You are on page 1of 366

Innovative Systems for Arch Bridges using

Ultra High-Performance Fibre-Reinforced Concrete

by

Jason Angeles Salonga

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Graduate Department of Civil Engineering

University of Toronto

© Copyright by Jason Angeles Salonga (2010)


Innovative Systems for Arch Bridges using Ultra High-Performance Fibre-Reinforced Concrete

Jason Angeles Salonga

Doctor of Philosophy

Department of Civil Engineering

University of Toronto

2010

Abstract

In this thesis, new design concepts for arch bridges using ultra high-performance fibre-reinforced concrete

are developed for spans of 50 to 400 m. These concepts are light-weight and efficient, and thus have the

potential to significantly reduce the cost of construction. Lightness is achieved by the thinning of structur-

al components and the efficient use of precompression in the arch, rather than by the decrease of bending

stiffness. Using the advanced properties of the material, the design concepts were shown to reduce the

consumption of concrete in arch bridges by more than 50% relative to arches built using conventional con-

crete technology. In addition to span length, other design parameters including span-to-rise ratio and

deck-stiffening were considered, resulting in a total of seventy-two design concepts.

Other important contributions made in this thesis include: (1) the development of a simple analytical

model that describes the transition of shallow arches between pure arch behaviour and pure beam beha-

viour, (2) a comprehensive comparative study of 58 existing concrete arch bridges that characterizes the

current state-of-the-art and serves as a valuable reference design tool, and (3) the development and experi-

mental validation of general and simplified methods for calculating the capacity of slender ultra high-per-

formance fibre-reinforced concrete members under compression and bending. The research presented in

this thesis provides a means for designers to take full advantage of the high compressive and tensile

strengths of the concrete and hence to exploit the economic potential offered by the material.

ii
Acknowledgments

I would like to thank, first and foremost, Professor D. P. Gauvreau, for all the guidance and encouragement

he has given me as a supervisor, teacher, and mentor over the past five years. His passion for bridges, with

respect to their design, construction, and aesthetics has certainly been contagious and will forever be an

inspiration to me.

I would also like to thank Professor F. J. Vecchio, Professor P. C. Birkemoe, Professor S. A. Sheikh, and Pro-

fessor V. Sigrist for serving on my Ph.D. Examination Committee and for reviewing this thesis.

Financial assistance during the majority of my graduate studies was provided in large part by the Natural

Sciences and Engineering Research Council of Canada, and the University of Toronto. For this, I am

sincerely grateful.

Many others have given their time selflessly to help me along the way. Experimental work would not have

been possible without the help of my colleagues Ivan Wu, Jimmy Susetyo, Kris Mermigas, Brent Visscher,

Davis Doan, Lulu Shen, Nabil Mansour, Boyan Mihailov, James Liu, and Serguei Bagrianski, and laborat-

ory staff Joel Babbin, Renzo Bassett, Giovanni Buzzeo, and John MacDonald. I am also indebted to others

who have accompanied and helped me along the way, including: Billy Cheung, Negar Elhami Khorasani,

Jeff Erochko, Eileen Li, Jamie McIntyre, Michael Montgomery, Talayeh Noshiravani, Sandy Poon, Carlene

Ramsay, Nick Zwerling, and especially Lydell Wiebe. A special thanks is dedicated to Catherine Chen who

assisted me over two summers in compiling the database of concrete arch bridges.

Last, I thank my wife Sarah, my parents, Danilo and Elizabeth, and my siblings, Michael, Elaine, and

Rodell, for their steadfast love and support.

iii
Table of Contents
Abstract, ii

Acknowledgments, iii

Table of Contents, iv

List of Figures, x

List of Tables, xvii

List of Symbols, xix

List of Uncommon Terms, xxix

Chapter 1. Introduction, 1

1.1 Motivation, 1

1.2 State-of-the-art, 5

1.2.1 Highway girder bridges, 5

1.2.2 Pedestrian arch bridges, 7

1.2.3 Highway arch bridges, 8

1.3 Objectives and content of thesis, 12

Chapter 2. Ultra High-Performance Fibre-Reinforced Concrete, 15

2.1 Material testing at the University of Toronto, 15

2.1.1 Mix design, 17

2.1.2 Batching and casting, 17

2.1.3 Material behaviour in compression, 18

2.1.4 Material behaviour in tension, 26

2.1.5 Shrinkage behaviour of the material, 30

2.1.6 Creep behaviour of the material, 31


iv
Chapter 3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns, 32

3.1 Response of eccentrically loaded columns, 33

3.2 General analysis method, 37

3.2.1 Methodology, 37

3.2.2 Calculating moment curvature diagrams, 38

3.2.3 Calculating column deflection curves, 42

3.2.4 Calculating member capacity interaction diagrams, 46

3.2.5 Calculating load-deflection response of a given column and eccentricity of load, 52

3.3 Overview of computer program, 54

3.4 Experimental validation, 56

3.4.1 Batching, casting, and curing, 57

3.4.2 Setup and instrumentation, 57

3.4.3 Load testing and failure, 60

3.4.4 Test results and discussion, 62

3.4.5 Validation of general method, 66

3.5 Simplified design method, 69

3.5.1 Menn’s simplified design method, 69

3.5.2 Simplified method for the design of slender members, 75

v
Chapter 4. Statical Analysis of Arch Bridges, 82

4.1 Relevant literature, 82

4.2 Definition of structural system, 84

4.3 Comparison of three-hinged, two-hinged, and fixed arches, 85

4.3.1 Flexural buckling of arches, 86

4.3.2 Setting the shape of the arch, 89

4.3.3 Response of arches to nonuniform, vertical loads, 91

4.3.4 Effects of restrained deformation on arches, 96

4.3.5 Other load cases, 98

4.4 Structural response of fixed arch systems, 98

4.4.1 Simplified statical system, 99

4.4.2 Force method for non-shallow arches, 107

4.4.3 Force method for shallow arches, 110

4.4.4 Flexural buckling of shallow arches, 118

4.4.5 Second-order analysis: geometric nonlinearity, 120

4.4.6 Second-order analysis: material nonlinearity, 123

4.4.7 Fixed system moments, 126

4.5 Critical load combinations, 127

4.5.1 Long-term loads, 130

4.5.2 Short-term loads, 134

4.5.3 Second-order analysis of combined long and short-term loads, 135

4.5.4 Maximum sectional forces at quarter-points, 138

4.6 Summary of arch analysis, 140

4.7 Critique of Billington’s arch stress analysis, 144

vi
Chapter 5. Comparative Study of 58 Concrete Arch Bridges, 153

5.1 Objective and utility, 153

5.2 Database of concrete arch bridges, 156

5.2.1 Sources of information, 156

5.2.2 Acceptance criteria, 157

5.2.3 Drawing database, 158

5.2.4 Recorded data, 160

5.3 Trends in recorded data and geometric ratios, 165

5.3.1 Geographic location, 166

5.3.2 Methods of construction, 167

5.3.3 Structural depths of arches and decks, 171

5.3.4 Moments of inertia of arches, decks, and columns, 174

5.3.5 Arch slenderness ratios, 179

5.3.6 Span-to-rise ratios, 180

5.3.7 Modified slenderness ratios, 184

5.3.8 Equivalent slab thickness, 187

5.3.9 Historical evolution, 189

5.3.10 Horizontal reactions caused by dead load, 190

5.3.11 Effects of combined dead and live loads, 191

5.4 Summary of observed trends, 194

vii
Chapter 6. New Concepts for Arch Bridges, 197

6.1 Parametric design study, 198

6.1.1 Loads, load factors, and material resistance factors, 201

6.1.2 Longitudinal proportioning of self-stiffened arch systems, 205

6.1.3 Longitudinal proportioning of deck-stiffened arch systems, 207

6.1.4 Longitudinal proportioning of partially deck-stiffened arch systems, 210

6.1.5 Wall slenderness, 212

6.1.6 Proportioning of spandrel columns, 215

6.2 Preliminary design concepts, 217

6.2.1 Parametric design solutions, 217

6.2.2 Sample design calculations including shear, 227

6.2.3 Concept 80, 228

6.3 Discussion of design concepts, 237

6.3.1 Proportions of self-stiffened arch concepts, 237

6.3.2 Proportions of deck-stiffened arch concepts, 239

6.3.3 Proportions of partially deck-stiffened arch systems, 240

6.3.4 Design interaction diagrams, 241

6.3.5 Proportions of spandrel columns, 243

6.3.6 Material efficiency, 243

6.3.7 Distribution of concrete volume, 245

6.3.8 System slenderness and shallowness, 247

6.3.9 Structural demands, 252

6.3.10 Summary of results and conclusions, 257

6.4 Recommendations, 259


viii
Chapter 7. Conclusions and Summary, 266

7.1 Conclusions, 266

7.2 Material behaviour, 269

7.3 Sectional behaviour, 271

7.4 Slender column behaviour, 273

7.5 Arch system behaviour, 275

7.6 Trends observed from existing concrete arch bridges, 280

7.7 Innovative systems for concrete arches, 282

References, 288

Appendices, 294

A. Integrated Database of Concrete Arch Bridges, 295

B. Details of Computer Program QULT, 299

C. Slender Column Drawings, 305

D. Design Calculations for Concept 80, 315

Curriculum Vitae, 335

ix
List of Figures
Figure 1-1. Flow of forces and sectional forces in arches and beams, 2

Figure 1-2. Construction of Sandö Bridge (top) and Colorado River Bridge (bottom), 3

Figure 1-3. Cross-section of ultra high-performance fibre-reinforced concrete girder used in Virginia Bridge, 6

Figure 1-4. Sun-Yu Pedestrian Bridge, 7

Figure 1-5. Reinforcement-free micro-anchorage detail, 8

Figure 1-6. Elevation, cross-section, and joint detail views of WILD Bridge concept, 9

Figure 1-7. Bakar Bridge concept, 10

Figure 1-8. Bridge concept for a 1000 m crossing, 11

Figure 2-1. Mortar mixer and concrete bucket, 18

Figure 2-2. Load versus displacement data of cylinders from Set 1, tested 27 days after casting, 19

Figure 2-3. Load versus displacement data of cylinder from Set 2, tested 60 days after casting, 19

Figure 2-4. Ultra high-performance fibre-reinforced concrete cylinder at failure, 22

Figure 2-5. Compressive strength versus age of concrete after casting, 23

Figure 2-6. Material model in compression, 24

Figure 2-7. Tangent modulus of elasticity versus compressive strength of tested cylinders, 25

Figure 2-8. Modulus of rupture test setup and load-displacement results, 26

Figure 2-9. Material model in tension, 28

Figure 2-10. Shrinkage strains as a function of days after casting, 30

Figure 2-11. Creep strain of an untreated ultra high-performance fibre-reinforced concrete specimen over time, 31

Figure 3-1. Structural model for eccentrically loaded columns, 33

Figure 3-2. Load deflection and sectional response of a slender column given different analytical assumptions, 35

Figure 3-3. Contours of equal curvature and corresponding moment-curvature diagrams, 40

x
Figure 3-4. Numerical integration procedure, 43

Figure 3-5. Column deflection curves and sectional responses of mid-length sections for various column lengths, 47

Figure 3-6. Member capacity interaction diagrams for various slenderness ratios, 52

Figure 3-7. Overview of computer program QULT, 55

Figure 3-8. Column drawings, details, and measurements, 56

Figure 3-9. Placing concrete directly into plywood forms, 57

Figure 3-10. Clamped steel end plate (left), spherical bearing at base (right), 58

Figure 3-11. Installing specimen with forms into test frame, 59

Figure 3-12. Position of cohesive laser radar scanner targets at zero load, 59

Figure 3-13. Creep behaviour of Specimen C, 61

Figure 3-14. Failure surface of Specimen C, 61

Figure 3-15. Load-deflection response of specimens and photographs at each load stage, 63

Figure 3-16. Load test results of four slender columns, 65

Figure 3-17. Comparison of load test results and general method predictions, 67

Figure 3-18. Conventional reinforced concrete box section, 70

Figure 3-19. Conventional strain limits and Menn’s reduced strain limits, 71

Figure 3-20. Interaction and moment-curvature diagrams for a 30 MPa concrete box, 72

Figure 3-21. Secant stiffness at yield as a function of axial force, 74

Figure 3-22. Ultra high-performance fibre-reinforced concrete box section, 76

Figure 3-23. Sectional analysis of an ultra high-performance fibre-reinforced concrete box, 76

Figure 3-24. Reduced interaction diagram, 79

Figure 4-1. Components of a typical arch bridge, 84

Figure 4-2. Location of hinges in typical arches, 85

xi
Figure 4-3. Arch and analogous column buckling models, 87

Figure 4-4. Comparison of effective arc length factors with stability analysis results obtained by Rambøll (1944), 88

Figure 4-5. Pressure lines for uniformly distributed and concentrated loads, 90

Figure 4-6. Bending moments caused by concentrated loads in fixed, two-hinged, and three-hinged arches, 93

Figure 4-7. Bending moments caused by distributed loads in fixed, two-hinged, and three-hinged arches, 94

Figure 4-8. Bending moment influence lines for fixed, two-hinged, and three-hinged arches, 95

Figure 4-9. Comparison of bending moment envelopes and variations in depth along the arch, 96

Figure 4-10. Bending moments caused by restrained deformations in fixed, two-hinged, and three-hinged arches, 97

Figure 4-11. Moment distribution factors for the Colorado River Bridge, 100

Figure 4-12. Arch flexural rigidity, 101

Figure 4-13. Classification of fixed arch systems, and their respective simplified and analogous structural models, 104

Figure 4-14. Stiffened arch model and moments and axial forces caused by unit redundant forces, 105

Figure 4-15. Effect of arch, deck, and column flexural rigidities on the distribution of bending moment, 106

Figure 4-16. Force method solutions for shallow arches, 113

Figure 4-17. Transition between fixed arch and fixed beam, 116

Figure 4-18. Square of slope of parabolic arches at their springing lines, 118

Figure 4-19. Buckling of shallow arches, 120

Figure 4-20. Analytical model for second-order analysis of arches, 122

Figure 4-21. Short-term and long-term load combinations, 127

Figure 4-22. Flat hydraulic jacks at crown of Rio Parana Bridge, 132

Figure 4-23. Free body diagrams of cantilevered arch during crown jacking, 132

Figure 4-24. Influence of deck-stiffening on flexural stresses in the arch and deck, 145

Figure 4-25. Schwandbach Bridge and Broadway Bridge, 145

xii
Figure 4-26. Billington’s graph of arch stress versus stiffness ratio between arch and deck, 147

Figure 4-27. Swiss concrete arches designed by Menn, 150

Figure 4-28. Modified flexural stress analysis, 151

Figure 5-1. Distribution of span lengths of bridges in database, 158

Figure 5-2. Drawings of bridges in database, 159

Figure 5-3. List of sources ordered by bridge ID, 160

Figure 5-4. World map showing location of bridges in database, 166

Figure 5-5. Year of completion and method of construction versus span length, 167

Figure 5-6. Sandö Bridge construction using timber scaffolding, 167

Figure 5-7. Gladesville Bridge construction using steel centering, 168

Figure 5-8. Steel pilot truss method of construction, 169

Figure 5-9. Colorado River Bridge construction using tower and stay method, 169

Figure 5-10. Krk Bridge construction using trussed cantilever method, 170

Figure 5-11. Argentobel Bridge construction using the cantilever rotation method, 171

Figure 5-12. Section type and structural depths of arches and decks versus span length, 172

Figure 5-13. Structural depth of arch and deck versus span length, 173

Figure 5-14. Span-to-depth ratios versus span length, 174

Figure 5-15. Moment of inertia of arch, deck, and system versus span length, 175

Figure 5-16. System radius of gyration versus span length, 177

Figure 5-17. Ratio of arch inertia to system inertia versus span length, 178

Figure 5-18. Ratio of average column inertia to system inertia versus span length, 179

Figure 5-19. System slenderness ratio versus span length, 180

Figure 5-20. Ground and arch profiles of selected bridges in the database, 181

xiii
Figure 5-21. Variations of span-to-rise ratio based on the Colorado River Bridge, 182

Figure 5-22. Span-to-rise ratios versus span length, 183

Figure 5-23. Modified slenderness ratio versus span length, 184

Figure 5-24. Efficiency threshold curve based on average system radius of gyration trend line, 185

Figure 5-25. Efficiency threshold curve based on subjective high estimate of system radius of gyration, 187

Figure 5-26. Equivalent slab thickness by volume versus span length, 188

Figure 5-27. Equivalent slab thickness, span length, and year of construction, 189

Figure 5-28. Normalized horizontal reaction versus span length, 190

Figure 5-29. Eccentricity of resultant compressive force caused by live load bending moments, 191

Figure 5-30. Normalized live load eccentricity versus span length, 193

Figure 6-1. Design stress-strain curves, 199

Figure 6-2. Superimposed dead load and traffic lanes, 202

Figure 6-3. Proportioning of self-stiffened arch systems, 205

Figure 6-4. Proportioning of deck-stiffened arch systems, 208

Figure 6-5. Buckling of slender arches over interior spans, 209

Figure 6-6. Proportioning of partially deck-stiffened arch systems, 210

Figure 6-7. Plate buckling in thin-walled concrete box members, 213

Figure 6-8. Spandrel column load capacity design curves, 216

Figure 6-9. Generalized column load capacity curves, 217

Figure 6-10. Interaction diagrams of all self-stiffened arch concepts, 223

Figure 6-11. Interaction diagrams of all deck-stiffened arch concepts, 224

Figure 6-12. Interaction diagrams of all partially deck-stiffened arch concepts, 225

Figure 6-13. Sections of selected design concepts and existing arch bridges, 226

xiv
Figure 6-14. General arrangement drawing, 230

Figure 6-15. Longitudinal section and anchorage diaphragm section, 231

Figure 6-16. Proposed match-casting method for abutting end ribs, 233

Figure 6-17. Proposed method for precasting segments, 235

Figure 6-18. Renderings of precast segments proposed in Concept 80, 236

Figure 6-19. Comparison of simplified and general method interaction diagrams, 242

Figure 6-20. Comparison of equivalent slab thicknesses of new arch concepts and existing arches in database, 245

Figure 6-21. Comparison of span-to-system-depth ratios, 247

Figure 6-22. Comparison of normalized system moment of inertia and system radius of gyration, 249

Figure 6-23. Comparison of system slenderness ratios, 250

Figure 6-24. Comparison of modified slenderness ratios, 251

Figure 6-25. Comparison of extra weight factors, 253

Figure 6-26. Comparison of governing moments at ultimate limit states, 255

Figure 6-27. Comparison of first-order live load portions of governing moments, 256

Figure A-1. Arch database: drawings, maps, and references, 296

Figure A-2. Arch database: geometry, structural types, and ratios, 297

Figure A-3. Arch database: geometrical trends, 298

Figure C-1. Slender column tests, 307

Figure C-2. End plate detail, 308

Figure C-3. Pin connection detail, 309

Figure C-4. Top and bottom tongues, 310

Figure C-5. Formwork, 311

Figure C-6. Formwork sections, 312

xv
Figure C-7. Instrumentation, 313

Figure C-8. As-built measurements, 314

xvi
List of Tables
Table 2-1. Mix design for ultra high-performance fibre-reinforced concrete, 17

Table 2-2. Results of concrete cylinder tests, 20

Table 2-3. Compressive strengths of cylinders organized by batch, age, and curing regime, 22

Table 4-1. Relative comparison between fixed, two-hinged, and three-hinged arches, 86

Table 4-2. Expressions for axial strains and flexural rigidities in arches, 97

Table 4-3. Bending moment diagrams for fixed arch systems, 109

Table 4-4. Calculation of maximum sectional forces at midspan, 128

Table 4-5. Calculation of maximum sectional forces at quarter-point, 128

Table 5-1. Concrete arch bridge data compiled from reference databases, 155

Table 5-2. Journal article search and results, 157

Table 5-3. Location and global geometry of arch bridges in database, 161

Table 5-4. Sectional geometry of arch bridges in database, 163

Table 5-5. Geometric ratios of arch bridges in database, 164

Table 5-6. Number of bridges in the database by country, 166

Table 5-7. Load parameter values, 193

Table 6-1. Design concept ID numbers and design matrix, 198

Table 6-2. Design values for material properties, 200

Table 6-3. Design live loads, 203

Table 6-4. Design load factor combinations, 204

Table 6-5. Sequence of trial dimensions used in the design of hollow box sections, 207

Table 6-6. Summary of quantities to be designed for parametric design study, 212

Table 6-7. Maximum plate widths for thin ultra high performance fibre-reinforced concrete plates, 215

xvii
Table 6-8. Proportions of design concepts, 219

Table 6-9. Geometric ratios and structural demands of design concepts, 220

Table 6-10. Sample comparison matrix, 237

Table 6-11. Proportions of self-stiffened arch concepts, 238

Table 6-12. Proportions of deck-stiffened arch concepts, 239

Table 6-13. Proportions of partially deck-stiffened arch concepts, 240

Table 6-14. Proportions of spandrel columns among all concepts, 243

Table 6-15. Equivalent slab thicknesses based on volume of longitudinal concrete among all concepts, 244

Table 6-16. Quantity of longitudinal prestressing strands in deck girder among all concepts, 245

Table 6-17. Relative volume of arch and deck among all concepts, 246

Table 6-18. Relative volume of spandrel columns to total volume, 246

Table 6-19. Span-to-system-depth ratios among all concepts, 247

Table 6-20. Normalized system moment of inertia among all concepts, 248

Table 6-21. System radius of gyration among all concepts, 248

Table 6-22. System slenderness ratios among all concepts, 250

Table 6-23. Modified slenderness ratios among all concepts, 251

Table 6-24. Maximum deflections at ultimate limit states among all concepts, 252

Table 6-25. Extra weight factors among all concepts, 253

Table 6-26. Governing live load cases among all concepts, 254

Table 6-27. Maximum flexible system moments at ultimate limit states among all concepts, 254

Table 6-28. Portions of governing moments caused by first-order live loads among all concepts, 256

Table 6-29. Most common geometrical quantities used among design concepts, 257

Table B-1. Description of program functions and calculation steps, 301

xviii
List of Symbols
§ reference to clause

! proportion of flexible system moment carried by arch

!D load factor for dead load

!K load factor for restrained deformation

!L load factor for live load

!T coefficient of thermal expansion

!V extra weight factor, equal to total weight of system divided by weight of continuous longitudinal concrete

" arch-beam parameter

# unit weight of concrete

$ deflection of member; deformation of member; total second-order deflection

$0 first-order deflection

$arch deflection of arch

$deck deflection of deck

$i0 deformations of primary system caused by applied loads or dimensional changes

$ij flexibility coefficients, or deformations of primary system caused by unit redundant forces

$eq,0 long-term, second-order deflection expressed as an equivalent short-term, first-order deflection

$long long-term, second-order deflection

$long,0 long-term, first-order deflection

$M bending deformation

$N axial deformation

$short,0 short-term, first-order deflection

$V shear deformation
xix
ΔL change in span length

ΔT change in temperature

Δf crown deflection

Δfcamb camber at arch crown section

Δfcreep crown deflection caused by creep

Δfdead crown deflection caused by dead load

Δfshrink crown deflection caused by concrete shrinkage

Δx element length for numerical integration

%c compressive strain

%c! strain at peak compressive stress of concrete

%cf strain on the flexural compression face

%cu crushing strain of concrete

%j real strain caused by unit redundant force j

%K strain caused by dimensional changes (i.e. creep, shrinkage, temperature)

%sh concrete shrinkage strain

%su rupture strain of steel

%t tensile strain

%t! strain at peak tensile stress of concrete

& influence line for bending moment; eccentricity constant

' angle of inclination of arch

κ system stiffness factor, or ratio between system secant bending stiffness and system flexural rigidity

( slenderness ratio

xx
(f modified slenderness ratio

) buckling coefficient

ρs steel reinforcement ratio in flexural tension zone

ρs! steel reinforcement ratio in flexural compression zone

σarch flexural stress in arch

* Poisson’s ratio

+ aging coefficient

, curvature of section; creep coefficient

,c material resistance factor for concrete

,cr curvature at initial cracking of concrete

,j real curvature caused by unit redundant force j

,p material resistance factor for prestressing steel

,y curvature at initial yielding of steel

,u curvature at ultimate capacity of section

,u, long-term curvature at ultimate capacity of section

a horizontal distance of concentrated load from support

A gross cross-sectional area

Aarch gross cross-sectional area of arch

Acol gross cross-sectional area of spandrel column(s)

Adeck gross cross-sectional area of deck girder

Ap cross-sectional area of prestressing steel

As cross-sectional area of reinforcing steel

b width of section or member; wall length


xxi
bdeck width of bridge deck

c integration constant resulting from virtual work integral

d depth of section or member

darch depth of arch

ddeck depth of deck girder

d flexibility matrix containing deformations of primary system caused by redundant forces

d0 deformation matrix containing deformations of primary system caused by applied loads

e initial eccentricity of load; end eccentricity of column

elive eccentricity of resultant axial force caused by live load

E modulus of elasticity

Eadj age-adjusted, effective modulus

Ec modulus of elasticity of concrete in compression

Ec! secant modulus at peak compressive stress

Ei modulus of elasticity at the time of initial stress

Etan tangent modulus of elasticity

EA axial rigidity

EAarch axial rigidity of arch

EAcrown axial rigidity of arch crown

EI flexural rigidity

EI! secant bending stiffness (secant to the moment-curvature diagram calculated for a given point)

EI* constant reference flexural rigidity

EI, long-term effective secant stiffness

EI0 initial uncracked flexural rigidity

xxii
EIarch flexural rigidity of arch

EIcol flexural rigidity of column

EIcr secant bending stiffness at initial cracking of concrete

EIcrown flexural rigidity of arch crown; prime symbol indicates secant stiffness

EIdeck flexural rigidity of deck girder; prime symbol indicates secant stiffness

EIsys flexural rigidity of system (sum of EIdeck and EIcrown); prime symbol indicates secant stiffness

EIy secant bending stiffness at initial yielding of steel

EIu secant bending stiffness at ultimate limit states

f rise of arch

fc compressive stress

fc! compressive strength of concrete, or peak compressive stress of concrete

fcr tensile cracking strength of concrete

fE critical buckling stress of thin plates

fflex flexural stress

fp! effective prestress after all losses

fpu ultimate stress of prestressing steel

ft tensile stress

ft! tensile strength of concrete, or peak tensile stress of concrete

F force vector containing unknown redundant forces

Fp! effective prestressing force after all losses

h depth of solid, rectangular section

hi height of ith spandrel column

hI normalized system moment of inertia

xxiii
hV equivalent slab thickness, or normalized concrete volume of system

H horizontal reaction of arch

Hcreep horizontal reaction caused by creep

Hdead horizontal reaction caused by dead load

HE buckling resistance of arch

HE,f buckling resistance of shallow arch

HE,long long-term buckling resistance

HE,short short-term buckling resistance

Hlive horizontal reaction caused by live load

Hlong long-term horizontal reaction

Hshrink horizontal reaction caused by concrete shrinkage

Hshort short-term horizontal reaction

Htemp horizontal reaction caused by changes in temperature

I initial, uncracked moment of inertia of section

Iarch initial, uncracked moment of inertia of arch section

Ideck initial, uncracked moment of inertia of deck section

Isys initial, uncracked system moment of inertia

k effective length factor; effective arc length factor

k, creep stiffness factor

kdyn dynamic amplification factor

kE plate buckling coefficient

kf shallowness factor

kL effective span length

xxiv
(kL)arch effective length of arch leg between spandrel columns

kS effective arc length

l length of element

L length of column; span length of arch

Ln interior span length

m normalized bending moment; virtual bending moment; multilane modification factor

m* normalized ultimate moment

mi virtual bending moment caused by unit redundant force i

mN=0 normalized bending moment resistance in pure flexure (axial force is zero)

M bending moment; second-order bending moment; flexible system moment

M* ultimate moment, or bending moment carried by the critical section of a column at ultimate load Q*

M0 first-order bending moment

M1 additional bending moment caused by second-order effects

March bending moment in arch

Mbeam bending moment in a fixed-fixed supported beam

Mclose bending moment at closure of the structural system

Mcr bending moment at initial cracking of concrete

Mcreep bending moment caused by creep of sustained load

Mdead bending moment caused by dead load

Mdeck bending moment in deck girder

Mfalse bending moment if structural system was built on falsework and then released

Mj real bending moment caused by unit redundant force j

MK redundant end moment caused by restrained deformation

xxv
Mlane bending moment caused by multilane lane load

Mlive bending moment caused by live load

Mneg maximum negative bending moments

Mpos maximum positive bending moments

MR moment capacity, or maximum moment of resistance

MS bending moment in simply supported beam

Mshrink bending moment caused by concrete shrinkage

Mshort short-term bending moment

Mtemp bending moment caused by uniform change in temperature

Mtruck bending moment cause by multilane truck load

My bending moment at initial yielding of steel

Mu bending moment resistance at ultimate limit states

Mu, long-term bending moment resistance at ultimate

n integer value; number of design traffic lanes; normalized axial force; virtual axial force

ncr normalized axial force at initial cracking

nE; nE linearity modular ratio between Ec and Ec!; normalized axial force at Euler buckling

ni virtual axial force caused by unit redundant force i

nt normalized axial force at peak tensile stress

n-m normalized design interaction diagram

n-m* normalized reduced, member capacity interaction diagram

N axial force

Nb axial force at the balanced point

Nj real axial force caused by unit redundant force j

xxvi
N-M sectional capacity interaction diagram

N-M* reduced, member capacity interaction diagram

q distributed load

qdead distributed dead load

qlive distributed live load

Q concentrated load

Q* ultimate load, or maximum axial load of column of given length, cross-section, material, and eccentricity

QE Euler buckling load

Qlive concentrated live load

r radius of gyration

rsys system radius of gyration

R2 coefficient of determination

s curved axis of the arch

S arc length of arch

t elapsed time (i.e. after casting, after initial stress); wall thickness

t0 reference starting time

V total longitudinal concrete volume of system, including the deck girder, spandrel columns, and arch

w second-order column deflection

w! slope of column deflection

w" second derivative of column deflection

w* ultimate deflection, or deflection of column at ultimate load Q*

w0 first-order deflection of column

x longitudinal axis

xxvii
x1, x2 redundant end moments of fixed arch system

x3 redundant horizontal reaction of fixed arch system

y vertical or transverse axis; ordinates of pressure line

ybot distance to the bottom extreme fibre

ytop distance to the top extreme fibre

z lever arm between flexural tensile force and flexural compressive force

xxviii
List of Uncommon Terms
arch-beam parameter—a non-dimensional parameter that situates a given fixed arch system within a transition

between pure arch behaviour and pure beam behaviour

balanced point—in the context of ultra high-performance fibre-reinforced concrete, the sectional failure mode

associated with the simultaneous failure of concrete in compression and tension at the extreme fibres

deck-stiffened arch system—a structural system comprised of an arch, spandrel columns, and a deck girder, in which

the flexural rigidity of the arch is much less than the flexural rigidity of the deck

eccentricity of load—the distance between the centroidal axis of a column and line of action of applied load

eccentricity constant—the ratio of the maximum deflection of a member to the maximum eccentricity of load

effective secant stiffness—conservative values of secant bending stiffness selected according to the highest bending

moment expected or allowed at a given limit state

effective arc length—similar to the concept of effective length in slender columns, this term is equal to the arc length

of the arch times a factor that accounts for the boundary conditions and buckled shape of the arch

equivalent slab thickness—a normalized volume quantity, equal to the total volume of longitudinal concrete divided

by the surface area of bridge deck

extra weight factor—the ratio of the total weight of the system to the weight of longitudinal concrete

fixed system moment—the moments that arise from local deformations of the deck and arch between spandrel

columns, assuming that the joints at each end of the columns do not translate

flexible system moment—the moments that arise from the global deformations of the deck and arch between

springing lines; these moments are typically much greater than fixed system moments

interior span length—the horizontal distance between spandrel columns

member capacity interaction diagram—an envelope of maximum sectional forces carried by the critical section of a

slender member of given length, cross-section, and material model; this envelope is a collection of

ultimate limit states of the member, corresponding to all possible values of initial eccentricity of load

xxix
modified slenderness ratio—a non-dimensional ratio related to the arch-beam parameter that situates a given fixed

arch system within a transition between pure arch behaviour and pure beam behaviour; this ratio also

determines the primary buckling mode of shallow arches

partially deck-stiffened arch system—a structural system comprised of an arch, spandrel columns, and a deck girder, in

which the flexural rigidity of the arch is roughly the same as the flexural rigidity of the deck

secant bending stiffness—a secant to the moment-curvature diagram calculated for a given point; this quantity is often

used to account for the softening of a member as curvatures are increased

sectional capacity interaction diagram—an envelope of maximum sectional forces carried by a given section and

material model; this envelope corresponds to the sectional failure (i.e. concrete crushing or fibre pull-out)

self-stiffened arch system—a structural system comprised of an arch, spandrel columns, and a deck girder, in which

the flexural rigidity of the arch is much greater than the flexural rigidity of the deck

shallow arch—an arch that is sensitive to axial deformations; in indeterminate arches, these axial deformations induce

bending moments in the system even under permanent load

system depth—the sum of depths of the arch and deck

system flexural rigidity—the sum of flexural rigidities of the arch and deck

system moment of inertia—the sum of moments of inertia of the arch and deck

system radius of gyration—the square root of (system moment of inertia divided by cross-sectional area of arch

crown)

system stiffness factor—a reduction factor less than one that accounts for material nonlinearity, equal to effective

system secant stiffness divided by system flexural rigidity

ultimate deflection—in the context of slender members, the deflection at ultimate load

ultimate load—in the context of slender members, the maximum compressive axial load that can be sustained by a

column of given length, cross-section, material, and eccentricity of load

ultimate moment—in the context of slender members, the bending moment at the critical section of a column at

ultimate load

xxx
“The form controls the forces; and the more clearly
the designer can visualize those forces the surer he is of his form”
—D. P. Billington

xxxi
Chapter 1. Introduction

This thesis concerns the development and validation of new structural systems for arch bridges using ultra

high-performance fibre-reinforced concrete. To design structures effectively using new materials, it is im-

portant that designers understand how the material affects structural behaviour at a variety of scales, in-

cluding that of the material, cross-section, member, and overall arch system. Correspondingly, this re-

search study consists of five parts: (1) an experimental investigation of the stress-strain response of ultra

high-performance fibre-reinforced concrete, (2) an analytical and experimental investigation of slender

members under compression and bending, (3) a review and synthesis of the statical analysis of arch

bridges, (4) an empirical study of fifty-eight concrete arch bridges built over the last century, and (5) a

parametric design study of arch bridge concepts using ultra high-performance fibre-reinforced concrete.

