db(cl. 8.4.2.1.1). Limiting the bar size, the steel required for

the deviation within the blister has been chosen to be bundled to allow for a larger Lb between

deviation forces.

For the critical bottom anchorage corresponding to a VSL ES6-19 tendon, the design force

corresponding to 80% of the ultimate strength of the tendon is 3958kN. For a bottom flange

thickness and web thickness of 150mm and 120mm, respectively, the eccentricities ef and ew are

380mm and 305mm, respectively (as shown previously in Figure 6-6). Nine hoops of 20M bars are

required to satisfy each of the tie forces, T1 and T2. Three sets of bundled bars consisting of 3-20M

bars each have been provided at the locations of T1 and T2. The blister length between the centroid

of steel providing the deviation forces is Lb = 1160mm which allows for 30mm cover for the

outermost hoop. Nine 20M bars are required for the bursting force in the web, which have been

evenly distributed at 50mm behind the anchorage (shown in Figure 6-11). Longitudinal tension

steel is required to resist the longitudinal component of the inclined compressive struts in the slabs

at the segment face. The longitudinal reinforcement described by T3 is required to be effective at

the location of the anchorage, which has been designed to be at the segment face. Due to congestion

difficulties, headed reinforcement has been provided in the longitudinal direction. Headed

reinforcement reduces congestion of bars in the concrete, and provides full development of the bar

at the location of the head.

Figure 6-11. Anchorage zone reinforcement for critical 19 strand bottom tendon

For a constant anchorage blister geometry, the change in required anchorage reinforcement and

the change in anchorage force for smaller tendon units are linearly related (since tension forces are

proportional to P). A similar reinforcing arrangement is required for all other anchored tendons.

The area of steel required for anchorage zone reinforcement on the basis of the truss model for all

136

Page 153: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

chosen prestressing arrangements is summarized below. Table 6-1 and Table 6-2 give the area of

required reinforcement steel corresponding to the major tension demands of the truss model in

Figure 6-4 for top cantilever tendons and bottom continuity tendons, respectively. All values in the

table have been generated using the formulas given from Equation 6-1 to Equation 6-5.

Table 6-1. Area of steel required for local spreading reinforcement for top tendonsa (in mm2)

a. corresponds to: jacking stress in tendon = 0.8fpu, allowable steel reinforcement stress = 240MPa

Tendon Size T1 T2 T3 T4 T5

8 strand 980 980 1920 960 980

12 strand 1470 1470 2880 1430 1470

Table 6-2. Area of steel required for local spreading reinforcement for bottom tendonsa (in mm2)

a. corresponds to: jacking stress in tendon = 0.8fpu, allowable steel reinforcement stress = 240MPa

Tendon Size T1 T2 T3 T4 T5

8 strand 1140 1140 2290 1140 910

12 strand 1710 1710 3430 1710 1370

19 strand 2700 2700 5440 2700 2170

137

Page 154: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

138

7.0 FIXED END ABUTMENT

In this chapter, a conceptual abutment design will be presented for the validation of the

cantilever-constructed bridge. The chapter begins by discussing the design decisions made for the

superstructure and the implications they have on the substructure design. A reference bridge for a

similar structural system is discussed which provides design guidelines on the conceptual layout

of the proposed abutment. Several conceptual design alternatives are briefly presented and

described. A conceptual layout of the recommended system is briefly discussed; however, the

design and analysis of the abutment structure is outside the scope of this thesis.

7.1 Introduction

For the proposed constant depth cantilever-constructed bridge, a monolithic connection at the

midspan was found to be most economical as it promotes the redistribution of moment to the

midspan region. Large positive moment at the midspan is desirable since a) the constant depth box

girder has significant capacity for positive bending resistance due to the large internal lever arm

and wide compression flange, and b) the redistribution of moment relieves large negative moments

at the support which decreases the demands on the substructure. Alternatively, providing an

expansion joint at the midspan while promoting the redistribution of positive moment in the

midspan region is less desirable since a) a continuity beam and interior diaphragms are required

for a moment connection across the joint, and b) detailing of continuity tendon anchorages in the

vicinity of the midspan expansion joint is crowded and complex. Thus, a monolithically connected

superstructure with no expansion joints is desired since it allows for the most favourable

distribution of forces, simplifies detailing, and provides better serviceability performance with

respect to durability and driver comfort. A jointless bridge, however, requires longitudinal

flexibility for the expansion and contraction of superstructure due to change in temperature.

7.2 Bras de la Plaine Bridge, France (Tanis, 2003)

The Bras de la Plaine Bridge, situated in the south of Reunion Island, France, is a modern

example of a light weight cantilever-constructed bridge erected from fixed abutments. The long

span of 280m is embedded in ballasted abutments, with a typical abutment length of 45m exterior

of the main span (Figure 7-1). This results in a mainspan to backspan ratio of approximately 6.2.

Page 155: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The abutments are 11.9m wide which matches the width of the bridge deck and the abutment height

ranges from the girder depth of 18m to a maximum height of 21m. The abutments economically

place counterweight material where the lever arm is greatest, as it is composed of a back

compartment 11m long filled with ballast, and a hollow front compartment (Figure 7-2). The

bottom slab of the abutment has a constant thickness of 1m to transfer the large compression force

due to the cantilever moment of the superstructure to the counterweight at the back compartment.

Figure 7-1. Bras de la Plaine Bridge, Reunion Island, France (adapted from Tanis, 2003)

Although the superstructure of the Bras de la Plaine Bridge has some dissimilarities from the

proposed cantilever girder, the rotational resistance at the abutment provided by the foundations is

similar to the boundary conditions required for the proposed cantilever girder. The top slab of the

Bras de la Plaine Bridge is prestressed with a combination of both internal and external prestressing

tendons; however, the prestress arrangement is similar since the cantilever tendons are stressed at

the face of each cantilevered segment and anchored at the back of the abutment. This prestress

arrangement and abutment proportions provides valuable information to use for reference in the

proposed bridge structure for the conceptual design of the abutment.

Figure 7-2. Ballasted abutment of Bras de la Plaine Bridge (adapted from Tanis, 2003)

3,0119%

1000

LEST

700

4000

joint

45,347

Struts

0,600

19,7

94

0,500

B AC27

section

11,900 35,973waterpipe5,950 5,950

Elevation

139

Page 156: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

7.3 Flexible Supports

One main advantage of cantilever-constructed bridges is that they are self-supporting during

erection, and do not require the placement of falsework below the superstructure for construction.

This advantage is useful for the proposed bridge since it requires no temporary supports which

would interfere with traffic below the structure, and for many cantilever bridges this advantage is

useful since it allows for ease of construction on tall slender piers. The Felsenau bridge, in Bern,

Switzerland, is one example of a cantilever-constructed bridge built on tall slender two-leg piers to

provide adequate moment resistance during construction, and longitudinal flexibility for service

due to thermal movements of the superstructure in the final state. For smaller span jointless bridges

that are built on abutments, longitudinal flexibility is achieved by the incorporation of integral

abutments.

Bridge structures with no intermediate joints and integral abutments do not contain traditional

expansion joints between the abutments and the superstructure (Abendroth, 1989). Instead, the

superstructure and the abutments are monolithically connected and founded on flexible steel piles.

The piles are subject to horizontal movement as the bridge contracts and expands due to variation

in temperature. This horizontal displacement of the superstructure induces a moment in the piles

in addition to the axial load applied from the vertical applied loads. The piles may be designed as

beam-columns, either from a conventional elastic design approach, or an inelastic design approach

that accounts for redistribution principles (Abendroth, 1989).