1.1 Motivation

Arches are intrinsically efficient because they carry a large share of their applied loads in compression. The

geometric form of an arch is typically chosen such that the centroidal axis of the arch coincides with the

line of action of force, or pressure line, of a given set of permanent loads (see Figure 1-1). Shaping arches in

this way allows them to carry permanent loads in a state of pure compression. Because concrete is strong

in compression and weak in tension, carrying permanent loads in compression is considered to be a more

economical use of material than carrying permanent loads in bending. To carry tensile stresses that arise

due to bending, concrete members must rely on longitudinal steel reinforcement to maintain equilibrium.
1
1. Introduction

Thus the required amount of steel and concrete is generally lower in arches than in beams for permanent

loads.

arch subjected to beam subjected to


permanent uniformly permanent uniformly
distributed load distributed load
pressure line of load
arch in compression top chord in compression
diagonals in compression
verticals in tension
structural model
bottom chord in tension
and flow of forces

bending moment diagram 0

shear force diagram 0

axial force diagram 0

Figure 1-1. Flow of forces and sectional forces in arches and beams

Despite their material efficiency, arches have seldom been built in the past fifty years due to the high cost

of labour and temporary structures. In general, construction costs increase with increases in complexity

and duration of construction and with increases in weight of temporary and permanent structures. Up un-

til the mid-20th century, concrete arches were usually built using elaborate timber scaffolding. For ex-

ample, the Sandö Bridge in Sweden was built this way in 1943, as shown in Figure 1-2. This method of con-

struction is time-consuming, generates a large amount of throwaway material, and requires the work of

highly skilled carpenters. In the second half of the twentieth century, modern cantilevered methods of

construction led to significant cost savings and greater span lengths for arches. The Colorado River Bridge

on the border of Arizona and Nevada was built this way in 2009, as shown in Figure 1-2. These cantilever

methods significantly reduced construction costs by reducing the amount of falsework and throwaway ma-

terial and by reducing the amount of manual labour through increased mechanization of work. The use of

precast concrete segments has allowed for further improvements in the speed and quality of construction.

Parallel advances in the construction technology of competing structural systems, such as variable-depth

girder and cable-stayed systems, were also made. The construction of cable-stayed bridges requires essen-

tially zero throwaway materials, which remains a lot less than arch bridges built using cantilever methods.

2
1. Introduction

Thus, despite innovations in arch construction techniques, concrete arches have remained, in most cases,

less economical than alternative structural systems.

Figure 1-2. Construction of Sandö Bridge (top) and Colorado River Bridge (bottom).
Adapted from Mondorf (2006) and usdot-fhwa (2003). Illustrations are not shown at the same scale.

Advances in material technology have produced concretes that have compressive strengths in excess of 150

MPa. This represents a significant increase in strength relative to the conventional 35 MPa concretes that

were common twenty years ago. In addition to exceptional strength, this new class of concrete called ultra

high-performance fibre-reinforced concrete also tends to exhibit superior durability, long-term stability

(i.e. reduced creep and shrinkage strains), stiffness, and tensile strength when compared to conventional

concrete (AFGC 2002). These enhanced properties are achieved through the use of steel fibres, water redu-

cers, a high binder content, and a special selection of fine aggregates. Steel fibres are used to improve the

ductility of concrete, to provide crack control, and to carry tensile forces, potentially reducing the need for

conventional steel reinforcing. In mixes with high fibre content, tensile strengths up to 20 MPa have been

observed (Habel 2004).

Using ultra high-performance fibre-reinforced concrete in arch bridges may lead to significant reductions

in concrete consumption and construction cost. The benefits of higher compressive strength are compoun-

3
1. Introduction

ded in terms of material efficiency: increases in member capacity allow for thinner members, thinner

members are lighter and so decrease dead load demand. Because arches carry loads primarily in compres-

sion, these benefits are applicable and should reduce concrete consumption.

In conventionally reinforced concrete members, minimum slab and flange thicknesses are often governed

by minimum concrete cover and spacing requirements for mild and prestressing steel. Minimum cover

depths to reinforcing steel ensure a minimum level of durability of the structure, including protection

against the corrosion of reinforcing steel. Where mild or internal prestressing steel is used in conjunction

with ultra high-performance fibre-reinforced concrete, minimum cover requirements can be reduced due

to the low permeability of the concrete. Where steel fibres are designed to carry tensile forces in lieu of

mild or internal prestressing steel, conventional cover requirements are less applicable because steel fibres

are not spatially continuous. The spatial discontinuity of the fibres prevents the flow of chloride ions in a

closed circuit, effectively limiting the the undesirable effects of corrosion. As cover requirements are re-

duced or eliminated through the use of ultra high-performance fibre-reinforced concrete, greater thinness

of components can be achieved. Thinness of components are instead governed by local stability problems,

by space requirements related to obtaining random fibre-orientation during concrete placement, and by

member strength requirements. Based on the amount of additional compressive and tensile strength

afforded by ultra high-performance fibre-reinforced concrete over conventional concrete, this research

study aims to reduce the consumption of concrete in arch bridges by half.

If the weight of concrete in arches can be reduced by half, the cost of labour, equipment, and temporary

structures will also be reduced. Because lighter structures are more easily lifted and manoeuvred, load ca-

pacities of cranes and temporary structures can be reduced. Further, construction should proceed more

quickly and total hours of labour reduced. The benefits of tensile strength provided by steel fibres can be

used to reduce the quantity of passive reinforcing, reducing the labour associated with assembling rebar

cages. All these factors should offset the higher unit cost of ultra high-performance fibre-reinforced con-

crete relative to conventional concrete. If these cost savings equal or exceed the economic differences

4
1. Introduction

between arch bridges and other structural systems, then concrete arches will be renewed as viable and eco-

nomically competitive structural systems for a wide range of spans.

In addition to economic considerations, the use of ultra high-performance fibre-reinforced concrete in

arches may create unique opportunities in aesthetic expression, leading to greater achievements in visual

slenderness, shallowness, and transparency than were previously possible. Reductions in concrete con-

sumption also have benefits in terms of environmental impact, such as reduced use of natural resources

and reduced carbon emissions in the manufacturing and transporting of concrete. Improvements in dur-

ability would likely increase the service life of the bridge and decrease the cost and frequency of

maintenance.

1.2 State-of-the-art

In this section, the state-of-the-art of bridges using ultra high-performance fibre-reinforced concrete is re-

viewed. The currently under-researched aspects of this body of knowledge, which are addressed by this

thesis, are identified and discussed.

1.2.1 Highway girder bridges

Real-world progress has been made over the past decade in the development and implementation of new

structural systems for vehicular bridges using ultra high-performance fibre-reinforced concrete. In her

doctoral thesis, Spasović (2008) reviews and compares five short-span highway bridges built between 2001

and 2006 which utilize girders made from ultra high-performance fibre-reinforced concrete. These bridges

were built in Virginia, USA, in Shepherds Creek, Australia, in Bourg-lès-Valence, France, in Saint-Pierre-

la-Cour, France, and over Highway A51 in France. The cross-section of the precast, π-shaped girders used

in the Virginia road bridge is shown in Figure 1-3 (Graybeal and Hartmann 2005). The girders span 21.3

metres and are simply supported. At their thinnest, the girders have slab thicknesses of 76 mm and web

thicknesses of 58 mm.

5
1. Introduction

Figure 1-3. Cross-section of ultra high-performance fibre-reinforced concrete girder used in Virginia Bridge.
Adapted from Graybeal and Hartmann (2005). Dimensions are in mm.

Spasović (2008) concluded that the ultra high-performance fibre-reinforced concrete girders among the

five bridges reviewed were all designed to function in a conceptually similar way: prestressing strands were

used to carry flexural tensile forces and ultra high performance fibre-reinforced concrete was used to carry

flexural compressive forces. Spasović observed, however, that the high compressive capacities of the con-

cretes were not being used to their fullest potential. In general, these girders were precast and assembled

in-situ using dry or wet joints and prestressing. In some of the five bridges, transverse moments were car-

ried by ultra high-performance fibre-reinforced concrete alone, without conventional reinforcing. Conven-

tional stirrups for shear reinforcement were also omitted in the designs due to the capacity of the concrete

and fibres to carry shear stresses. According to the designers, these bridges were economically competitive

in comparison to conventional concrete girder systems. These project studies demonstrate that structural

members with thin slabs, webs, and walls are capable of transmitting localized forces, such as wheel loads,

to primary longitudinal structural elements of bridges. They also set precedence for the successful applica-

tion of ultra high-performance fibre-reinforced concrete in short-span highway girder bridges.

6
1. Introduction

1.2.2 Pedestrian arch bridges

Figure 1-4. Sun-Yu Pedestrian Bridge. Adapted from Huh and Byun (2005). Dimensions are in mm.

The Sun-Yu Bridge in Korea, built in 2002, is the first and longest pedestrian bridge to use ultra high-per-

formance fibre-reinforced concrete in an arch form (Huh and Byun 2005). The arch, which spans 120 m

and rises 15 m, consists of a single π-shaped arch rib, as shown in Figure 1-4. The arch rib is prestressed

transversely with 13 mm monostrands at each transverse stiffening rib spaced at 1225 mm along the length

of the arch. Longitudinally, the rib is prestressed with two tendons of 9-15 mm strands in the lower webs,

and one tendon of 12-15 mm strands in the upper web. The specified compressive strength of the reactive

powder concrete used was 200 MPa.

Instead of conventional spiral reinforcement details, special reinforcement-free micro-anchorages (see Fig-

ure 1-5) were used to confine the concrete and to control the high tensile splitting forces near the an-

chorage zone. Using these micro-anchorages remove the slenderness limits normally imposed by the cover

requirements of spiral reinforcement, and allow for unprecedented thinness in the slabs, webs, and ribs of

structural members. These micro-anchorages were first developed for the Sherbrooke Bridge in Quebec,

7
1. Introduction

which is a trussed pedestrian bridge that has struts made from ultra high-performance fibre-reinforced

concrete (Aitcin et al. 1998). The Sun-Yu and Sherbrooke Bridges demonstrate that thin, ultra high-per-

formance fibre-reinforced concrete slabs, webs, and ribs can be prestressed and detailed without mild

reinforcing.

Figure 1-5. Reinforcement-free micro-anchorage detail. Adapted from Aitcin et al. (1998).

1.2.3 Highway arch bridges

To date, there has been no real-world application of ultra high-performance fibre-reinforced concrete in

highway arch bridges. Conceptual design studies, however, have been done by Freytag et al. (2009), Can-

drlic et al. (2001), and Radić et al. (2005) for arch spans of 69 m, 432 m, and 1000 m, respectively.

Freytag et al. (2009) investigate a conceptual design for the wild Bridge, which is designed for a 69 m

crossing in Austria using 165/185 MPa ultra high-performance fibre-reinforced concrete. The design,

shown in Figure 1-6, proposes two, hollow concrete ribs for the arch. These ribs are slender and thin, with

structural depths of 1200 mm (corresponding to a span-to-arch-depth ratio of 58:1) and wall thicknesses of

60 mm. Each leg in the polygonal arch is built as individual precast segments, which are later joined by

special knee-joints that form the angle breaks of the polygon. The knee-joints share the same exterior di-

mensions as the arch segments, but are made inwardly thicker to make space for embedded deviators for

prestressing tendons. The prestressing in the arch ribs are external and unbonded within the arch seg-

8
1. Introduction

ments, and are distributed along the inside perimeter of the section to obtain concentric prestress. The

proposed bridge deck consists of a conventional 600 mm thick slab made from 40/50 MPa concrete.

Figure 1-6. Elevation, cross-section, and joint detail views of WILD Bridge concept.
Scale is shown at 80% of original. Adapted from Freytag et al. (2009).

Freytag et al. (2009) conducted compression and bending tests on a prototype prestressed arch segment

and knee-joint assembly, which was built at full scale. The specimens were subjected to the forces expected

at all critical design states, and were found to be adequate. Results were also compared to finite element

models and calculations using simple mechanical models.

Candrlic et al. (2001) propose a conceptual design for a 432 m crossing over the Straits of Bakar in Croatia,

between Rijeka and Senj. The study is intended as a probe into the possibilities offered by using ultra high-

performance fibre-reinforced concrete with compressive strength of 200 MPa in the design an arch bridge.

The elevation and sections of the design are shown in Figure 1-7. Both the deck and arch are proposed as

three-cell box girders, with top and bottom slabs of 120 mm and web thicknesses of 200 mm. These mem-

bers are given oblong, aerodynamic shapes to reduce drag forces caused by transverse wind loading. The

arch rib is 6500 mm deep, corresponding to a span-to-arch-depth ratio of 66:1. Precast, segmental canti-
9
1. Introduction

lever construction is proposed, whereby match-cast segments, each 3800 mm in length, are joined in-situ

using epoxy resin and temporary external prestressing. Transverse ribs along the top and bottom slabs of

the box girders are provided at the ends of each segment. These ribs are used as a form of anchorage block

for temporary and permanent longitudinal prestressing. Using linear-elastic analysis, Candrlic et al. (2001)

calculated stresses caused by dead load, live load, wind, earthquake, and temperature, and found that con-

crete stresses were in all cases below allowable limits.

Figure 1-7. Bakar Bridge concept. Adapted from Candrlic et al. (2001). Dimensions are in mm.

10
1. Introduction

Radić et al. (2005) propose a conceptual design for a fictitious 1000 m crossing. Elevation and cross-sec-

tion views of their design are shown in Figure 1-8. The arch is a three-cell box girder and is 16.5 m deep,

corresponding to a span-to-arch-depth ratio of 61:1. The arch rib is composed of 5000 mm long segments,

each weighing 5.6 MN. A truss cantilevered method of construction is proposed, whereby the arch forms

the bottom chord, spandrel columns form the vertical struts, and steel ties form the diagonal ties and top

chord. A statical analysis was performed for all critical stages of construction. Maximum compressive and

tensile stresses were found to be 61 MPa and 14 MPa, respectively. Beyond this, very few details of the ana-

lysis and its assumptions are provided in the paper.

Figure 1-8. Bridge concept for a 1000 m crossing. Adapted from Radić et al. (2005). Dimensions are in mm.

The three arch bridge concepts described above begin to reveal the possibilities offered by using ultra high-

performance fibre-reinforced concrete in arch bridges. For all concepts, most bridge components are pro-

posed as modular precast segments with thin slabs and webs. Embedded steel fibres are used to displace

11
1. Introduction

most, if not all, of the conventional mild steel reinforcement. Prestressing steel is used temporarily to con-

nect segments during construction and also used as permanent reinforcement to provide flexural tensile

strength. All proposed segments incorporate some form of transverse stiffening through the local thicken-

ing of the section or the use of transverse ribs. Only Freytag et al. (2009) provide any experimental valida-

tion of their design.

Although these concept studies put forth are new and innovative ideas, they are narrow in scope and are

only directly applicable to the spans and rises specified in their designs. The calculated stresses and design

checks given in the studies are not helpful to designers who wish to design their own arch systems. Instead,

general guidelines and simplified methods of analysis are needed to aid designers. This would include a

thorough exposition on member behaviour and system behaviour, and how they differ from convention

reinforced concrete structural systems. This would also include simplified design methods for calculating

maximum sectional forces in the system, and recommendations on the proportion and size of members.

Structural efficiency needs to be quantified and compared with existing concrete arch technology to furth-

er substantiate the material and cost savings offered by these new systems. This thesis undertakes these

tasks in hope of educating designers and bringing these innovative structural systems one step closer to

real-world application.

1.3 Objectives and content of thesis

The primary objectives of this thesis are:

1. to design and validate new structural systems for arch bridges that reduce concrete consumption by

at least 50% relative to arch bridges built using conventional reinforced concrete technology, and

2. to provide designers with guidance on the structural analysis and design of these structural systems.

The secondary objectives of this thesis, which support the primary objectives are:

3. to characterize a local ultra high-performance fibre-reinforced concrete mix and to propose material

models that are sufficiently accurate for design,


12
1. Introduction

4. to identify how differences between ultra high-performance fibre-reinforced concrete and conven-

tionally reinforced concrete affect the axial force-moment-curvature behaviour of sections,

5. to determine the capacity of slender members subjected to compression and bending,

6. to review, synthesize, and illustrate the statical analysis of concrete arch bridges,

7. to identify thresholds for efficient arch behaviour related to the shallowness of concrete arch bridges,

8. to characterize the state-of-the-art of arch bridges made from conventional concrete, and

9. to determine which load combinations cause maximum sectional forces in these arch systems.

The content of this thesis reflects the primary and secondary objectives stated above. Topics are discussed

in sequence of scale of structural behaviour, starting with behaviour of the material, then of slender mem-

bers, and last of arch systems. In Chapter 2, the material behaviour of a local mix design of ultra high-per-

formance fibre-reinforced concrete is investigated experimentally with respect to uniaxial compression, in-

direct tension, creep, and shrinkage. From these tests, simple materials models are proposed for use in

analysis and design. In Chapter 3, the behaviour of slender members subjected to compression and bend-

ing is investigated. A general analysis method, which accounts for material and geometric nonlinearities, is

described and is implemented as a computer program that calculates load-deflection response of eccent-

rically-loaded columns in compression. This method is validated experimentally with load tests of slender

columns made from ultra high-performance fibre-reinforced concrete. Simplified design methods for cal-

culating member capacity are also proposed. In Chapter 4, the behaviour of arch structural systems is re-

viewed, synthesized, and illustrated. Using the force method, simplified methods for calculating sectional

forces caused by dead loads, live loads, and volumetric changes are presented and solved. The effects of

arch slenderness (sensitivity to flexural buckling) and arch shallowness (sensitivity to axial deformations)

are discussed in detail. In Chapter 5, a comparative study of fifty-eight concrete arch bridges built over the

past century is presented. Trends of various geometrical ratios and quantities among bridges in the data-

base are identified and discussed. These trends represent the current state-of-the-art of concrete arch

13
1. Introduction

bridges. In Chapter 6, a parametric design study of arch bridges using ultra high-performance fibre-rein-

forced concrete is described. Span length, arch rise, and degree of deck-stiffening are the primary design

parameters considered. Dimensions and sections of satisfactory design solutions are presented. Insights re-

lated to the efficiency, consumption of materials, slenderness, and shallowness are discussed. Design con-

cepts are also compared with the database of existing concrete arch bridges. In the final chapter, the most

important results and conclusions from each of the preceding chapters are summarized. Potential research

topics for future work are also identified.

The most important contributions of this thesis are:

• the validation of new light-weight and efficient arch bridge design concepts using ultra high-per-

formance fibre-reinforced concrete for spans of 50 to 400 m,

• the development of a simple analytical model that describes the transition of shallow arches between

pure arch behaviour and pure beam behaviour,

• a comprehensive comparative study of 58 existing concrete arch bridges that characterizes the current

state-of-the-art and serves as a valuable reference design tool, and

• the development and experimental validation of a general and simplified method for calculating the

capacity of slender ultra high-performance fibre-reinforced concrete members under compression

and bending.

The research presented in this thesis provides a means for designers to take full advantage of the high com-

pressive and tensile strengths of the concrete and hence to exploit the economic potential offered by the

material.

14
Chapter 2. Ultra High-Performance Fibre-Reinforced
Concrete

The purpose of this chapter is to describe the production of ultra high-performance fibre-reinforced con-

crete and to develop workable stress-strain models based on experiments conducted at the University of

Toronto. These stress-strain models will be used in Chapter 3 to investigate the behaviour of slender mem-

bers subjected to compression and bending and in Chapter 6 to design concepts for new arch bridge

systems.

French Interim Recommendations on the use of ultra high-performance fibre-reinforced concrete have

been previously developed (AFGC 2002). The first set of international design rules for the material is under

development by the International Federation for Structural Concrete (fib) Task Group 8.6 (Walraven

2008). However, because ultra high-performance fibre-reinforced concrete mixes are not standardized and

are not yet widely produced, a universal material model for ultra high-performance fibre-reinforced con-

crete does not exist (Graybeal 2006). Thus, mechanical properties of these concrete mixes are sensitive to

differences in local raw materials, mixing procedures, and testing procedures. It was therefore necessary to

develop a stress-strain model appropriate for the concrete mix used at the University of Toronto.

2.1 Material testing at the University of Toronto

An ultra high-performance fibre-reinforced concrete mix was developed by Habel et al. (2008) at the Uni-

versity of Toronto and École Polytechnique de Montréal using locally available materials. The mix exhib-
15
2. Ultra High-Performance Fibre-Reinforced Concrete

ited a compressive strength of 128 MPa and a tensile strength of 11.0 MPa (Habel and Gauvreau 2008).

High compressive strength is achieved by using water reducers, a low water-to-cement ratio, and no coarse

aggregate. High tensile strength is achieved by incorporating steel fibre reinforcement into the concrete.

These fibres are intended to provide crack control and member strength, and offer the potential to reduce

the need for steel reinforcing.

The French Interim Recommendations (AFGC 2002) gives three criteria for differentiating ultra high-per-

formance fibre-reinforced concrete from other high-performance cementitious materials: (1) material has

compressive strength that is systematically greater than 150 MPa, (2) material systematically uses fibres to

improve ductility and to modify conventional requirements for mild and prestressing steel, and (3) materi-

al uses high binder content and special selection of aggregates.

The University of Toronto mix used in this study satisfies the last two criteria but does not satisfy the com-

pressive strength criterion. Habel et al. (2008), who developed this mix based on an existing European

mix, attributes the lower compressive strength of the Toronto mix to the composition of the selected local

cement. The cement composition was found to be less appropriate for the production of concretes with low

water-to-binder ratio than those used in the European mix. They also suspect that the higher specific sur-

face area of the silica fume (15 to 20 g/m2 as opposed to 12 g/m2) might have led to higher water demand in

the Toronto mix as compared to the European mix. The compressive strength could be increased to satisfy

the strength criteria by further optimizing the selection and mix proportions of materials, by improving

the equipment and procedure used for mixing and casting, or by modifying the curing method through

the use of pressure or heat treatment. While these changes could be made, the concrete mix still exhibits

excellent material properties that can be used in its present formulation to design efficient, light-weight

structures. Increasing the strength of the concrete above 128 MPa and beyond 150 MPa does not necessar-

ily reduce the weight of the structure, since member geometry becomes increasingly governed by local sta-

bility problems rather than requirements for member strength.

16
2. Ultra High-Performance Fibre-Reinforced Concrete

2.1.1 Mix design

The ultra high-performance fibre-reinforced concrete mix developed by Habel et al. (2008) is presented in

Table 2-1. Fibre content is 5.5% of total volume and water-to-cement ratio is 0.25 by mass. This mix exhib-

ited good workability, low permeability, and high strength. This mix design was used as the basis for all ex-

periments described in Sections 2.1.3, 2.1.4, and 3.4.

Table 2-1. Mix design for ultra high-performance fibre-reinforced concrete

weight 48 L batch 52 L batch


constituent type
kg/m3 kg kg

cement general usage from Bath, Ontario 967 46.4 50.3


silica fume white, specific surface 15-18 m2/g 251 12.0 13.0
sand silica sand, grain size < 0.5 mm 675 32.4 35.1
steel fibres straight, 10 mm in length and 0.2 mm in diam. 430 20.6 22.3
superplasticizer polycarbonate, solid content 35 mass % 35 1.68 1.82
water tap water 244 11.7 12.7

2.1.2 Batching and casting

Six batches of ultra high-performance fibre-reinforced concrete were prepared according to the mix design

shown in Table 2-1. The first three batches were 48 litres in volume, while the remaining batches were 52

litres in volume. Materials were mixed in a 0.17 m3 mortar mixer (see Figure 2-1) in the following order: (1)

the cement, silica fume, and sand were dry mixed until they were homogeneous, (2) the superplasticizer

and water, which were pre-mixed, were added and mixed until no dry materials remained, and (3) the steel

fibres were added and mixed in. The mixing process took 5 to 10 minutes per batch.

The finished batches were poured into a concrete bucket (see Figure 2-1) and then placed into a wheelbar-

row. Cylinder moulds were filled using hand shovels. Since the concrete was self-consolidating, no vibra-

tion was done during placement.

17
2. Ultra High-Performance Fibre-Reinforced Concrete

Figure 2-1. Mortar mixer and concrete bucket

2.1.3 Material behaviour in compression

Two sets of cylinder tests were conducted to investigate the stress-strain behaviour of ultra high-perform-

ance fibre-reinforced concrete in compression. All cylinders were nominally 100 mm in diameter and 200

mm in length. Cylinders from Set 1 were moist-cured at room temperature until they were tested 27 days

after casting. Cylinders from Set 2 were moist-cured at room temperature for 7 days and tested 60 days

after casting. The ends of all cylinders were saw-cut and left uncapped. All cylinders were tested using an

mts testing machine, with initial displacement rates of 4.44 μm/sec. To produce this displacement rate,

machine head displacements were continuously used to control the application of load during each test.

For cylinders from Set 1, vertical displacements of the cylinder were measured by two linear variable di-

fferential transformers, with gage lengths of 153.5 mm. For cylinders from Set 2, vertical displacements

were measured from machine head displacements. Load versus vertical displacement data of the tested

cylinders are shown in Figure 2-2 and Figure 2-3.

18
2. Ultra High-Performance Fibre-Reinforced Concrete

specimens: tested 27 days


100 mm in diam. after casting
200 mm in length
bracket
clamped to
gage length cylinder
826 kN 835 kN
154 mm
linear variable
differential
transformer
measuring
load applied by displacement
MTS testing machine

batch U2 batch U5
specimen 6 specimen 5
Load
in kN

1000
780 kN 865 kN 866 kN

500

0
0 0.5 1 1.5 batch U5 batch U5
displacement data is displacement in mm specimen 4 specimen 6
averaged from two batch U1
transformers specimen 1

Figure 2-2. Load versus displacement data of cylinders from Set 1, tested 27 days after casting

specimens: tested 60 days


100 mm in diam. after casting
200 mm in length
1081 kN
1015 kN
approximately
displacement
190 mm with
measured from
ends removed
machine head

load applied by
MTS testing machine

batch U5 batch U6
specimen 7 specimen 7
Load
in kN

1000 1024 kN
951 kN 950 kN

500

0
0 0.5 1 1.5 batch U5 batch U6
displacement in mm specimen 8 specimen 8
batch U4
specimen 4

Figure 2-3. Load versus displacement data of cylinder from Set 2, tested 60 days after casting

19
2. Ultra High-Performance Fibre-Reinforced Concrete

The cylinders exhibited nearly linear behaviour until their peak load, except for Cylinders U2-6 and U5-7

(cylinder id numbers correspond to batch number and specimen number, respectively). The linearity of a

material can be quantified by calculating the ratio of the tangent modulus of elasticity in compression to

the secant modulus at peak compressive stress. The tangent modulus of elasticity is given by the stress at

40% of peak compressive stress divided by the strain of the cylinder at this stress. The secant modulus at

peak compressive stress is given by the peak compressive stress divided by the strain of the cylinder at this

stress. Stress in the concrete is given by the applied load divided by the cross sectional area of the cylinder

(7850 mm2). The strain in the concrete is given by the average displacement of the linear variable differen-

tial transformers divided by their gage lengths of 153.5 mm. The average ratios between the tangent and

secant moduli for cylinders from Set 1 was 1.28, as shown in Table 2-2. This average ratio value is consistent

with results obtained from cylinder tests conducted by Graybeal (2006). Strain and modulus values of cyl-

inders from Set 2 cannot be calculated accurately based on machine head displacements, and so are not

tabulated in Table 2-2. Machine head displacements were not linearly related to transformer displacement

in load tests of cylinders from Set 1.

Table 2-2. Results of concrete cylinder tests


batch- age peak peak strain 40% strain at tangent secant tangent
specimen load stress at peak peak 40% peak modulus of modulus at to secant
stress stress stress elasticity peak stress modular ratio

days kN MPa 10-3 MPa 10-3 GPa GPa GPa/GPa

U1-1 27 780 99 2.49 39.7 0.82 48.3 39.9 1.21


U2-6 27 826 105 4.35 42.1 0.97 43.5 24.2 1.80
U5-4 27 865 110 2.52 44.1 0.89 49.6 43.7 1.13
U5-5 27 835 106 2.38 42.5 0.55 76.9 44.7 1.72
U5-6 27 866 110 2.67 44.1 1.08 40.8 41.3 0.99

average 834 106 42.5 48.1 37.3 1.28

Calculated averages exclude values from U5-5 because of its inordinately high tangent modulus of elasticity.

20
2. Ultra High-Performance Fibre-Reinforced Concrete

Because cylinder tests were displacement controlled, post peak ductility could be investigated. Most cylin-

ders exhibited ductility beyond their peak compressive load. As shown in Figures 2-2 and 2-3, the post-

peak response of each cylinder differed. Cylinders U5-4 and U5-6 from Set 1 lost about half their load ca-

pacity soon after sustaining their peak load, while the other three cylinders from this set exhibited long

yield plateaus. Cylinders from Set 2 all exhibited some decrease in load beyond the peak load. In all but

one case (U4-4), this was followed by a significant increase in displacement.

Based on the test results, the load-carrying capacities of the cylinders beyond their peak loads are high, but

inconsistent. The sudden loss of strength in Cylinders U5-4, U5-6, and U4-4 may be a result of: (1) subop-

timal mixing procedures and equipment, (2) the small size of production, (3) misalignments of the saw-cut

ends of the cylinders causing non-uniform stresses, or (4) the use of machine head displacements as the

feedback for displacement control. Calibrating and standardizing the mixing procedure to the local mater-

ials and equipment can help to improve the consistency of results. This optimization process is most suit-

able in an industrial setting, where the size of production can be increased. Using machine displacements

to control the application of load on cylinders is not ideal because the displacement of the machine head is

not perfectly correlated with the displacement of the cylinder. This discrepancy may cause the machine to

apply loads that are not in sync with the desired cylinder displacements, causing Cylinders U5-4, U5-6,

and U4-4 to be unloaded as they would in a load-controlled test. If this is indeed the case, then the sudden

loss of strength observed in these cylinders may be an artifact of the test equipment. Improving the test

method would likely increase the observed ductility of these cylinders.

Given that the inherent and observed consistency and reliability of the concrete can be improved through

the means discussed above, it is assumed that the University of Toronto ultra high-performance fibre-rein-

forced concrete mix is inherently more ductile in compression than conventional concrete. This limited

ductility is incorporated into the material model proposed later in this section.

21
2. Ultra High-Performance Fibre-Reinforced Concrete

Figure 2-4. Ultra high-performance fibre-reinforced concrete cylinder at failure

All cylinders tested did not fail by brittle crushing, which is the mode of failure of conventional concrete

cylinders. Instead, the cylinders continued to carry load at strains higher than conventional concrete due

to the ability of the dispersed fibres to keep the concrete matrix together. A photograph of a failed ultra

high-performance fibre-reinforced concrete cylinder is shown in Figure 2-4.

Table 2-3. Compressive strengths of cylinders organized by batch, age, and curing regime

batch- age curing peak stress batch- age curing peak stress
specimen days MPa specimen days MPa
U2-1 7 moist 90.1 U6-4 28 moist 90.3
U2-2 7 moist 84.2 U6-5 28 moist 96.1
U5-1 7 moist 89.2 U6-6 28 moist 111.2
U5-2 7 moist 83.7 U1-1 28 moist 99.3
U5-3 7 moist 90.6 U1-2 48 air 124.3
U2-3 14 moist 94.7 U1-3 48 air 138.3
U2-4 14 moist 99.3 U2-7 49 air 111.4
U6-1 14 moist 89.6 U2-8 49 air 118.4
U6-2 14 moist 80.7 U2-9 49 air 118.6
U6-3 14 moist 75.3 U3-4 54 air 127.3
U2-5 27 moist 106.5 U3-5 54 air 134.4
U2-6 27 moist 105.2 U3-6 54 air 134.6
U5-4 27 moist 110.1 U5-7 60 air 137.7
U5-5 27 moist 106.3 U5-8 60 air 130.4
U5-6 27 moist 110.3 U4-4 60 air 121.1
U3-1 28 moist 113.1 U6-7 60 air 129.2
U3-2 28 moist 113.4 U6-8 60 air 121.0
U3-3 28 moist 101.1

22
2. Ultra High-Performance Fibre-Reinforced Concrete

A total of thirty-five ultra high-performance fibre-reinforced concrete cylinders were tested for compress-

ive strength. Cylinders differed by batch, age of concrete after casting, and curing regime. Test results are

tabulated in Table 2-3.

Age after casting and compressive strength of the tested cylinders are plotted in Figure 2-5. The results

show an increase in strength with age between 7 and 60 days. The average strength of the cylinders tested

27 or 28 days after casting was 107 MPa. The average strength of the cylinders tested at 7 days after casting

was 87.6 MPa, which is 82% of the 27 or 28 day strength. This shows that this material gains strength

quickly. The average strength of the air-cured cylinders tested 60 days after casting was 128 MPa, which is

120% of the 27 or 28 day strength of the cylinders moist-cured for 7 days.

Stress moist cured moist cured for 7


specimens: days, then air cured Load
in MPa
100 mm in diam. in kN
150
200 mm in length
7850 mm2 in area 1000

100

500
50
load applied by
MTS testing machine
0 0
0 28 56
age after casting in days

Figure 2-5. Compressive strength versus age of concrete after casting

The proposed compressive stress-strain models for the concrete mix under investigation is shown in Fig-

ure 2-6 and is labelled as “material model.” These model curves are based on average values of modulus of

elasticity in compression Ec, peak compressive stress fc!, and strain at peak stress %c!, observed from the cyl-

inder tests described earlier in this section. The model curve is defined by four points: (0, 0), (0.7fc!/Ec,

0.7fc!), (%c!, fc!), and (1.4%c!, 0.7fc!). The strain at peak stress %c! can also be expressed as fc!/Ec!, where Ec! is

the secant modulus at peak compressive stress. Stress-strain results from the cylinder tests from Set 1 are

shown in the figure to show their variance, and the degree of simplification that the model curve repres-

ents. The proposed strain limit of 1.4%c! represents a practical limit for calculation rather than the complete

loss of load-carrying capacity by concrete crushing. Straining the extreme fibres of a section beyond this

23
2. Ultra High-Performance Fibre-Reinforced Concrete

strain limit tends to result in bending moments with curvatures greater than the curvature at maximum

moment. Since designers are typically interested in determining the moment capacity of a section, then

this strain limit of 1.4%c! is considered to be acceptable.

Material model Compressive


stress in MPa Set 1 (5 cylinders)
Compressive 200 100 mm ! 200 mm
moist-cured for 27 days
stress
fc! ( fc!÷Ec′, fc′ ) avg. material properties:
150 Ec = 48.1 GPa
Ec! = 37.3 GPa
( 1.4 fc!÷Ec!, 0.7fc! )
( 0.7 fc!÷Ec, 0.7fc! ) fc ! = 106 MPa
100

50
0
0 !c′ Compressive strain cylinder test data
material model
0
0 4·10-3 8·10-3
Compressive strain

Figure 2-6. Material model in compression

According to this compressive stress-strain model, three input variables fc!, Ec!, and Ec are required. Cylin-

der tests are often conducted without measuring displacement. In these cases, only ultimate load is meas-

ured, which gives the concrete strength fc! of the cylinder. Compressive stress-strain models for these tests

can be estimated by expressing Ec and Ec! as functions of fc!.

Empirical relationships between Ec and fc! have been proposed by Ma and Orgass (2004) and Graybeal

(2007) using other ultra high-performance fibre-reinforced concrete mixes. Many cylinders were tested in

these studies at different ages after casting, resulting in a wide spectrum of compressive strengths and

elastic moduli. From these tests, Ma and Orgass fitted their data to a one-third power equation form

(Equation 2-1) and Graybeal fitted his test data to a square root equation form (Equation 2-2):

Ec = 8820 3 fc! in MPa Equation 2-1

Ec = 3840 fc! in MPa Equation 2-2

24
2. Ultra High-Performance Fibre-Reinforced Concrete

Modulus of elasticity 80
data points are results from 5 cylinder tests
in compression
in GPa
curve proposed by Graybeal (2007):
60
Ec = 3840 fc! in MPa

40
curve proposed by Ma and Orgass (2004)

Ec = 8820 3 fc! in MPa


20
Note:
Only a small set of stress-strain cylinder tests
were done for this thesis. The curves proposed by
0 Graybeal (2007) and Ma and Orgass (2004) are
0 50 100 150 200 based on large sets of ultra high-performance
fibre-reinforced concrete cylinder tests. The mix
Compressive strength designs used in these studies are similar to the
fc! in MPa one used at the University of Toronto, but not the
same. Data for different compressive strengths
were obtained by changing the age at which the
specimens were tested.