7.4 Design Approach

The abutment system requires the following features: a) anchorage of all external unbonded

cantilever tendons at the back of the abutment, b) capacity to resist the large negative bending

moment of the cantilever-constructed main span due to factored dead load and live load, and c)

longitudinal flexibility to allow for thermal expansion and contraction of the 90m main span. A

large force couple in the abutment is required to establish equilibrium for the resistance of the

cantilever moment. The Bras de la Plaine bridge provides this force couple due to the heavy ballast

placed at the back of the abutment which levers against the front abutment wall. For the Felsenau

bridge, this force couple is achieved by a force differential in the two-leg pier system.

140

Page 157: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

7.4.1 Preliminary Design Alternatives

Several preliminary abutment design alternatives were considered in the development of a

feasible abutment concept for construction. A ballasted abutment design provides the cantilever

moment resistance that is needed; however, longitudinal flexibility can not be provided. Due to the

large counterweight force of the ballast material, a significant normal stress is present on the

bottom surface of the abutment. Using polyethylene sheeting as an anti-friction coating between

the abutment and soil reduces the friction coefficient, , to 0.5 (VSL, 1990); however, the sliding

force required to cause movement in the abutment is still excessive.

Figure 7-3. Counterweight abutment concept (superseded)

A similar problem exists when considering a tie-down abutment system, where the

counterbalancing force is provided by soil anchors on the backspan instead of ballast material

(Figure 7-4). The post-tensioning soil anchors are required to provide a proof load to ensure that

adequate capacity can be provided to resist the ultimate moment demand. Thus, the large normal

stresses between the bottom abutment surface and the soil are large and do not allow for

longitudinal sliding. This concept presents additional concerns since the stability of the structure

for the entire life of the bridge is dependent on soil anchor reliability.

µ

141

Page 158: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 7-4. Tie-down abutment concept (superseded)

7.4.2 Recommended Design

The recommended design uses ballast material to provide uplift resistance while incorporating

the longitudinal flexibility of a two-leg system to allow for horizontal movement due to expansion

and contraction. Similar to the Bras de la Plaine bridge, the backspan to mainspan ratio has been

designed to be 0.6, which results in an abutment backspan of 15m. The abutment system requires

the excavation of soil on each side of the highway to an adequate depth for the casting of a strong

slab to provide a firm foundation for bridge piles. The piles are intended to be encased in

corrugated steel pipe (CSP) filled with loose sand to allow longitudinal expansion/contraction of

the bridge with minimum passive soil pressure acting on the pile. Granular backfill with a density

of is assumed to provide ballast for the backspan. The calculated depth of soil, H,

required to resist uplift due to the maximum factored load on the superstructure is 8.5m. The

backspan of the abutment includes a solid anchorage zone to allow spreading of forces required for

cantilever tendons, and a front hollow compartment to minimize material consumption (Tanis,

2003). The backspan is tied down by using flexible HP piles which act as massive headed

reinforcing bars. Mechanical bond of the HP section to the concrete should be provided by a

sufficient number of shear studs welded to the flanges. The studs should be designed to effectively

develop the full cross sectional strength of the tension piles above the cantilever prestressing

anchors.

All cantilever tendons are arranged in a single horizontal plane in the superstructure, and

deviate vertically at the abutment wall. At a distance 5m from the abutment wall, all tendons are

arranged on three distinct elevations with adequate vertical clearance between tendon ducts. A

vertical wall is required at this location to provide horizontal deviation of tendons. The tendons

20kN/m3

142

Page 159: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

continue on the same vertical trajectory but are arranged horizontally into vertical columns

consisting of three anchorages each at the back face of the abutment. Providing three rows of

anchorages at the back of the abutment minimizes the angle breaks in the tendon as it minimizes

the amount of fanning of tendons required to provide sufficient anchorage space. The centrally

anchored tendons also provides a more direct flow of forces into the components of the box girder.

Figure 7-5. Proposed conceptual abutment design

7.5 Sizing of Flexible Piles

Abendroth (1989) has provided a rational approach for the design of integral abutment bridge

piles. One of the two design alternatives he discusses is based on elastic behaviour assuming no

plastic redistribution of internal forces. For HP piles in compression, the steel sections can be

designed as beam-columns undergoing axial force due to weight of the applied vertical loads and

moment caused by horizontal displacement due to thermal variations. Following the maximum and

minimum design temperatures defined in the CAN/CSA-S6-06 for a structure of this type located

near Toronto, Ontario, the design temperature range is approximately . For a 90m span, the

total horizontal displacement due to temperature, assuming a coefficient of thermal expansion to

be / , is approximately 52mm. Assuming that the bridge is built in temperature

γ

58°C

10x10 6– °C 15°C

143

Page 160: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

and that each abutment deflects equally, the maximum expansion and contraction at each abutment

is 8mm and 18mm, respectively. The elastic moment and horizontal force for a pile fixed at the

ends due a horizontal deflection, , is given by:

(7-1)

(7-2)

The maximum factored shear force and the maximum factored bending moment at the

abutment are conservatively assumed to occur simultaneously for the design of the compression

piles. For Vf = 9810kN, Mf = 144000kNm and d = 15m, the maximum compression force required

at the abutment wall is 20280kN. Assuming that the section used for compression piles is

HP310x125, the bending moment and horizontal force induced due to the 18mm thermal deflection

is 63kNm and 23kN, respectively. For an unbraced length of 5500mm and an effective length

factor of 0.65 for design, 6 - HP310x125 piles are required to resist the axial compression due to

the total vertical load and moment due to maximum thermal contraction, assuming the loads are

evenly distributed over the piles. For the backspan tie, 4 - HP310x95 piles are required to provide

the necessary tie down force, limiting the steel stress to 240Mpa.

7.6 Future Work

A feasible abutment concept for the proposed cantilever-constructed bridge design has been

described; however, several design considerations are yet to be considered. To name a few: a) the

design of the end zone reinforcement must be investigated for the spreading of forces for all

cantilever tendons, b) the webs must be designed to transfer the constant shear force to the

abutment wall, c) the top slab must be checked for live loads, d) the strong slab must be designed

to rigidly support the factored uplift load, and e) the front toe of the strong slab must be designed

as a spread footing for the large compression force at the abutment wall.

M 6EI∆

L2-------------=

H 12EI∆

L3----------------=

144

Page 161: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

145

8.0 MATERIAL UTILIZATION

This chapter begins by describing two reference bridges. These bridges provide a basis for

comparison to evaluate the economy in material consumption for the proposed box girder bridge.

Material quantity estimates are presented and compared for the reference bridges and the proposed

box girder bridge. Concrete consumption, mild steel utilization, longitudinal prestressing strand

utilization, and transverse prestressing strand utilization will be identified. Since the abutment

presented in this thesis has not been designed, only the superstructure quantities will be presented

and compared within this chapter.

8.1 Reference Bridges

In this section, a brief description will be given on the chosen bridge structures used for a

comparison on material consumption. The bridges are representative of two different types of

bridge design: a classical precast segmental cantilever-constructed bridge, and a typical box girder

highway overpass bridge built using current practice of the Ontario Ministry of Transportation. The

precast segmental bridge reference will provide a comparative basis for the savings in material that

may be realized by building a cantilever-constructed bridge using external unbonded tendons

instead of traditional internally bonded tendons. The highway overpass bridge reference will

provide a comparative basis for the savings in material that may be realized by adapting a new

cantilever-constructed type of overpass bridge in place of the current methods used.

8.1.1 Windward Viaduct, Interstate Route H-3, Hawaii, USA (Hawaii DOT, 1991)

The Windward Viaduct is a precast segmental cantilever-constructed multiple span bridge with

typical span lengths of 85.3m (280ft) and is located in the Haiku Valley of Hawaii. The elevation

of a typical span is shown below in Figure 8-1. The precast segments are made with typical

37.9MPa (5500psi) concrete and reinforced with ASTM 615 Grade 60 reinforcing steel. The

haunched girder has a depth of 4.88m (16ft) at the pier and a depth of 2.44m (8ft) at the mid-span.

These girder depths correspond to span-to-depth ratios of 17.5:1 and 35:1, respectively.

Page 162: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Figure 8-1. Partial elevation view of Windward Viaduct (adapted from Hawaii DOT, 1991)

The cross section is a single cell box girder with a vertical web and a constant web thickness

of 381mm (15in). The typical deck width is 12520mm (41ft 1in) and the web spacing is 6320mm

(20ft 9in) centre-to-centre. The top flange is constant for all girder depths. It has a typical thickness

of 229mm (9in) and tapers to a haunch of 508mm (20in) at the webs. The bottom flange varies in

thickness from 229mm (9in) at the midspan to 406mm (16in) at the pier. The typical cross section

is shown below in Figure 8-2.

Figure 8-2. Typical cross section of Windward Viaduct (adapted from Hawaii DOT, 1991)

All cantilever prestressing tendons are internal within the top flange thickness and all

continuity prestressing tendons are internal within the bottom flange thickness. In the top flange,

146

Page 163: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the cantilever tendons are anchored at the junction of the top flange and the web; therefore, the

haunch in the top flange is present to accommodate the anchorage and spacing of tendon ducts. The

continuity tendons are anchored at corner blisters on the bottom flange and follow the vertical

profile of the bottom slab haunch. The prestressing arrangement consists of 17-strand and 19-

strand tendons using 13mm (0.5in) strands. A typical prestressing arrangement is shown in

Figure 8-3 which displays a half plan for the top flange prestressing tendons and a half plan for the

bottom prestressing tendons.

Figure 8-3. Post-tensioning arrangement for typical span of Windward Viaduct (adapted from Hawaii DOT, 1991)

8.1.2 Hwy 407 - Islington Avenue Underpass, Toronto, Canada (MTO, 1990)

The Islington Avenue Underpass is a two-span structure with equal span lengths of 50.5m built

over Highway 407 in Toronto, Canada. The general arrangement elevation is shown below in

Figure 8-4. The superstructure is a cast-in-place post-tensioned multiple cell box girder. The

superstructure is cast with 35MPa concrete and reinforced with Grade 400 reinforcing steel. The

147

Page 164: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

constant depth box girder has a girder depth of 2m which corresponds to a span-to-depth ratio of

25.2:1.

Figure 8-4. General arrangement of Islington Avenue Underpass (adapted from MTO, 1990)

The cross section is a multiple cell box girder with 6 webs having a typical web thickness of

840mm each. Each web contains 6 multi-strand draped post-tensioning tendons. The deck width is

20.41m, and the typical top flange thickness is 225mm. The cantilever wing extends 2m beyond

the web and tapers to a flange thickness of 450mm at the web. The typical bottom flange thickness

is 175mm. A typical cross section is shown below in Figure 8-5.

Figure 8-5. Typical cross section of Islington Avenue Underpass (adapted from MTO, 1990)

8.2 Concrete Consumption

The concrete consumption for the bridges described above and the proposed segmental box

girder have been summarized in Table 8-1, Table 8-2, and Table 8-3 below. Since all bridges

considered have different bridge widths and bridge lengths, the concrete consumption for each

bridge has been normalized to the volume of concrete consumed per square meter of bridge deck.

The concrete consumption for the Windward Viaduct has been estimated based on the variable

depth typical cross section shown in Figure 8-2, and the pier segment volume which is substantially

heavier. The concrete consumption for the Islington Avenue Underpass has been estimated based

on the typical cross section shown in Figure 8-5, and solid concrete sections which are located over

148

Page 165: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

the abutments and the centre pier. The concrete consumption for the proposed box girder bridge is

based on the typical 3.6m deep segment.

It is observed that there is greater economy in concrete consumption for traditional cantilever-

constructed bridges than conventionally constructed bridges built by the Ontario Ministry of

Transportation (MTO). Comparing the unit volume consumption of the Windward Viaduct to the

Islington Avenue Underpass, the Islington Avenue Underpass consumes approximately 44% more

concrete per square meter of bridge deck than the Windward Viaduct. The design of multiple cell

girders with thick webs for typical MTO overpass bridge construction has a significant impact on

the overall concrete consumption. The proposed box girder bridge consumes approximately 44%

of concrete compared to classical cantilever constructed bridges, and approximately 31% of

concrete compared to traditional MTO overpass bridge construction per square meter of bridge

deck.

Table 8-1. Concrete consumption for one span of the Windward Viaduct (Hawaii DOT, 1991)

Item Area[m2]

Width[m]

Length[m]

Volume[m3]

Unit Volume[m3/m2]

Typical Segment varies from 7.18 to 9.61 12.5 81.1 637 0.628

Pier Segment 13.7 12.5 4.20 58 1.10

Total 85.3 695 0.651

Table 8-2. Concrete consumption for Islington Avenue Underpass (MTO, 1990)

Item Area[m2]

Width[m]

Length[m]

Volume[m3]

Unit Volume[m3/m2]

Typical Section 16.2 20.41 84 1360 0.794

Solid sections over supports 33.6 20.41 17 571 1.65

Total 101 1931 0.937

Table 8-3. Concrete consumption for proposed cantilever box girder

Item Area[m2]

Width[m]

Length[m]

Volume[m3]

Unit Volume[m3/m2]

Typical Section 3.43a

a. smeared area accounts for transverse ribs and corner blisters

12 90 308 0.286

149

Page 166: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

8.3 Mild Reinforcing Steel Utilization

The mild reinforcing steel utilization for the reference bridges and the proposed segmental box

girder have been summarized in Table 8-4, Table 8-5, and Table 8-6 below. Since concrete

consumption per square meter of bridge deck is vastly different between the structures compared,

the classical reinforcement utilization ratio expressed on the basis of mass of steel per cubic meter

of concrete is less meaningful. For a more consistent comparison, reinforcing steel utilization on

the basis of mass of steel per square meter of bridge deck has also been given. The Windward

Viaduct mild steel reinforcement has been estimated based on the reinforcing schedule for the

typical segment. For the pier segment, the quantity of steel has been estimated by assuming the

calculated reinforcing ratio, ρ = 0.0089, based on the typical section. The Islington Avenue

Underpass mild steel reinforcement has been estimated based on the reinforcing schedule for the

typical cross section. For the solid sections over the supports, the quantity of steel has been

estimated by assuming the calculated reinforcing ratio, ρ = 0.0088, based on the typical section.

The mild steel reinforcement for the proposed box girder bridge accounts for the required web

reinforcement for global spreading of forces and the anchorage zone reinforcement required for

local spreading.