Figure 2-7. Tangent modulus of elasticity versus compressive strength of tested cylinders

Figure 2-7 shows that the empirical equations proposed by Ma and Orgass (2004), and Graybeal (2007)

underestimate the measured moduli of elasticity of the cylinders tested at the University of Toronto. Thus,

in most cases using these empirical equations will result in design calculations that lead to conservative

member capacities.

Using the linearity modular ratio nE between the tangent modulus of elasticity in compression and the sec-

ant modulus at peak compressive stress, Equation 2-2 can be modified to relate fc! and Ec!:

1
Ec! = " 3840 fc! in MPa Equation 2-3
nE

Using Equation 2-3, together with Equations 2-1 or 2-2 results in a material model that requires only com-

pressive strength fc! as an input parameter. It is assumed that this material model is sufficiently accurate for

the conceptual design studies presented in Chapter 6. Because the shape of the material model was de-

termined from the cylinder test results shown in Figure 2-2 and Figure 2-3, it is prudent to restrict the ap-

plicability of this design material model to compressive strengths between 100 MPa and 140 MPa. These

strengths are roughly those which were observed among all cylinders tested at the University of Toronto.

Extending the model beyond this range or to different concrete mixes may result in gross model errors.

25
2. Ultra High-Performance Fibre-Reinforced Concrete

2.1.4 Material behaviour in tension

Modulus of rupture tests were conducted at the University of Toronto1 to investigate the behaviour of ultra

high-performance fibre-reinforced concrete in tension. All specimens were square prisms, 150 mm by 150

mm in section, spanning 900 mm between roller supports. Elongations along the tension side of the prism

were recorded with two longitudinally oriented linear variable differential transformers with gage lengths

of 900 mm. Three specimens were moist-cured for 7 days at room temperature and tested 54 days after

casting under four-point loading. Load points are located at the third-points of the span. Load-displace-

ment results are shown in Figure 2-8.

load applied by 1000 kN linear variable differential trans-


Prism specimens were moist-cured at room MTS testing machine
spherical former measuring displacement
temperature and tested 54 days after casting.
bearing 75

50
steel spreader
free rotation about axis of pin free rotation about axis of pin
beam cross-section

150
mount

roller support (no translation) roller support


150
bearing (no translation)
concrete prism
pin cross-section

300 300 300

span and gage length = 900 mm

Load
in kN
80 79.3 kN
74.5 kN 72.9 kN

60

40

20

0
0 4 8 12 0 4 8 12 0 4 8 12
displacement data is displacement in mm displacement in mm displacement in mm
averaged from two trans- batch U4 batch U5 batch U6
formers, one on each side

Figure 2-8. Modulus of rupture test setup and load-displacement results

1. These prism tests were part of a related research project conducted by Ivan Wu.
26
2. Ultra High-Performance Fibre-Reinforced Concrete

Because the transformers shown in Figure 2-8 do not bend along with the prism, there is some discrep-

ancy between the relative horizontal displacements of the ends of the beam and the average change in

length of the longitudinal fibres of the beam at the same level. The error associated with this second-order

effect at the initial loss of stiffness (or first significant change in slope on the diagrams) is about one-hun-

dredth of a percent, which is small and therefore can be neglected. This initial loss of stiffness corresponds

to the initiation of distributed microcracking. Data recorded after initial microcracking is not considered

to be representative of the uniaxial stress-strain behaviour of ultra high-performance fibre-reinforced con-

crete and will not be used to characterize the stress-strain behaviour of the material.

Based on elasticity, the flexural tensile stress fflex at the extreme fibre of the prism between the third-point

loads can be calculated using Equation 2-4 (Neville 1995), where Q is the total applied load, L is the span

length, and b and d are the width and depth of the prism:

QL
fflex = Equation 2-4
bd 2

Because Equation 2-4 assumes linear-elastic behaviour, only the stresses of the prism before cracking can

be calculated accurately. These flexural tensile stresses are further reduced by a factor of 0.727 to estimate

axial tensile stresses ft according to Equation 2-5 (afgc 2002), where d is the prism depth of 150 mm, and

d0 is the reference depth of 100 mm determined from empirical analysis:

2 ( d / d0 )
0.7

ft = fflex Equation 2-5


1 + 2 ( d / d0 )
0.7

Equation 2-5 accounts for the overestimation of stresses in small-scale flexural tests due to strain gradient

effects and is recommended in the French Interim Recommendations (afgc 2002). After applying the cor-

rection factor of 0.727, the average cracking stress of the three prisms tested were found to be 7 MPa.

Fibre strains at the level of the transformers were obtained by dividing the average displacement of the

transformers by their gage length. By assuming plane sections remain plane, bottom fibre strains were es-

27
2. Ultra High-Performance Fibre-Reinforced Concrete

timated by multiplying the strains at the level of the transformers by two, since the transformers are at one

fourth the depth from the bottom. The average cracking strain of the three prisms tested were 0.15·10-3.

Dividing the average cracking stress by the average cracking strain results in a tangent modulus of elasti-

city in tension of 46.7 GPa. This value falls within the range of moduli obtained from direct tension tests

conducted by Habel and Gauvreau (2008) who used the same ultra high-performance fibre-reinforced

concrete mix design shown in Table 2-1.

Three split cylinder tests were conducted to investigate the peak tensile stress, or post-cracking strength, of

the material. All three cylinders were cast from batch U4, moist-cured for 7 days at room temperature, and

then air-cured for 21 days until they were tested. The peak tensile stresses of the cylinders were 18.1, 12.0,

and 13.8 MPa. The peak tensile stress ft! was calculated by Equation 2-6 (Neville 1995), where Q! is the peak

applied load, l is the cylinder length of 200 mm, and d is the cylinder diameter of 100 mm.

2Q !
ft! = Equation 2-6
" ld

Average measured values from the prism and split-cylinder tests, and the direct-tension tests by Habel and

Gauvreau (2008) were used to define a bilinear curve representing the tensile stress-strain response of the

ultra high-performance fibre-reinforced concrete mix being investigated at the University of Toronto (see

Figure 2-9). The material model for tension is only directly applicable to this particular mix design.

Tensile 3 prisms tested


Material model 150 mm ! 150 mm ! 900 mm
stress in MPa
moist-cured for 7 days
Tensile 30 air-cured for 47 days
stress tested at 54 days
ft! avg. material properties:

Obsolete
Et = 46.7 GPa
20 prism test data only !cr = 0.15·10-3 fcr! and !cr! based on
valid before cracking averages from prism data
fcr = 7.0 MPa
fcr
!t! = 1.5·10-3 !t! estimated using results
ft! = 14.6 MPa from direct tension tests by
10
Habel and Gauvreau (2008)

0 ft! estimated using results


0 !cr !t! prism test data
Tensile material model
from 3 split cylinder tests
strain 0
0 1·10-3 2·10-3

Tensile strain

Figure 2-9. Material model in tension

28
2. Ultra High-Performance Fibre-Reinforced Concrete

The initial linear portion of the material model curve represents the uncracked, linear-elastic behaviour of

the material in tension. As the material is loaded in tension, stress increases linearly with strain until the

cracking stress fcr is reached. Upon cracking, the material exhibits strain-hardening, as represented by the

second linear portion of the material model. In this domain, strain-hardening occurs due to: (1) the forma-

tion of multiple uniformly distributed cracks that are invisible to the naked eye, and (2) the progressive

pull-out of fibres (Habel 2004). Because the damage to the concrete matrix is gradual, there is a gradual

decrease in secant modulus. This strain-hardening behaviour concludes with the peak tensile stress ft! of

the material, after which a localized, visible crack forms. After this point, the material exhibits strain-

softening behaviour, in which the stress-carrying capacity of the material diminishes as a function of the

opening of the localized crack. Because of the continued, progressive pull-out of fibres, an abrupt loss of

stress does not occur. This strain-softening domain of behaviour of ultra high-performance fibre-rein-

forced concrete is neglected in the material model proposed in this thesis. This leads to conservative estim-

ates of member capacity.

For the purposes of this study and the analytical models developed later in this thesis, this tensile material

model will be assumed to remain constant. Tensile properties of conventional concrete are often related to

the compressive strength of concrete fc!. There is, however, insufficient data available to correlate tensile

cracking stresses with compressive strength. The steel reinforcement fibres, whose properties do not

change with the maturity of the concrete, provide a large share of the tensile capacity of ultra high-per-

formance fibre-reinforced concrete. Thus, assuming that the tensile behaviour of the cured material re-

mains constant is not unreasonable.

29
2. Ultra High-Performance Fibre-Reinforced Concrete

2.1.5 Shrinkage behaviour of the material

Concrete shrinkage is caused by autogenous shrinkage and drying shrinkage. Autogenous shrinkage is

caused by self-desiccation, in which water is withdrawn from capillary pores by the continued hydration of

concrete after it has set. Drying shrinkage is caused by the movement of water out of the non-rigid, porous

body of concrete.

Shrinkage strain 0.6·10-3

total shrinkage

0.4·10-3
autogenous shrinkage

0.2·10-3

0
0 20 40 60

Days after casting

Figure 2-10. Shrinkage strains as a function of days after casting. Adapted from Habel et al. (2008).

Shrinkage tests on six specimens made from the concrete mix design shown in Table 2-1 were done by

Habel et al. (2008) at the University of Toronto. Specimens were 305 mm in length and 76 mm in width

and height and were tested according to astm c 157. Autogenous shrinkage was measured from the three

specimens that were sealed with two layers of self-adhesive aluminum foil tape. Total shrinkage (the sum

of drying and autogenous shrinkage) was measured from the three unsealed specimens. These strains are

shown in Figure 2-10 as functions of time.

The French Interim Recommendations on ultra high-performance fibre-reinforced concrete (AFGC 2002)

suggest using a long-term shrinkage strain of 0.550·10-3. This estimate of long-term shrinkage strain correl-

ates well with the shrinkage strain data shown in Figure 2-10 obtained by Habel et al. (2008). Shrinkage

tests done by Graybeal (2006) on a similar material resulted in a long-term shrinkage strain of 0.555·10-3.

30
2. Ultra High-Performance Fibre-Reinforced Concrete

2.1.6 Creep behaviour of the material

Creep is the increase of strain over time caused by a constant stress in a concrete member. No tests were

done at the University of Toronto to study the creep behaviour of ultra high-performance fibre-reinforced

concrete. Results from other studies are presented below.

Creep strain 2.0·10-3


(excludes initial elastic strain)

1.5·10-3 initial elastic stress


compressive strength

1.0·10-3 initial elastic strain = 2.06·10-3


(not shown in plot)

final creep strain = 1.60·10-3


0.5·10-3
creep coefficient = 0.78

0
0 100 200 300 400

Elapsed days under load

Figure 2-11. Creep strain of an untreated ultra high-performance fibre-reinforced concrete specimen over time.
Adapted from Graybeal (2006).

Creep is normally quantified using the creep coefficient, which is defined as the ratio of additional creep

strain with time to the initial elastic strain caused by a constant applied stress. The French Interim Recom-

mendations on ultra high-performance fibre-reinforced concrete (AFGC 2002) suggest using a long-term

creep coefficient of 0.8 if no heat treatment is applied to the concrete. Creep tests conducted by Graybeal

(2006) resulted in a long-term creep coefficient of 0.78 for specimens without heat or steam treatment, as

shown in Figure 2-11. The long-term creep coefficient and shrinkage strains presented above will be used in

Chapter 6 in the design of new concepts for concrete arch bridges.

31
Chapter 3. Slender Ultra High-Performance Fibre-Reinforced
Concrete Columns

In this chapter, an analytical model for determining the ultimate capacity of slender ultra high-perform-

ance fibre-reinforced concrete members in combined flexure and axial compression is developed. This

model will be used to calculate member capacities for all the primary load-carrying members in arch

bridges, including the arch, spandrel columns, and prestressed deck. The methods and insights described

in this chapter will be used in Chapter 6 to design new concepts for arches using ultra high-performance

fibre-reinforced concrete.

The content of this chapter is divided into four sections. In Section 3.1, the structural model of an eccent-

rically loaded slender column is described and the importance of considering nonlinear behaviour is illus-

trated. In Section 3.2, a general method of analysis is proposed for estimating the capacity and load-deflec-

tion response of given ultra high-performance fibre-reinforced concrete columns, considering material

and geometric nonlinearities. The general method is validated experimentally with a set of column tests, as

described in Section 3.4. In the final section, a simplified design method is proposed for calculating ulti-

mate member capacity. This approximate method is shown to be accurate and conservative relative to the

more rigorous general method of analysis. All test specimens described in this chapter were made using

the concrete mix design given in Table 2-1. All sectional analyses described in this chapter use the material

model proposed in Section 2.1 adjusted according to the assumed or observed compressive strengths.

32
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

The research described in this chapter is the first known study of the behaviour of slender ultra high-per-

formance fibre-reinforced concrete members in combined flexure and axial compression. It offers a means

for designers to take advantage of the high compressive strength of ultra high-performance fibre-rein-

forced concrete and hence to exploit the economic potential offered by this material.

3.1 Response of eccentrically loaded columns

Q Q Q

w +
x M
line of action
of load N

Q Q M = Qe + Qw(x)

structural deformed sectional second-


model shape forces order
moment
Figure 3-1. Structural model for eccentrically loaded columns

A structural model for eccentrically loaded columns is shown in Figure 3-1. In this system, moments M are

caused by axial load Q times the distance between its line of action and the centroidal axis of the member.

As load Q is applied, the centroidal axis of the member deforms away from the line of action, producing

additional moments in the system. These additional moments can be accounted for by second-order ana-

lysis, in which states of equilibrium are calculated on the geometry of the deformed structure. Moments M

calculated using second-order analysis are called second-order moments and are given by Equation 3-1,

where e is eccentricity of load and w the deflection of the column:

M (x) = Qe + Qw(x)
Equation 3-1
M (x) = M 0 + M 1 (x)

The first term Qe in Equation 3-1 is called the first-order moment M0, because it corresponds to first-order

analysis, which is based on a state of equilibrium calculated on the geometry of the undeformed structure.

The second term Qw represents the additional moments M1 caused by member deformation.

33
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

The relative magnitude of first-order moments M0 and additional moments M1 is related to the bending

stiffness of the member. When members are stiff, deflections w are small and therefore additional mo-

ments M1 are small relative to first-order moments M0. When members are slender, deflections w are large

and therefore additional moments M1 can be large relative to first order moments M0. Thus, it is important

to calculate second-order moments M accurately when designing slender members, so that moment de-

mands of the member are not underestimated.

Calculating second-order moments M accurately requires an analysis that considers both geometric and

material nonlinearities. Geometric nonlinearity refers to the nonlinear load-deflection behaviour of the

column. One way to account for geometric nonlinearity is to use Vianello’s method of successive approx-

imations (Vianello 1898, Menn 1990). According to this method, deflections w can be calculated approxim-

ately using Equation 3-2 (Menn 1990), where w0 is first-order deflection, Q is applied load, and QE is Euler

buckling load:

1
w(x) = w0 (x) Equation 3-2
1 ! Q / QE

The Euler buckling load QE is given by Equation 3-3, where EI is flexural rigidity, k is effective length factor

(k=1 for pin-ended members), and L is length of column:

EI
QE = ! 2 Equation 3-3
( kL )2

Material nonlinearity refers to nonlinear stress-strain behaviour of the composite material, such as caused

by the cracking of concrete or the yielding of reinforcing steel. When used in sectional analysis, nonlinear

stress-strain curves produce nonlinear moment-curvature diagrams. One way to account for material non-

linearity when calculating deflections w is to replace flexural rigidity EI in Equation 3-3 with secant bend-

ing stiffness EI!. Secant stiffness is given by Equation 3-4 (Menn 1990), where , is curvature of the section:

34
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

M
EI !(N ) = Equation 3-4
"

Secant stiffness EI! varies with axial force N and moment M and is calculated as a secant to the moment-

curvature diagram calculated for a given point.

Load Q A
Axial force N A
sectional capacity
Q Q interaction diagram
A—geometric linearity, material linearity
B—geometric nonlinearity, material linearity
C—geometric nonlinearity, material nonlinearity
e
ultimate limit state ultimate limit state
B
w 1
e B
M ultimate N*
N load Q*
C C

stable unstable ultimate


Q equilibrium equilibrium moment M*

Deflection w response of mid-height section Bending moment M

Figure 3-2. Load deflection and sectional response of a slender column given different analytical assumptions

Comparing the linear and nonlinear response of a given eccentrically loaded slender column illustrates the

importance of considering both geometric and material nonlinearities. The column considered is 2300

mm long, is square in section with side lengths of 100 mm, and is made of ultra high-performance fibre-

reinforced concrete without conventional reinforcing steel. The applied load has an initial eccentricity e of

10 mm to the centre of the column. Figure 3-2 shows the load-deflection and axial force-moment response

of the column according to three sets of analytical assumptions that follow, corresponding to Curves A, B,

and C:

Curve A: This response considers geometric linearity and material linearity. Equilibrium is calculated

based on undeformed geometry, thus only first-order moments M0 are considered. The stress-strain beha-

viour of the material is assumed to be linear-elastic. First order deflections w0 are calculated using an elast-

ic modulus E of 42000 MPa, and uncracked moment of inertia of 833 cm4.

Curve B: This response considers geometric nonlinearity and material linearity. Equilibrium is calcu-

lated based on deformed geometry, thus additional moments M1 are considered and first-order deflections

35
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

w0 are magnified according to Equation 3-2. Because a linear-elastic material is assumed, first-order deflec-

tions w0 and the Euler buckling load QE are calculated using an elastic modulus E of 42000 MPa and un-

cracked moment of inertia of 833 cm4. Deflections w from Curve B are larger than those from Curve A for

any given load Q. The ratio between additional moments M1 and first-order moments M0 increases as the

applied load Q is increased. For the slender column considered, this ratio increases beyond two. This

shows that neglecting second-order effects, as done for Curve A, can lead to gross underestimates of bend-

ing moment demand in slender members.

Curve C: This response considers both geometric and material nonlinearities and is calculated based

on the general method described in Section 3.2. This method does not use the approximations considered

by Equations 3-2, 3-3, and 3-4, which are purely sectional calculations. Instead, the general method con-

siders the overall stability of the member. It uses the nonlinear stress-strain model proposed in Chapter 2

(assuming a compressive strength of 120 MPa) to calculate nonlinear moment-curvature diagrams, which

are then used to determine curvatures and deflections along the length of the column. Because the materi-

al behaves elastically at first, Curve C closely follows Curve B until the ultimate load Q* is reached. Ulti-

mate loads Q* are defined as the maximum applied axial loads that can be sustained by columns of given

length, cross-section, material, and eccentricity. Beyond this ultimate limit state, the applied load Q must

be decreased to maintain a state of equilibrium. Any attempt to increase load beyond the ultimate load will

rapidly increase deflections and moments causing the member to become unstable. As such, the ultimate

load Q*, which cannot be calculated using sectional calculations as in Curve B, marks an ultimate limit

state that occurs before sectional failure. The bending moment that is sustained by the critical section of

the column when the ultimate load Q* is reached will be called the ultimate moment M* (this is not the

same as the moment at sectional failure).

As illustrated by the different analyses described in the previous example, it is important to consider geo-

metric and material nonlinearities and the overall stability of the system when designing slender members.

36
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

For slender members loaded in compression and bending, ultimate limit states can be reached before sec-

tional failure occurs anywhere along the column.

3.2 General analysis method

The primary objective of the general analysis method described in the following section is to determine

the ultimate loads Q* of a given slender ultra high-performance fibre-reinforced concrete column for all

all possible values of initial eccentricity of load e. The structural model considered has been previously dis-

cussed and is shown in Figure 3-1. The model has a constant cross section, pins at each end of the column,

and an axial load, which is applied eccentric to the centre of the column.

From external quantities Q* and w*, the limiting sectional forces N and M* of the critical section of the

column can be calculated (see Figure 3-2). Ultimate moment M* is given by Equation 3-5, where e is the

initial eccentricity of load:

M * = Q * (e + w*) Equation 3-5

Expressing slender column capacity in terms of limiting sectional forces of the critical section allows for

the use of N-M* interaction diagrams, which correspond to the various ultimate limit states (Q=Q*) of the

column for all possible values of initial eccentricity. These ultimate N-M* interaction diagrams are reduced

versions of traditional N-M interaction diagram, which are instead calculated based on sectional failure.

Calculating N-M* interaction diagrams for a given column is the secondary objective of the general ana-

lysis method.

3.2.1 Methodology

A rational method for calculating N-M* interaction diagrams for slender prestressed concrete columns

(arranged as in Figure 3-1) is presented by Nathan (1985). Nathan’s method requires the following input:

material stress-strain curves, cross sectional geometry, and pin-to-pin length of the column. The method

has three steps:

37
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Step 1. A set of sectional analyses is used to calculate contours of equal curvature on an axial force-

moment diagram. These contours are used to calculate a set of nonlinear moment curvature diagrams,

each diagram associated with a different axial force N.

Step 2. Column deflection curves w(x), which are column deflections along their length for a given

axial load Q and mid-length deflection w(0), are calculated by solving Equation 3-1 numerically. A set of

column deflection curves are calculated by changing the mid-length deflection. The ultimate deflection

w*, corresponding to Q=Q*, is taken as the mid-length deflection w(0) of the column deflection curve

with the largest end eccentricity e. From w*, the ultimate moment M* of the critical section is calculated.

Step 3. The member capacity N-M* interaction diagram of the slender member is determined by re-

cording ultimate moments M* of the critical section for all values of axial force (N=Q) up to the maximum

axial force of the section under pure compression.

In the following sections, this method will be adapted for use with ultra high-performance fibre-reinforced

concrete by using the material model proposed in Chapter 2. The interactions among axial force, moment,

and curvature of a given section, which are unique to the stress-strain curve and geometry of the section

assumed, will be discussed. In addition to proposing a general method for calculating N-M* interaction

diagram for slender columns, a method for calculating load-deflection response of slender columns will be

presented, which is not included in Nathan’s formulations. This latter calculation method predicts the ulti-

mate load Q* and corresponding ultimate deflection w* of columns of given length, sectional geometry,

and eccentricity of load. Load-deflection response will be used to validate the general method

experimentally.

3.2.2 Calculating moment curvature diagrams

The first step in the general method is to calculate a set of nonlinear moment-curvature M-ϕ diagrams,

each diagram corresponding to a different axial force N. An effective way to do this is to calculate contours

of equal curvature on an axial force-moment diagram and then interpolate M-ϕ diagrams from those con-

tours. This computational method offers two advantages: (1) it does not require iterative calculations, and
38
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

(2) the contours of equal curvature give insight into the sectional behaviour of ultra high-performance

fibre-reinforced concrete sections. A given contour of equal curvature is found by calculating sectional

forces corresponding to a set of planes of strain, each with the same curvature ϕ but different strains on the

flexural compression face of the column %cf. Contours of equal curvature can be determined by dividing

the ranges of %cf and ϕ into intervals as in Equation 3-6, where %t! is strain at peak stress in tension, %c! is

strain at peak stress in compression, and d is depth of section. The limits of %cf and ϕ correspond to the

stress-strain limits proposed in the material model.

(N, M ) = f (! cf , " ) sectional analysis


! # % 1.4 ! c# Equation 3-6
1.4 ! c# $ ! cf $ 0 and 0 $ " $ t
d

Contours of equal curvature for a 100 mm by 100 mm square section using the material model presented

in Chapter 2 and a compressive strength of 130 MPa are shown in Figure 3-3. Figure 3-3b illustrates how

contours of equal curvature are calculated. First, various planes of strain with the same value of curvature

are chosen. Second, stress profiles for each plane of strain are calculated using the constitutive laws of the

material. Third, axial forces and moments are calculated through the integration of stresses. Last, the res-

ulting sectional forces from each plane of strain are plotted in axial force-moment space (see Figure 3-3a).

The envelope of all contours of equal curvature forms the axial force-moment N-M interaction diagram

corresponding to various states of sectional failure.

For this section, setting the strain on the flexural compression face %cf to 1.2%c! and setting the strain on the

flexural tension face %tf to %t! gives a good estimate of the plane of strain that results in the maximum mo-

ment of resistance of the section. This state is analogous to the balanced point in a conventionally rein-

forced concrete section, in which the section fails simultaneously by the concrete crushing at the flexural

compression face and by the yielding of steel near the flexural tension face. This state is labelled as the “bal-

anced point” in Figure 3-3a.

39
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

(a) Axial (b) planes of strain with equal !


force N
–1000 kN one contour of equal corresponding
curvature in rad/km stresses are integrated
! = 10 rad/km ... ...
to get N-M point on
–800
diagram

–600 100 mm
! = 10 20 30 40 50

100 mm
! = 30 rad/km ... ...
–400

balanced point ultra high-


–200
performance
fibre reinforced
! = 50 rad/km ... ... concrete section
0
0 4 8 12 16 kN·m
Bending moment M

(c) Axial (d)


force N Moment M
–1000 kN 16 kN·m 1
EI0
N = -500 kN
–800 N = -400 kN
12
N = -300 kN
C
–600
!=10 20 30 40 50 B
N = -500 kN 8 N = -200 kN
cracking point
–400 N = -400 kN
A B C
N = -300 kN N = -100 kN
4 A
–200 N = -200 kN
N = -100 kN
N = 0 kN
0 0
0 4 8 12 16 kN·m 0 30 60 90 120 rad/km
Moment M Curvature !

(e) Axial (f)


force N Moment M
–1000 kN 16 kN·m
N = -900 kN 1
N = -600 kN
D E F EI0
–800 N = -800 kN N = -700 kN
12 F
N = -700 kN
N = -800 kN
–600 N = -600 kN
cracking point
8 E N = -900 kN
–400
!=10 20 30 40

4 D
–200

0 0
0 4 8 12 16 kN·m 0 30 60 90 120 rad/km
Moment M Curvature !

Figure 3-3. Contours of equal curvature and corresponding moment-curvature diagrams

Moment-curvature diagram points for a given axial force N can be found by interpolating values of mo-

ment M from each contour of equal curvature. In Figures 3-3c through 3-3f, ten moment-curvature dia-

grams are calculated for ten different values of axial force N using contours of equal curvature. Points A

through F illustrate how moment-curvature diagrams for N=–300 kN and N=–800 kN are calculated. For
40
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

example, Point B lies on the intersection of N=–300 kN and contour of equal curvature with ϕ=30 rad/km.

The moment M associated with Point B is 9.5 kN·m, which is plotted on the moment-curvature diagram

shown in Figure 3-3d. Interpolating values of moment from each contour of equal curvature in this way

creates a complete nonlinear moment-curvature diagram for a given section and axial force N.

The axial force-moment-curvature N-M-ϕ relationship depicted in Figure 3-3 reveals trends in the mo-

ment-curvature behaviour of ultra high-performance fibre-reinforced concrete sections. The horizontal

spacing of the contours of equal curvature at a given axial force N is related to the linearity of the moment-

curvature diagram. Constant spacing corresponds to linear behaviour, while decreasing spacing corres-

ponds to decreases in secant stiffness EI!. The shape of the moment-curvature diagram depends on wheth-

er the axial force N is less than or greater than the axial force at the balanced point Nb, as described below:

Case 1: For states of strain where N<Nb, the spacing between contours of equal curvature get smaller

and smaller until they converge as curvature is increased (see Figure 3-3c). This corresponds to sectional

failure caused by the pull-out of steel fibre reinforcement. The initial slope of the resulting moment-

curvature diagram, shown in Figure 3-3d is equal to the uncracked, elastic flexural rigidity EI0, where E is

the elastic modulus of the concrete, and I0 is the moment of inertia of the gross uncracked section. The

cracking point, which occurs when the flexural tension face reaches the cracking strain %tf=%cr, is indicated

for the curve labelled N=–300kN in Figure 3-3d. For states of strain where N<Nb, the secant stiffness at ini-

tial cracking EI!cr is equal to EI0. After the section cracks, the secant stiffness EI! begins to decrease and

small increases in moment require large increases in curvature. The difference between cracking moments

and peak moment for the N<Nb curves shown in Figure 3-3d are relatively large.

Case 2: For states of strain where N≥Nb, the contours of equal curvature fold back on one another

(see Figure 3-3e), suggesting that straining the section beyond 1.4%!c leads only to states of equilibrium on

the descending post-peak branch of moment-curvature diagrams. The non-convergence of the contours

corresponds to sectional failure caused by concrete crushing. The ascending branch of moment-curvature

diagrams in Figure 3-3f is bilinear before cracking due to the bilinear form assumed for the compressive

41
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

stress-strain model. For states of strain where N≥Nb, the secant stiffness at cracking EI!cr decreases as axial

force N is increased. The difference between cracking moments and peak moments of the N≥Nb curves

shown in Figure 3-3f is relatively small.

3.2.3 Calculating column deflection curves

With the axial force-moment-curvature N-M-ϕ relationship of the section established, the equilibrium of a

given member can now be studied. Equating moment demand (Equation 3-1) and moment resistance

(Equation 3-4) gives Equation 3-7, where Q is applied load, e is eccentricity of load, w(x) is deflection of the

column along its length, EI! is secant stiffness, N is axial force of the section, and ϕ(x) is curvature of the

column along its length:

Qe + Qw(x) = EI !(N ) " # (x) Equation 3-7

By approximating curvature ϕ(x) as the second derivative of the column deflection w"(x) using a second-

order, small-displacement approach, and by equating axial force N with applied load Q, the following di-

fferential equation is obtained:

!EI "(x) # w""(x) + Qw(x) + Qe = 0 Equation 3-8

Solving Equation 3-8 for w(x) for a given applied load Q, mid-length moment M(0), and column length

produces one column deflection curve. Because secant stiffness and corresponding nonlinear moment-

curvature diagram cannot be expressed in functional form, Equation 3-8 has to be integrated numerically

over the length of the column. Only half the column needs to be considered, since the column and load

configurations are symmetric. To perform the numerical integration, the half-length of the column is di-

vided into n elements of equal length Δx as shown in Figure 3-4.

42
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

–EI!(x)·w"(x)+Qw(x)+Qe=0

e+w(x)
e
Q Q

start end
column column
mid-length L end
2

elem. 1 elem. 2 elem. i


elem. n-1
..... elem. n
.....

w(0) w M w w(L/2)
w!(0) w! w" w! w!(L/2)

!x !x !x !x !x !x
2 2
x

Figure 3-4. Numerical integration procedure

One column deflection curve w(x) is calculated given the following input parameters: column length L,

axial load Q, moment-curvature diagram M-ϕ, and mid-length moment M(0). The numerical integration

procedure begins at mid-length (x=0) and proceeds toward the end of the column (x=L/2). The value of

initial eccentricity e is solved at the end of the integration procedure, rather than being specified as an in-

put parameter. Listed below are the steps for calculating one column deflection curve w(x), adapted from

Nathan’s method (1985):

1. The total eccentricity w(0)+e between the line of action of load and the centroid of the column at

mid-length is calculated based on the specified mid-length moment M(0) and axial load Q:

M (0)
w(0) + e = Equation 3-9
Q

2. The mid-length slope w!(0) is set to zero (condition of symmetry).

3. The total eccentricity w+e and slope w! at mid-length are assigned to the total eccentricity and slope

at the beginning of the first element:

43
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

w(x0 ) + e = w(0) + e Equation 3-10

w!(x0 ) = w!(0) Equation 3-11

4. The total eccentricity w+e and moment M at the middle of the current element are approximated:

w(x0 + 12 !x) + e = w(x0 ) + e " w#(x0 ) 12 !x Equation 3-12

M (x0 + 12 !x) = Q " ( w(x0 + 12 !x) + e ) Equation 3-13

5. The curvature ϕ at the middle of the current element is found by dividing moment M by secant stiff-

ness EI! using the given moment-curvature diagram. The second derivative of deflection w" is as-

sumed to be equal to the curvature:

M (x0 + 12 "x)
w!!(x0 + 12 "x) # $ (x0 + 12 "x) = Equation 3-14
EI !

6. The total eccentricity w+e at the end of the current element is approximated using the first three

terms of a Taylor series expansion of deflection w:

w(x0 + !x) + e = w(x0 ) + e + w"(x0 )!x + 12 w""(x0 + 12 !x)!x 2 Equation 3-15

7. The slope w! at the end of the current element is approximated using the first two terms of its Taylor

series expansion:

w!(x0 + "x) = w!(x0 ) + w!!(x0 + 12 "x)"x Equation 3-16

8. The total eccentricity w+e and slope w! at the end of the current element are assigned to the total ec-

centricity and slope at the beginning of the next element:

w(x0 )i+1 + e = w(x0 + !x)i + e Equation 3-17

w!(x0 )i+1 = w!(x0 + "x)i Equation 3-18

9. Steps 4 through 8 are repeated for all elements until the end of the column is reached.

44
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

10. Because the deflection of the column is zero at its end, the total eccentricity w+e at the end of the

last (nth) element is equal to the initial eccentricity e:

e = w(x0 + !x)n + e Equation 3-19

11. The column deflection curve w(x) can be determined by subtracting the initial eccentricity e from

all recorded values of total eccentricity w(x)+e.

The purpose of calculating column deflection curves w(x) is to identify which combinations of axial load Q

and initial eccentricity e cause the column to fail. To do this, a family of column deflection curves needs to

be calculated, each with the same axial load Q, but with different specified mid-length moments M(0). The

mid-length moment of the first curve in the family is assigned the peak moment from the given moment-

curvature diagram. Subsequent curves in the family are assigned mid-length moments that are success-

ively less than the peak moment, until a mid-length moment of zero is reached. As the number of column

deflection curves in the family is increased, the resolution of the analysis is increased. The higher the resol-

ution is, the higher the accuracy is in determining the ultimate load. Some of the resulting column deflec-

tion curves will have initial eccentricities e that are negative. These curves represent unstable column

deflection states that occur after the ultimate load has been reached. The process of identifying the ulti-

mate limit state among deflection curves in a family is explained in the next section.

45
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

3.2.4 Calculating member capacity interaction diagrams

This section describes how member capacity N-M* interaction diagrams are calculated. Each point on an

N-M* interaction diagram corresponds to a unique ultimate limit state at which axial load Q has reached

the ultimate load Q*. One point on a member capacity N-M* interaction diagram is calculated by identify-

ing which curve in a family of column deflection curves corresponds to an ultimate limit state.

Column deflection curves for an ultra high-performance fibre-reinforced concrete member subjected to

an axial load Q1 of 600 kN are shown in Figure 3-5a,c,e. The dashed line directly below the column deflec-

tion curves is the lines of action of the applied load. Three column lengths are considered: 1200 mm, 2050

mm, and 2800 mm. The nonlinear moment-curvature diagram used to calculate the column deflection

curves is based on a 100 mm by 100 mm square section and the material model presented in Chapter 2 us-

ing a compressive strength of 130 MPa. Curves A, D, and G correspond to mid-length moments causing

sectional failure (peak moment from moment-curvature diagram). All other curves correspond to mid-

length moments less than the moment causing sectional failure.