Table 8-4. Mild reinforcing steel utilization for one span of the Windward Viaduct (Hawaii DOT, 1991)

Item SteelVolume

[m3]

SteelMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Top Flange (typ.) 2.22 17500 25.2 16.4

Webs (typ.) 1.13 8880 12.8 8.30

Bottom Flange (typ.) 2.23 17500 25.2 16.3

Pier Segment c

c. assumes equal reinforcing ratio as the typical segment, ρ = 0.0089

0.52 4070 5.8 3.81

6.10 48000 69.0 44.9

150

Page 167: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

On the basis of mild steel consumption per square meter of bridge deck, it is observed that there

is greater economy in steel consumption for classical cantilever construction than for

conventionally constructed bridges built by the MTO. However, both structures are reinforced with

approximately the same reinforcing ratio on the basis of steel consumption per cubic meter of

concrete. Since the overall concrete consumption for the Windward Viaduct is less than the

Islington Avenue Underpass, the steel consumption for the Windward Viaduct is also less. For the

proposed box girder bridge, the low concrete consumption results in a high steel utilization factor

on the basis of steel consumption per cubic meter of concrete. However, the economy in the

proposed girder design can be seen when comparing the material utilization factor based on the

bridge deck area. The proposed box girder bridge consumes only 50% of mild reinforcing steel in

comparison to the Windward Viaduct, and only 29% mild reinforcing steel in comparison to the

Islington Avenue Underpass on the basis of bridge deck area.

Table 8-5. Mild reinforcing steel utilization for Islington Ave Underpass (MTO, 1990)

Item SteelVolume

[m3]

SteelMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Top Flange (typ.) 5.49 43100 22.3 25.1

Webs (typ.) 1.63 12800 6.63 7.46

Bottom Flange (typ.) 4.85 38100 19.7 22.2

Solid sections over supportsc

c. assumes equal reinforcing ratio as the typical section, ρ = 0.0088

5.02 39400 20.41 23.0

17.0 133000 69.0 77.8

Table 8-6. Mild reinforcing steel utilization for proposed box girder

Item SteelVolume

[m3]

SteelMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Web Reinforcement 1.17 9190 29.8 8.51

Anchorage Zone Reinforcementc

c. includes bursting reinforcement in slabs

1.90 14900 48.4 13.8

3.07 24100 78.2 22.3

151

Page 168: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

8.4 Longitudinal Post-Tensioning Utilization

The longitudinal post-tensioning strand utilization for the reference bridges and the proposed

segmental box girder have been summarized in Table 8-7, Table 8-8, and Table 8-9 below. For

consistency in steel utilization, the amount of strand consumed has been expressed in terms of mass

of strand per square meter of bridge deck. The Windward Viaduct strand consumption and number

of required anchorage units has been estimated based on one typical span. The Islington Avenue

Underpass strand consumption and number of required anchorages has been determined based on

the full bridge span. The proposed box girder bridge strand consumption and number of required

anchorage units accounts for all anchorages and strand within the span and the abutment.

Table 8-7. Post-tensioning utilization for one span of the Windward Viaduct (Hawaii DOT, 1991)

Item Number of Anchorages

Strand Volume

[m3]

StrandMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Cantilever Tendons 60 2.30 18100 26.0 16.9

Continuity Tendons 32 1.14 8920 12.8 8.35

92 3.44 27000 38.8 25.3

Table 8-8. Post-tensioning utilization for Islington Ave. Underpass (MTO, 1990)

Item Number of Anchorages

Strand Volume

[m3]

StrandMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Draped Tendons 72 13.2 104000 53.8 50.4

Table 8-9. Post-tensioning utilization for proposed cantilever box girder

Item Number of Anchorages

Strand Volume

[m3]

StrandMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Cantilever Tendons 120 1.03 13300 43.2 12.3

Continuity Tendons 32 1.69 13200 42.9 12.2

152 2.72 26500 86.1 24.6

152

Page 169: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The longitudinal post-tensioning strand utilization for the proposed girder is comparable to the

strand utilization of a traditional cantilever constructed bridge per square meter of bridge deck.

Within the single span of the proposed girder, the strand utilization is less than the traditional

cantilever girder; however, this economy is offset by the significant extension of cantilever strand

length that is required to the back face of the abutment. The required number of anchorage units

for the proposed box girder is also significantly increased since the proposed box girder requires

twice the number of anchorage units for cantilever post-tensioning for one span (accounting for

anchors at the stressing face and at the abutment). In comparison to conventional overpass

construction, the proposed box girder consumes approximately 49% post-tensioning strand, but

requires approximately 2.1 times the number of anchorages.

8.5 Transverse Post-Tensioning Utilization

The transverse post-tensioning strand utilization for the reference bridges and the proposed

segmental box girder have been summarized in Table 8-10, Table 8-11, and Table 8-12 below. The

Windward Viaduct top flange is transversely post-tensioned with 4-15mm multi-strand post-

tensioning flat slab tendons. The Islington Avenue Underpass is not designed with a transversely

post-tensioned deck; however, transverse post-tensioning tendons are used in the solid cross

section regions in the vicinity of the supports. The proposed box girder is transversely prestressed

with 15mm pretensioning strands.

Table 8-10. Transverse prestressing steel utilization for one span of the Windward Viaduct (Hawaii DOT, 1991)

Item StrandVolume

[m3]

StrandMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Post-tensioning tendons 1.17 9210 13.2 8.62

Table 8-11. Transverse prestressing steel utilization for Islington Avenue Underpass (MTO, 1990)

Item StrandVolume

[m3]

StrandMass[kg]

Utilization Ratioa

[kg/m3]

a. per cubic meter of concrete

Utilization Ratiob

[kg/m2]

b. per square meter of deck

Post-tensioning tendonsc

c. transverse post-tensioning is only at the supports. deck is not transversely prestressed.

0.346 2720 1.41 1.32

153

Page 170: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Comparing the transversely prestressed deck of the proposed box girder to the Windward

Viaduct, it is observed that the proposed box girder requires approximately 58% more prestressing

strand. The deck slab design for the proposed girder has been carried out using high strength FRC

and thin section components. The major design requirements for the top flange proportioning

includes that: a) the slab thickness satisfies strength requirements for longitudinal global bending,

b) the slab thickness provides a minimum of 40mm top cover to bonded steel and 30mm bottom

cover to bonded steel, c) the transverse stiffness provides an acceptable maximum allowable

deflection based on a chosen reference structure, and d) the SLS cracking and fatigue requirements

are satisfied. This design minimizes the consumption of high strength concrete; however, the

minimization of concrete consumption and the resulting light weight design comes at the cost of

an increased need for transverse prestressing steel.

8.6 Summary

Using the material utilization factors derived above for material consumption per square meter

of bridge deck, a relative savings in materials can be determined for the proposed bridge design.

For example, the savings in material of the proposed girder with respect to the Windward Viaduct

may be expressed as:

The relative material savings are given in Table 8-13 below.

Table 8-12. Transverse prestressing steel utilization for proposed box girder

Item StrandVolume

[m3]

StrandMass[kg]

Utilization Ratioa

[kg/m3]

Utilization Ratiob

[kg/m2]

Pretensioning strands 1.86 14600 47.5 13.6

a. per cubic meter of concreteb. per square meter of deck

Table 8-13. Relative material consumption of proposed box girder

Item Windward Viaduct Islington Avenue Underpass

Concrete 0.44 0.31

Mild Reinforcing Steel 0.50 0.29

Longitudinal Prestressing 0.97 0.49

Transverse Prestressing 1.58 n/a

Utilization FactorproposedUtilization FactorWindward Viaduct-----------------------------------------------------------------------------

154

Page 171: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

155

9.0 CONCLUSIONS

This chapter summarizes the principal findings related to the bridge concept described in this

thesis and the application of external unbonded prestressing for cantilever construction.

9.1 Longitudinal Design

The increase in stress for external unbonded tendons in a cantilever-constructed girder is small

at the ultimate limit state. The increased rotational capacity achieved through the ductile nature of

high-strength fibre reinforced concrete does not significantly improve the economy of prestressing

steel consumption. This is due to the fact that only a small plastic hinge length forms in negative

bending regions. However, economy in prestressing steel consumption is achieved by reducing the

dead load of the girder through the minimization of concrete consumption. The minimization of

concrete consumption is achieved through the design of thin section components. In comparison

to conventional precast cantilever construction, the demand for longitudinal prestressing steel is

similar. However, in comparison to conventional overpass bridge construction, the proposed box

girder requires only 49% of prestressing strand per square meter of bridge deck.

The moment development length for a box girder with external unbonded prestressing tendons

anchored at intermediate corner blisters is much larger than a girder with internal tendons. In

addition to the typical spreading of compressive stresses in a cross section to achieve moment

resistance, a local spreading of forces is first required to deviate the eccentrically applied prestress

force into the box section. While the CAN/CSA-S6-06 requires that continuity tendons should be

anchored one segment beyond where they are theoretically required, external unbonded tendons

should be anchored two segements beyond where they are theoretically required.

A constant depth girder appears to be the most economical solution for the intended

standardized precast box girder cantilevered from fixed abutments. There are four reasons for this

result. First, the reduction in cantilever dead load moment at the fixed end is small due to a decrease

in web depth at midpsan. Second, reducing the internal lever arm to the continuity tendons at

midspan increases the demand for continuity prestressing steel. Third, a larger redistribution of

positive bending moment to the midspan is desirable since it relieves the overturning demand on

the substructure. Lastly, for the intent of precasting and standardization, economy is optimized by

Page 172: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

maintaining repetitive and standard detailing of struts for a constant depth girder. A span-to-depth

ratio of 1:25 appears to be a reasonable girder slenderness for design using nominal 80MPa FRC.

This girder slenderness satisfies strength requirements, deflection requirements, and provides

reasonable visual slenderness.

9.2 Transverse Design

For the design of thin-slabs, a pretensioning transverse prestress concept is necessary.

Pretensioning strands can be placed more compactly in the top slab and the well-distributed

arrangement of prestressing strands provides a more uniform introduction of prestress force at the

cantilever tip. The high tensile strength of FRC facilitates excellent bond strength on 15mm

strands, reducing the transfer length to allow for a sufficient effective prestress at the inside face

of the barrier.

The design of thin slabs using high-strength FRC markedly reduces overall concrete

consumption and reduces dead load; however, the decrease in flange thickness results in an

increased demand for transverse prestressing steel to satisfy serviceability requirements. A

comparison of transverse prestressing utilization showed that the proposed box girder requires 58%

more prestressing strand (per square meter of bridge deck) than a conventional box girder with free

cantilever wings.

9.3 Shear Design

The flow of forces in the webs of box girders prestressed with curtailed tendons at intermediate

anchorages at the top and bottom flanges is complex. The flow of forces can be separated into a

local spreading of forces in the vicinity of the anchorage and a global flow of forces in the cross

section to transfer vertical applied load to the supports. As a consistent methodology for modelling

the flow of forces in the girder, a parallel chord truss model with chords at the elevation of the

flanges is developed for the design of web reinforcement. For the design of thin webs, a single layer

of web reinforcement is placed in the central plane of the web axis.

9.4 Anchorage Zone Design

The flow of forces in anchorage zones for girders prestressed with external unbonded tendons

at corner blisters is fundamentally different from girders prestressed with internal tendons at corner

blisters. Additional mild steel reinforcement is required for girders with unbonded tendons to

156

Page 173: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

deviate the prestress force into the cross section. This anchorage zone steel reinforcement can be

reduced by minimizing the distance from the anchorage to the junction of the flange and web. The

anchorage zone steel reinforcement can also be reduced by increasing the length of the anchorage

blister which allows the deviation to occur over a longer distance. An increase in deviation length,

however, also increases the moment development length for the global bending resistance of the

cross section. The anchorage blister length for continuity tendons is governed by jack clearance

requirements for the tendon stressing operation.

As a practical limitation, the maximum tendon unit size for use in external post-tensioned

bridges is 19 strands of 15mm diameter. Larger tendon units result in an increased demand for local

spreading reinforcement which increases congestion of reinforcing bars.

9.5 Economy

Based on material consumption estimations, a large savings in material for the superstructure

may be realized for overpass bridge construction by adapting a cantilever-constructed box girder

design using external unbonded tendons. In comparison with conventional construction for

overpass bridges in Ontario, the proposed bridge design consumes only 31% of concrete, 29% of

mild reinforcing steel, and 49% of longitudinal post-tensioning strand per square meter of bridge

deck. The proposed bridge, however, requires a significant amount of prestressing steel for the

transverse prestressing system. In comparison to a conventional box girder bridge with a

transversely prestressed concrete deck, 58% more prestressing strand consumption is required for

the proposed box girder.

The material consumption for the proposed box girder superstructure indicates that a savings

in material is possible; however, a full bridge design including abutment and approach materials is

necessary to understand the overall savings of this concept.

9.6 Future Work

The current work has identified the primary design considerations for longitudinal bending,

transverse bending, shear, and anchorage zone detailing. However, additional force effects due to

thermal loads and torsion have not been considered in this work. For final design, creep and

shrinkage should be considered in a more explicit manner. The use of small concrete covers has

been assumed in this study, but this assumption has not been supported with a definitive proof.

Further work is required to validate the use of small concrete covers for high-performance

157

Page 174: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

concrete. Finally, a complete substructure analysis needs to be performed to validate the proposed

box girder concept for cantilever construction.

158

Page 175: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

159

10.0 REFERENCES

Abendroth, R.E., and Greimann, L.F. 1989. Rational Design Approach for Integral Abutment Bridge Piles. Transportation Research Record. 1223: 12-23.

American Association of State Highway and Transportation Officials. 1996. Standard Specifications for Highway Bridges, 16th Ed. Washington, D.C.

American Concrete Institute. 1999. Building Code Requirements for Structural Concrete, ACI 318-99. ACI, Detroit, Mich.

American Concrete Institute Committee 318. 2005. Building Code Requirements for Reinforced Concrete, ACI 318-05. ACI, Detroit, Mich.

American Concrete Institute Committee 363. 1992. Report on High-Strength Concrete ACI 363R-92. ACI, Farmington Hills, Mich., 55 pp.

Beaupre, R.J., Powell, L.C., Breen, J.E., Kreger, M.E. 1990. Deviator Behavior and Design for Externally Post-Tensioned Bridges. External Prestressing in Bridges, SP-120. American Concrete Institute, Detroit. 257-288.

Canadian Precast/Prestressed Concrete Instittute. 2007. Spliced Girder Bridges. <http://cpci.ca/?sc=bs&pn=splicedgirderbridges>.