46
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

e+w(x)
e
Q Q

Domain of Column deflection curves


slenderness for Q1 = 600 kN Response of mid-length section

(b) sectional capacity


(a) Axial N-M interaction
Slender column diagram
This 1200 mm column A force N
w(0)
reaches sectional failure B C B A
when axial load Q1 is N=Q1
C e1
applied at an eccentricity Q1 Q1
e1, as represented by
deflection Curve A. response of mid-
600 mm length section for
line of action of load e1 e=e1
mid-length section
1
Moment M=Q(e+w(0))

(c)
D

Very slender column E member capacity


(d)
N-M* interaction diagram
This 2050 mm column F Axial
N-M interaction
reaches ultimate load when Q1 Q1 force N
axial load Q1 is applied
F D
with eccentricity e2, as 1025 mm N=Q1
E
represented by deflection
Curve E. The deflection Curve E represents an ultimate
state because it has the highest
curve with the highest end eccentricity in this family of curves. E
D
eccentricity corresponds to Curves D and F have the same F e2
an ultimate limit state. eccentricity. Only Curve F e3
represents a stable state of Q1
equilibrium. 1 Moment M=Q(e+w(0))
e3

(f)
Axial N-M interaction
Unstable column (e)
force N
This 2800 mm column is G
unstable at this axial load H N=Q1
Q1, since all column
I N=QE (Euler buckling)
deflection curves in the Q1
Q1
family extend below the
line of action of load. N-M* interaction
1400 mm

Moment M=Q(e+w(0))

Figure 3-5. Column deflection curves and sectional responses of mid-length sections for various column lengths

In Figure 3-5a,c,e, the left sides of the column deflections curves are at mid-length, while the right sides are

at the pinned ends of the columns. These two boundaries define the length over which the half-column is

allowed to bend. Mathematically, this corresponds to the bounds over which Equation 3-8 is integrated.

Since the deflection of the columns is zero at the pinned supports w(L/2)=0, the distance from the line of

47
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

action to the end of the column is equal to the initial eccentricity of load e. The value of e is different for

each column deflection curve in the family of curves. In general, the column deflection curve with the

greatest eccentricity at the end of the column is the curve that corresponds to ultimate load Q*. This will

be explained in greater detail according to the three domains of slenderness described below:

Slender Column: The curve with the greatest end eccentricity e is the curve with the greatest mid-

length eccentricity w(0)+e. The mid-length moment M(0) of this curve is equal to the peak moment of the

moment-curvature diagram, and thus corresponds to sectional failure. This means that the column reaches

sectional failure without first reaching ultimate load Q*. This type of slender column behaviour is illus-

trated by the column deflection curves shown in Figure 3-5a. All column deflection curves in the diagram

are subjected to an axial load Q1 of 600 kN. Curve A is the critical column deflection curve with end ec-

centricity e1. Curves B and C are noncritical column deflection curves because their end eccentricities are

less than e1.

(b) sectional capacity


(a) Axial N-M interaction
Slender column diagram
This 1200 mm column A force N
w(0)
reaches sectional failure B C B A
when axial load Q1 is N=Q1
C e1
applied at an eccentricity Q1 Q1
e1, as represented by
deflection Curve A. response of mid-
600 mm length section for
line of action of load e1 e=e1
mid-length section
1
Moment M=Q(e+w(0))

Reprint of Figure 3-5a,b. Slender column deflection curves

The diagram in Figure 3-5b shows possible axial force-moment responses of the mid-length section that

lead to the column deflection curves represented by Curves A, B, and C. Each axial force-moment re-

sponse corresponds to a unique eccentricity of load. Only Curve A, which has the highest eccentricity of

load and lies on the section capacity N-M interaction diagram, represents a state of failure when Q=Q1.

Very Slender Column: The curve with the greatest end eccentricity e does not correspond to the curve

with the greatest mid-length eccentricity w(0)+e. In this case the curve with the greatest end eccentricity

has a mid-length moment M(0) that is less than the peak moment of the moment-curvature diagram. This

48
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

means that the column reaches ultimate load Q* before reaching sectional failure. This type of very slender

column behaviour is illustrated by the column deflection curves shown in Figure 3-5c. All column deflec-

tion curves in the diagram are subjected to an axial load Q1 of 600 kN. Curve E is the critical column

deflection curve with end eccentricity e2. Curves D and F are noncritical column deflection curves because

their end eccentricities are less than e2. Curve D, which has the highest mid-length eccentricity w(0)+e,

folds back on to the curves with lower mid-length eccentricities. This means that two column deflection

curves may have the same end eccentricity e and may be part of the same axial force-moment response.

This is the case for Curves D and F, which share the same end eccentricity e3. This pair of column deflec-

tion curves represents two viable solutions to Equation 3-8.

(c)
D

Very slender column E member capacity


(d)
N-M* interaction diagram
This 2050 mm column F Axial
N-M interaction
reaches ultimate load when Q1 Q1 force N
axial load Q1 is applied
F D
with eccentricity e2, as 1025 mm N=Q1
E
represented by deflection
Curve E. The deflection Curve E represents an ultimate
state because it has the highest
curve with the highest end eccentricity in this family of curves. E
D
eccentricity corresponds to Curves D and F have the same F e2
an ultimate limit state. eccentricity. Only Curve F e3
represents a stable state of Q1
equilibrium. 1 Moment M=Q(e+w(0))
e3

Reprint of Figure 3-5c,d. Very slender column deflection curves

The diagram in Figure 3-5d shows possible axial force-moment responses of the mid-length section that

lead to the column deflection curves represented by Curves D, E, and F. The points representing the mid-

length sectional forces of Curves F and D lie on the same axial force-moment response. The former point

lies on the ascending branch of the sectional response, while the latter point lies on the descending branch.

Only the former point (Curve F) represents a stable state of equilibrium for which the ultimate load Q*

has not been reached. The point representing the mid-length sectional forces of Curve E lies at the apex of

the second axial-force moment response, and thus corresponds to an ultimate limit state. The mid-length

moment of this curve is the ultimate moment M* for this value of applied load Q=Q1. The ultimate mo-

49
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

ment M* is less than the moment causing sectional failure (sectional failure is marked by the dashed N-M

interaction diagram).

A second envelope is marked in Figure 3-5d with a dotted line. This envelope, called the member capacity

N-M* interaction diagram, is a locus of all possible ultimate limit states for the specific length of column

considered. At each point along this envelope, different combinations of ultimate load Q* and ultimate

moment M* are reached, similar to the ultimate state represented by Curve E.

Unstable Column: At the end of the column, all curves are below the line of action of applied load,

corresponding to negative end eccentricities e<0. This is the case for all column deflection curves shown

in Figure 3-5e. This means that the column is unstable under axial load Q1. As shown in Figure 3-5f, the

dotted N-M* interaction diagram for this specific length of column is less than Q1. The upper boundary of

this interaction diagram is equal to the Euler buckling load of the column, as calculated by Equation 3-3.

(f)
Axial N-M interaction
Unstable column (e)
force N
This 2800 mm column is G
unstable at this axial load H N=Q1
Q1, since all column
I N=QE (Euler buckling)
deflection curves in the Q1
Q1
family extend below the
line of action of load. N-M* interaction
1400 mm

Moment M=Q(e+w(0))

Reprint of Figure 3-5e,f. Unstable column deflection curves

With the three domains of slenderness described above, ultimate moments M* for any applied axial load Q

can be determined. The axial load Q equals the ultimate load Q* when the moment of the mid-length sec-

tion M(0) reaches ultimate moment M*. Ultimate moments M*, thus, represent the highest moment M

that the given member can carry safely at the specified axial load Q. Member capacity N-M* interaction

diagrams are the collection of ultimate moments M* for all different values of axial force N. By equating

axial force N with axial load Q, sectional limits of the critical mid-length section can be defined, represent-

ing states at which ultimate loads Q* of the member have been reached. Constructing the member capa-

city N-M* interaction diagram proceeds as follows:

50
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Step 1. A family of column deflection curves is calculated for a given axial load Q.

Step 2. The axial load Q and mid-length moment M of the deflection curve with the highest eccentri-

city e at the end of the column are recorded (these are the ultimate load Q* and ultimate moment M*, re-

spectively). The mid-length moment recorded is zero if all curves are below the line of action of the load at

the end of the column (this corresponds to the unstable column condition).

Step 3. The axial load Q is incremented and Steps 1 and 2 are repeated until the failure load in pure

axial compression Q! is reached. Q! is given by Equation 3-20, where fc! is the compressive strength of con-

crete in compression and A is the cross-sectional area of the column:

Q ! = fc! " A Equation 3-20

Step 4. Recorded values are plotted on an axial force-moment diagram, setting recorded values of Q

equal to N. This locus of points represents all possible ultimate limit states of the given column.

The member capacity N-M* interaction diagram for a 100 mm × 100 mm ultra high-performance fibre re-

inforced concrete column with a length of 2310 mm is shown in Figure 3-6a (dotted curve). This length

corresponds to a slenderness ratio ( of 80. Slenderness ratios ( are normalized measures of column length

L as defined by Equation 3-21, where r is radius of gyration of the section and k is effective length factor:

kL
!= Equation 3-21
r

Each point on the N-M* interaction diagram corresponds to a unique combination of ultimate load Q*

(N=Q*) and ultimate moment M*. If the given column were loaded axially with initial eccentricities e of

0.1, 0.5, 1, 2, 4, 8, 16, and 32 mm, the sectional response of the mid-length section of the column would fol-

low the response curves shown in Figure 3-6a. These response curves show that each point on the N-M*

interaction diagram corresponds to unique initial eccentricities of load e.

51
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

member capacity N-M* sectional capacity N-M


interaction diagram interaction diagram member capacity N-M*
Axial interaction diagrams for capacity of a member
force N ultimate limit states Q=Q* various slenderness ratios ! decreases as slender-
1000 sectional response at ness ! increases
N
mid-length for various e: !=60
Q Q e=0.1mm
0.5mm
750 1mm
!=80 2mm !=80
e 4mm
8mm
w 500
!=100
M
!=120
N 250 !=140
16mm
Q 32mm
0
0 4 8 12 16 M
Bending moment M
(a) (b)

Figure 3-6. Member capacity interaction diagrams for various slenderness ratios

Member capacity N-M* interaction diagrams define the sectional capacity of the critical section of the

column. If the axial-force and second-order moment for given load cases are within the space bounded by

N-M*, then the member can safely carry the load. Load cases with N-M combinations that lie outside the

envelope cannot be safely be carried by the column.

The member capacity N-M* interaction diagram changes with the slenderness ratio ( of the column. Fig-

ure 3-6b shows the member capacity interaction diagrams for different slenderness ratios of the 100 mm ×

100 mm column considered previously. Decreasing ( increases the allowable axial forces and moments

that can be sustained by the mid-length section of the column. At some value of (, the N-M* interaction

diagram utilizes the full capacity of the section. In this case, this occurs when ( is decreased below 60. Dis-

playing the trade-off between slenderness and member capacity visually as in Figure 3-6b can help design-

ers select member sizes that are slender, safe, and structurally efficient.

3.2.5 Calculating load-deflection response of a given column and eccentricity of load

The load-deflection Q-w response of slender concrete columns with given initial eccentricities of load e

can be calculated using the column deflection curves described in the previous section. The calculation

process is similar to the way ultimate moments are calculated, except that the column deflection curve of

52
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

interest within a family of curves is the one that produces the given eccentricity at the end of the column

(previously, the curve with the highest end eccentricity was selected among those in the family).

Figure 3-5c, which shows column deflection curves for a very slender column, can be used to illustrate this

process. The column arrangement considered is a 100 mm × 100 mm ultra high-performance fibre-rein-

forced concrete column with a length of 2050 mm and a constant eccentricity of load e3. Based on the fig-

ure, the mid-length moments of both Curves D and F would be recorded for this value of axial load Q1 and

eccentricity e3. Setting axial force equal to axial load N=Q1, two points, labelled D and F, can be plotted as

a part of the axial force-moment response of the mid-length section (see Figure 3-5d). Other N-M points

can be plotted by repeating this process with different families of column deflection curves corresponding

to different axial loads Q.


(c)
D

Very slender column E member capacity


(d)
N-M* interaction diagram
This 2050 mm column F Axial
N-M interaction
reaches ultimate load when Q1 Q1 force N
axial load Q1 is applied
F D
with eccentricity e2, as 1025 mm N=Q1
E
represented by deflection
Curve E. The deflection Curve E represents an ultimate
state because it has the highest
curve with the highest end eccentricity in this family of curves. E
D
eccentricity corresponds to Curves D and F have the same F e2
an ultimate limit state. eccentricity. Only Curve F e3
represents a stable state of Q1
equilibrium. 1 Moment M=Q(e+w(0))
e3

Reprint of Figure 3-5c,d. Very slender column deflection curves

Once the complete axial force-moment N-M response of the mid-length section has been calculated, the

load-deflection Q-w response of the column can be calculated. In this column arrangement, mid-length

deflection w(0) is given by the following equation, where axial load Q is taken as axial force N:

M (0)
w(0) = !e Equation 3-22
Q

Responses calculated using this method assume that the loads applied are deflection-controlled and that

they can be reduced once ultimate load Q* has been reached. These load-deflection responses will be used

in Section 3.4.4 to validate the proposed analytical model experimentally.

53
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

3.3 Overview of computer program

The computer program qult was written by the author to model the behaviour of ultra high-performance

fibre-reinforced concrete slender members under combined compression and bending, according to the

general method described in Section 3.2. The program considers nonlinear material behaviour in sectional

analysis and calculates equilibrium based on deformed geometry (i.e. second-order analysis). For a given

stress-strain curve, constant sectional geometry, and column length, qult returns the member capacity N-

M* interaction diagram of the column, which bounds all admissible combinations of axial force and mo-

ment that can be carried safely by the critical section of the member (see Section 3.2.4). For a given stress-

strain curve, constant sectional geometry, column length, and eccentricity of load, qult returns the load-

deflection Q-w response of a column subjected to an eccentrically applied axial load (see Section 3.2.5).

From this, the ultimate load, or maximum load sustained by the column, is determined.

The overall structure of the program is based on a similar program developed by Nathan in his paper en-

titled "Rational Analysis and Design of Prestressed Concrete Beam Columns and Wall Panels" (1985). As

shown in Figure 3-7, qult is organized as a hierarchy of functions: matmodel defines the stress-strain re-

sponse of the material, geom defines the geometry of the section, sectforce performs sectional analysis,

contourcurv calculates contours of equal curvature of the section, momcurv calculates the axial force-

moment-curvature response of the section, deflection calculates column deflection curves of the

column, response calculates the load-deflection response of the column for a given eccentricity of load,

and capacity calculates the member capacity interaction diagram of the column. The function ultimate

returns the ultimate load Q*, corresponding deflection w*, and ultimate moment M* of the column for a

given eccentricity of load. Functions at any calculation stage in the program can be executed to obtain in-

termediate results. For example, a set of column deflection curves can be calculated using deflection giv-

en that the following input parameters are defined: stress-strain curve, sectional geometry, member length,

and applied load.

54
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

sectional
member
geometry & sectional member
response
material analysis deflection
& capacity
response

input parameters stress-strain curve stress-strain curve stress-strain curve stress-strain curve
sectional geometry sectional geometry sectional geometry sectional geometry
member length member length
applied load applied load
eccentricity of load
FUNCTIONS GEOM
superscripts refer CENTROID1
to the sample output SECTPROP
diagrams below
MATMODEL2 SECTFORCE3
GETSTRESS NMBOUND
MFACTOR CONTOURCURV4 MOMCURV5 RESPONSE7
ULTIMATE
DEFLECTION6 CAPACITY8
depth
ENVELOPES

plane stress
width of strain  

element strip
moment

axial moment
force
1. CENTROID
curvature
axial load
3. SECTFORCE ultimate load

stress
5. MOMCURV
 
contour of
axial equal curvature
force   7. RESPONSE
strain + eccentricity
column
 

2. MATMODEL sectional failure


moment
mid-length moment limit
axial
load end moment limit
force
4. CONTOURCURV
half length of member

6. DEFLECTION
moment

8. CAPACITY

Figure 3-7. Overview of computer program QULT

qult is written in matlab r2007a and is included with this thesis digitally as a set of matlab m-files.

Table B-1, which is in Appendix B, lists all the major functions used in qult and describes their input

parameters, calculation steps, and output data.

55
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

3.4 Experimental validation

Four eccentrically-loaded ultra high-performance fibre-reinforced concrete columns were tested at the

University of Toronto for the purpose of validating the general analysis method described in Section 3.2.

The primary objectives of this test program were (1) to study the effect of slenderness on ultimate load Q*,

(2) to assess the accuracy of the load-deflection response predicted by the general method, and (3) to ob-

serve the mode of failure of the columns. Details of the test specimens and connections are shown in Fig-

ure 3-8, along with their as-built measurements. The complete set of column drawings, including measure-

ments can be found in Appendix C. The pin-to-pin lengths of the specimens ranged from 1737 mm to 3466

mm. The longest column had a slenderness ratio ( of 117.8.

Figure 3-8. Column drawings, details, and measurements

56
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

3.4.1 Batching, casting, and curing

Three 48 litre batches of ultra high-performance fibre-reinforced concrete were prepared according to the

mix design in Table 2-1 on page 17. Materials were mixed in a 0.17 m3 mortar mixer in the following order:

(1) the cement, silica fume, and sand were dry mixed until they were homogeneous, (2) the superplasti-

cizer and water (pre-mixed) were added and mixed until no dry materials remained, and (3) the steel fibres

were mixed in. The mixing process took 5-10 minutes per batch.

The finished batches were poured into a concrete bucket and then placed directly into the horizontally ori-

ented forms as shown in Figure 3-9. Specimens A and B each had their own batch, while Specimens C and

D shared the same batch. Because the concrete is self-consolidating, no vibration was done during place-

ment. The specimens and cylinders were left in the laboratory at approximately 20°C and moist-cured for 7

days after concrete was placed. The average compressive strengths fc! of the cylinders on the days of

column testing are shown in Figure 3-8.

Figure 3-9. Placing concrete directly into plywood forms

3.4.2 Setup and instrumentation

The desired eccentricity of load was built into steel end plates as shown in Figure 3-8. Welded steel studs

were offset 10 mm from the centre of each plate. After the studs were manually aligned with the centroid of

the form void, the steel end plates were clamped to the plywood form (see Figure 3-10), which ensured a

clean load transfer from plate to column.

57
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Figure 3-10. Clamped steel end plate (left), spherical bearing at base (right)

Spherical bearings were built into receiving steel plates at the two ends of the column, allowing for rotation

about all axes (see Figure 3-10). This eliminated any directional bias of the column deflection, except for

that of the eccentricity of load. Friction between the bearing and pins were not measured and were not ac-

counted for in calculations done according to the general method.

A few days before testing, 60 mm surface strain gauges were installed along the centreline of the flexural

compression face of each specimen at a centre-to-centre spacing of 300 mm. Along the centreline of one

adjacent face, 6 mm tooling balls targets were glued to the column at a spacing of 150 mm. Tooling ball tar-

gets were also glued to the centre of each pin. A metris cohesive laser radar scanner was used to measure

the three spatial coordinates of each tooling ball with respect to a predefined frame of reference. With

these measurements, column deflection curves w(x) could be recorded directly at any given load stage.

In order to prevent premature concrete cracking, forms were designed to be stiff enough to support the

columns during their transportation, erection, and installation into the test machine (see Figure 3-11).

Forms were removed after each column was pinned to the machine head. A linear variable differential

transformer was installed at the middle of the column to measure mid-length deflections of the column in

the direction of the eccentricity of load.

58
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Figure 3-11. Installing specimen with forms into test frame

Specimen A Specimen B Specimen C Specimen D


3466 mm 2892 mm 2317 mm 1737 mm pin-to-pin length
Height of targets 4000
above bottom pin
in mm —9.4 mm —11.5 —8.9 mm —5.8 mm dashed line values
mm mark average
3500 eccentricity of load,
top pin excluding top and
bottom pin

3000
top pin

2500

top pin

2000

A 9.4 mm top pin

1500
12.3 mm A Point A marks the
mid-length target
of each column
A 8.8 mm
1000
A 5.3 mm

500

all measurements
taken at zero load
bottom pin at (0,0) 0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20 0 5 10 15 20

Horizontal distance of centreline targets from a line passing through the top and bottom pin targets in mm

Figure 3-12. Position of cohesive laser radar scanner targets at zero load

For each column, initial readings of all tooling ball targets were taken before load was applied. Figure 3-12

shows the position of each target on the column relative to the line of action of load, which is taken as a

59
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

line passing through the top and bottom pin targets. For each column, the as-built initial eccentricity of

load was measured as the horizontal distance between the middle target (Point A) and the line of action of

load. The error in manually positioning each target using a ruler is assumed to be ±1 mm. The estimated

margin of error in measuring the initial eccentricity of load is thus ±2 mm. Pin-to-pin length was taken as

the vertical distance between the top and bottom pin targets.

The nominal offset of 10 mm (built into the steel end plates) was chosen to produce a load-deflection re-

sponse in which second-order effects could be observed. The resulting as-built eccentricities of load varied

between 5.3 mm and 12.3 mm. The variations are not ideal, but are acceptable because they can be accoun-

ted for in general method calculations. Increasing the overall scale of the specimens to reduce the effects of

construction error was not possible due to restrictions on the height clearance of the testing machine.

3.4.3 Load testing and failure

A large-scale universal mts mobile testing machine with a load capacity of 2650 kN was used to exert a

downward load on the top pin of each specimen. The load tests were displacement-controlled using a com-

puter-controlled, electro-hydraulic servo testing system. Specimens were loaded in 10 to 15 stages. At each

load stage, the machine displacement was held so that the scanner could measure the position of each tar-

get. Because the tests were vertical displacement-controlled, the columns would slowly lose load and

deflect while scanner measurements were being taken. Figure 3-13 shows the effect of this creep behaviour

on the load-deflection response of Specimen C between load stages. The plots of tangent slopes dQ/dw and

secant slopes Q/w reveal that the stiffness fluctuations caused by the discontinuous application of load did

not adversely affect the overall active load-deflection response of the column.

Due to safety concerns, close examination of the columns during the post-peak load stages of the test was

not possible. At failure, each specimen developed a full depth crack within 300 mm of the middle of the

column. The failure surface of Specimen C is shown in Figure 3-14. Fibres were left protruding from the

failure surface. The concrete matrix that had held these fibres together had disintegrated into powder. Ex-

cept for the failure crack, all other cracks that had formed in the specimen were too small to be seen.

60
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Q Applied load 600 machine displacement


Q in kN held at end of load stage

e 400

w
200

Q 0
w
100 spikes correspond to
Tangent slope creeping of column
dQ/dw in kN/mm between load stages
Q tan shape of curve with
50 continuous loading

w upon reactivating the load,


the tangent slope returns
0 to curve with continuous loading

–50
w
Secant slope 60
fluctuation of secant
Q/w in kN/mm slope between load
stages are minimal
c
se

Q 40

w
20

0
0 10 20 30

Horizontal deflection of
column at mid-length w in mm

Figure 3-13. Creep behaviour of Specimen C

Figure 3-14. Failure surface of Specimen C

61
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

3.4.4 Test results and discussion

Load-deflection responses of all specimens are shown in Figure 3-15 along with photographs of the

columns at different load stages. Load stages are marked on the load-deflection plot with small solid

circles. Load stage 0 refers to the unloaded state of the column. At each subsequent load stage, photo-

graphs were taken each time the vertical displacement of the column was held to allow for scanner meas-

urements. This figure presents a visual record of the deflected shape of the column deflection w(x) as a fun-

ction of applied load Q. The lines superimposed on the photographs indicate the location of the mid-

length section of the column and the line of action of load, defined by a line passing through the centre of

the top and bottom pins. These lines create a frame of reference that can help the reader compare subtle

changes in the deflection of the column. The last photograph in each set shows the location of the full

depth crack in each failed column, except for Specimen B for which a photograph at failure was not taken.

The displacement measurements taken by the laser radar scanner at each load stage were fitted to 4th order

polynomial equations w(x). These polynomials were then differentiated twice to obtain approximate ex-

pressions for curvature ϕ(x) along the column. These curvatures ϕ along with the surface strain gauge

measurements on the flexural compression face of the column were used to calculate the planes of strain of

several cross-sections along the column. Strains on the flexural tension face %tf of the column were calcu-

lated according to Equation 3-23, where %cf is strain of the flexural compression face and d is the depth

measured between the two faces.

! tf = ! cf + " # d Equation 3-23

62
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

load stage 0 1 2 3 4 5 6 7* 8 9 10 11 Fail


Specimen A
eccentricity e = 9.4 mm
pin-to-pin length = 3466 mm

axial load
Q in kN
800

600

400

200
w*
0
0 20 40 60
deflection w at
mid-length in mm

load stage 0 1 2 3 4 5 6 7 8 9 10* 11


Specimen B line of action of load
Q eccentricity e = 12.3 mm
pin-to-pin length = 2892 mm
800

600

400 mid-length

200
w*
w
0
0 20 40 60

load stage 0 1 2 3 4 5 6 7 8 9 10 11 12 13* Fail


Specimen C
Q eccentricity e = 8.8 mm
pin-to-pin length = 2317 mm
800

600

400
w*
200

w
0
0 20 40 60

Q load stage 0 1 2 3 4 5 6 7 8 9* Fail


Specimen D
800 eccentricity e = 5.3 mm
pin-to-pin length = 1737 mm

600
w*

400

200

w
0
0 20 40 60

Figure 3-15. Load-deflection response of specimens and photographs at each load stage

63
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Figure 3-16 presents all key measurements recorded during each load test. From top to bottom, the follow-

ing data sets are shown for each column load test: (1) photographs at the unloaded, ultimate, and failed

states, (2) as-built measurements, (3) ultimate loads Q*, deflections w*, and moments M*, (4) load-deflec-

tion Q-w response, (5) axial force-moment N-M response of the mid-length section, (6) strain of the flex-

ural tension face %tf and strain of the flexural compression faces %cf of the mid-length section versus mo-

ment M, and (7) curvature-moment ϕ-M response of the mid-length section.

The effect of slenderness on ultimate load can be observed from the load-deflection diagrams in Figure 3-

16. For columns with the same initial eccentricity of load e, the ultimate load of the column decreases as

column slenderness is increased. Eccentricity of load and sectional geometry also affect the value of ulti-

mate load according to the analytical model developed in Section 3.2. These parameters were not held per-

fectly constant in the load tests due to construction tolerances, but are at least close in value.

Observations on the behaviour of the mid-length section can be made by comparing the bottom three sets

of plots in Figure 3-16. The dashed lines that link the bottom three plots mark the value of the observed

ultimate moment M* for each load test. On the diagrams of extreme fibre strains versus moment, the

shaded regions represent the typical elastic stress range based on the material model developed in Chapter

2, using a compressive strength of 131 MPa. It can be seen that the mid-length sections behave inelastically

(data points move outside the shaded area) before the ultimate loads of the columns have been reached.

Thus, it would be unconservative to assume a purely elastic stress-strain response to simplify the calcula-

tion of ultimate load.

64
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Q=0 Q=Q* Fail


note: column
top pin
photos resized
Q Q to 1:50 scale

e Q=0 Q=Q*

w
M Q=0 Q=Q* Fail
N

mid-length
Q Q=0 Q=Q* Fail

Specimen A B C D
mid-length eccentricity in mm 9.4 12.3 8.8 5.3
mid-length depth in mm 102.2 101.7 102.5 102.6
pin-to-pin length in mm 3466 2892 2317 1737
compressive strength in MPa 132.1 116.1 131.3 131.3

Results
ultimate load Q* in kN 205 258 393 661
ultimate deflection w* in mm 24.3 21.1 21.0 10.0
ultimate moment M* in kN·m 6.89 8.64 11.7 10.1

axial load Q in kN 800 Q Q Q Q


indicates when the
vertical deflection of the 600 w*
columns were held
between load stages so 400
that the tooling ball w*
targets could be scanned 200 w*
w*
w w w w
deflection w at 0
mid-length in mm 0 20 40 60 0 20 40 60 0 20 40 60 0 20 40 60

load stages load stages load stages load stages


0 12 0 11 0 13 0 9

axial force N at 800


N N N N
mid-length in kN
600

400
typical elastic stress
range is shaded , 200
inelastic response
occurs before M* M M M M
0
M* M* M* M*
extreme fibre strains 2 ! ! ! !
at mid-length in 10-3 0
M M M M
strain on flexural
tension face of column –2 !!depth
strain on flexural com-
pression face of column –4

M M M M
0
curvature ! at mid-
length in rad/km 10
EI*=0.77·EI 0.82·EI
curvature based on 20 0.75·EI
laser scanner data 1
30 1 0.64·EI 1
EI is taken as the bending
stiffness of the uncracked 40
section " " " 1 "
50

moment M at 0 5 10 15 0 5 10 15 0 5 10 15 0 5 10 15
mid-length in kN·m apparent drop in stiff- secant stiffness at
ness after ultimate load ultimate load EI*

Figure 3-16. Load test results of four slender columns.


Top to bottom: Photos, as-built measurements, ultimate loads and deflections, load-deflection response,
axial force-moment response, extreme fibre strain-moment response, and curvature-moment response.

65
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

The moment-curvature diagram in Figure 3-16 for Specimen D (bottom-right plot) shows a bending mo-

ment of 1.8 kN·m and a nearly zero curvature at the first load stage. At this stage, the linear variable differ-

ential transformer measured a displacement of 1.03 mm, while the cohesive laser radar scanner measured a

displacement of 0.31 mm at the middle target of the column. The scanner measurement, which is used in

the calculation of curvature, may be the source of the uncharacteristically small curvature at the first load

stage. Another possible source of error is friction at the pins, which would cause some rotational fixity at

the ends of the column. If friction was present, the resultant line of action of load would be shifted toward

the centroidal axis of the column, resulting in much less curvature, as was observed during the test. Sub-

sequent load stages had much larger mid-length curvatures at higher loads, similar to those observed in

the other load tests. Measurements between the scanner and transformer agreed well after the first load

stage, suggesting that the displacement error did not continue past the first load stage.

The curvature-moment ϕ-M diagrams in Figure 3-16 are similar to conventional moment-curvature dia-

grams except that axial force N is not held constant in each diagram. Secant stiffnesses at ultimate limit

state EI* are shown on the diagram as secant slopes and are expressed in terms of the initial uncracked

flexural rigidity of the column EI0. The lowest recorded secant stiffness EI* was 0.64EI0, which gives furth-

er evidence that the column was well into its nonlinear stress range at ultimate limit state. Specimens A

and B exhibited a distinct drop in secant stiffness after their ultimate loads were reached. This was prob-

ably true for Specimens C and D, but no curvature data was recorded after their ultimate loads were

reached.

3.4.5 Validation of general method

The input parameters required by the general analysis method are column length, geometry of section, ec-

centricity of load, and stress-strain curve. Pin-to-pin column length and eccentricity of load at mid-length

were measured by the cohesive laser radar scanner from the zero-displacement scans described in Section

3.4.2. Average depth and average width of section were taken as the average measured value along the

length of column as shown in Figure C-8 on page 314. Compressive strengths of each specimen were based

66
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

on the average strength of cylinders that were cast from the same concrete batch, cured under the same

conditions, and tested at the same age as their respective specimens. The elastic modulus E for all speci-

mens was taken as 46.7 GPa, as determined by the average elastic modulus of the three prism tests de-

scribed in Section 2.1.4. The average elastic modulus of 48.1 GPa obtained from the 28-day cylinder tests

was not used because the prisms tested were more similar to the column tests in terms of age of concrete

and stress gradients. The linearity ratio between the secant modulus at peak stress Ec! and elastic modulus

Ec was taken as 1.3 based on the cylinder test results in Table 2-2. From these material properties, a stress-

strain curve based on the material model proposed in Chapter 2 was used as input for the general method.

Specimen | Model A A B B C C D D
mid-length eccentricity in mm 9.4 9.4 12.3 12.3 8.8 8.8 5.3 5.3
average depth in mm 101.9 101.9 101.4 101.4 102.4 102.4 101.3 101.3
average width in mm 103.2 103.2 103.0 103.0 101.7 101.7 102.9 102.9
pin-to-pin length in mm 3466 3466 2892 2892 2317 2317 1737 1737
compressive strength in MPa 132.1 132.1 116.1 116.1. 131.3 131.3 131.3 131.3
elastic modulus in GPa 46.7 46.7 46.7 46.7 46.7 46.7 46.7 46.7
percent
Results | Prediction difference
ultimate load Q* in kN 205 218 +6.1% 258 260 +0.8% 393 450 +14.5% 661 728 +10.1
ultimate deflection w* in mm 24.3 25.1 +3.3% 21.1 23.5 +11.4% 21.0 18.4 -12.4% 10.0 13.3 +33.0
ultimate moment M* in kN·m 6.89 7.51 +9.0% 8.64 9.31 +7.8% 11.7 12.2 + 4.3% 10.1 13.5 +33.7

Q axial load 800 Q Q Q Q


Q in kN
600 w*
e

w 400
† w*

200 w*
w*

Q deflection 0
w w w w
w in mm 0 20 40 60 0 20 40 60 0 20 40 60 0 20 40 60


Q axial force 800
N N N N
N in kN
600 M*

400
M M*
N
200 M*
M*

M M M M
moment 0
M in kN·m 0 5 10 15 0 5 10 15 0 5 10 15 0 5 10 15

† material model neglects post-peak ‡ model over-estimates the ultimate


experimental results ductility in tension, thus model predicts loads of Specimens C and D
prediction using e much less deflection capacity
prediction using e ± 2mm

Figure 3-17. Comparison of load test results and general method predictions

67
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Figure 3-17 shows a comparison between load test results and general method predictions. The table at the

top of the figure shows the measured values that were used as input for the general method. Percent differ-

ences between actual and predicted ultimate loads Q* of Specimens A, B, C and D were +6.1%, +0.8%,

+14.5%, and +10.1%, respectively. Hence, all predictions moderately overestimated the observed ultimate

loads of the column. The ratio between assumed error and measured initial eccentricity ranges between

16.2% and 37.7%. Thus, considering the ±2 mm error in measuring initial eccentricities of load, the general

method predictions were fairly good.

The load-deflection Q-w and axial force-moment N-M diagrams shown in Figure 3-17 include both ob-

served test results and general method predictions. The sensitivity of the ±1 mm error associated with the

manual positioning of the tooling balls is captured by two boundary prediction curves. These curves use

initial eccentricity of load values of 2 mm greater than and 2 mm less than the measured value. Test results

of all specimens, except for Specimen C, stay within the error bounds. Overall, the predictions and experi-

mental results agree reasonably well, attesting to the validity of the analytical model described by the gen-

eral method.

The measured deflections beyond the ultimate load of Specimens A, B, and C extend further than the

those predicted by the general method. This discrepancy may be accounted for by the tensile ductility of

the material beyond peak tensile stress. This post-peak tensile ductility is neglected in the stress-strain

models used for calculation. Neglecting this ductility is thus conservative in terms of estimating deflection

capacities of slender columns. Further, this omission does not appear to adversely affect the accuracy of

predicting the ultimate load of the tested columns.

Specimen D failed prematurely according to the response predicted by the general method. One possible

explanation is that there was a zone of weakness in the column where it failed, about 200 mm above the

middle of the column. Throughout the load test of Specimen D, the strain gauge at mid-length recorded

compressive strains smaller in magnitude than those above and below it. This means the section at mid-

length was not the critical section. The load-deflection response of Specimen D shows an abrupt change in

68
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

stiffness at 150 kN. This suggests that there may have been some non-uniformity in the column or some

friction at the pins, neither of which were accounted for in general method calculations.

3.5 Simplified design method

The general method presented in Section 3.2 requires a large amount of computation, and thus cannot be

done by hand in a reasonable amount of time. General method calculations are only practical with the use

of a computer, as done in the program qult written by the author. In order to reduce the analytical effort

required by the general method, approximate simplified methods for designing slender ultra high-per-

formance fibre-reinforced concrete members are proposed in this section.

The objective of the simplified design method is to determine whether a given slender column is capable of

sustaining the combined axial compression and bending caused by a given load. Given the following input

parameters: length of column, sectional geometry, and stress-strain curve, an approximate member capa-

city N-M* interaction diagram is produced. This interaction diagram serves as an envelope for all com-

binations of axial force and moment that can safely be carried by the critical section of the member.

The proposed simplified method is inspired by a simple and elegant design method proposed by Menn

(1990) for the design of slender reinforced concrete members. Menn uses a reduced N-M interaction dia-

gram based on a sectional analysis that is limited by the yield strain of steel. Designing slender members

using this reduced interaction diagram ensures that curvatures and second-order deflections remain small.

Menn’s simplified design method is particularly useful because it is simple, concise, and can be calculated

by hand. Menn’s method will be reviewed in Section 3.5.1 and then used as a model for the simplified

method proposed in Section 3.5.2.

3.5.1 Menn’s simplified design method

Menn’s design method (1990) is best described within the context of an example. The hollow, reinforced

concrete box section shown in Figure 3-18 will be used to illustrate Menn’s design method. A concrete

compressive strength of 30 MPa is assumed. The box is lightly reinforced with two layers of fifty-five 10M

69
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

bars in each of the top and bottom slabs, corresponding to top and bottom reinforcement ratios ρs! and ρs

of 0.15%. Various material and section properties are given in Figure 3-18.

Cross-section and properties Stress-Strain


dimensions in mm in MPa and 10-3
400 5200 400
fc

30
400

gross section 2.0 3.5 !c


properties
3200

concrete in compression
A = 7.36 m2
steel I = 17.8 m4 fs
reinforcement r = 1.55 m 400
As = 0.011 m2 E = 25100 MPa
!s = 0.15% EIo = 447000 MN·m2
400

Reinforced concrete section


2.0 !s
steel in tension

Figure 3-18. Conventional reinforced concrete box section

Sectional capacities of reinforced concrete sections are typically calculated based on the following modes

of failure: (1) the crushing of concrete in compression at a strain of -0.0035, and (2) the rupturing of rein-

forcing steel at a strain of about +0.010. Straining reinforced concrete members to these limits tend to sig-

nificantly reduce the stiffness of the system and cause large plastic deformations. Slender members subjec-

ted to axial compression are particularly sensitive to these effects and can reach their ultimate load at

strains much lower than these conventional strain limits. Thus, in many cases, it is unconservative to rely

upon the full moment capacity of the section, as calculated by these strain limits. As previously discussed

in this chapter, the member capacities of slender members are less than their sectional capacities.

To reduce the severity of second-order effects in slender members, Menn proposes that sectional capacity

be calculated based on the initial yielding of reinforcing steel. The yield strain of conventional reinforcing

steel is typically ±0.002, which is significantly less than the crushing strain of concrete or the rupture strain

of steel. The ranges of admissible planes of strain based on conventional material failure strain limits and

based on Menn’s yield strain limits are illustrated in Figure 3-19. The figure shows that curvature and axial

strain are significantly reduced when the lower yield strain limits are imposed.

70
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Conventional !cu !cu !cu 0 !su


material failure
strain limits
!cu = -0.0035 !su !su
!cu 0 !su
!su = +0.010

!sy !sy !sy 0 !sy


Menn's reduced
yield strain limits
!sy = ±0.002
!sy 0 !sy !sy !sy

Figure 3-19. Conventional strain limits and Menn’s reduced strain limits

Menn’s method is intended for solving specific design problems related to the use of slender members in

bridges, particularly in the design of piers and arches. The choice of yield strain limits is foremost a design

choice intended to limit deflections and second order-effects. The reduced strain limits effectively limit the

moment capacity of the section from that of the sectional moment of resistance Mu to the yield moment of

the section My. In most practical slender member designs, the ultimate moment M*, as defined and de-

scribed in Section 3.1, occurs somewhere between the two moments: My<M*<Mu. Because the calculation

of M* is complex, My is used as a conservative estimate of the ultimate moment capacity of the member.

The difference between My and Mu, which affects the degree of conservatism, varies with the shape of the

cross-section and the arrangement of longitudinal steel. The additional moment capacity gained by strain-

ing the section beyond the yield strain of steel is typically small. Once the steel has yielded, the moment

resistance of the section can be increased by increasing the lever arm between the resultant force of rein-

forcing steel in tension and the resultant force of concrete in compression. For hollow box sections, the

resultant compression force is typically confined within the thickness of the top slab. Since the thickness of

the top slab is very small relative to the overall depth of member, only small increases in lever arm and

moment capacity are possible. Thus, the differences between My, M*, and Mu are small, as is the degree of

conservatism. For solid sections and for sections with multiple layers of reinforcing steel spread over a sig-

nificant portion of the depth, the degree of conservatism is higher, since the difference between My and Mu

is larger. This is because there is (1) a greater capacity for the lever arm to increase after the reinforcing

71
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

steel yields, and (2) the inside layers of steel reinforcing have not yet reached their yield strength. This

means that in some cases, Menn’s design method may lead to overly conservative designs. However,

without a simple way to calculate the ultimate moment M*, there is no choice but to use My as an estimate

of the moment capacity of the member to ensure that the design is satisfactory. If designers are unsatisfied

with restricting the capacity of their slender member to My, then they must resort to more complex analyt-

ical methods, such as the one presented in Section 3.2.

Post-yield increases in moment resistance can also be attributed to increases in stress in the reinforcing

steel due to strain-hardening. Accounting for strain-hardening using manual calculation, however, is com-

plex. Hence, the over-strength provided by strain-hardening is normally neglected in design, and is instead

considered to be a form of built-in reserve strength (Menn 1990).

Axial force- Moment-curvature diagram


moment diagram in MN·m and rad/km N=-55.2 MN
in MN and MN·m

N M

150
!200
My Mu

Mcr
!150

!100
N-My 0
N-Mu !
!cr !y !u
!50 N=-55.2 MN
EI10 EIcr

0
0 50 100 150 M EIy
EIu
0
N-Mu interaction diagram based 0 0.5 1 1.5 !
on material failure strain limits
Secant stiffness-curvature diagram
N-My interaction diagram based on in EI0 and rad/km N=-55.2 MN
Menn's reduced yield strain limits

(a) (b)

Figure 3-20. Interaction and moment-curvature diagrams for a 30 MPa concrete box

Results from a sectional analysis of the example reinforced concrete box section are shown in Figure 3-20a.

Two interaction diagrams are shown, one based on conventional strain limits (dashed N-Mu curve) and

one based on yield-strain limits (solid N-My curve). For the given section, the difference between the two
72
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

interaction diagrams are minimal. This means that for any value of axial force N there is very little penalty

in using yield moments My as design moments of resistance rather than using ultimate moments Mu.

The behaviour of the reinforced concrete box section is further explored in Figure 3-20b, which shows the

moment-curvature diagram of the section for N=–55.2 MN. According to this diagram, the section has

three distinct domains of behaviour. The transitions between domains are marked by: the initial cracking

of concrete at the flexural tensile face of the section (Mcr), the yielding of the bottom layer of steel rein-

forcement in tension (My), and the rupture of steel or crushing of concrete (Mu). Beyond the yield moment

My, small increases in moment resistance require large increases in curvature.

The softening of the concrete box section is illustrated by the secant stiffness-curvature diagram shown in

Figure 3-20b. Secant stiffness EI! is calculated by dividing each value of moment M on the moment-

curvature diagram by its corresponding value of curvature ϕ. Between zero curvature and the curvature at

cracking ,cr, the section has a secant stiffness higher than the gross flexural rigidity EI0 of the section be-

cause of the stiffness contributed by the reinforcing steel. After cracking, the section loses stiffness, with

secant stiffness dropping down to about 33% of EI0, at which point reinforcing steel begins to yield. This

value is called the secant stiffness at yield EIy. Beyond the yield condition, secant stiffness continues to

drop as the reinforcing steel continues yielding until sectional failure is reached.

The secant stiffness at yield EIy is an important quantity because it represents a lower limit of secant stiff-

ness for members designed according to Menn’s reduced N-My interaction diagram. Provided that moment

demands at ultimate limit states are less than the yield moment, then it is always conservative to assume

EIy as the effective stiffness of the member being designed. Once calculated, this effective stiffness can be

used in the calculation of defections, buckling resistances, and moments caused by restrained deformation.

In lieu of detailed calculations, EIy can roughly be taken as one-quarter of the gross flexural rigidity of the

section EI0 (Menn 1990):

EI y ! 41 EI 0 Equation 3-24

73
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

The accuracy of this estimate of EIy varies with axial force N, as shown by the axial force-secant stiffness

diagram of the 30 MPa concrete section in Figure 3-21b. Three curves are shown corresponding to secant

stiffnesses at cracking, yielding, and ultimate (calculated assuming no strain-hardening). The EIy curve

shows that detailed calculations of EIy give values that are close to "EI0. For this section, it is unconservat-

ive to use Equation 3-24 at low values of axial force N.

Axial force- Axial force-


Moment interaction Secant stiffness
in MN and MN·m in MN and EI0

N N

!200

!150

!100 EIcr

Mcr
EIy
!50
My

EIu
0
0 50 100 150 M 0 !EI 10
EI EI!
0

(a) (b)

Figure 3-21. Secant stiffness at yield as a function of axial force

With means of calculating effective secant stiffness EIy and reduced N-My interaction diagrams, Menn’s

simplified design method can now be fully described. The method proceeds as follows:

Step 1. A reduced N-My design interaction diagram is calculated for the trial section based on planes

of strain that are limited to the yield strain of steel %sy at the outer layers of steel reinforcing (Figure 3-19).

Step 2. Axial force demand N is calculated based on the applied design loads. Based on this axial

force, yield moment My and yield curvature ,y is calculated using sectional analysis. From these values,

effective secant stiffness EIy=My/,y is calculated.

Step 3. Second-order deflections and moments are calculated using Vianello’s method using Equa-

tions 3-2 and 3-3 on page 34. In the calculation of deflections, moments, and buckling resistance, flexural

rigidity EI is replaced with effective secant stiffness EIy.

74
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Step 4. The adequacy of the slender member is determined by plotting axial force-moment demands

onto the reduced N-My design interaction diagram. If the points lie outside the interaction diagram, then a

new trial section is selected, and Steps 1 through 4 are repeated.

Menn’s simplified design method is validated experimentally by load tests of reinforced concrete beams in

compression and bending. Original work is reported in a university report by Gruber and Menn (1978).

3.5.2 Simplified method for the design of slender members

This section describes a simple method for designing slender members made from ultra high-performance

fibre-reinforced concrete. The method is inspired by Menn’s design approach, and involves using a reduced

interaction diagram for member capacity, and also an effective secant stiffness in the calculation of second-

order moments. The key difference that prevents the direct application of Menn’s method to ultra high-

performance fibre-reinforced concrete members without conventional steel reinforcing is that there is no

distinct yield point to use as a strain limit. An alternative means of defining a reduced member capacity in-

teraction diagram will be proposed.

As done for Menn’s method, the proposed simplified design method will be described within the context

of an example. The ultra high-performance fibre-reinforced concrete box considered is shown in Figure 3-

22. The outside dimensions of the box are the same as the conventionally reinforced concrete box analyzed

in the previous section. The strength of concrete assumed is 120 MPa, which is four times the strength of

the concrete in the other box. The webs and slabs of the higher strength box have specified thicknesses of

100 mm. The thicknesses are chosen such that the product of cross-sectional area and concrete strength

are roughly equal among the high-strength and conventional-strength boxes. This improves the validity of

comparisons made between the behaviour of the two sections.

75
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

Cross-section and properties Stress-Strain


dimensions in mm in MPa and 10-3

100 5800 100


fc
120

100
84

2.0 3.7 5.2 !c


gross section
3800

properties concrete in compression

A = 1.96 m2 ft
I = 5.48 m4
r = 1.67 m
E = 42000 MPa 14.6
EI0 = 230000 MN·m2
100

7
Ultra high-performance
fibre-reinforced concrete section 0.15 1.5 !t
concrete in tension

Figure 3-22. Ultra high-performance fibre-reinforced concrete box section

Moment-Curvature
in MN·m and rad/km N=-55.2 MN
Axial force- Axial force-
Moment interaction Secant stiffness M
in MN and MN·m in MN and EI0
150
Mu
N N
Mcr
post-peak response
!200 depends on the control
of localized cracking

EIu
!150
0
"cr "u "
!100 EIcr
EI10 EIcr
Mcr
Mu EIu
!50

0
0 0 0.5 1 1.5 "
0 50 100 150 M 0 10
EI EI!
!EI0
Secant stiffness-Curvature
(a) (b) in EI0 and rad/km N=-55.2 MN

(c)

Figure 3-23. Sectional analysis of an ultra high-performance fibre-reinforced concrete box

The ultra high-performance fibre-reinforced concrete box section considered has three domains of beha-

viour, as shown by the moment-curvature diagram in Figure 3-23c. The transitions between domains are

marked by: initial cracking of concrete at the flexural tensile face of the section (Mcr) and the localization

of cracks after strain-hardening (Mu). The moment at crack localization is assigned to be the ultimate mo-

76
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

ment Mu because (1) the post-peak tensile stress-strain behaviour of the material is not precisely known,

and (2) any increase in moment capacity after crack localization will be at best, minimal. A distinct yield

moment for this section cannot be characterized because the fibre reinforcement embedded in the materi-

al is smeared over a large area rather than concentrated in bar form.

The amount of post-peak ductility (often measured by the area under the moment-curvature diagram)

depends on the ability of the embedded steel fibres to bridge the the localized crack while being pulled out

from the concrete matrix. Volume and orientation of steel fibres are thus the most important factors that

affect post-peak ductility. The ductility of the ultra high-performance fibre-reinforced concrete mix used

in this thesis can be indirectly observed from the load-deflection results of the prism tests shown in Figure

2-8 on page 26. The post-peak curves of two out of the three prisms show gradual decreases in load with

increasing deflection. This type of ductility behaviour is only exhibited if the test is deflection-controlled.

The secant stiffness-curvature behaviour of the 120 MPa concrete section is shown below the moment-

curvature diagram in Figure 3-23c. The diagram shows that the section has relatively high stiffness at ulti-

mate limit state. For the given axial force N=–55.2 MN, the section maintains 68% of the gross flexural ri-

gidity of the section, which is much higher than the 30 MPa concrete section based on percentage of EI0.

Comparing the actual values of secant stiffness, the 120 MPa concrete section has EIu=156000 MN·m2 at

ultimate and the 30 MPa concrete section has EIy=148000 MN·m2 at yield. Thus, even though the 120 MPa

concrete section is about four times lighter and thinner than the 30 MPa concrete box section, both sec-

tions have effective secant stiffnesses at the design limit state that are more or less the same.

As shown in Figure 3-23b, secant stiffnesses of the 120 MPa box section vary with axial force N. For most

values of N below the balanced condition, secant stiffness at cracking EIcr is roughly equal to the gross flex-

ural rigidity of the section EI0. Thus, if the demand moment is less than the cracking moment M<Mcr, then

it is reasonable to assume that the design secant stiffness is equal to EI0. This design criterion is most ap-

propriate for serviceability limit state calculations, for which it is favourable to have the structure remain

crack-free. For most values of N, secant stiffness at ultimate EIu ranges between 50% and 75% of EI0. If the

77
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

demand moment is less than the ultimate moment M<Mu, then it is conservative to assume that the design

secant stiffness is EIu. In lieu of more detailed calculations, EIu can be estimated using Equation 3-25,

where N is axial force, and N! is maximum force sustained by the section in pure axial compression:

+ # N &
- 1
EI 0 % + 1( 0 ) N ) 12 N "
-
2
$ N" '
EI u ! , Equation 3-25
- # N&
1
EI 0 % 2 * ( 1
N" ) N ) N"
- 2
$ N"'
2
.

Estimates of EIu using Equation 3-25 were found to be in close agreement with detailed calculations of EIu

for many different ultra high-performance fibre-reinforced concrete sections that are symmetric about the

axis of bending. These estimates are considered to be sufficiently accurate for preliminary design.

The N-Mu interaction diagram shown in Figure 3-23a is not a reduced interaction diagram, and is thus un-

conservative for slender members whose ultimate loads Q* are reached before sectional failure (as exhib-

ited by the columns tested in Section 3.4). A strain-limited, reduced interaction diagram is not possible be-

cause a distinct yield point does not exist. Instead, a reduced interaction diagram is proposed as a

conservative approximation of the member capacity N-M* interaction diagrams produced by the general

method. The proposed reduced interaction diagram is shown in Figure 3-24 in terms of normalized axial

force n and normalized moment m.

78
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

design interaction diagram


n 1
admissible forces for
serviceability limit states
n admissible forces for
ultimate limit states

0.5
n —n

design contour of cracking


design interaction diagram

0
n 0.5 1 1.5
n
m
mn=

Figure 3-24. Reduced interaction diagram

Normalized axial force n and moment m are given by Equations 3-26 and 3-27, where A is cross sectional

area, fc! is compressive strength of concrete, d is depth of section, and I is moment of inertia of the un-

cracked section:

N 1
n= ! Equation 3-26
A fc"

Md 1
m= ! Equation 3-27
I fc"

The outside boundary of the lightly shaded area in Figure 3-24 represents the proposed interaction dia-

gram for serviceability limit states. The interaction diagram is defined by three coordinates: (m=0, n=0.5),

(m=1, n=0.5), and (m=0, n=–ncr).The normalized axial force at cracking ncr is equal to the cracking

strength of concrete fcr divided by the compressive strength of concrete fc!. The upper horizontal boundary

ensures that stresses in the slender member remain in the linear range of the compressive stress-strain

curve. The lower diagonal boundary approximates the contour of cracking, which is a set of sectional

forces that correspond to the initial cracking of the flexural tensile face of the section. These limits ensure

that the member remains in the linear-elastic range under service loads, and that secant stiffness is equal

79
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

to or slightly less than the gross flexural rigidity of the section. These limits also ensure that some nominal

separation is maintained between service and ultimate load conditions.

The outside boundary of the darkly shaded area in Figure 3-24 represents the proposed reduced interac-

tion diagram for ultimate limit states. The interaction diagram is defined by five coordinates: (m=0, n=nE),

(m=1, n=0.5), (m=1, n=0.5–(nt–ncr)), (m=mN=0, n=0), and (m=0, n=–nt). The normalized axial force at

Euler buckling nE is equal to the buckling resistance of the member divided by the maximum axial force of

the section in pure compression: QE/N!, or equal to 0.8, whichever is less. The normalized axial force at

peak tensile stress nt is equal to the tensile strength of concrete ft! divided by the compressive strength of

concrete fc!. The normalized moment resistance in pure flexure mN=0 (zero axial force) is calculated as the

ultimate moment of the section for N=0.

The proposed reduced n-m interaction diagram is validated in Section 6.3.4. This section discusses the sec-

tional force demands and capacities of the concept arches that were designed as part of the parametric

design study. In Figure 6-19 on page 242, the member capacity N-M* interaction diagrams for 72 concept

arch members are displayed and superimposed on to the proposed reduced interaction diagram. These 72

interaction diagrams are calculated based on trial arch sections that were found to be satisfactory under

the imposed bridge loads. Each of the 72 concepts were designed for a unique set of span length, span-to-

rise ratio, and degree of deck-stiffening. For the arch cross-sections considered, the proposed reduced n-m

interaction diagram was found to be conservative and accurate in comparison to the member capacity N-

M* interaction diagrams calculated using the general method.

Only hollow and solid-filled, rectangular ultra high-performance fibre-reinforced concrete sections were

considered in the parametric design study. The reduced interaction diagrams for these two types of section

differ in terms of their degree of conservatism. Reduced interaction diagrams for hollow sections tend to

follow the diagrams calculated using the general method reasonably well, resulting in a nominal degree of

conservatism. As the walls of the hollow section are made inwardly thicker, the degree of conservatism in-

80
3. Slender Ultra High-Performance Fibre-Reinforced Concrete Columns

creases. Thus, solid-filled sections tend to have actual ultimate moment M* capacities that are moderately

higher than those predicted by the simplified method.

With means of calculating effective secant stiffness EIu and reduced n-m interaction diagrams, the pro-

posed simplified design method for slender ultra high-performance fibre-reinforced concrete members

can now be fully described. For the design at ultimate limit states, the method proceeds as follows:

Step 1. A reduced N-M* design interaction diagram for the trial section is calculated based on the

normalized n-m interaction diagram given in Figure 3-24.

Step 2. Axial force demand N is calculated based on the applied design loads. Based on this axial

force, effective secant stiffness EIu is calculated using Equation 3-25.

Step 3. Second-order deflections and moments are calculated using Vianello’s method using Equa-

tions 3-2 and 3-3 on page 34. In the calculation of deflections, moments, and buckling resistance, flexural

rigidity EI is replaced with effective secant stiffness EIu.

Step 4. The adequacy of the slender member is determined by plotting axial force-moment demands

onto the reduced N-M* design interaction diagram. If the points lie outside the interaction diagram, then a

new trial section is selected, and Steps 1 through 4 are repeated.

81
Chapter 4. Statical Analysis of Arch Bridges

In this chapter, the statical analysis of concrete arch bridges is presented. The analytical methods described

in this chapter give estimates of maximum sectional forces in the structural system that are sufficiently ac-

curate for preliminary design. Further, this chapter examines how slenderness, shallowness, and distribu-

tions of flexural rigidity affect the structural behaviour of arches. Insights drawn from these studies along

with the proposed analytical methods are used to validate new design concepts for concrete arch bridges in

Chapter 6.

4.1 Relevant literature

Early literature on the statical analysis of arch bridges includes work by Ritter (1877) and Culmann and

Ritter (1906). The former is a paper that discusses stiffening trusses in arch and suspension bridges, while

the latter is a book on graphic statics applied to arches (Billington 2003).

Robert Maillart, a student of Ritter, was one of the early designers of reinforced concrete arch bridges. His

design methods are described by Maillart (1934), Billington (1979), and Laffranchi and Marti (1997). One

important contribution of Maillart’s is the deck-stiffened arch form, for which he assumed the deck girder

to carry all bending stresses caused by nonuniform loads.

Classical methods of arch analysis neglect the deformations of the structure when formulating equilibrium

equations. These methods are suitable for designing and analyzing heavy arches since deflections tend to

82
4. Statical Analysis of Arch Bridges

be small in comparison to the depth of the arch ribs. As arches become more slender, considerable addi-

tional bending moments can be caused by second-order deflections of the arch. Deflection theory analysis

of arches, which accounts for arch displacements in formulating equations of equilibrium, was studied by

Melan and Steinman (1913) and later by Asplund (1963). From these formulations, approximate methods

were developed based on amplifying moments and deflections calculated from classical methods, similar

to those developed for straight, slender members subjected to compression and bending.

Slender arches are also sensitive to stability problems. Chapters on arch stability edited by Johnston (1976)

and Galambos (1998) give comprehensive reviews of the topic in the context of metal structures based on

studies by Gjelsvik and Bodner (1962), Schreyer and Masur (1966), Austin (1971), Dickie and Broughton

(1971), and Kennedy and Aggarwal (1971). Many of these studies on the stability of arches are based on

elastic analysis and thus can be applied to concrete structures with some adjustment for material

properties.

Of particular interest to the objectives of this thesis are studies on the buckling of shallow arches. In shal-

low arches, axial deformations under uniform load cause significant crown deflections. These deforma-

tions cause shear and bending stresses in the arch similar to those in beam-type structures. Thus, perman-

ent loads, which are normally carried in pure compression along the axis of the arch, are partially carried

by shallow arches in bending. Studies on the stability of shallow concrete arches have been done by Wang

et al. (2006) and Cai et al. (2009). The former work includes shrinkage and creep deformations in formu-

lating stability equations to describe the long-term buckling of shallow arches with hinged supports, while

the latter work deals with shallow arches with fixed supports.

Recent publications that review the analysis of concrete arches include works by Menn (1990), Mondorf

(2006), and Benaim (2007). On-going research in the area of arch bridges, including the analysis of arches,

is presented at the International Conference on Arch Bridges, which has been held every three years since

1995.

83
4. Statical Analysis of Arch Bridges

The arch analysis methods proposed in this chapter are based on the works briefly described above. Al-

though the proposed analysis methods do not introduce anything fundamentally different from current

literature, they do attempt to synthesize all the simplified analysis techniques from literature into one con-

sistent method for calculating second-order sectional forces in arch systems. The study of the transition of

shallow arches between pure arch behaviour and pure beam behaviour has not been previously solved or

explained using the force method to the extent presented by the author in Section 4.4.3. This chapter also

provides many original figures and diagrams that are designed to help readers improve their understand-

ing of arch behaviour at an intuitive level.

4.2 Definition of structural system

The main structural components and geometrical quantities of typical concrete arch bridges are illustrated

in Figure 4-1. The structural system consists of a deck girder, an arch, and spandrel columns. The supports

at each end of the arch are referred to as springing lines. The midspan section of the arch is referred to as

the arch crown. The arch span is the horizontal distance between springing lines, and the arch rise is the

vertical distance between the crown and a line intersecting the springing lines. The spans that the deck

girder traverses between spandrel columns are called interior spans.

interior span deck


approach spans
crown arch

rise spandrel column

springing line springing line


span

Figure 4-1. Components of a typical arch bridge

Most concrete arches built to date are arranged with the arch supporting the bridge deck from below as il-

lustrated in Figure 4-1. Arches that suspend the deck from above usually require the arch to be split into

two arch ribs that rise from either side of the bridge deck. These ribs are often braced together above the

deck to provide lateral stability. Tension members are used instead of spandrel columns to transmit forces

from the deck to the arch. Arches in this arrangement are usually designed in structural steel or composite

84
4. Statical Analysis of Arch Bridges

concrete-filled steel tubes as in the Lupu Bridge in China (Lin et al. 2004). One exception to this is the re-

cently completed Third Millennium Bridge in Spain (Kremer 2008), in which the deck is supported from

above by a concrete arch. Although advanced construction techniques and high-strength concrete may al-

low for concrete arches to be built economically above the deck, the scope of this thesis is limited to the

study of arches that support the deck from below.

4.3 Comparison of three-hinged, two-hinged, and fixed arches

Three types of arch systems are used in practice: three-hinged arches, two-hinged arches, and fixed arches.

As their classification names imply, the types of arches are distinguished by the number of hinges in the

system. Fixed arches have no hinges, two-hinged arches have hinges at each springing line, and three-

hinged arches have an additional hinge at the crown, as shown in Figure 4-2.

three-hinged arch two-hinged arch fixed arch

Figure 4-2. Location of hinges in typical arches

The relative differences between the three types of arch systems are described in Table 4-1. In broad terms,

these systems differ in terms of degree of indeterminacy, cost of construction and maintenance, flexibility

and stability, and bending moment demand. The degree of indeterminacy is highest in fixed arches, with

three redundant forces. While fixed arches offer the highest level of redundancy, which is favourable in

terms of structural reliability, fixed arches also require the most computational effort when calculating sec-

tional forces. Hinges in concrete structures require details that must be able to safely transmit high shear

and axial stresses while providing a rotational degree of freedom. Fabricating and maintaining mechanical

steel hinges can be costly due to their complexity and high exposure to the environment. Concrete hinges,

which are typically made by reducing the depth of section locally and providing sufficient steel reinforce-

ment, are less complex than mechanical hinges and can be constructed relatively cheaply.

85
4. Statical Analysis of Arch Bridges

Table 4-1. Relative comparison between fixed, two-hinged, and three-hinged arches
fixed two-hinged three-hinged
relative comparison arches arches arches
degree of indeterminacy 3 1 0
cost of construction and maintenance lower medium higher
magnitude of deflections lower medium higher
buckling strength higher medium lower
arch axis geometry same same same
moments caused by nonuniform loads lower medium higher
moments caused by restrained deformations higher lower none

Flexibility of the system is also determined by the fixity of the arch, with the highest flexibility associated

with three-hinged arches. Hence, deflections in hinged arches tend to be larger than those in fixed arches,

which make them more sensitive to long-term deflections. Higher flexibility also leads to higher susceptib-

ility to stability problems. Fixed arches have the highest resistance against flexural buckling, while three-

hinged arches have the lowest. Methods for calculating the buckling load of arches are discussed in the

next section.

4.3.1 Flexural buckling of arches

Primary flexural buckling modes for each type of arch system are shown in Figure 4-3. These systems are

subjected to vertical, uniformly distributed loads, which cause pure axial compression along the axis of the

arch. As the applied loads are increased, the arches will eventually reach their critical load HE and buckle

according to the deflected shapes shown in the figure. Three-hinged arches may buckle in a symmetric,

snap-through mode or in an antisymmetric, bifurcation mode. Two-hinged and fixed arches buckle only

in antisymmetric, bifurcation modes. Analogous straight column buckling models can be constructed us-

ing the same member length and support conditions as the arch models they describe. Additional lateral

supports at mid-length are needed in the analogous column models to account for the two-wave primary

buckling modes that arches tend to exhibit. In these column models, the axes of the arches are straightened

out, and horizontal reactions of the arch are applied as concentric axial loads on the columns. Using this

86
4. Statical Analysis of Arch Bridges

analogy, the effective arc length factor k of each arch buckling model can be determined using convention-

al stability analysis of straight compression members. Values of k are given below (Austin 1971):

! 0.54 for three-hinged arches


#
k = " 0.50 for two-hinged arches Equation 4-1
# 0.35 for fixed arches
$

arch model
arch flexural EI
HE  / 2 S f
buckling load  kS 2
L
L2 £ 4 f
L
16 f ¥
2 2
1 2
arc length S L
16 f 2
ln ² ´ analogous column model
2 8f ¤ L ¦
S

three-hinged arch three-hinged arch


symmetric buckling antisymmetric buckling

initial shape
deflected shape

HE HE HE HE

k  0.54 k  0.50

two-hinged arch fixed arch


antisymmetric buckling antisymmetric buckling

HE HE HE HE

k  0.50 k  0.35

Figure 4-3. Arch and analogous column buckling models. Partially adapted from Austin (1971).

The buckling resistance of arches HE is given by the following equation, where EI is flexural rigidity of the

arch and S is arc length (Austin 1971):

EI
HE = ! 2 Equation 4-2
( kS )2

87
4. Statical Analysis of Arch Bridges

Arc lengths of parabolic arches can be calculated as a function of span L and rise f using Equation 4-3. This

expression is based on the segment of a parabola, as given by Spiegel and Liu (1999).

L2 ! 4 f + L2 + 16 f 2 $
S= 1
2 L2 + 16 f 2 + ln # & Equation 4-3
8f " L %

The effective arc length factors given in Equation 4-1 can be verified by comparing their values with other

stability analyses of arches. Rambøll (1944) used a virtual work calculation method to calculate buckling

loads of three-hinged, two-hinged, and fixed arches. Results from Rambøll’s analysis are shown in Figure

4-4a as a plot of buckling coefficient ) versus rise-to-span ratio f/L. By reworking Rambøll’s buckling coe-

fficient results in terms of effective arc length factors k, the curves shown in Figure 4-4b are obtained. In

comparison to Rambøll’s analysis, the k values given in Equation 4-1 are accurate at low rise-to-span ratios,

but are slightly unconservative at rise-to-span ratios higher than 0.1. Other arch stability studies have been

done analytically by Dinnik (1955) and Stussi (1935), and experimentally by Gaber (1934) and Kolibrunner

(1936), all of which are in good agreement with the analogous column method (Austin 1971). All these

studies, however, neglect the effects of axial shortening, which can reduce buckling resistance considerably

in shallow, slender arches, as will be discussed in Section 4.4.4.

HE = ! EI2 HE = "2 EI 2
L (kS)

! 90 k 1

0.8
3-hinged arch
60 2-hinged arch
0.6
fixed arch k=0.54 effective arc length
k=0.50
factors given by
0.4
k=0.35 analgous column model
30
2-hinged arch
3-hinged arch 0.2 fixed arch

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
f f
(a) L
(b) L

Buckling coefficient curves calculated by In this plot, Rambøll's data is


Rambøll (1944) for different arch types shown in terms of effective
using virtual work calculations are shown arch length factor k
above (see Mondorf (2006) for English
description).

Figure 4-4. Comparison of effective arc length factors with stability analysis results obtained by Rambøll (1944)

88
4. Statical Analysis of Arch Bridges

4.3.2 Setting the shape of the arch

All arch systems should be designed to carry dead loads in pure compression along their arch axis. In or-

der to achieve this, the shape of the arch should coincide as closely as possible with the pressure line of the

sustained load. For a given set of loads, the shape of the pressure line is the same for all arches, whether

they are hinged or fixed. The shape of the pressure line is the same as the bending moment diagram for a

simply supported beam subjected to the same loading (see Figure 4-5a,b). In order to determine a unique

pressure line, the rise of the arch f must be specified. Once f is specified, the ordinates of the pressure line

y(x) can be calculated using Equation 4-4, where MS(x) is the bending moment diagram in a simply sup-

ported beam caused by the given set of loads, and MS(L/2) is the moment of the midspan section of the

simply supported beam (Favre and de Castro San Román 2001):

M S (x)
y(x) = f Equation 4-4
M S ( 12 L)

Since most of the sustained loads are applied to the arch through spandrel columns, the loads can be rep-

resented as a set of concentrated loads, as shown in Figure 4-5c,d. Representing the loads in this way res-

ults in a pressure line that is polygonal, rather than smoothly curved. Using a polygonal shape for the arch

simplifies the arch geometry, making it easier to build than a smoothly curved arch. In this form, devia-

tions of the arch from the pressure line are greatest between spandrel columns, since the self-weight of the

arch will always be distributed rather than concentrated at the angle breaks of the arch.

As an alternative to Equation 4-4, a graphical method of calculation can also be used to determine the

pressure line for a set of concentrated loads, as illustrated by Figure 4-5d,e. In this method, forces are

drawn as force vectors, with vector length representing magnitude of force, and vector direction represent-

ing direction of force. Equilibrium about Joint 1 in Figure 4-5d is expressed in Figure 4-5e by three force

vectors arranged in a triangle. Vector ST represents the load from the spandrel column and Vectors TO

and OS represent reacting forces along the pressure line. Assembling triangulated forces from each joint in

the pressure line results in global equilibrium Triangle VOS, where Vectors VO and OS are the right and

89
4. Statical Analysis of Arch Bridges

left support reactions, respectively, and Vector SV is the sum of applied loads. The horizontal length H of

Triangle VOS represents the horizontal support reactions and the constant horizontal component of force

along the pressure line. Alternate states of global equilibrium can be made by moving Point O horizontally.

Moving this point to the left increases the horizontal reaction H and reduces the rise f. Thus, Point O can

be adjusted to obtain the desired rise in the arch.

(a) bending moment for a simply supported beam (b) parabolic pressure line for
caused by a uniformly distributed load uniformly distributed load

y
f
x
H x H

4f
M S ( L2 )  H u f y(x)  (x)(x < L)
L2

(c) bending moment for a simply supported beam (d) polygonal pressure line for
caused by a set of concentrated loads a set of concentrated loads

y
f
1
x H x H
M (x)
y(x)  S L f
MS ( 2 )
M S ( L2 )  H u f

(e) graphical representation of forces

H
S
tion
e ac T
tr
1 lef
S O
rig
T ht
re
ac
tio
n
O V
equilibrium of equilibrium of
forces at joint 1 forces at all joints

Figure 4-5. Pressure lines for uniformly distributed and concentrated loads

The true distribution of self-weight, which is needed to determine the pressure line, is not known until the

final geometry of the arch is defined. Thus, the process of finding the pressure line and setting the arch on

it is iterative. For preliminary design, it is sufficient to approximate the sustained load as a uniformly dis-

tributed load, as in Figure 4-5b. With this assumption, the pressure line is described by the parabola y(x)

given by Equation 4-5, where f is rise, and L is span:

90
4. Statical Analysis of Arch Bridges

4f
y(x) = (x)(L ! x) Equation 4-5
L2

The horizontal reaction H caused by a uniformly distributed load q is:

qL2
H= Equation 4-6
8f

Axial forces along the arch N(x) are given by Equation 4-7, where '(x) is angle of inclination of the arch:

H
N (x) = Equation 4-7
cos ! (x)

The axial force N is equal to the horizontal reaction H at the crown, and is greatest at the springing lines, as

given by:

2
! f$
N (x = 0, L) = H 1 + 16 # & Equation 4-8
" L%

Equations 4-5 through 4-8 were calculated based on the static equilibrium of a three-hinged parabolic

arch. They also apply to statically indeterminate arches because the additional rotational restraints are ir-

relevant to this uniform load condition (only axial compression stresses are caused). This can be verified

using the force method, as described later in Section 4.4.2.