Canadian Precast/Prestressed Concrete Instittute. 1996. Precast/Prestressed Concrete Infrastructure Construction - Technical Guide. Ottawa, Ontario. 40 pp.

Canadian Standards Association. 2006. Canadian Highway Bridge Design Code, CAN/CSA-S6-06. CSA. Mississauga, Ontario.

Canadian Standards Association. 2004. Design of concrete structures, Standard CAN/CSA A23.3-04. CSA, Mississauga, Ont.

Chao, S., Naaman, A.E., and Parra-Montesinos, G.J. 2006. Bond Behaviour of Strand Embedded in Fiber Reinforced Cementitious Composites. PCI Journal. 51, 6: 56-71.

Collins, M. P., and Mitchell, D. 1997. Prestressed Concrete Structures. Toronto: Response Publications.

Comité Européen du Beton. 1990. High Strength Concrete, State of the Art Report. CEB Bulletin 197, Aug., Lausanne, Switzerland, 61 pp.

Page 176: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

den Uijl, J.A., 1998. Bond Modeling of Prestressing Strand. Bond and Development of Reinforcement, SP-180. American Concrete Institute, Detroit. 145-169.

Dilger, W., and Sherif, A. 2003. Stability of Slender Webs of Prestressed Concrete Box-Girder Bridges. Journal of Bridge Engineering. 8, 3: 138-143.

Dunker, K.F., and Rabbat, B.G. 1992. Performance of Prestressed Concrete Highway Bridges in the United States - the First 40 Years. PCI Journal. 37, 3: 48-64.

Dywidag-Systems International. 2006. Dywidag Post-Tensioning Systems. Brochure.

El-Tawil, S., Severino, E., and Fonseca, P. 2005. Vehicle Collision with Bridge Piers. Journal of Bridge Engineering. 10, 3: 345-353.

ENR. 2003. Nebraska overpass will be rebuilt with fewer piers. Engineering News Record, June 9. 25.

Figg, E.C. 1997. Proposed AASHTO Standards for Segmental Bridges Represent a Growing Market for the Precast Concrete Industry. PCI Journal. 42, 5: 30-31.

Freyermuth, C.L. 1997. AASHTO-PCI-ASBI Segmental Box Girder Standards: A New Product for Grade Separations and Interchange Bridges. PCI Journal. 42, 5: 32-42.

Gauvreau, P. 1993. Ultimate Limit State of Concrete Girders Prestressed with Unbonded Tendons. Institute of Structural Engineering, Swiss Federal Institute of Technology, Report No. 198. Basel, Boston, Berlin: Birkhäuser Verlag.

Graybeal, B.A. 2007. Compressive Behaviour of Ultra-High Performance Fiber-Reinforced Concrete. American Concrete Institute. 86, 2: 146-152.

Habel, K., and Gauvreau, D.P. 2006. Impact Behaviour of Concrete Bridge Deck Slabs with Composite Ultra-High Performance Fiber Reinforced Concrete (UHPFRC) Overlay. Proceedings to the 7th International Conference on Short & Medium Span Bridges. Montreal, Quebec. 10pp.

Haroon, S., Yazdani, N., and Tawfiq, K. 2006. Posttensioned Anchorage Zone Enhancement with Fibre-Reinforced Concrete. Journal of Bridge Engineering. 11, 5: 566-572.

Hota, S., and Naaman, A. 1997. Bond Stress-Slip Response of Reinforcing Bars Embedded in FRC Matrices Under Monotonic and Cyclic Loading. ACI Structural Journal. 94, 5: 525-537.

Homberg, H. 1968. Fahrbahnplatten mit veränderlicher Dicke (Deck Slabs with Variable Depth). Heidelberg, Berlin, Germany, Springer-Verlag, New York, NY.

160

Page 177: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Kahn, L.F., Dill, J.C., Reutlinger, C.G. 2002. Transfer and Development Length of 15-mm Strand in High Performance Concrete Girders. Journal of Structural Engineering. 128, 7: 913-921.

Khayat, K.H., and Aïtcin, P.C. 1993. Silica Fume in Concrete – An Overview. Proceedings of the 4th International Conference on the Use of Fly Ash, Silica Fume, Slag and Natural Pozzolans in Concrete, ACI SP-132. American Concrete Institute, Detroit. 2: 835-872.

LoBuono, J. 1997. Design and Construction of Segmental Concrete Bridges. American Segmental Bridge Institute. Springfield.

Lounis, Z., Mirza, M.S., and Cohn, M.Z. 1997. Segmental and Conventional Precast Prestressed Concrete I-Bridge Girders. Journal of Bridge Engineering. 2, 3: 73-82.

Ma, J., Dehn, F., Tue, N. V., Orgass, M., and Schmidt, D. 2004. Comparative Investigations on Ultra-High Performance Concrete with and without Coarse Aggregates. Proceedings of the International Symposium on Ultra-High Performance Concrete, Kassel, Germany. 205-212.

Marti, P. 1985. Basic Tools of Reinforced Concrete Design. ACI Journal. 82, 1: 46-56.

Mehta, P.K. 1986 Concrete: Structure, Properties and Materials. New Jersey: Prentice-Hall Inc.

Menn, C. 1990. Prestressed Concrete Bridges. Translated and Edited by P. Gauvreau. Basel, Boston, Berlin: Birkhäuser Verlag.

Mizon, D.H. 1997. Second Severn Crossing - Viaduct Superstructure and Piers. Proceedings of the Institute for Civil Engineers - Civil Engineering. 120 (Special Issue): 35-48 pp.

Muller, J., and Gauthier, Y. 1990. Ultimate Behaviour of Precast Segmental Box Girders with External Tendons. External Prestressing in Bridges, SP 120. American Concrete Institute, Detroit. 355-373.

Mutsuyoshi, H., and Witchukreangkrai, E. 2004. Recent Techniques of Prestressed Concrete Bridges in Japan. Proceedings of the Ninth National Convention on Civil Engineering (NCCE9), Thailand.

Naaman, A. 2006. Stresses in External Tendons at Ultimate - Discussion. ACI Structural Journal. 103, 1: 150-152.

Neville, A.M. 1995. Properties of Concrete. Fourth Edition. New Jersey: Prentice-Hall Inc.

Noshiravani, T. 2007. Shear Resistance of Concrete Girders Prestressed with Unbonded Tendons. Master’s thesis, University of Toronto, Toronto.

Ontario Ministry of Transportation. 1990. Islington Avenue Underpass, Design Drawings.

161

Page 178: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Ontario Ministry of Transportation. 2002. Geometric Design Standards for Ontario Highways, Revision Information Sheet.

Paultre, P, and Mitchell, D. 2003. Code Provisions for High-Strength Concrete - an International Perspective. Concrete International. ACI, 25, 5: 76-90.

Podolny, W., and Muller, J. 1982. Construction and Design of Prestressed Concrete Segmental Bridges. New York: John Wiley & Sons.

Popovics, S. 1970. A Review of Stress-Strain Relationships for Concrete. ACI Journal. 67, 3: 243-248.

Post-Tensioning Institute. 2006. Post-Tensioning Manual - Sixth Edition. Phoenix.

Precast/Prestressed Concrete Institute. 1997. Precast Prestressed Concrete Bridge Design Manual. Illinois.

Pucher, A. 1977. Einflussfelder elastischer Platten (Influence Surfaces of Elastic Plates), Fifth Edition. Springer-Verlag, Vienna, Austria, and New York, NY.