4.3.3 Response of arches to nonuniform, vertical loads

Nonuniform loads such as partial live loads cause arches to deflect off the pressure line. At each arch sec-

tion, bending moments are generated equal to the axial force times the eccentricity between the centroid

of the arch and the displaced line of action of force. Thus, nonuniform loads are carried by arches primar-

ily in bending.

Figure 4-6 shows the bending moment diagrams M(x) and bending moment envelopes in fixed, two-

hinged, and three-hinged arches caused by a single, vertical concentrated load Q. Each row of diagrams

represents one position of Q. The last row shows the maximum positive and negative moments caused by
91
4. Statical Analysis of Arch Bridges

applying Q at any point along the arch. Bending moments were obtained from the structural analysis pack-

age sap2000 using simple, elastic arch models. Only first-order analysis was considered, implying that

equilibrium was calculated based on the original undeformed geometry of the structure. All structural ele-

ments were made weightless and inextensible. In order to prevent frame action in the system, the arch was

isolated by modelling deck elements as simply supported between adjacent spandrel columns.

The data presented in Figure 4-6 show that the highest bending moments are carried near the quarter-

points of arches. One exception to this are the large negative moments carried by fixed arches near their

springing lines. Fixed arches are often deepened toward the springing lines to accommodate for these large

moments. Comparing moments within the middle three-quarters of the span, three-hinged arches carry

62% more positive bending moments than fixed arches, and 115% more in negative bending. These premi-

ums in bending moment are also reflected when partial distributed loads are considered, as shown in Fig-

ure 4-7. For partial distributed loads, three-hinged arches carry twice as much as the bending moments

carried in fixed arches near the quarter-points. Again, this reduction in moments in fixed arches comes

with the penalty of large moment demands near the supports.

92
4. Statical Analysis of Arch Bridges

Arch system bending moment diagrams M(x) caused by concentrated load Q


Each diagram depicts M(x) when Q is applied at x=a

L L L
a a a
Q Q Q

f f f
M/QL in 10-3 fixed 2-hinged 3-hinged
a/L
-39.4
-2.0 -11.2 -6.2
0.05 3.32
8.6
41.6 42.7

-60.6
-21.6 -12.5
-6.8
0.10 11.3
25.6
67.9 72.0
-67.7
-30.6 -18.7
-12.9
0.15
21.5
42.0
81.6 89.2
-64.0 -37.6
-19.2 -25.0
0.20
31.9
53.6
85.6 96.0
-52.9 -42.1 -31.2
-24.5
0.25
40.8
59.1
82.9 93.7
-44.4 -37.5
-37.0 -28.0
0.30
46.9
59.3
76.3 84.0
-43.7 -43.7
-29.1
-18.9
0.35
49.3
56.0 68.3 68.2

-39.8 -50.0
-28.2 -12.0
-13.4 -7.1
0.40 -0.46
47.5 48.0
51.5 61.1
-56.2
-33.8 -34.9
-15.4 -24.6 -16.1
0.45
16.5
41.2 24.7
47.9 56.0
-62.5 -62.5
-19.9 -19.9 -25.3 -25.3
0.50
30.7 30.7
46.6 54.3

-67.7 -62.5
-29.1 -44.4
-14.3 -12.9 -16.2
Envelope
49.3 25.5
46.6 54.3
59.3
85.6 96.0

Figure 4-6. Bending moments caused by concentrated loads in fixed, two-hinged, and three-hinged arches

93
4. Statical Analysis of Arch Bridges

Arch system bending moment diagrams M(x)


caused by uniformly distributed load q
M/qL2 in 10-3

application of q fixed arch 2-hinged arch 3-hinged arch

-17.0 q
-12.4
-6.7 -10.0
q over
four-tenths
of span
9.1 11.4
16.3 18.7
q -15.6 -15.6
-15.5
-8.8
q over
half of span
8.8
15.5 15.6 15.6

q
-11.1 -11.1
-4.9 -4.9 -6.8 -6.8
q over middle
third of span
8.0 5.2 8.0 7.2

-16.4 -18.7
-17.0
-9.3 -7.2
-4.1 -5.4
q envelope
4.1 5.4
9.3 7.2
17.0
16.4 18.7

Figure 4-7. Bending moments caused by distributed loads in fixed, two-hinged, and three-hinged arches

The three partial distributed load arrangements shown in Figure 4-7 are typically used to estimate maxim-

um bending moments in arches. The arrangements of partial distributed loads that cause maximum bend-

ing moments at given sections along the arch can be found using influence lines &, as shown in Figure 4-8.

The positive and negative areas under the influence lines are tabulated in the first column to the right of

each diagram. The second column tabulates the locations where influence lines are zero. Based on these

values, applying distributed loads from the support x=0 to x=0.36L, or x=0 to x=0.46L result in maxim-

um moments in the sections near the quarter-points. Thus, using distributed loads over four-tenths of the

span provides reasonable estimates of maximum first-order moments. Relative to the four-tenths span dis-

tributed load case, half span loading has slightly lesser maximum moments, but has greater moments in

the unloaded half of the arch. Also, the moments and deflections caused by half span loading are symmet-

rical, which better reflect the primary buckling modes illustrated in Figure 4-3. Hence, half span loading is

the most suitable load case when second-order effects and stability are being considered in arches.

94
4. Statical Analysis of Arch Bridges

Arch system influence lines for bending moment !


Each diagram depicts ! of section at x

L L L
x x x

f f f
!=0 at
!/L in 10-3 fixed 2-hinged 3-hinged
± area/L2
x/L
in 10-3
-67.7

0.00 ±17.0 0.04L ±0 ±0

49.3

-37.5
-15.7 -22.5
0.07L
0.05 ±8.90 ±6.68 0.34L ±7.37 0.35L
0.44L
26.6
41.6 42.7

-28.0 -40.0
-18.1
0.10 ±4.19 0.17L ±11.6 0.36L ±12.9 0.36L
0.54L
25.6
67.9 72.0
-52.5
-36.7
-14.3
0.15 ±4.66 0.28L ±14.7 0.38L ±16.5 0.37L

42.0
81.6 89.2
-60.0
-41.9
-20.6
0.20 ±7.41 0.36L ±16.3 0.40L ±18.5 0.39L

53.6
85.6 96.0
-62.5
-44.4
-26.4
0.25 ±9.00 0.40L ±16.4 0.43L ±18.8 0.40L

59.1
82.9 93.7
-60.0
-43.7
-29.1
0.30 ±9.34 0.44L ±15.3 0.46L ±17.5 0.42L

59.3
76.3 84.0
-52.5
-28.5 -39.8
0.35 ±8.57 0.48L ±13.1 0.49L ±14.8 0.44L

56.0 68.2
68.3

-33.9 -40.0
-25.4
0.16L
0.40 ±7.08 ±13.2 0.53L ±10.9 0.46L
0.52L
51.5 61.1 48.0

-19.5 -25.6 -22.5


0.28L 0.25L
0.45 ±5.84 ±7.94 ±5.89 0.48L
0.57L 0.59L
24.7
47.9 56.0

-12.9 0.37L -16.2 0.35L


0.50 ±5.41 ±7.24 ±0
0.63L 0.65L
46.6 54.3

Figure 4-8. Bending moment influence lines for fixed, two-hinged, and three-hinged arches

95
4. Statical Analysis of Arch Bridges

fixed arch 2-hinged 3-hinged

Shape of bending
moment envelope for
concentrated loads

Shape of bending
moment envelope for
distributed loads

note: bridges
not drawn to
same scale Sandö Bridge (1943) Caracas (1953) Salginatobel (1930)

Figure 4-9. Comparison of bending moment envelopes and variations in depth along the arch

Many designers shape their arches to reflect the bending moment demands of the chosen arch type, as

shown in Figure 4-9. For example, the three-hinged arch in Maillart’s Salginatobel Bridge is deepest at the

quarter-points and thinnest near the springing lines and arch crown (Billington 2003). The deepening of

the arch increases the section modulus of the arch ribs where it is needed most, improving the structural

efficiency of the structure. A typical strategy for fixed arch bridges is to increase the depth and width of the

arch gradually toward the springing lines, as done in the Sandö Bridge. Most designers use symmetric arch

sections since positive and negative bending moment demands are usually of similar magnitude.

4.3.4 Effects of restrained deformation on arches

The effects of restrained deformation increase with the degree of statical indeterminacy of any system. Axi-

al deformations in arches, such as caused by shrinkage, creep of sustained loads, elastic axial stresses, hori-

zontal yielding of supports, and uniform changes in temperature induce bending moments in statically in-

determinate arches. These redundant moments MK are proportional to the imposed axial strain %K and

flexural rigidity EI of the arch and are inversely proportional to rise f. Using the force method (as done in

Section 4.4.2), it can be shown that the redundant moments MK at the ends of fixed, parabolic arches are

given by:

EI
M K = ! 152 " K Equation 4-9
f

96
4. Statical Analysis of Arch Bridges

Expressions for various imposed axial strains %K are given in Table 4-2, where H is horizontal reaction, EA

is average axial rigidity of the arch, ,(t) is creep coefficient, %sh(t) is concrete shrinkage strain, #T is change

in temperature, #L is change in span length, and L is span.

Table 4-2. Expressions for axial strains and flexural rigidities in arches
effective
action axial strain %K flexural rigidity
elastic shortening H/EA EI
creep of sustained load ,(t)·H/EA Eadj·I
concrete shrinkage %sh(t) Eadj·I
uniform change in temperature -T/L EI
horizontal movement of supports -L/L EI

Because of the effects of creep, the effective flexural rigidity of the arch can be reduced if the imposed axial

strains are applied gradually over a long period of time. Trost (1967) and Bazant (1972) developed an ap-

proximate method that uses an age-adjusted, effective modulus to calculate long-term stresses caused by

long-term strains. The age-adjusted, effective modulus Eadj is given by Equation 4-10, where Ei is elastic

modulus at the time of initial stress, ,(t) is creep coefficient, and +(t) is aging coefficient. Typical values for

,(t) and +(t), which are functions of time and age at initial loading, are given in Bazant (1972).

Ei
Eadj (t) = Equation 4-10
1 + !" (t)

Figure 4-10 shows bending moment diagrams caused by restrained deformations in different arch types.

Fixed arches have strain-induced moments that are up to four times as large as moments in two-hinged

arches. Three-hinged arches are determinate and so do not develop these strain-induced bending stresses.

fixed arch 2-hinged arch 3-hinged arch


-MK

0
!·MK
!·MK

Figure 4-10. Bending moments caused by restrained deformations in fixed, two-hinged, and three-hinged arches

97
4. Statical Analysis of Arch Bridges

Since the effects of restrained deformations are proportional to the flexural rigidity of the system, the

effects tend to decrease as critical sections of the arch become nonlinear. As redundant moments in fixed

arches are increased, the arch system should evolve into a two-hinged arch and then into a three-hinged

arch, provided that there is enough plastic rotation capacity to form plastic hinges at the critical sections of

the arch. As the system softens, moments caused by restrained deformation will be limited, and load-in-

duced moments will redistribute themselves toward the quarter-points. Thus, statically indeterminate

arches that are sensitive to restrained deformations will probably not undergo sectional failure due to large

redundant moments alone. These arches, however, can fail due to the formation of plastic hinges, which

will redistribute and increase load-induced moments and increase deflections, and reduce the resistance of

the system against buckling. In other words, the softened system will be more prone to buckling as plastic

hinges are formed.

4.3.5 Other load cases

Other loads, such as wind loading and earthquake loading, may cause sectional forces in the arch that ex-

ceed those caused by vertical loads (dead loads and live loads) and restrained deformations. These other

loads, along with unbalanced multilane lane loading, also affect the out-of-plane behaviour of the system.

Because this thesis deals primarily with the longitudinal behaviour and proportioning of arch systems,

transverse loading will not be considered. It is therefore assumed that these other loads and their load

combinations do not increase the structural demands imposed on the concept arches proposed in Chapter

6. In places where wind and earthquake loads are severe, appropriate modifications should be made.

4.4 Structural response of fixed arch systems

Section 4.3 described the primary differences between three-hinged, two-hinged, and fixed arches. Only

fixed arches, however, will be considered in the development of new concepts for arches in Chapter 6. The

primary reasons why only fixed arches will be considered are: (1) fixed arches have higher resistances

against buckling than hinged arches, which allows for lighter and more slender members to be used (2)

fixed arches have higher degrees of redundancy than hinged arches, which is favourable in terms of struc-
98
4. Statical Analysis of Arch Bridges

tural reliability, and (3) fixed arches are more commonly built than hinged arches (as observed later from

the database of concrete arches, Table 5-3 on page 161), which suggests that they may have some intrinsic

economic advantage over hinged arches. The following section will expand on the structural analysis of

fixed arch systems, and discuss the effects of deck-stiffening, shallowness, slenderness, system softening,

long-term behaviour, and method of construction.

4.4.1 Simplified statical system

When acting alone, fixed arches are statically indeterminate to the third degree. When spandrel columns

and deck girders are integrated with the arch, the degree of indeterminacy of the system increases with the

number of fixed joints between the structural members. Analyzing such highly indeterminate systems pre-

cisely is only practically feasible through the use of computer frame analysis packages. Such packages are

generally ill-suited for preliminary design calculations of arches because: (1) they tend to inhibit creativity

by separating design from insights of structural behaviour drawn from simple manual calculation, (2) the

inter-relationships between redundant moments, distributions of stiffness in the system, and system

softening are not easily observed, and (3) the geometry of an arch is sensitive to the pressure line of the

load, and thus must be updated with each design change to minimize dead load moments. Thus, approx-

imate analysis methods, which allow for the identification of important design parameters and critical load

cases are best suited for preliminary design.

The degree of indeterminacy of fixed arched systems and the complexity of computation can be signific-

antly reduced by using the following simplifying assumptions:

Assumption 1: Spandrel columns are pin-connected to the arch and deck. This assumption reduces the

indeterminacy of the system by removing the redundant moments associated with the ends of each span-

drel column. In some cases, spandrel columns are physically pin-connected to the deck using special bear-

ings. In other cases cases, spandrel columns are monolithically built into the arch and deck. For these

cases, Assumption 1 is still reasonable because the bending stiffness of the column is typically much less

than the connected deck or arch members. According to Hardy Cross’ moment distribution method

99
4. Statical Analysis of Arch Bridges

(Cross 1932), unbalanced fixed-end moments at joints are successively distributed to each connected ele-

ment in proportion to the ratio of its bending stiffness to the sum of bending stiffnesses of all elements

connected to the joint. As an example, moment distribution factors have been calculated for several

column joints in the Colorado River Bridge and are shown in Figure 4-11. Distribution factors were calcu-

lated using equations for non-prismatic members, as derived by Gere (1963). The highest spandrel column

distribution factor is 0.14, which is small. The other spandrel column joints have even smaller distribution

factors at their bases. The distribution factors equal to zero indicate that the joint connections between

column and deck are effectively pinned.

fixed connection pin connection expansion bearing


0.43 0.43 0.5 0.5 0.5 0.5

0.14 0 0
fixed
connection
0.06

0.48
0.46
fixed
connection
0.03
1
0.5
6
0.4

fixed
connection
0.03
51
0. moment distribution factors
46
0.

Figure 4-11. Moment distribution factors for the Colorado River Bridge.
Drawing adapted from us dot fhwa (2003).

Assumption 2: The flexural rigidity of the arch projected on to the horizontal axis is constant over the

entire span and is equal to the flexural rigidity of the arch at the crown. This assumption is illustrated in Fig-

100
4. Statical Analysis of Arch Bridges

ure 4-12 and is expressed algebraically in Equation 4-11 (Menn 1990), where ' is the angle of inclination of

the arch, and EIarch and EIcrown are flexural rigidities of the arch and crown section of arch, respectively.

EI arch (s) ! cos " (s) = EI crown Equation 4-11

y tanƧ
s 1

x
EI arch (s)
flexural rigidity perpendicular to curved axis
EI arch (s)u cosƧ(s)  EI crown
horizontal component of flexural rigidity EI arch (x)  EI crown

Figure 4-12. Arch flexural rigidity

Assumption 2 simplifies deflection calculations by making arches act as straight beams with constant flex-

ural rigidity. By changing the variable of integration (Equation 4-12), deflections $ calculated by the meth-

od of virtual work can be integrated with respect to the horizontal axis x (Equation 4-14), rather than the

curved axis of the arch s (Equation 4-13). In these equations, S is arc length, L is span length, m is virtual

moment, , is curvature, and M is bending moment.

ds cos ! (s) = dx Equation 4-12

S S M (s)
! = $ m(s) " # (s)ds = $ m(s) " ds Equation 4-13
0 0 EI arch (s)

L M (x)
! = # m(x) " dx Equation 4-14
0 EI crown

To ensure the validity and accuracy of Assumption 2, designers may choose to shape the arch such that

Equation 4-11 is true. Stiffening the arch towards its ends, as required by the assumption, may also come

naturally from the higher sectional force demands at the ends of the arch. If the distribution of flexural ri-

gidity departs significantly from Equation 4-11, then some further redistribution of moments among the

arch and deck will needed, especially near the springing lines.

101
4. Statical Analysis of Arch Bridges

Assumption 3: Vertical deflections of the deck and arch are equal along the span. This compatibility as-

sumption is valid at the locations where the spandrel columns connect with the deck and girder. The ac-

curacy of solutions based on this assumption increases as the number of spandrel columns are increased.

Assuming equal deck and arch deflections, $deck(x)=$arch(x), leads to simple expressions for the distribution

of system moments among arch and deck, as derived by the equations below.

Deflections of the deck are given by Equation 4-15 and Equation 4-16, where c is an integration constant

that result from solving virtual work integrals for given curvature diagrams ,(x), Mdeck is deck moment,

and March is arch moment (Billington 1973).

M deck (x) 2
! deck (x) = c(x) " L Equation 4-15
EI deck

M arch (x) 2
! arch (x) = c(x) " L Equation 4-16
EI crown

L
c = # m(x) ! " (x)dx ÷ " (x) ! L2 Equation 4-17
0

Equating arch and deck deflections $deck=$arch results in Equation 4-18 (integration constant c and span L

are the same for both arch and deck). This equation can be rearranged to obtain an expression for deck

moment, as given by Equation 4-19 (Billington 1973).

M arch (x) M deck (x)


= Equation 4-18
EI crown EI deck

EI deck
M deck (x) = M arch (x) Equation 4-19
EI crown

Flexible system moments M are defined as the sum of arch and deck moments (Billington 1973):

102
4. Statical Analysis of Arch Bridges

M (x) = M arch (x) + M deck (x)


! EI $ Equation 4-20
= M arch (x) # 1 + deck &
" EI crown %

The flexural rigidity of the system EIsys is defined as the sum of arch and deck flexural rigidities, EIdeck and

EIcrown. Substituting EIsys in Equation 4-20 and rearranging results in expressions for moments distributed

to the arch and deck:

EI crown
M arch (x) = M (x) Equation 4-21
EI sys

EI deck
M deck (x) = M (x) Equation 4-22
EI sys

These expressions show that system moments are distributed to the arch in proportion to the ratio of its

flexural rigidity to the flexural rigidity of the system (the same applies to the deck). One interpretation of

these results is that arch and deck participate together like a pair of parallel beams in carrying moments

imposed on the system, as shown in the right-side column of Figure 4-13.

In general, there are four ways that flexural rigidity can be distributed in the system, corresponding to the

four classes of fixed arch systems shown in Figure 4-13. The first class is the unstiffened arch system, whose

decks are discontinuous over the spandrel columns. Because of the discontinuity, the global flexural rigid-

ity of the deck can be taken as zero, regardless of the section properties of the deck. In this case, arches act

alone in resisting moments caused by nonuniform loads or restrained deformations. The second class is

self-stiffened arch system, or arch-stiffened system. In this case, the deck is continuous over the spandrel

columns, but its flexural rigidity is much smaller than the flexural rigidity of the arch. Because of this, only

a small portion of the system moments is distributed to the deck. Reversing the flexural rigidity of the arch

and deck results in the third class, called the deck-stiffened arch system. In this system, the flexural rigidity

of the deck is much greater than the flexural rigidity of the deck. This system was pioneered by Robert

Maillart in the 1920s and 1930s, during the latter years of his career as a bridge designer in Switzerland
103
4. Statical Analysis of Arch Bridges

(Billington 1979). His analysis of this system was simple and approximate, assuming that: (1) the arch axis

is taken as the pressure line of the dead weight plus half of the traffic load, (2) the arch carries all uniform

loads by axial forces only, and (3) all bending moments in the vertical plane are carried by the stiffening

deck girder alone (Billington 1973). The fourth class is the partially deck-stiffened arch system, whose deck

and arch have flexural rigidities of similar magnitude. In this case, both deck and arch carry significant

portions of the flexible system moments.

arch system structural model assumed primary system parallel beam analogy
classification of arches and redundant forces
Classification based Without simplification, Only three independent redundant System moments are distributed
on stiffness distribution. these arch models are forces are considered in each system. to deck and arch in proportion to
highly indeterminate. System moments caused by redundant the ratio of their flexural rigidity to
forces are identical among systems. the flexural rigidity of the system.

EI deck  0
unstiffened
arch system x1 x3
x2 EI crown

EI deck  EI crown
self-stiffened
arch system
x1 x3 EI crown
x2

(1 < Ơ ) u x1 (1 < Ơ ) u x2
EI deck
partially deck-stiffened
arch system
x3
EI crown
Ơ x1 Ơ x2

x1 x2 EI deck
deck-stiffened
arch system
x3
EI crown  EI deck

Figure 4-13. Classification of fixed arch systems, and their respective simplified and analogous structural models

The middle column of diagrams in Figure 4-13 shows the primary systems and redundant forces x1, x2, and

x3 for each class of fixed arch system. The sum of virtual bending moment diagrams caused by each set of

redundant forces is the same among the four classes. Thus, the procedure for calculating bending moments

for any arch system is as follows: (1) loads or restrained deformations are applied to the arch alone, (2) the

104
4. Statical Analysis of Arch Bridges

properties of the arch are assumed to be identical to the arch in the original system, except that its flexural

rigidity includes the flexural rigidity of the deck, (3) uniform loads are carried by the arch along the pres-

sure line in compression, and (4) bending moments are distributed to the arch and deck according to

Equations 4-21 and 4-22. The stiffened arch model used to calculate system moments is shown in Figure 4-

14, along with the virtual moment and virtual axial forces caused by the three redundant forces.

L
redundant virtual bending virtual axial
forces moment diagram force diagram
f

equivalent fixed arch system x1


1
EIsys
x2
x2
x1 x3 1

-f
primary system x3
and redundant forces -1

Figure 4-14. Stiffened arch model and moments and axial forces caused by unit redundant forces

Figure 4-15 investigates the validity of Assumptions 1 and 3 described previously. Fixed arch systems sub-

jected to uniformly distributed loading over half the span have been analyzed using the structural analysis

package sap2000. Three parameters have been varied: the flexural rigidities of the arch EIcrown, deck EIdeck,

and spandrel columns EIcol. All flexural rigidities are expressed as fractions of a constant reference rigidity

EI*. Each row of bending moment diagrams in the figure corresponds to a constant value of EIcol, but di-

fferent proportions of EIcrown and EIdeck. Each column of diagrams corresponds to constant proportion of

EIcrown and EIdeck, but different values of EIcol. Spandrel columns were modelled as monolithically connected

to the arch and deck, except for the last row of diagrams, where they are pin-connected. The flexural rigid-

ity of the arch EIarch was varied along the span according to Equation 4-11, such that the horizontal com-

ponent of EIarch remained constant and equal to EIcrown.

105
4. Statical Analysis of Arch Bridges

span The proportion of bending moment Bending moments in the arch and deck
distributed to the arch is proportional to increase as the flexural rigidity of the
span!= 200 m the ratio of the flexural rigidity of the columns decrease.Thus, it is conservatve
q EIdeck rise != 25 m arch to the flexural rigidity of the to assume that the columns are pin-
q != 35 kN/m system. This is illustrated by connected when designing the deck and
f EIcol I*!= 16.6 m4 comparing diagrams in the same row. arch.This is illustrated by comparing
E != 25000 MPa diagrams in the same column.

EIarch(s) · cos !(s) = EIcrown

class of partially
arch system deck-stiffened deck-stiffened self-stiffened

10000 10000 10000 10000 10000

400
1930
2590
3090
0.25·EI*
3500

deck section 0.75·EI* 0.5·EI* 0.003·EI*


EI*

3500
3090
2590
EI*
1930

arch section 0.003·EI* 0.25·EI* 0.5·EI* 0.75·EI*


400

flexural moment diagram of deck


rigidity of moment diagram of columns
spandrel moment diagram of arch
columns all moments in MN·m
-9.6 -4.9 -3.7 -2.5 -0.4

EI*
-4.2 -8.0
-1.9 -3.0
-0.1

-14.8 -7.9 -5.7 -3.3 -0.4

0.1·EI*
-7.1 -12.4
-2.9 -5.0
-0.1

-19.6 -11.9 -8.3 -4.6


-0.4

0.01·EI* -11.2 -17.1


-7.8
-4.3
-0.1

-24.2 -16.1 -11.1 -6.1


-0.4

0.001·EI* -15.9 -21.8


-6.2 -11.2
-0.1

-26.1 -17.6 -12.1 -6.7


pin- -0.4
connected
columns -17.6 -23.7
-6.9 -12.3
-0.1

Figure 4-15. Effect of arch, deck, and column flexural rigidities on the distribution of bending moment

Two observations can be made from the analytical results presented in Figure 4-15. First, bending mo-

ments in the arch and deck increase as the flexural rigidity of the columns are decreased. This implies that

assuming pin-connected spandrel columns is conservative, since the moments in the arch and deck are

highest when the columns are pin-connected. The moments in spandrel columns, however, increase as

their flexural rigidity increases, and should be checked for strength and ductility once the overall propor-

tions of the system are specified. The second observation is that the moment distribution between arch and

106
4. Statical Analysis of Arch Bridges

deck deduced from Assumption 3 is reasonably accurate, regardless of the fixity and flexural rigidity of the

spandrel columns. This can be seen by comparing the end moments of the arch and deck in the figure.

Assumption 4: When using the method of virtual work for calculating sectional forces in arches, only

bending deformations and their internal work terms need be considered. Figure 4-14 shows the virtual bend-

ing moment and virtual axial force diagrams caused by each unit redundant force. The internal work terms

associated with axial deformation . nN/EA dx can be neglected in non-shallow arches, which are arches

whose crown deflections due to axial shortening are small relative to the initial rise of the arch. Using

high-strength materials in arches may allow for greater shallowness than has been previously achieved.

Thus, Assumption 4, which has typically been valid for arches made from conventional concrete, will not

be used for the parametric design study in Chapter 6.

4.4.2 Force method for non-shallow arches

In this section, the analysis of fixed arches using the force method will be presented using the simplified

statical system and the four assumptions presented in Section 4.4.1. The inclusion of axial deformations in

internal work formulations, which is important for the statical analysis of shallow arches, will be con-

sidered in Section 4.4.3.

The force method uses compatibility equations to solve redundant forces, as given by:

[d][F] + [d 0 ] = 0 Equation 4-23

In this equation, deformation matrix [d0] contains deformations of the statically determinate primary sys-

tem caused by loads or dimensional changes, force vector [F] contains unknown redundant forces, and

flexibility matrix [d] contains deformations caused by unit redundant forces. Total deformations $ are the

sum of bending deformations $M, axial deformations $N, and shear deformations $V:

! = ! M + ! N + !V Equation 4-24

For the type of structural members used in arches, shear deformations are typically small and therefore

will be assumed to be zero ($V=0). Using Assumption 4, axial deformations will be neglected ($N=0). Thus,

107
4. Statical Analysis of Arch Bridges

only bending deformations $M will be considered in the calculation of flexibility coefficients $ij as given by

Equation 4-25, where mi is virtual moment caused by unit redundant force i, and ,j is real curvature

caused by unit redundant force j:

S
! ij = $ mi (s) " # j (s)ds Equation 4-25
0

These quantities are integrated with respect to the curved arch axis s, over the total arc length of the arch S.

Assuming linear-elastic behaviour, real curvature ,j can be expressed in terms of bending moment Mj di-

vided by flexural rigidity EI (Equation 4-26). The resulting expression can be simplified further by apply-

ing Assumptions 1, 2, and 3 and changing the variable of integration to a horizontal variable, using the

transformation: ds cos ' = dx (Equation 4-26).

S M j (s)
! ij = " mi (s) ds
0 EI (s)
Equation 4-26
L M j (x)
= " mi (x) dx
0 EI sys

According to the redundant forces shown in Figure 4-14, the flexibility matrix [d] of the system is:

" !11 !12 !13 % " 13 1


( 13 f %
L $ 1 '
6
$ '
[d] = $ ! 21 ! 22 ! 23 '= $ 6 1
3 ( 13 f ' Equation 4-27
$ ! ' EI sys $ 1 2 '
$# 31 ! 32 ! 33 $# ( 3 f ( 13 f 8
'& 15 f '&

The inverse of the flexibility matrix [d]-1 is:

" 9 3 15
f !1 %
EI sys $ '
2

[d]!1 = $ 3 9 15
2 f !1 ' Equation 4-28
L $ '
!1 !1
$# 15
f 15
f 45
f !2 '
2 2 4
&

108
4. Statical Analysis of Arch Bridges

Force vector [F] can be solved by using Equation 4-29, where xi are redundant forces, and $i0 are deforma-

tions of the primary system caused by loads or dimensional changes calculated using Equation 4-25.

[F] = ![d]!1[d 0 ]
" x1 % " 9 3 15
f !1 % " )10 %
$ ' EI sys $ 2
' $ ' Equation 4-29
$ x2 ' = ! $ 3 9 15
2 f !1 ' ( $ ) 20 '
$ x ' L $ ' $ '
$# 3 '& $# 15
f !1 15
f !1 45
f !2 ' $# ) 30 '&
2 2 4
&

Table 4-3. Bending moment diagrams for fixed arch systems

load case moment diagram algebraic expression for moment

distributed load q
M (x) = 0
over full span

# q
q
% (
141x 2 ! 60Lx + 4L2 ) 0"x"
L
% 486 3
distributed load q %
over middle third span
M (x) = $ !
q
(
102x 2 ! 102Lx + 23L2 ) L 2
<x" L
% 486 3 3
% q
(
% 486 141x ! 222Lx + 85L
2 2
) 2
3
L<x"L
&
q
# q L
% ! ( 8x ! L ) ( 2x ! L ) 0 " x "
distributed load q % 64 2
M (x) = $
over half span % q ( 8x ! 7L ) ( 2x ! L ) L < x " L
%& 64 2

Q
) 27 " 10 2 % L
+ Q $ x + 6x ! L ' 0(x(
concentrated load Q + 512 # L & 4
M (x) = *
at quarter-point + 1 " 270 % L
x 2 ! 350x + 101L ' <x(L
+ 512 $# L
Q
& 4
,

Q
) Q " 30 2 % L
+ $ x ! 14x + L ' 0(x(
concentrated load Q + 32 # L & 2
M (x) = *
at midspan + Q " 30 2 % L
$ x ! 46x + 17L ' <x(L
+ 32 # L & 2
,

strain
EI # # x & &
2
15 x
axial shortening M (x) = ! K " " % 6 % ( ) 6 + 1(
2 f $ $ L' L '

109
4. Statical Analysis of Arch Bridges

With the redundant forces solved, final bending moment diagrams can be found by superimposing mo-

ment diagrams caused by external loads and redundant forces. These moments can then be distributed

among deck and arch according to Equations 4-21 and 4-22. Table 4-3 summarizes bending moment dia-

grams for six load cases calculated by the author using the force method.

4.4.3 Force method for shallow arches

In this section, the analyses of fixed arch systems will be repeated, this time with the inclusion of axial de-

formations in internal work formulations. This more complex analysis reveals the effects of axial rigidity of

the arch on the overall behaviour of the system.

The cross sectional area of the arch typically changes along the arch axis. Similar to the assumption made

for flexural rigidity, the axial rigidity of the arch is assumed to vary according to Equation 4-30, such that

its horizontal component is constant and equal to the axial rigidity of the arch at the crown EAcrown (Melan

and Steinman 1913):

EAarch (s) ! cos " (s) = EAcrown Equation 4-30

Flexibility coefficients $ij including axial deformations are given by Equation 4-31, where ni is virtual axial

force caused by unit redundant force i, and %j is real strain caused by unit redundant force j. Real strain is

equal to real axial force Nj divided by arch axial rigidity EAcrown. The first term in Equation 4-31 is identical

to Equation 4-26, for which the variables have already been defined.

L L
! ij = $ mi" j dx + $ ni # j dx
0 0

L Mj L Nj Equation 4-31
= $ mi dx + $ ni dx
0 EI sys 0 EAcrown

The flexibility matrix [d] of the system based on the redundant forces shown in Figure 4-14 is:

110
4. Statical Analysis of Arch Bridges

" 13 1
( 13 f %
" !11 !12 !13 % $ 6
'
$ ' L $ 16 1
( 13 f '
[d] = $ ! 21 ! 22 ! 23 '= $
3
' Equation 4-32
$ ! ' EI sys $ ( 1 f EI sys '
$# 31 ! 32 ! 33 ( 13 f f +
8 2
'& $ 3 15
EAcrown '
# &

The inverse of the flexibility matrix [d]-1 is:

# 6
f 2 Acrown + 6I sys 2
f 2 Acrown ! 3I sys fAcrown &
EI sys 30 % 5 5
(
!1
[d] = " % 2
f 2 Acrown ! 3I sys 6
f 2 Acrown + 6I sys fAcrown ( Equation 4-33
L 4 f 2 Acown + 45I sys % 5 5
(
% fAcrown fAcrown 2 Acrown (
3
$ '

When uniformly distributed load q is applied over the entire span, the primary system deformations [d0]

are given by Equation 4-34, where $10 and $20 are rotations at each end of the arch, and $30 is horizontal

displacement:

"!10 % " %
T
qL3 qL3 qL3
[d 0 ] = $$! 20 '' = $ ( 241 ( 1
24
1
15 f ' Equation 4-34
$ EI sys EI sys EI sys '
$#! 30 '& # &

Solving for force vector [F], the redundant forces shown in Equation 4-35 are obtained, where x1 is mo-

ment at the left support ML, x2 is moment at the right support MR, and x3 is horizontal reaction H:

! x1 $ ! $
T
( ( qL2
[F] = ## x2 && = # ' 121 qL2
1
' qL1 2 1
& Equation 4-35
1+ ( 1 + 45 ( f 1 + 45 ( &
12 8
#" x3 &% #" %

These expressions have been simplified by introducing a non-dimensional ratio ", which is defined as:

45 EI sys
!= " Equation 4-36
4 f 2 EAcrown

111
4. Statical Analysis of Arch Bridges

Non-dimensional ratio " will be referred to as the arch-beam parameter because it describes the transition

of shallow arches between fixed arch and fixed beam behaviour. At the limit "=0, fixed arch systems exhib-

it pure arch behaviour, responding to loads and deformations as described in Section 4.4.2. At this limit,

arches are not shallow, since crown deflections are small relative to the initial rise of the arch. At the limit

"=!, fixed arch systems exhibit pure beam behaviour, responding to loads and deformations as would

straight fixed beams. At this limit, arch crown deflections are large relative to the rise of the arch.

The force method was repeated for several different load cases, including: partial distributed load over half

span and middle third, concentrated loads at midspan and quarter-point, and restrained axial deforma-

tion. The resulting support reactions are tabulated in Figure 4-16. In addition to showing support reac-

tions, the figure also shows bending moments and bending deflections at critical locations (x=a) for the

limiting cases of "=0 and "=!.

112
L moments caused by axial forces caused by
! ! redundant forces ! redundant forces ! flexural rigidity EI = EI sys ! axial rigidity EA = EAcrown
f -H·f -H 45 EI
! ! ! ! arch-beam parameter ! = " = 45 # $2
f !
fixed arch system 4 f 2 EA
2
ML $ H ' $ L 3
'
! ! ! ! axial deflections at crown " n! dx !f " & 64
+ # ) * & 15 + 4 f)
% EA ( % f (
ML MR
primary system H MR
and redundant forces

load case! full span q ! middle third q ! half span q ! midspan Q ! quarter-point Q ! shrinkage
q q q Q Q
strain

primary system moment


!>0 is compressive
2 2 2
qL 1 47 qL 1 qL 1 QL 1 135 QL 1 EI 1
horizontal reaction, H! + 18 ! ! + 648 ! ! + 161 ! ! + 15
64 ! ! + 1024 ! ! ! 454 " # # !
f 1+ " f 1+ " f 1+ " f 1+ " f 1+ " f2 1+ $

" 2
1 " 398 # 1 ! 53 " 1 " 4# 27
1 + 83 # EI 1
left moment reaction, ML! ! 121 qL2 ! + 243 qL2 ! ! ! 641 qL2 ! + 321 QL ! ! ! 512 QL " ! ! 152 " # #
1+ " 1+ # 1+ " 1+ # 1+ # f 1+ $

113
" 2
1 " 398 # 1 + 113 ! 1 " 4# 21
1 " 87 # EI 1
right moment reaction, MR! ! 121 qL2 ! + 243 qL2 ! ! + 641 qL2 ! + 321 QL ! ! + 512 QL ! ! ! 152 " # #
1+ " 1+ # 1+ ! 1+ # 1+ # f 1+ $

L L L L L L
critical location x=a! ! ! ! ! !
2 2 4 2 4 2

57 32
! 5
1 + 10 " 1
1 + 373 ! 1 + 83 " 243
1 + 27 " EI 1
moment at x=a ! + 241 qL2 ! + 972 qL2 ! ! + 128 qL2 ! + 643 QL ! ! + 4096 QL ! ! + 154 ! " "
1+ ! 1+ " 1+ ! 1+ " 1+ " f 1+ #

Figure 4-16. Force method solutions for shallow arches


case ! = 0 (pure fixed arch)
moment diagram!

2
! 49 qL4 $ 2
! 11 qL4 $ 1
! 1 QL3 $ 19
! 9 QL3 $
bending deflections " m! dx at x=a! 0 ! 49 #" 31104 EI &% ! 11 #" 12288 EI &% ! 16 #" 192 EI &% ! 64 #" 4096 EI &% ! 0

case ! = " (pure fixed beam)


moment diagram
1 qL4 49 qL4 11 qL4 1 QL3 9 QL3
bending deflections " m! dx at x=a! 384 ! 31104 ! 12288 ! 192 ! 4096 ! 0
EI EI EI EI EI
4. Statical Analysis of Arch Bridges
4. Statical Analysis of Arch Bridges

Another important system parameter is modified slenderness ratio (f, which has been used in stability

models for shallow arches by Cai et al. (2009) and Bradford et al. (2007). Modified slenderness ratio is de-

fined as two times the arch rise f divided by radius of gyration r of the arch (Equation 4-37). Because the

deck contributes to the bending stiffness of the systems being considered, it is appropriate to replace r with

an equivalent system radius of gyration rsys (Equation 4-38).

EAcrown 2 f
!f = 2 f = Equation 4-37
EI sys rsys

EI deck + EI crown
rsys = Equation 4-38
EAcrown

Because they share the same parameters, arch-beam parameter " can be expressed explicitly in terms of

modified slenderness ratio (f without introducing additional variables:

! = 45 " #2
f Equation 4-39

The transition between pure arch and pure beam behaviour is best illustrated by comparing moments car-

ried by fixed arch systems with those carried by fixed beams subjected to the same loading. From Equation

4-35, arch end moments ML and MR caused by uniform load q distributed over the full span L are given by:

"
M arch (x = 0, L) = ! 121 qL2 Equation 4-40
1+ "

Under the same loading, end moments in fixed beams are given by:

M beam (x = 0, L) = ! 121 qL2 Equation 4-41

In terms of arch-beam parameter " and modified slenderness ratio (f, the ratio between fixed arch and

fixed beam moments caused by uniformly distributed loading is:

114
4. Statical Analysis of Arch Bridges

M arch (x = 0, L) !
=
M beam (x = 0, L) 1 + !
Equation 4-42
45 " #2
=
f

1 + 45 " #2
f

The horizontal reaction in shallow arches H caused by uniform loading q is given by:

qL2 1
H= !
8 f 1+ "
Equation 4-43
qL2 1
= !
8 f 1 + 45 # $2
f

The first factor in Equation 4-43 is the horizontal reaction in non-shallow arches (Equation 4-6). The

second factor 1/(1+") is a ratio between 0 and 1 that represents the amount of arch action that is present in

the system, 0 being no arch action (pure beam action), and 1 being pure arch action.

The ratio of horizontal reactions (Equation 4-43) and ratio of end moments (Equation 4-42) are plotted in

Figure 4-17 as functions of modified slenderness ratio (f. The transition point, where arch systems behave

equally as arch and beam occurs when (f =9.49. Although it would be possible to design arch systems with

(f =9.49, the system would not be structurally efficient. The primary advantage of the arch form is that per-

manent loads can be carried in pure compression, which requires less material than carrying permanent

loads in flexure. This advantage is lost when the modified slenderness ratio of the system is low enough

such that significant bending moments are present even under permanent loads whose pressure line fol-

lows the shape of the arch. Such moments would better be resisted using girder-type systems rather than

shallow arch systems.

115
4. Statical Analysis of Arch Bridges

Ratio of moments in M(x)arch 1


fixed arches to M(x)beam
moments in fixed
beams under
uniformly distributed
load 0.5 Fixed arches with !f > 20
Proposed threshold for
efficient arch behaviour: have bending moments
!f > 20 that are less than 10% of
those in fixed beams
with the same span and
uniform loading
0
0 20 40 60
Modified slenderness ratio !f

Ratio of horizontal H(! !f ) 1


reaction in shallow fixed H(! !f = " )
arches to horizontal
reaction in non-shallow
arches under uniformly Fixed arches with !f > 20
distributed load: !f > 20 resist more than 90% of
0.5
uniform loads in axial
compression

0
0 20 40 60
Modified slenderness ratio !f

Figure 4-17. Transition between fixed arch and fixed beam

To ensure that given arch designs behave efficiently, a minimum modified slenderness ratio (f of 20 is pro-

posed (see Figure 4-17). This threshold ensures that dead load moments in fixed arches are less than 10% of

those in fixed beams, and that more than 90% of uniform loads are carried in axial compression along the

axis of the arch. As (f is decreased beyond the proposed threshold, there is a rapid increase in bending mo-

ment and a rapid decrease in axial force in the arch, which are both undesirable.

The transitional behaviour between arch and beam is not limited to the case of uniformly distributed load-

ing. All applied load cases shown in Figure 4-16 have moments that are functions of the factor "/(1+").

These factors account for the effects of axial deformation on the system. When "=0, axial deformations do

not cause additional bending moments in the system above that of non-shallow arches. As " is increased

above 0, axial deformations introduce additional sectional forces that move the system toward fixed beam

behaviour.

116
4. Statical Analysis of Arch Bridges

Crown deflections #f, which will later be used to calculate second-order moments, can be calculated using

Equation 4-44, where H is horizontal reaction caused by dead and live loads, EAcrown is axial rigidity of the

arch crown, % is strain caused by creep, shrinkage, and uniform drops in temperature, L is span, f is rise,

and " is arch-beam parameter:

$ H ' $ L2 3 '
!f " & + # ) * & 15 +4 f) Equation 4-44
% EAcrown ( % f (
64

This equation is calculated by integrating ∫n(x)·%(x) dx, where n is virtual axial force caused by a unit load

applied at the crown, and % is real axial strain in the system. This equation is similar to the approximate

equation given by Menn (1990):

H $ 1 L2 3 '
!f " # 4 +4 f) Equation 4-45
EA &% f (

The last four rows of diagrams and equations in Figure 4-16 depict the differences of the limiting cases of

pure arch behaviour "=0 and pure beam behaviour "=!. For each type of load case, including distributed

load q, concentrated load Q, and restrained deformation, bending moment diagrams for arches with "=0

and arches with "=! are drawn to the same scale. By inspection, pure arch systems have significantly

lower moments than pure beam systems. Bending deflections in the systems are calculated using virtual

work: ∫m(x)·,(x) dx, where m is virtual moment caused by a unit load applied at the location of and in the

same direction as the desired bending displacement and , is real curvature caused by the given loading.

The results show that pure arch systems have significantly lower bending deflections than pure beam sys-

tems. The case of distributed loading over half the span, for example, causes deflections in pure arches that

are 18% of those in pure beams. These deflections, however, exclude vertical deflections caused by axial de-

formations of the arch. Thus, the expressions given in the figure for bending deflections at midspan must

be added to the crown deflections caused by axial deformation (Equation 4-44).

117
4. Statical Analysis of Arch Bridges

4.4.4 Flexural buckling of shallow arches

The flexural buckling of arches has previously been discussed in Section 4.3.1. Recent analytical studies by

Cai et al. (2009) have shown that the buckling resistance of shallow arches can be significantly lower than

those predicted by the effective arc length method presented in Section 4.3.1. Using virtual work, Cai et al.

formulated nonlinear equilibrium and compatibility equations, and then solved them to obtain solutions

for symmetric snap-through and antisymmetric bifurcation buckling loads. In their formulations, they as-

sumed that: (1) the form of the arch is parabolic, (2) the square of slopes of the parabola are much less than

unity: (dy/dx)2<<1, and (3) the change in horizontal displacements of the arch with respect to the hori-

zontal axis are much less than unity: du/dx<<1.

The distinction between when arches are shallow and when they are not shallow is not specified by Cai et

al. (2009). An inspection of their second assumption (dy/dx)2<<1 gives some insight into when their stab-

ility analysis is valid. The square of the slope (dy/dx)2 of parabolic arches at their springing lines is given by

Equation 4-46, where L/f is the span-to-rise ratio of the arch.

'2
! dy(x = 0, L) $ ! L$
2

#" &% = 16 # & Equation 4-46


dx " f%

Figure 4-18, which illustrates Equation 4-46, shows that the assumption of (dy/dx)2<<1 is only valid at the

springing lines at high span-to-rise ratios. Arch springing lines with span-to-rise ratios of 12:1 or higher

have values of (dy/dx)2 that are less than 0.1.

square of slope 1
at springing line
d2y
dx2
0.5
12:1

0.1
0
0:1 4:1 8:1 12:1 16:1 20:1
span-to-rise ratio L/f

Figure 4-18. Square of slope of parabolic arches at their springing lines

118
4. Statical Analysis of Arch Bridges

Cai et al. (2009) found that modified slenderness ratio (f (Equation 4-37) was the governing parameter of

their virtual work analysis. They identified three domains of buckling behaviour, determined by the value

of (f. Their results (adjusted according to the erratum by Cai and Feng (2009)) are summarized below.

Domain 1: When (f >19.8, shallow arches buckle anti-symmetrically at critical buckling loads HE, f :

$ "2 '
H E, f = & 0.6 ± 0.4 1 ! 30.7 2 ) * H E # f > 19.8 Equation 4-47
% #f (

HE is the second mode, buckling resistance of fixed columns under uniform compression, as given by

Equation 4-48. When arc length S is substituted for L, HE is equal to the buckling resistance of non-shallow

fixed arches, as given by Equation 4-2 on page 87.

EI
HE = ! 2 2
Equation 4-48
# L&
%$ 0.7 " ('
2

Domain 2. When 9.87 $ (f $ 19.8, shallow arches buckle symmetrically at critical buckling loads:

( )
H E, f ! 0.00079 " 2f # 0.009 " f + 0.314 $ H E 9.87 % " f % 19.8 Equation 4-49

Cai et al. (2009) wrote a computer program to solve two equations iteratively to obtain symmetric buck-

ling loads for given values of modified slenderness. Because the iterative procedure was complex and did

not result in closed form solutions, Cai et al. used a least squares approximation of the results to calculate

the quadratic coefficients in Equation 4-49. Their solution showed good agreement with results from

numerical buckling analyses done in the finite element analysis package ansys.

Domain 3. When (f < 9.87, shallow arches do not buckle because they do not carry axial forces large

enough to cause buckling. Thus, these shallow arches behave more like curved beams with small initial

curvature than as arches.

119
4. Statical Analysis of Arch Bridges

no buckling
symmetric buckling
antisymmetric buckling
q
HE,f
1
H E, f H E, f
HE
0.8

anti-symmetric buckling mode 0.6

0.4
q
0.2
H E, f H E, f
0
0 9.87 19.8 60
symmetric buckling mode
Modified slenderness ratio !f

Figure 4-19. Buckling of shallow arches. Adapted from Cai et al. (2009).

Equations for buckling loads, or buckling resistances HE, f for shallow arches are plotted in Figure 4-19. The

figure shows that the efficiency threshold of (f > 20 proposed in Section 4.4.3 has relevance to the buckling

mode of shallow arches. Proportioning arches such that their modified slenderness is greater than 20 en-

sures that their primary buckling mode remains anti-symmetric and that their buckling resistance is at

least 80% of HE. As (f is decreased beyond the proposed threshold, there is a rapid decrease in arch buck-

ling resistance due to a distinct change in the primary buckling mode of the system.

An additional effective arc length factor kf can be used in Equation 4-2 on page 87 to account for the re-

duction of buckling resistance in arches due to the effects shallowness. The shallowness factor can be cal-

culated using Equation 4-50, or can be conservatively taken as 1.12, if modified slenderness ratio is greater

than 20.

' 12
! H E, f $
kf = # Equation 4-50
" H E &%

4.4.5 Second-order analysis: geometric nonlinearity

The force method solutions presented in Section 4.4.2 and Section 4.4.3 do not account for arch displace-

ments in formulations of equilibrium, and are thus considered first-order analyses. Because arches carry

loads in compression and bending, they are susceptible to additional bending moments that are propor-

120
4. Statical Analysis of Arch Bridges

tional to the eccentricities between the centroid of the arch and line of action of load. Second-order ana-

lyses, in which these additional moments are considered, have been previously discussed in Section 3.1 for

eccentrically-loaded columns. The second-order behaviour of arches is very similar to the second-order

behaviour of straight columns. Vianello’s deflection magnification method (1898) will be used to calculate

second-order deflections and moments approximately. Second-order arch deflections w(x) are given by

Equation 4-51, where w0(x) is first-order arch deflection, H is horizontal reaction, and HE is buckling resist-

ance of the arch.

1
w(x) = w0 (x) Equation 4-51
H
1!
HE

First-order deflections at critical locations along the arch w0(x=a) can be calculated using the equations

shown in Figure 4-16 on page 113. When considering crown deflections w0(x=L/2), deflections caused by

both flexural deformations ∫m·, dx and axial deformations #f =∫n·% dx (Equation 4-44) need to be con-

sidered. The component of #f that is caused by sustained loads can be excluded if the arch is properly

cambered during construction. In other words, the camber of the arch should be designed such that the

pressure line of the sustained load and the axis of the arch are concentric once the arch is closed, eliminat-

ing any eccentricity that would cause additional moments. When considering quarter-point deflections

w0(x=L/4), deflections caused only by flexural deformations ∫m·, dx need be considered, since axial short-

ening of the arch causes only minimal eccentricities of load near the quarter-points.

Horizontal reactions of the arch H can be calculated by summing all applicable equations for H in Figure

4-16, including the effects of restrained deformations and axial shortening which can reduce compression

in the arch. The largest portions of H are typically contributed by permanent dead loads qdead and uni-

formly distributed live loads qlive, which can be calculated using Equation 4-43.

The critical buckling load of fixed arches HE can be calculated using Equation 4-2 on page 87, in which

effective arc length factor k is taken as 0.35, and EI is taken as the flexural rigidity of the system EIsys. If the

121
4. Statical Analysis of Arch Bridges

span-to-rise ratio of the system is 12 or higher then the buckling resistance of shallow arches HE, f (Equa-

tion 4-47) should be used instead of HE.

At the critical section, the eccentricity e(x=a) between the centroid of the arch and the deformed line of

action of the load is proportional to the second-order deflection w(x=a), as given by Equation 4-52, where

& is eccentricity constant (Menn 1990).

e(x = a) = ! " w(x = a) Equation 4-52

Eccentricity constants & can be assumed to be the same as those in straight columns with similar deflected

shapes. Arch deflections caused by uniformly distributed loading over half the span are similar to the

second mode shape of straight, fixed-fixed beams, as shown in Figure 4-20 (Menn 1990). In this case, the

eccentricity between the deformed line of action of load and the centroid of the critical section of the de-

formed member is 0.733 times the maximum deflection of the member (& = 0.733). For load cases in which

the crown section is critical, arch deflections are similar to the first mode shape of straight, fixed-fixed

beams, as shown in Figure 4-20. In this case, the eccentricity between the deformed line of action of load

and centroid of the critical section of the deformed member is 0.5 times the maximum deflection of the

member (& = 0.5).


distributed loading over distributed loading over
left half of arch span middle third of arch span

q q

H H H H
Z = second-order deflection
Z Z at critical section
H
H H H H undeformed line of action
H H H deformed line of action
e=ƦÃZ e = second-order eccentricity
e=ƦÃZ
e=0.733ÃZ of load at critical section
e=0.5ÃZ

Figure 4-20. Analytical model for second-order analysis of arches. Partially adapted from Menn (1990).

The total second-order moment M at the critical section (x=a) is the sum of first-order moment M0 and

additional moment M1=H·&·w, as given below:

M (x = a) = M 0 (x = a) + H ! " ! w(x = a) Equation 4-53

122
4. Statical Analysis of Arch Bridges

4.4.6 Second-order analysis: material nonlinearity

Vianello’s method for calculating second-order moments provides a simple and rational means of account-

ing for geometric nonlinearity. The nonlinear response of the material caused by the cracking of concrete,

the yielding of steel, and the pull-out of fibres can be accounted for by replacing flexural rigidity EI with

secant stiffness EI!, which is a secant to the moment-curvature diagram calculated for a given point as giv-

en by Equation 4-54, where M is bending moment, , is curvature of the section, and N is axial force

(Menn 1990):

M
EI !(N ) = Equation 4-54
"

The calculation of EI! is iterative because the second-order moment M is a function of deflection, and

deflection is a function of EI!, and EI! is a function of M. To simplify calculations, secant stiffness can be

taken as one of three discrete values: secant stiffness at cracking EIcr = Mcr ÷ ,cr, secant stiffness at yielding

of reinforcing steel EIy = My ÷ ,y, and secant stiffness at ultimate state EIu = Mu ÷ ,u.

The various limit states corresponding to cracking, yielding, and ultimate have previously been discussed

in Section 3.5.1 for conventional reinforced concrete sections and in Section 3.5.2 for ultra high-perform-

ance fibre-reinforced concrete sections. These limit states were examined within the context of two ex-

ample hollow box sections whose geometries were calibrated so that their maximum axial forces in pure

axial compression were about the same. From this study, it was observed that high-strength concrete sec-

tions do not exhibit a distinct yield point corresponding to distinct changes in flexural stiffness. This is due

to the difference in form factor of reinforcing steel as fibres, rather than bars. It was also observed that

effective secant stiffnesses of high-strength concrete sections at ultimate limit states are typically 50% to

75% of the gross flexural rigidity EI0 of the section. This stiffness is quite high compared to the 25% of EI0

secant stiffness values typical of conventionally reinforced concrete sections.

For the statical analysis of ultra high-performance fibre-reinforced concrete systems, the effective secant

stiffness for serviceability limit states can be assumed to be equal to the secant stiffness at cracking EIcr,

123
4. Statical Analysis of Arch Bridges

which is equal to, or slightly less than the gross flexural rigidity of the section EI0. The effective secant stiff-

ness at ultimate limit states EIu can be approximated using the following equation, where N! is maximum

axial force of the section in pure axial compression (same as Equation 3-25):

+ # N &
- 1
EI 0 % + 1( 0 ) N ) 12 N "
-
2
$ N" '
EI u ! , Equation 4-55
- # N&
1
EI 0 % 2 * ( 1
N" ) N ) N"
- 2
$ N"'
2
.

System flexural rigidity of systems EIsys is defined as the sum of the rigidities of the deck and arch. Simil-

arly, system secant stiffness EI!sys is defined as the sum of secant stiffnesses of the arch and deck:

! = EI deck
EI sys ! + EI crown
! Equation 4-56

The value of EI!sys at ultimate limit states can conservatively be taken as the sum of effective secant stiff-

nesses EIu of both the arch and deck. In some cases, only one of the two elements will have exceeded its

cracking moment under a given set of factored loads. For example, in a deck-stiffened arch, it is possible

that the deck girder has exceeded its cracking moment, while the arch remains uncracked at ultimate limit

states. In this case, it is appropriate to calculate the effective system secant stiffness as:

! = EI u,deck + EI 0,crown
EI sys Equation 4-57

Using secant stiffnesses to account for nonlinear moment-curvature behaviour also affects how flexible

system moments are distributed to the arch and deck. Updating Equations 4-21 and 4-22 on page 103 result

in the following expressions:

!
EI crown
M arch (x) = M (x) Equation 4-58
!
EI sys

!
EI deck
M deck (x) = M (x) Equation 4-59
!
EI sys

124
4. Statical Analysis of Arch Bridges

For convenience, a system stiffness factor κ can be defined as the ratio between the effective system secant

stiffness EI!sys and the system flexural rigidity EIsys:

"
EI sys
!= Equation 4-60
EI sys

Hence, all calculations that use system flexural rigidity EIsys can be modified by multiplying EIsys by system

stiffness factor κ. When the arch and deck remain uncracked, such as under service loads, system stiffness

factor κ is equal to one. At ultimate limit states, κ is less than one.

There are four important quantities that depend on the system stiffness factor κ: (1) deflections, which are

inversely proportional to κ, (2) buckling resistances, which are proportional to κ, (3) system radii of gyra-

tion, which are proportional to the square root of κ, and (4) moments caused by restrained deformation,

which are proportional to κ. Using estimates of system stiffness factors κ that are lower than the value ob-

tained more precisely through iteration (Equation 4-54) is conservative for the first two quantities: deflec-

tions are increased and buckling resistances are reduced. Using low estimates of system stiffness factors κ is

unconservative for the last two quantities. Decreases in system radii of gyration rsys causes increases in

modified slenderness ratios (f, which moves the system toward arch behaviour (as seen in Figure 4-17 on

page 116), causing moments in the system to be reduced. Moments caused by restrained deformation are

highly sensitive to the assumed value of system flexural rigidity. Significant underestimates of κ would res-

ult in significant underestimates of these moments. Thus, particular attention must be given to the calcula-

tion of these moments at serviceability limit states, where structures typically remain uncracked (κ=1), and

at ultimate limit states, where the extent of system softening (κ<1) is governed by the rotation capacities of

the arch and deck.

Arch-beam parameter " (Equation 4-36) can be reformulated to account for material nonlinearity by using

system stiffness factor κ:

125
4. Statical Analysis of Arch Bridges

45 # EI sys
!= " Equation 4-61
4 f 2 EAcrown

This formula assumes that the axial rigidity EA of the member remains unchanged even at ultimate limit

state. This assumption is valid as long as the axial forces in the member are kept below 70% of the maxim-

um compressive strength of the section: N < 0.7N!. Below this limit, mid-depth strains along the member

will remain in the linear-elastic range of the material, as shown by the stress-strain curve in Figure 2-6 on

page 24.

4.4.7 Fixed system moments

Up to this point, only moments arising from the global deformations of arch systems have been con-

sidered. These moments are collectively known as flexible system moments (Menn 1990). Fixed system

moments, which are defined as the moments that arise from local deformations of the deck and arch

between spandrels columns, must be added to the flexible system moments to obtain the total bending

moment demand. The magnitudes of fixed system moments are related to interior span lengths. Decks

must safely carry both dead and live loads to supporting spandrel columns in bending before loads are car-

ried by the system as a whole. If the arch form is polygonal in shape, then self-weight moments of the arch

will also cause local bending between the angle breaks of the polygon. As a conservative approximation,

fixed system moments can be accounted for by increasing the positive and negative flexible system mo-

ment demands Mpos and Mneg by the maximum moments caused by distributed load q and concentrated

load Q fixed-fixed beams with spans equal to the interior span length Ln:

M neg = M ! 121 qLn2 ! 18 QLn Equation 4-62

M pos = M + 241 qLn2 + 18 QLn Equation 4-63

126
4. Statical Analysis of Arch Bridges

4.5 Critical load combinations

Section 4.4 presented methods for calculating second-order moments in fixed arches caused by dead loads,

live loads, and restrained deformations applied individually. The following sections will describe methods

for which these loads can be combined to estimate maximum sectional forces in the system. Five load

combinations are considered: three that cause maximum sectional forces at midspan and two that cause

maximum sectional forces near the quarter-points, as shown in Figure 4-21. When the effects of restrained

deformation (i.e. creep, shrinkage, and changes in temperature) are non-negligible, maximum sectional

forces in fixed, shallow arch systems tend to occur at the springing lines and at midspan. Calculations for

these load combinations are summarized in Table 4-4. When the effects of restrained deformation are neg-

ligible, maximum sectional forces tend to occur near the quarter-points. Calculations for these load com-

binations are summarized in Table 4-5.

critical section midspan midspan midspan quarter-point quarter-point


ULS 2 ULS 2 ULS 2 ULS 1 ULS 1
load combination live load case 1: live load case 2: live load csae 3: live load case 4: live load case 5:
lane load over lane load over truck load* lane load over truck load*
full span middle third at midspan half span at quarter
q q Q q Q
load duration load type

short-term
live load
moments

temperature
induced moments

creep, shrinkage, and


temperature effects not
considered for these load cases
long-term dead load and
creep induced
moments

concrete shrinkage
induced moments

short-term
and additional
long-term secondary
moments

*truck load idealized as single axle

Figure 4-21. Short-term and long-term load combinations

127
4. Statical Analysis of Arch Bridges

Table 4-4. Calculation of maximum sectional forces at midspan.


Calculation includes effects of restrained deformation, and are applicable to S6-06 load combinations SLS 1 and ULS 2.
LOAD
calculation component symbol factors equation

long-term moment elastic shortening of arch Mdead !D 4-66, 4-71


creep of sustained load Mcreep !K 4-68, 4-71
shrinkage of arch1 Mshrink !K 4-69

long-term deflection horizontal reaction1 Hlong !D, !K 4-73


buckling resistance1 HE,long — 4-81
first-order deflection1 $long,0 — 4-74
second-order deflection $long — 4-80
equivalent short-term deflection $eq,0 — 4-84

short-term moment temperature drop, and: Mtemp !K 4-77


—l1. lane load over full span, or Mlive !L (Figure 4-16)
—l2. lane load over middle third, or Mlive !L 4-76
—l3. truck load at midspan Mlive !L 4-76

short-term deflection horizontal reaction Hshort !K, !L 4-78


buckling resistance HE,short — 4-85
first-order deflection $short,0 — 4-79

eccentricity at midspan total second-order deflection $ — 4-87


total eccentricity of load e — 4-52
total second-order moment M — 4-88

1use EIsys=Eadj·Isys to account for creep

Table 4-5. Calculation of maximum sectional forces at quarter-point


Calculation excludes effects of restrained deformation, and are applicable to S6-06 load combinations SLS 1 and ULS 1.

calculation component symbol factors equation

moments —l4. lane load over half span, or Mlive !L 4-90


—l5. truck load at quarter-point Mlive !L 4-90

deflections horizontal reaction, dead load Hdead !D 4-89


horizontal reaction, live load Hlive !L 4-89
buckling resistance HE,short — 4-85
first-order deflection $0 — 4-91

eccentricity at quarter-point second-order deflection $ — 4-92


total eccentricity of load e — 4-92
total second-order moment M — 4-93

128
4. Statical Analysis of Arch Bridges

Horizontal reactions H and bending moments M are divided into two categories: forces caused by long-

term loads and forces caused by short-term loads:

H (t) = H long (t) + H short


Equation 4-64
M (x,t) = M long (x,t) + M short (x)

The response of fixed arches to long-term loads and short-term loads will be discussed respectively in Sec-

tion 4.5.1 and Section 4.5.2. Calculations are presented in the context of live load 2 (partial distributed load

over middle third span) and live load 3 (concentrated load at midspan) from Figure 4-21, combined with

dead load, creep, shrinkage, and uniform temperature drop to cause maximum sectional forces at mid-

span. Section 4.5.3 describes the calculation of total second-order moments, which requires that long-term

deflections are modified as equivalent short-term deflections. The last section deals with live load 4 (partial

distributed load over half the span) and live load 5 (concentrated load at the quarter-point) from Figure 4-

21. These calculations are less complex because creep, shrinkage, and temperature are not considered.

In all these calculations, the effects of axial deformation and arch shallowness are considered. Thus, equa-

tions in Figure 4-16 on page 113, which were determined using the force method, are used in calculating

bending moments and deflections at critical sections of the system. These calculations are outlined in

Tables 4-4 and 4-5. Load factors are also given in these tables according to the types of action being con-

sidered. The symbols used for load factors are !D for dead load, !L for live load, and !K for restrained de-

formations. As discussed in Section 4.4.6, material nonlinearity is accounted for by using system stiffness

factor κ which accounts for the softening of the system at ultimate limit states. When appropriate, age-ad-

justed effective modulus of elasticity Eadj is used to account for creep effects.

The middle four-fifths of the span can be designed for the greater of the of the calculated quarter-point or

midspan moments. The remaining outer portions of the span will have larger sectional force demands than

the middle four-fifths of the span. Sectional forces at the arch springing lines are not explicitly described in

the rest of this chapter because they can be estimated from the maximum sectional forces at the quarter

point and midspan. For preliminary design, bending moments at the springing lines can be taken as twice
129
4. Statical Analysis of Arch Bridges

the maximum moments found at the quarter-points or at midspan. Among the five critical live load cases

considered in Figure 4-21, end moments are less than or equal to two times the maximum quarter-point or

midspan moment. Alternatively, a more detailed calculation can be made by using the various redundant

moment equations given in Figure 4-16 on page 113. Once the deck and arch have been designed to meet

the force demands at the quarter-point and midspan, their proportions can be gradually increased as

needed toward the springing lines. The additional amount of concrete needed to increase the capacity of

the deck and arch at the springing lines is typically small relative to the total weight of the superstructure.

4.5.1 Long-term loads

Long-term moments Mlong in arch systems are caused by dead loads Mdead, by concrete creep Mcreep, and by

concrete shrinkage Mshrink:

M long (x,t) = M dead (x) + M creep (x,t) + M shrink (x,t) Equation 4-65

The moments caused by dead loads Mdead are greatly influenced by the camber of the arch and method of

construction. Arches should be cambered such that their geometries after elastic deformation are aligned

as closely as possible with the pressure line of the permanent loads, adjusted to the desired arch rise. Even

if arches are perfectly cambered, bending stresses will still occur if the system is restrained while the arch

shortens elastically. Such stresses occur when dead loads are transferred to arches by the systematic lower-

ing of falsework. In this case, axial stresses and axial deformations are introduced while arches are already

in their final, indeterminate state. The bending moments that arise due to uniformly distributed dead load

qdead are given by Equation 4-66:

# " &
M dead (x = 12 L) = +! D % 241 qdead L2 Equation 4-66
$ 1 + " ('

According to Bazant (1972), the effect of concrete creep can be accounted for by using Equation 4-67,

which gives the ratio of long-term stress /(t) to initial applied stress /(t0) as functions of creep coefficient

,(t) and ageing coefficient +(t):

130
4. Statical Analysis of Arch Bridges

! (t) # (t)
= 1" Equation 4-67
! (t 0 ) 1 + $ (t) % # (t)

This ratio represents the reduction by creep of stresses caused by restrained deformation. Combining the

second term in Equation 4-67 with Equation 4-66 leads to an expression for the reduction by creep of mo-

ments Mdead caused by dead loads:

# (t)
M creep (x = 12 L,t) = !" K M dead (x = 12 L) Equation 4-68
1 + $ (t) % # (t)

Moments caused by the autogenous and drying shrinkage of concrete are given by Equation 4-69, where

%sh is shrinkage strain and Eadj is age-adjusted, effective modulus:

$ # Eadj (t) # I sys 1


M shrink (x = 12 L,t) = + 154 ! K " sh (t) # Equation 4-69
f 1+ %

An improved method for decentering arches from their forms was developed by Freyssinet for the con-

struction of the Praireal bridge in 1907 (Ordóñez and Antonio 1979). Freyssinet used flat jacks at the arch

crown joint to gradually introduce axial stresses into the arch before it was closed. As hydraulic pressure

was increased in the jacks, abutting sides of the joint pressed apart, causing the two sides of the arch to

shorten while they were still statically determinate. Eventually, the arch would lift off of its centering and

become self-supporting (Mondorf 2006). This method of transferring dead loads to the arch has been ad-

opted in the construction of most large concrete arch bridges ever since. In the construction of the Rio

Parana Bridge in 1965, for example, twenty-eight hydraulic jacks (see Figure 4-22) were activated at the

crown of the arch to introduce a force of 100 MN (Stellmann 1966). This caused the joint to open by 60

mm. Once the thrust in the arch was fully developed, the open joint was wedged so that the jacks could be

deactivated. Then the joint was filled in with a new placement of concrete.

131
4. Statical Analysis of Arch Bridges

Figure 4-22. Flat hydraulic jacks at crown of Rio Parana Bridge. Adapted from Stellmann (1966).

Using this method of releasing the arch from its forms by jacking significantly reduces the bending stresses

in the arch. If the procedure is executed with high precision and control, it is possible to eliminate virtually

all bending moments at closure Mclose (Favre and de Castro San Román 2001):

M close (x) ! 0 Equation 4-70

Applying a concentrated jacking force at the tip of a curved cantilever normally causes bending in the

member. The shape of the two cantilevered arch halves and the arrangement of loads and reactions, how-

ever, allow the unclosed arch to remain free from moment. Free body diagrams of one-half of the unclosed

arch during the crown jacking procedure are shown in Figure 4-23. The weight of the arch q is assumed to

be uniformly distributed. At the initial stage, the weight of the arch q is carried entirely by the falsework.

As the force in the jacks Fjack is slowly increased, the reacting pressure of the falsework qfalse is decreased.