Rabbat, B.G., Bhide, S.B. 1999. Evolution of Prestressed Concrete Bridges in the United States. Bridge Talk. National Bridge Research Organization, Lincoln, Nebraska. 3-4.

Ralls, M., and Tang, B. 2003. Laying the Groundwork for Fast Bridge Construction. Public Roads, Federal Highway Administration. 67, 3.

Ramos, G., and Aparlcio, A. 1996. Ultimate Analysis of Monolithic and Segmental Externally Prestressed Bridges. Journal of Bridge Engineering. 1, 1: 10-17.

Rapoport, Irwin. Sept 2006. Highway Bridge Construction Turns to Pre-casting. Daily Commercial News and Construction Record. Markham, Ontario.

Ritter, W. 1899. Die Bauwise Hennebique (Construction Techniques of Hennebique). Schweizerische Bauzeitung, Zürich.

Russell, B., and Burns, N. 1993. Design Guidelines for Transfer, Development, and Debonding of Large Diameter Seven Wire Strands in Pretensioned Concrete Girders. Texas Department of Transportation in Cooperation with Federal Highway Administration, Research Rep. No. 1210-5F.

Schlaich, J., Shäfer, K., and Jennewein, M. 1987. Towards a Consistent Design of Structural Concrete. PCI Journal. 32, 3: 74-150.

SIA. 2003. Concrete Structures, SIA 262. Swiss Society of Engineers and Architects. SIA Zurich.

162

Page 179: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Sivakumar, A., and Santhanam, M. 2007. A Quantitative Study on the Plastic Shrinkage Cracking in High Strength Hybrid Fibre Reinforced Concrete. Cement & Concrete Composites. 29: 575-581.

Sotelino, E.D., and Liu, J. 2004. Simplified Load Distribution Factor for Use in LRFD Design. Federal Highway Administration Final Report SPR-2477. Indiana.

State of Hawaii Department of Transportation. 1991. Windward Viaduct, Interstate Route H-3, Design Drawings.

Straight Crossing. 1994. Northumberland Straight Crossing, Design Drawings.

Susetyo, J. 2007. Unpublished Material. University of Toronto, Toronto, Ontario.

Tanis, J. 2003. Bras de la Plaine Bridge, Reunion Island, France. Structural Engineering International. 13, 4: 259-262.

Tinkess, Terry. Sept 2007. Ottawa Deploys New Bridge Using Rapid Technology. Daily Commercial News and Construction Record. Markham, Ontario.

Turmo, J., Ramos, G., and Aparicio, A. C. 2006. Shear Behaviour of Unbonded Post-Tensioned Segmental Beams with Dry Joints. ACI Structural Journal. 103, 3: 409-417.

Virlogeux, M. 1993. External Prestressing Historical and Modern Applications. External Prestressing in Structures. AFPC, Saint-Rémy-lès-Chevreuse, France. 13-35.

VSL. 2007. VSL Technical Data and Dimensions. Brochure.

Wollmann, G.P., Kreger, M.E., Roberts-Wollmann, C.L., and Breen, J.E. 1993. External Tendon Anchorage in Diaphragms and Intermediate Slab Blisters. Proceedings to the Workshop on Behaviour of External Prestressing in Structures, Saint-Remy-Les-Chevreuse, France. 113-121.

Xu, P., Yi, C., Fan, C.M., and Joshi, R.C. 1998. Performance of Fiber Reinforced Concrete with Respect to Frost Resistance: A Case Study. Proceedings of the International Conference on Cold Regions Engineering (ASCE). 479-488.

Zia, P., and Mostafa, T. 1977. Development Length of Prestressing Strands. Journal of the Prestressed Concrete Institute. 22, 5: 54-65.

163

Page 180: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

164

APPENDIX A DESIGN DRAWINGS

Page 181: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

165

Page 182: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

166

Page 183: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

167

Page 184: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

168

Page 185: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

169

Page 186: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

170

Page 187: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

171

Page 188: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

172

Page 189: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

173

Page 190: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

174

Page 191: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

175

APPENDIX B MATERIAL TESTING

This appendix describes the concrete mix design adapted from Susetyo (2007) and the

compression cylinder test results obtained for the 7-day strength tests and the 28-day compressive

behaviour tests. The stress-strain relationship obtained from these material tests was implemented

as the design material model for the proposed bridge concept.

B.1 Reference Concrete Mix

A nominal 80Mpa fibre-reinforced concrete mix design developed by Susetyo (2007) was

adapted for the specimen design. The concrete mix has a water-to-cement ratio of 0.27 and uses

high silica fume (HSF) cement. The hooked steel fibre volume ratio is 1.5%. The coarse and fine

aggregates are measured by weight in the saturated surface dry (SSD) condition, where the

aggregates neither absorb water from nor contribute water to the concrete mixture. The total

volume of concrete required for one test batch (for casting of two cylinders) is 0.013m3, which

includes a 20% allowance for spillage and losses. Scaling the proportions directly for the reference

mix and the batch mix, the mix proportions are described in Table B-1 below:

Since the reference design mix is calibrated for fine and coarse aggregate in the SSD condition,

the sand and limestone aggregates in the Galbraith Building concrete lab were tested for moisture

content and absorption rates. The moisture content test results are given below in Table B-2:

Table B-1. Reference concrete mix quantities (Susetyo, 2007) and scaled batch quantities

MaterialMeasured

Unit Ref. QuantityRef. Equivalent

VolumeScaled Batch

Quantity

HSF Cement kg 600 0.191 7.20

Sand (SSD) kg 1133 0.419 13.59

10mm limestones (SSD) kg 802 0.292 9.62

Water L 162 0.162 1.94

Water Reducer mL 4200 - 50.40

Superplasticizer mL 9600 - 115.20

Entrapped Air 0.02

Total Volume m3 1.083

Page 192: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

The water absorption rate of the sand and the stone are 0.99% and 1.22%, respectively. Based

on the measured moisture content and absorption rates of the fine and coarse aggregate, the batch

quantities were adjusted to match the designed water-to-cement ratio of 0.27. The adjusted batch

quantities are given below in Table B-3:

B.2 Material Batching and Preparation

Six compression cylinders 150mm diameter by 300mm in height were desired for testing the

FRC behaviour - three cylinders were intended to measure the 7-day strength, and three cylinders

were intended to measure the 28-day compressive stress-strain response. The small mixer in the

Galbraith Building concrete lab was used for the batching and mixing of the concrete. This required

three separate batches (for casting of two cylinder specimens each) for mixing due to the volume

limitation of the mixer. Specimens were cast in the cylinder moulds and left to harden for 1 day,

and then the cylinders were demoulded and placed in the moist-curing room to cure. The cylinders

were removed temporarily after 6 days for end grinding, and then replaced in the moist-curing

room.

Table B-2. Measured moisture content of fine and coarse aggregates

Sand Stone

Tray 200g 195g

Tray + Material (wet) 1590g 1700g

Tray + Material (oven dry) 1580g 1685g

Moisture Content 0.72% 1.01%

Table B-3. Adjusted batch quantities for compression cylinders

MaterialMeasured

Unit Single BatchTotal Material

(3 Batches)

HSF Cement kg/m3 7.20 21.60

Sand kg/m3 13.55 40.66

10mm limestones kg/m3 9.60 28.80

Water L/m3 2.00 6.00

Water Reducer mL/m3 50.40 151.20

Superplasticizer mL/m3 115.20 345.60

176

Page 193: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

B.3 7-day Compressive Strength Tests

One cylinder from each batch was tested for 7-day strength. No displacement measurements

were taken for the load deflection response. The cylinders were loaded at a constant rate of 4kN/s.

The test cylinders were observed to have a slightly lower compressive 7-day strength than expected

based on previous results obtained by Susetyo. The 7-day strength results are summarized below

in Table B-4, where the equivalent stress has been calculated based on the gross cross sectional

area:

B.4 28-day Compressive Strength Tests

The remaining cylinders from each batch was tested for 28-day strength. Results from the

analog plot based on the displacement of the load head of the test machine have been presented in

Chapter 2. The peak stress and peak strain results are tabulated below in Table B-5:

Table B-4. 7-day compressive strength results

Cylinder Peak Load (kN) Peak Stress (Mpa)

FRC 1b 979 55.4

FRC 2b 1164 65.9

FRC 3b 999 56.5

Average 1047 59.3

Table B-5. 28-day compressive strength results

Cylinder f’c (Mpa) (mm/mm) Ec (Mpa)

FRC 1a 72.1 0.002139 48200

FRC 1b n/a n/a n/a

FRC 1c 75.9 0.002050 39160

Average 74.6 0.00209 43680

ε’c

177

Page 194: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

178

APPENDIX C CAN/CSA-S6-06 SHEAR PROVISIONS

The sectional design approach of the CAN/CSA S6-06 shear provisions calculates the shear

resistance of concrete members based on an idealized cross section that is composed of a flexural

compression flange, a web, and a flexural tension flange. The total shear resistance of the web is

due to the additive contribution of the shear resistance provided by tensile stresses in concrete, Vc,

shear resistance provided by the transverse steel reinforcement, Vs, and the shear resistance

provided by the component of force in a prestress tendon acting in the direction of applied shear,

Vp.

(C-1)

The factored shear resistance due to the contribution of tensile stresses in the concrete is given

as:

(C-2)

where is the factor that accounts for the shear resistance of cracked concrete, fcr is the

cracking stress of concrete, bv is the width of the structural component resisting shear, and dv is

taken as 0.9d. The cracking stress is limited to 3.2MPa since observed tests have shown no

increased benefit in shear resistance with concrete strengths above 65MPa (CSA S6.1-06).

The factored shear resistance due to the contribution of transverse steel reinforcement is given

as:

(C-3)

where Av is the cross sectional area of each stirrup, fy is the yield stress of reinforcing steel,

is the angle of compressive stresses in the web, and s is the spacing of stirrups.

The angle of compressive stress, , and the ability of cracked concrete to transfer stress across

a crack, , are determined through convenient empirical equations based on analytical results of

MCFT elements. These parameters are given as:

(C-4)

Vr Vc Vs Vp+ +≥

Vc 2.5βφcfcrbvdv=

β

VsφsAvfydv θcot

s----------------------------------=

θ

θ

β

β 0.41 1500εx+--------------------------⎝ ⎠

⎛ ⎞ 13001000 sze+-------------------------⎝ ⎠

⎛ ⎞=

Page 195: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

(C-5)

where sze is the crack spacing parameter and is defined as . The parameter

accounting for nominal size of coarse aggregate, ag, is taken as zero for high strength concrete

above 70MPa and the spacing factor, sz, is taken as dv for girders with no longitudinal steel in the

webs.

Most importantly, the parameters required for the calculation of shear resistance are dependent

on the longitudinal strain in the concrete member taken at the mid-height of the section, due to

the combined actions of shear, moment, and prestressing. Conservative assumptions for the

calculation of are built into the code for a safe approximation that can be calculated in lieu of a

more accurate iterative calculation, which are described in cl.8.9.3.8. The strain at mid-height is

given as:

(C-6)

where Mf, Vf, and Nf are applied sectional forces due to moment, shear, and axial force,

respectively, and taken as positive quantities. The area of prestressing steel is given as Aps, and fpe

represents the long-term effective prestress for unbonded tendons, taken as 0.6fpu in this study. For

compressive mid-height strains, the denominator is modified to account for the axial stiffness of

the concrete flange on the flexural tension side.

The parallel chord model used for the CAN/CSA-S6-06 shear provisions does not account for

an arching contribution due to the flow of inclined compression forces following the funicular

shape, but rather it accounts for the presence of prestressing by simply offsetting the force in the

tension flange by including the compression force due to all prestressing tendons that are on the

flexural tension side of the girder. This is represented by the Apsfpe term that appears in the

numerator of the equation for .

θ 29 7000εx+( ) 0.88sze

2500------------+⎝ ⎠⎛ ⎞=

35sz 15 ag+( )⁄( )

εx

εx

εxMf dv⁄ Vf Vp– 0.5Nf Apsfpe–+ +

2 EsAs EpAps+( )------------------------------------------------------------------------------------=

εx

179

Page 196: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

180

CURRICULUM VITAE

Brent Tyler Visscher

Born September 6, 1981 in St. Thomas, Ontario, Canada

Education

Master of Applied Science (2008)University of Toronto, Toronto, Ontario, Canada

Bachelor of Engineering Science (2005)The University of Western Ontario, London, Ontario, Canada

Honours and Awards

2006, Excellence in Graduate Studies Award2005, Natural Sciences and Engineering Research Council, Canada Graduate Scholarship2005, Ontario Graduate Scholarship (Declined)2005, Dr. James A. Vance Gold Medal in Civil Engineering2005, 1st Place Civil and Environmental Engineering Thesis2005, 1st Place Team of the City of London Bridge Design Competition2004, London and District Construction Association Award2004, Dr. L. Stuart Lauchland Scholarship2003, Natural Sciences and Engineering Research Council Undergraduate Student Research Award2000, Governor General's Bronze Medal

Related Work Experience

September 2005 to December 2007Teaching Assistant, University of Toronto, Toronto, Canada

September 2003 to August 2004Engineer-In-Training, Buckland & Taylor Ltd. Bridge Engineers, North Vancouver, Canada

May 2003 to August 2003Project Researcher, Alan G. Davenport Wind Engineering Group, London, Canada

Page 197: INNOVATIVE PRE-CAST CANTILEVER CONSTRUCTED BRIDGE CONCEPT by Brent Tyler

Publications

Refereed Journal Papers

Visscher, B.T., Kopp, G.A. 2007. Trajectories of Roof Sheathing Panels Under High Winds, Journal of Wind Engineering and Industrial Aerodynamics, 95, 8: 697-713.

Conference Proceedings

Visscher, B.T., Kopp, G.A. 2005. Wind Tunnel Measurements of Trajectories of Roof Sheathing Panels Under High Winds, Proceedings of the 10th Americas Conference on Wind Engineering, Baton Rouge, USA (CD-ROM).

Other Publications - Reports

Visscher, B.T. 2005. Trajectories of Roof Sheathing Panels Under High Winds. Graduation Thesis, Department of Civil and Environmental Engineering, The University of Western Ontario, London, Canada.

Visscher, B., Kopp, G.A., and Vickery, P.J. 2004. Wind Loads on Houses: Destructive Model Testing of a Residential Gable Roofed House, Institute for Catastrophic Loss Reduction Research Paper 37, London, Canada.

Visscher, B. 2003. Destructive model of a low-rise residential gable roofed house, Alan G. Davenport Wind Engineering Group Rep. No. BLWT-1-2003, The University of Western Ontario, London, Canada.

181