The line of action of the jacking force is deviated downward by the portion of arch weight that has been

unloaded from falsework, resulting in a state of pure axial compression along the arch. When the jacking

force Fjack reaches a value of qL2/8f, the arch breaks free from the falsework and becomes self-supporting.

q ZLSM^LPNO[<+3 q q q

qL2 qL2 qL2


Fjack  101 Fjack  104 Fjack 
8f 8f 8f
qL2 qL2 qL2 qfalse  0
1 qfalse  109 q 4 qfalse  106 q
qfalse  q 10
8f
10
8f 8f

qL qL qL
R  101 R  104 R
2 2 2

‹JHU[PSL]LYM\SS` ‹QHJRH[ VMMPUHSWYLZZ\YL ‹QHJRH[ VMMPUHSWYLZZ\YL ‹QHJRYLHJOLZMPUHSWYLZZ\YL


Z\WWVY[LKI`MHSZL^VYR ‹MHSZL^VYRMVSSV^ZJHU[PSL]LY ‹MHSZL^VYRJVU[PU\LZ[V ‹HYJOPZUV^Z\WWVY[PUN
\W^HYKHZP[KLJVTWYLZZLZ MVSSV^JHU[PSL]LY\W^HYK P[ZV^U^LPNO[
 

Figure 4-23. Free body diagrams of cantilevered arch during crown jacking

132
4. Statical Analysis of Arch Bridges

The change of system at closure, from statically determinate to statically indeterminate, causes a redistribu-

tion of stresses with time. The sum of moments caused by dead loads and creep will be somewhere

between the moments at closure Mclose, and the moments that would have occurred had the bridge been

built on falsework, and released by the lowering of the falsework Mfalse (Favre and de Castro San Román

2001):

" (t)
M dead (x) + M creep (x,t) = ! D M close (x) + ! K $ ( M false (x) % M close (x))
1 + # (t) $ " (t)
Equation 4-71
" (t)
= +! K $ M false (x)
1 + # (t) $ " (t)

Equation 4-71 shows that although the moments at closure can be virtually eliminated through the use of

hydraulic jacks, indeterminate moments in the final system are inevitably reintroduced by creep. Delaying

the closure of the arch crown after the jacks have been activated has the benefit of decreasing the amount

of creep and shrinkage that remains after the system is made indeterminate.

The horizontal reaction Hdead caused by uniformly distributed dead load qdead is given by:

qdead ! L2 1
H dead = + 18 ! Equation 4-72
f 1+ "

The total long-term horizontal reaction Hlong is given by Equation 4-73. The terms representing the effects

of creep Hcreep and shrinkage Hshrink are negative because these actions reduce the compression in the arch.

H long (t) = H dead + H creep (t) + H shrink (t)


& % ) - ! Eadj (t) ! I sys 1 Equation 4-73
= +H dead ! ( " D # " K$ (t) + ! #" K , sh (t) ! 454
' 1+ % * f2 1+ %

The long-term, first-order crown deflection of arches $long,0 is given by Equation 4-74. Here, elastic, creep,

and shrinkage deformations all cause the crown of the arch to move downward.

133
4. Statical Analysis of Arch Bridges

! long,0 (x = 12 L,t) = "fdead + "fcreep (t) + "fshrink (t)


& H ) & L2 3 ) Equation 4-74
# ( dead (1 + $ (t)) + % sh (t)+ ( 15 +4 f+
' EAcrown *' f *
64

If the crown jacking method is used and the arch camber is equal and opposite to the crown deflection

caused by elastic deformations of the arch under dead load Δfcamb=–Δfdead, then Δfdead can be taken as zero

in Equation 4-74. In this case, only creep deflections Δfcreep and shrinkage deflections Δfshrink need to be con-

sidered in Equation 4-74, since only they cause the arch to move off the pressure line. It is possible to in-

crease the camber of the arch to account for the long-term deflections caused by creep and shrinkage. This

would ensure that the arch maintains some nominal clearance below the crown. The additional camber,

however, will not eliminate the bending moments caused by creep and shrinkage, since they are primarily

a function of the indeterminacy of the arch rather than its shape.

4.5.2 Short-term loads

The short-term moments Mshort that will be considered are those caused by live load Mlive and by uniform

changes in temperature Mtemp:

M short (x) = M live (x) + M temp (x) Equation 4-75

The maximum positive live load moment at midspan can be estimated by applying distributed live load

qlive over the middle third of the span and a concentrated live load Qlive at midspan, as given by:

1 + 10
57
# 1 + 83 #
M live (x = 12 L) = +! L 5
qlive L2 " + !L 3
Qlive L " Equation 4-76
1+ # 1+ #
972 64

The temperature-induced moment at midspan is given by Equation 4-77, where !T is coefficient of thermal

expansion, and ΔT is uniform change in temperature. A drop in temperature results in a positive moment

at midspan.

% EI sys 1
M temp (x = 12 L) = +! K 154 ( "! T # T ) $ $ Equation 4-77
f 1+ &

134
4. Statical Analysis of Arch Bridges

The short-term horizontal reaction Hshort is taken as the sum of horizontal reactions caused by live load

Hlive and temperature change Htemp, as given by Equation 4-78, where a drop in temperature ΔT causes a re-

duction of compressive forces in the arch:

H short = H live + H temp


qlive L2 1 Qlive L 1 & EI sys 1 Equation 4-78
= +! L 47
" + !L 15
" $ ! K ( $! T % T ) " 454
1+ # 1+ # f 2 1+ #
648 64
f f

The total short-term, first-order deflection $short,0 at midspan is given:

# 49 qlive L4 & 1 # 1 Qlive L3 & # H live & # L2 3 &


! short,0 (x = 12 L) = + 492 % 31104 ( + 16 % 192 ( +% + ( )* T + T )( % 15 + 4 f( Equation 4-79
" EI sys ' $ " EI sys ' $ EAcrown '$ f '
64
$

The first and second terms in Equation 4-79 correspond to bending deformations, while the third term

corresponds to axial deformations.

4.5.3 Second-order analysis of combined long and short-term loads

In the calculation of additional moments caused by second-order effects, it is overly conservative to dir-

ectly superimpose long-term and short-term deflections. The method for combining these deflections de-

scribed in the following section is based on a method proposed by Menn (1990) for straight columns.

The second-order, long-term deflection $long can be calculated using Vianello’s method as given by Equa-

tion 4-80, where $long,0 is first-order, long term deflection (Equation 4-74), Hlong is long-term horizontal re-

action (Equation 4-73), and HE,long is long-term buckling resistance (Equation 4-81).

1
! long (x,t) = ! long,0 (x,t) Equation 4-80
H long (t)
1"
H E, long

$ " Eadj (t) " I sys


H E,long (t) = !c " # 2 Equation 4-81
(k )
2
f " kS

135
4. Statical Analysis of Arch Bridges

In Equation 4-81, ,c is resistance factor of concrete, kf is shallowness factor (Equation 4-50), kS is effective

arc length, κ is system stiffness factor, and Eadj is age-adjusted, effective modulus of elasticity. For rein-

forced concrete members, Menn (1990) uses a different effective secant stiffness than shown in the nomin-

ator of Equation 4-81.

For long-term buckling resistance, Menn (1990) uses a long-term effective secant stiffness EI, equal to the

short-term secant stiffness at yield EIy times a creep stiffness factor k,, which accounts for the reduction of

stiffness due to creep:

EI! (t) = EI y " k! (t) Equation 4-82

The creep stiffness factor k% is given by Equation 4-83, where Mu, and ,u, are long-term moment of resist-

ance and long-term curvature at ultimate limit state, and My and ,y are short-term moment resistance and

short-term curvature at yield. These quantities vary with the axial force N in the section, and can be solved

using sectional analysis. Design charts of creep stiffness factor k, for typical circular and rectangular solid

reinforced concrete sections are given by Menn (1990).

M u! !y
k! (N ) = " Equation 4-83
M y !u!

Long-term sectional forces can be calculated using concrete stress-strain diagrams whose strains are

factored by 1++(t),(t). This results in “horizontally-stretched” long-term stress-strain diagrams that are

softer than those for short-term loading. Compared to short-term analysis, using this creep modified

stress-strain curve only slightly reduces the ultimate moment of resistance of the section Mu,≈Mu≈My be-

cause: (1) the assumed compressive strength of the concrete remains the same, and (2) the yield force in re-

inforcing steel is unaffected by creep. This means that the first factor in Equation 4-83 is equal to, or

slightly less than unity. Curvatures at ultimate limit state, however, do increase significantly ,u,>,y, which

in turn results in creep stiffness factors k, that are less than one. Hence, the long-term effective secant stiff-

ness at ultimate EI, is less than short-term secant stiffness at yield EIy.

136
4. Statical Analysis of Arch Bridges

For ultra high-performance fibre-reinforced concrete sections, the effective secant stiffness κ·Eadj·Isys shown

in Equation 4-81 will be used to calculate long-term buckling resistances of arches. Using this effective

stiffness is equivalent to multiplying short-term effective stiffnesses by creep stiffness factors k,, as in

Menn’s method. This is possible to do for ultra high-performance fibre-reinforced concrete members be-

cause: (1) the material is homogenous (i.e. reinforcement is smeared), (2) secant stiffnesses at ultimate limit

state EIu are already being used (as opposed to secant stiffness at yield EIy in reinforced concrete mem-

bers), and (3) using both short-term and long-term stress-strain curves produce the same ultimate mo-

ments of resistance Mu=Mu,.

Long-term deflections $long (Equation 4-80) can be expressed as equivalent short-term deflections $eq,0 us-

ing Equation 4-84 (Menn 1990), where Hlong is long-term horizontal reaction (Equation 4-73) and HE,short is

short-term buckling resistance (Equation 4-85):

# H long (t) &


! eq,0 (x,t) = ! long (x,t) % 1 " Equation 4-84
$ H E,short ('

$ " EI sys
H E,short = !c " # 2 Equation 4-85
(k )
2
f " kS

The total first-order deflection $0 is then the sum of the equivalent short-term deflection $eq,0 (Equation 4-

84) and actual short-term deflection (Equation 4-79):

! 0 (x,t) = ! eq,0 (x,t) + ! short,0 (x) Equation 4-86

The total first-order deflection $0 can be magnified using Vianello’s method to obtain the total second-or-

der deflection $ (Menn 1990):

1
! (x,t) = ! 0 (x,t) Equation 4-87
H long (t) + H short
1"
H E,short

137
4. Statical Analysis of Arch Bridges

As discussed in Section 4.4.5, the eccentricity e between the line of action of force and centroid of critical

sections (x=a) of a member can be approximated by multiplying the second-order deflection $ of the crit-

ical section by eccentricity constant &, as given by Equation 4-52 on page 122. When sectional forces are

greatest at midspan $(x=0.5L), eccentricity constant & is equal to 0.5, which is the same as in the analysis

of fixed-fixed columns (see Figure 4-20 on page 122).

The additional moments caused by geometric nonlinearity are the product of eccentricity e and total hori-

zontal reaction H. Adding this quantity to the long-term and short-term moments results in an expression

for total second-order moments M:

(
M (x,t) = M long (x,t) + M short (x) + e(x = a,t) ! H long (t) + H short ) Equation 4-88

These second-order moments are the moments carried by the flexible system. Once calculated, the distri-

bution of these moments among deck and arch can be determined using Equations 4-58 and 4-59 on page

124.

To evaluate the adequacy of given arch designs, second-order moments March and corresponding axial

forces N(x=0.5L)=H can be plotted on to member capacity N-M* interaction diagrams calculated using the

general analysis method or simplified design method described in Sections 3.2 and 3.5. The same can be

done for trial deck girder cross-section, where axial force N is taken as the long-term effective prestressing

force after all losses.

4.5.4 Maximum sectional forces at quarter-points

Only dead and live loads will be used to estimate maximum sectional forces at the quarter-points because

the effects of restrained deformations are near zero at the quarter-points. This method is more or less the

same as the method presented by Menn (1990), except that his method neglects the effects of arch shallow-

ness (" = 0).

The horizontal reaction H caused by uniformly distributed dead load qdead, distributed live load qlive over

half the span, and concentrated live load Qlive applied at the quarter-point is given by:

138
4. Statical Analysis of Arch Bridges

H = H dead + H live
qdead " L2 1 qlive L2 1 Qlive L 1 Equation 4-89
= +! D 1
" + !L 1
" + !L 135
"
1+ # 1+ # 1+ #
8 16 1024
f f f

The bending moment Mlive at the quarter-point caused by the assumed arrangement of live load is given by:

1 + 373 # 1 + 27
32
#
M live (x = 41 L) = +! L 128
1
qlive L2 " + !L 243
Qlive L " Equation 4-90
1+ # 1+ #
4096

The first-order deflection $0 at the quarter-point caused by these live loads is given by:

qlive L4 Q L3
! 0 (x = 41 L) = +1.63"10 #4 + 6.53"10 #4 live Equation 4-91
$ EI sys $ EI sys

The second-order eccentricity e between the line of action of force and centroid of the quarter-point sec-

tion can be calculated using Equation 4-92, where short-term buckling resistance HE,short is calculated from

Equation 4-85 and eccentricity constant & is 0.733, which is the same as in the analysis of straight fixed-pin

columns (see Figure 4-20 on page 122):

e(x = 41 L) = ! " # (x = 41 L)
1
e(x = 41 L) = 0.733" # 0 (x = 41 L) Equation 4-92
H
1$
H E,short

The total second-order moments M are the sum of first-order moments Mlive and additional secondary mo-

ments caused by the maximum displacement of the arch off the pressure line H·e as given by Equation 4-93

(Menn 1990). These second-order moments are the moments carried by the flexible system, and can be

distributed among the deck and arch using Equations 4-58 and 4-59 on page 124.

M (x) = M live (x) + H ! e(x = 41 L) Equation 4-93

For arches with low span-to-rise ratios L/f, axial forces N at the quarter-points can be significantly greater

than the horizontal reaction H. Axial forces N at the quarter-point can be calculated by:

139
4. Statical Analysis of Arch Bridges

2
!2f $
N (x = L) = H 1 + # &
1
Equation 4-94
4
" L %

To evaluate the adequacy of given arch designs, second-order moments March and corresponding axial

forces N(x=0.25L) can be plotted on to member capacity N-M* interaction diagrams calculated using the

general analysis method or simplified design method described in Sections 3.2 and 3.5. The same can be

done for trial deck cross-section, where axial force N is taken as the long-term effective prestressing force

after all losses.

4.6 Summary of arch analysis

In this chapter, a large amount of technical material has been presented regarding the statical analysis of

concrete arch bridges. The most important topics and insights are summarized below:

• Three-hinged, two-hinged, and fixed arches were compared in terms of: bending moment diagrams

caused by concentrated load, distributed loads, and restrained deformations (Figures 4-6, 4-7, and 4-

10 on pages 93-97), bending moment influence lines and envelopes (Figures 4-8 and 4-9 on pages 95-

96), and primary buckling modes and buckling resistance (Figure 4-3 on page 87).

• Only fixed arch systems will be considered for the development of new concepts in Chapter 6, be-

cause: (1) fixed arches have higher resistances against buckling than hinged arches, which allows for

lighter and more slender members to be used, (2) fixed arches have higher degrees of redundancy

than hinged arches, which is favourable in terms of structural reliability, and (3) fixed arches are more

commonly built than hinged arches, which suggests that they may have some intrinsic economic ad-

vantage over hinges arches.

• A simplified statical system consisting of a deck girder, spandrel columns, and an arch was defined

using the following three assumptions: (1) spandrel columns are pin-connected to the arch and deck,

(2) the flexural rigidity of the arch projected on to the horizontal axis is constant over the entire span

and is equal to the flexural rigidity of the arch at the crown, and (3) vertical deflections of the deck

140
4. Statical Analysis of Arch Bridges

and arch are equal along the span. Assumption 1 is shown to be conservative, based on the results of a

simple parametric study on systems with varying distributions of flexural rigidity among deck, arch,

and spandrel columns (Figure 4-15 on page 106). Assumption 2 simplifies the bending behaviour of

arches under pure antisymmetric loading to that of constant depth beams with pin supports at mid-

span. Assumptions 3 leads to expressions for distributing flexible system moments among arch and

deck:

EI crown EI
M arch (x) = M (x) and M deck (x) = M (x) deck
EI sys EI sys

• The force method was used to solve the three redundant forces in fixed arch systems. The method of

virtual work was used to calculate both bending and axial deformations. Analytical solutions for vari-

ous critical load cases are summarized in Figure 4-16 on page 113. From these results, a non-dimen-

sional ratio ", called the arch-beam parameter was identified as the parameter that governs the trans-

ition of shallow arch behaviour between pure arch action and pure beam action:

45 EI sys
!= "
4 f 2 EAcrown

• By comparing sectional forces in shallow arches with those in a fixed beams, a minimum modified

slenderness ratio (f threshold value of 20 was proposed. Fixed arch systems with (f >20 carry more

than 90% of permanent loads in pure axial compression along the pressure line:

EAcrown 2 f
!f = 2 f =
EI sys rsys

• Modified slenderness ratio (f also affects the buckling behaviour of shallow arches. Shallow arches

buckle in an anti-symmetric, bifurcation mode when (f >19.8, and in a symmetric, snap-through

mode when 9.87 $ (f $ 19.8. The buckling load HE, f for the anti-symmetrical mode is:

141
4. Statical Analysis of Arch Bridges

$ "2 '
H E, f = & 0.6 ± 0.4 1 ! 30.7 2 ) * H E # f > 19.8
% #f (

• Geometric nonlinearity is accounted for by using Vianello’s deflection magnification method. Eccent-

ricities e between the deformed line of action of load and centroid of critical section are calculated

from second-order deflections. Eccentricity constants & are assumed to be the same as those in

straight columns with similar deflected shapes, as shown in Figure 4-20 on page 122.

1
e(x = a) = ! " # 0 (x = a)
H
1$
HE

• Material nonlinearity is accounted for by substituting system flexural rigidity EIsys in linear-elastic

calculations with effective system secant stiffness EI!sys=κ·EIsys. At serviceability limit state, stiffness

factor κ is taken as unity. At ultimate limit state, stiffness factor is a function of axial force and is in

the range: 0.5 $ κ $ 0.75. These values are based on the sectional analysis of a hollow, ultra high-per-

formance fibre-reinforced concrete box, as shown in Figure 3-23 on page 76.

• Flexible system moments are defined as the moments that arise from the global deformations of the

integrated arch system. These moments are distributed to the arch and deck according to their relat-

ive flexural rigidities. Fixed system moments, which are defined as the moments that arise from local

deformations of the deck and arch between spandrels columns, must be added to flexible system mo-

ments to obtain the total bending moment demand.

• Critical load cases causing maximum sectional forces at midspan, at the quarter-points, and over the

springing lines are illustrated in Figure 4-21 on page 127. Step-by-step design calculations for two of

these load combinations are presented in Section 4.5. Calculations are separated into short-term and

long-term loads effects, which are then combined by expressing second-order, long-term deflections

142
4. Statical Analysis of Arch Bridges

as equivalent first-order, short-term deflections. Sample design calculations are provided in

Appendix D.

The simple arch analysis methods presented in this chapter are consistent with the methods presented in

previous works. Based on the author’s review of current literature, there is no single publication that covers

all the material presented in this chapter in a comprehensive way. The significance of this chapter lies in its

utility, consistency, and illustration of fundamental arch behaviour. The concepts and equations presented

in this chapter cover all the major design checks required for the preliminary design of concrete arches. As

such, this chapter provides engineers with a complete and consistent means of learning how to design con-

crete arches. Based on the author’s limited experience, the analysis of arches is not a topic that is com-

monly taught to undergraduate students in civil engineering. Based on what is taught, graduates of civil

engineering would probably approach the design of an arch by creating a numerical model in a frame ana-

lysis program, and then use the resulting force demands to proportion the structure. Because sectional

forces in arches are intrinsically linked to the assumed geometry, distribution of stiffness, and nonlinear

behaviour of arches, a computer model-driven design approach is both inefficient and flawed. The design

of an arch, including the selection of form, layout of geometry, and proportioning of members, should aim

to control the flow of forces, rather than respond to a predefined structural analysis. This chapter, thus,

provides the necessary means for understanding how arches work at an intuitive level, and how they

should be designed.

Before concluding this chapter on the statical analysis of arch bridges, a simple parametric analysis on the

effects of deck-stiffening will be discussed in the next and last section. This analysis attempts to determine

whether or not there are any advantages to using the deck girder to stiffen the arch.

143
4. Statical Analysis of Arch Bridges

4.7 Critique of Billington’s arch stress analysis

In his book Robert Maillart’s Bridges: The Art of Engineering, Billington (1979) discusses the differing views

of the analyst and the designer with regard to the use of deck-stiffening in concrete arch bridges. As an ex-

ample of the viewpoint of the analyst, Billington cites the doctoral work of El-Arousy at the ETH in Zurich

(1942), whose objective was to further improve and simplify the analytical calculation of stiffened arches.

El-Arousy’s work contained two ideas: (1) that generality is the essential basis for application, and (2) that

the structural design of forms is based upon forces predetermined by a prior structural analysis. According

to this viewpoint, the analysis of a structure precedes its design. Billington (1979) contrasted El-Arousy’s

analytical views, which were shared by many experts at the time (and to this day), with those of structural

designer Robert Maillart, whose interest was in the creation of specific bridges rather than the discovery of

general theories of nature. Maillart believed that the dimensions set by the designer controlled the forces,

and that analysis and calculation were subservient to design (Billington 1979). For many of his arch

bridges, Maillart used deck-stiffening to enable him to design very slender, flexible arches. In his design

approach, all nonuniform live loads are carried in bending by the deck girder and all permanent loads are

carried by the arch in pure compression along the pressure line.

To illustrate the validity of Maillart’s design approach to deck-stiffened arches, Billington (1979) presents a

simple analysis of flexural stresses in arch ribs. The objective of the analysis is to determine the optimum

arch-to-deck moment of inertia ratio needed to produce the lowest flexural stresses for a given flexible sys-

tem moment. For the analysis, the moment of inertia of the deck girder is held constant, while the depth of

the arch is increased from zero to infinite. The changing geometry is illustrated in Figure 4-24a. Assuming

linear-elastic behaviour, the flexural stress of an arch σarch caused by the arch portion of the flexible system

moment March is given by Equation 4-95, where y is distance away from the centroid of the section and Iarch

is arch moment of inertia:

M arch " y
! arch = Equation 4-95
I arch

144
4. Statical Analysis of Arch Bridges

Billington's flexural arch stress analysis (1979)


parameters • ddeck is constant
of analysis • darch is increased
• Isys increases without bound
(a)
deck section
changing ddeck
sectional
geometry arch section
darch

b b b

flexural
(b)
stress per !
unit M
bending stress in deck
moment
adapted from
Billington (1979)
stress in arch
2. Broadway
1. Schwand- Bridge (1932)
bach Bridge
(1933)

0
0.001 0.01 0.1 1 10 100
arch-to-deck Iarch
moment of Iarch 1
inertia ratio maximum arch = Ideck
stress occurs when Ideck 2

limitations arch needs some volume of arch is


minimum flexural rigidity unrealistically large
to resist buckling between
spandrel columns

Figure 4-24. Influence of deck-stiffening on flexural stresses in the arch and deck

Figure 4-25. Schwandbach Bridge and Broadway Bridge. Adapted from Bill and Maillart (1955) and Ostrander and
Oliver (1987). Elevation views are drawn at 1:1000 scale and sections views are drawn at 1:500 scale.

145
4. Statical Analysis of Arch Bridges

Assuming a rectangular arch section with width b and depth darch, distances to extreme fibres ytop and ybot,

and moment of inertia Iarch are given by:

darch
ytop = ybot = Equation 4-96
2

3
bd arch
I arch = Equation 4-97
12

Substituting Equations 4-20 (page 103), 4-96 and 4-97 into 4-95, and simplifying gives:

6darch
! arch = Equation 4-98
bd + 12I deck
3
arch

The derivative of arch flexural stress (/arch)! with respect to darch is given by:

" 6 ( bd arch
3
+ 12I deck ) # 6darch $ 3bd 2arch
( arch )
! = Equation 4-99
(bd arch + 12I deck )
3 2

Setting (/arch)! to zero and rearranging gives the following critical point (Billington 1979):

I arch 1
= Equation 4-100
I deck 2

The critical point is labelled in Figure 4-24b, corresponding to maximum flexural stresses in arches caused

by given flexible system moments M. The vertical axis measures flexural stress and increases upwards from

zero at the origin. The scale of the axis is unspecified because only the shapes of the stress curves is relev-

ant to this analysis. The horizontal axis measures arch-to-deck moment of inertia ratio Iarch/Ideck, which in-

creases logarithmically to the right.

Billington (1979) observed that the change in arch stress (solid curve in Figure 4-24b) with respect to the

ratio of arch-to-deck moment of inertia depends on the initial starting form of the design. If a flexible arch

(Iarch/Ideck < 0.5) is envisioned by the designer, then flexural stresses in the arch are reduced by decreasing the

146
4. Statical Analysis of Arch Bridges

depth of the arch. This is the case for Maillart’s deck-stiffened Schwandbach Bridge, built in Switzerland in

1933 (see Figure 4-25). If a stiff arch (Iarch/Ideck > 0.5) is envisioned by the designer, then flexural stresses in

the arch are reduced by increasing the depth of the arch. This is the case for the Broadway Bridge built in

Saskatchewan in 1932 (Figure 4-25). There are practical limitations on both the minimum and maximum

depth of arches. Minimum depths are governed by bucking of the arch between spandrel columns. Max-

imum depths are governed by excessively large arch volumes.

Although Billington’s (1979) analysis gives some validity to the deck-stiffened arch form, it is problematic

in several ways. The first problem is in his visual display of results, as shown in Figure 4-26. Although he

does not explicitly state that deck-stiffened arches are inherently more efficient than partially deck-sti-

ffened or self-stiffened arches, it is not difficult for readers to erroneously draw this conclusion due to the

way the results are presented. By omitting stress values between 0 and 6 kg/cm2, Billington’s graph overem-

phasizes the differences in arch stress between Maillart’s Schwandbach Bridge and “typical American

designs of of 1931” (the Canadian Broadway Bridge is substituted here as an example). Also, the horizontal

axis is cut-off prematurely at a arch-to-deck moment of inertia ratio of 1, which dismisses the self-stiffened

arch form as a viable solution. These concerns of visual presentation are addressed in Figure 4-24b by

changing the ranges of the axes.

Arch stress 9.5


in kg/cm2
9.0 typical American
design of 1931

8.5

8.0

7.5

7.0

6.5
Schwandbach
Bridge design
6.0
0.06 0.20 0.50 1.00
Stiffness ratio Iarch / Ideck

Figure 4-26. Billington’s graph of arch stress versus stiffness ratio between arch and deck.
Adapted from Billington (1979).

147
4. Statical Analysis of Arch Bridges

The second problem with Billington’s analysis is that the minimization of arch stress is assumed to be a val-

id measure of efficiency. Billington’s analysis presupposes that: (1) the minimization of arch stress will res-

ult in the direct minimization of arch weight, and that (2) the minimization of arch weight is favoured over

the minimization of deck weight. The first assumption is only valid when only solid sections are con-

sidered. For solid sections, the only way to increase flexural strength (i.e. reduce bending stress) is to make

arches deeper and hence to add weight. In this case, the increase in weight is equal to the change in depth

times the total width of the section. If more efficient hollow box sections are considered, then it is possible

to increase flexural strength (i.e. reduce bending stress) by adding much less weight. In this case, the in-

crease in weight is equal to the change in depth times the sum of web thicknesses.

Regarding the second assumption, the minimization of arch weight is most critical for cast-in-place on

falsework methods of construction: the lighter the arch, the lighter the required falsework. This method re-

quires that falsework must be designed to carry 100% of the weight of the forms and concrete before the

arch cures and starts to carry its own weight. Hence, there are clear economic incentives for reducing the

weight of the arch. Maillart was aware of these incentives, and designed many of his arches with the intent

of reducing load demands on falsework (Billington 1979). If modern precast, cantilever methods of con-

struction are considered, these economic incentives of minimizing arch weight (as opposed to the total

weight of structure) are less apparent. Precasting allows for stresses to be applied to the arch as soon as

they are erected, reducing the demands on temporary works relative to cast-in-place on falsework. Minim-

izing the flexural rigidity of the arch reduces its resistance against buckling during construction, which

might increase demands on temporary works. Hence, the optimal distribution of flexural stresses, weight,

and stiffness among the deck and arch cannot be determined through the optimization of any one single

parameter, such as arch weight.

Billington’s analysis can be extended by calculating stresses in the deck. If the deck girder is assumed to be

rectangular and to have the same width as the arch, flexural stresses in the deck can be calculated expli-

citly. Deck stresses are shown in Figure 4-24b as a dashed curve. This curve gives a new perspective on the

148
4. Statical Analysis of Arch Bridges

efficiency of the system, since deck stresses become larger as the moment of inertia of the arch is reduced.

Thus, there is a penalty associated with reducing stresses in the arch through the use of deck-stiffening.

These increases in flexural stress in the deck girder were acknowledged by contemporaries of Maillart,

such as Ketchum (1934). Maillart’s contemporaries saw these relatively high deck stresses as being an in-

herent inefficiency of the deck-stiffened arch form. In doing so, they failed to see the potential gain in con-

struction efficiency that Maillart saw in minimizing the weight of the arch.

The third problem with Billington’s analysis is that the effects of dead load are not considered. Dead loads

are designed to be carried in pure compression along the arch, causing uniform compressive stresses.

These stresses act as a form of effective prestress for the arch, which in turn increases the flexural capacity

of the arch. Flexural capacity is increased because precompression stresses in the arch delay the onset of

tensile stresses along the flexural tension face of the arch. Thus, the optimization of system geometry is

achieved not by minimizing arch stress, but by maximizing the use of the decompression force to carry

flexural stresses. The same can be done for the deck girder using prestressing steel. The amount of

prestressing steel needed to offset flexural stresses in the arch, however, may in some cases be overly large

and impractical. Deck prestressing requires additional materials, detailing, and labour. The natural

prestress in the arch, on the other hand, is obtained essentially for free. For these reasons, it is more effi-

cient to carry flexible system moments in arches rather than in deck girders.

This design strategy of using the arch prestress and prestressing the deck girder can be observed in some of

Menn’s Swiss concrete arch bridges built in the 1960s. Although he was aware of Maillart’s innovative and

economical deck-stiffened arch designs, Menn moved in a different direction, proportioning some of his

arches to be partially deck-stiffened, rather than fully deck-stiffened. For the design of the Nanin and

Rhine Bridges (see Figure 4-27), Menn chose to prestress the deck girder and proportioned the arch to

carry roughly half of the flexible system bending moments.

149
4. Statical Analysis of Arch Bridges

Figure 4-27. Swiss concrete arches designed by Menn. Adapted from Vogel and Marti (1997).
Elevation views are drawn at 1:2000 scale and sections views are drawn at 1:400 scale.

As discussed above, Billington’s arch stress analysis has limited applicability because its conclusions draw

upon assumptions that are not always valid. To broaden its applicability, a modified version of the analysis,

shown in Figure 4-28, is proposed. First, the variation in assumed geometry is modified (see Figure 4-28a).

Hollow box sections are chosen to better reflect the type of sections considered in modern concrete arch

systems. Arch depth is increased as deck depth is decreased, such that (1) the sum of cross-sectional areas

of arch and deck remains constant, and (2) the widths and thicknesses of both top and bottom slabs of arch

and deck remain the same. Effectively, the webs of the arch and deck sections are linearly exchanged

between the two sections. This improves the comparison between self-stiffened and deck-stiffened arch be-

cause the total volume of concrete is kept constant. In Billington’s analysis, the total volume and weight of

the system were rapidly increasing as the arch was made deeper. The second modification is that the hori-

zontal axis is expressed as the ratio of arch moment of inertia to system moment of inertia with a theoret-

ical range between 0 and 1. These lower boundary does not actually exist, since it implies that the arch has

zero stiffness. The upper boundary can exist if the deck girder is simply supported between spandrel

columns. In this case, the arch is completely unstiffened by the arch.

150
4. Statical Analysis of Arch Bridges

Modified stress analysis


parameters • ddeck is decreased
of analysis • darch is increased
• Aarch + Adeck remains constant
(a)
deck section
changing
ddeck
sectional
geometry
arch section
darch

b b b

flexural
(b)
!
stress per
unit M
bending
moment
stress in deck stress in arch
adapted from
Billington (1979)

0
0 0.2 0.4 0.6 0.8 1
arch-to-system
moment of Iarch
inertia ratio Isys

Figure 4-28. Modified flexural stress analysis

The modified stress analysis results show that arch stresses are minimum when Iarch/Isys & 0 and are maxim-

um when Iarch/Isys & 0.85. Deck stresses are minimum when Iarch/Isys & 1 and are maximum when Iarch/Isys &

0.15. These results still do not provide any definitive conclusions regarding the optimal distribution of

flexural rigidity among the arch and deck. Some broad conclusions can be drawn if the relative magnitudes

of axial stress and flexural stress are known:

• If compressive stresses in the arch caused by permanent load (i.e. arch prestress) are much higher

than flexural tensile stresses caused by maximum bending moments, then it is probably best to move

towards self-stiffened arch systems. This is because: (1) arch flexural stresses will be offset by precom-

pression in the arch, and (2) flexural stresses in the deck will be minimized. The penalty of increased

151
4. Statical Analysis of Arch Bridges

arch weight during construction is much less pronounced since efficient hollow, box sections are

used instead of solid, rectangular sections.

• If compressive stresses in the arch caused by permanent load are much lower than flexural tensile

stresses caused by bending moments, then it is probably best to move towards deck-stiffened arch

systems. Because of the low prestress, the flexural capacity of the arch will be low. Thus, it is advant-

ageous to protect the arch by distributing most of the flexible system moments to the deck. For the

deck, flexural capacity can be increased as needed by the use of prestressing steel.

152
Chapter 5. Comparative Study of 58 Concrete Arch Bridges

This chapter describes a comparative study of fifty-eight concrete arch bridges. These bridges were com-

piled to characterize a baseline of existing concrete arch technology. These bridges represent designs that

have been feasibly built using conventional reinforced concrete as the primary building material. These

bridges respond to a variety of geometrical site requirements, and represent design solutions that satisfy

the serviceability, strength, and stability requirements imposed on highway traffic bridges. Trends of sec-

tional properties and global geometries among bridges in the study will be compared with new concepts

for arch bridges in Chapter 6.

5.1 Objective and utility

The primary objectives of this study are (1) to describe and summarize the state-of-the-art of concrete arch

bridge technology in the form of a consistent and comprehensive database, and (2) to identify and ration-

alize trends among the geometries of the concrete arch bridges being studied.

The compiled database describes each concrete arch bridge graphically, quantitatively, and qualitatively.

These descriptions along with trends observed from them can be used as references and design tools for

designers. This database can also serve as a tool for evaluating structural efficiency and feasibility. New

design concepts for arch bridges, such as those presented in Chapter 6, can be compared to bridges in the

database to demonstrate their innovation, or lack thereof.

153
5. Comparative Study of 58 Concrete Arch Bridges

Databases of concrete arch bridges in current literature have limited utility for designers because of their

lack of information on the sectional properties and geometric ratios of each bridge. Most publications re-

viewed on this topic were concerned with record span lengths of arches and innovative construction tech-

niques. For example, Chen (2007) discusses concrete arch bridges built in China and provides a table with

the names, years of completion, spans, structural types, and construction methods of thirteen arch bridges.

Ewert (1999) provides a table with the names, countries, years of completion, spans, rises, arch depths, and

methods of construction of twenty-eight concrete arch bridges with spans over 200 metres. Further, di-

mensioned drawings of selected arch cross-sections accompany short descriptions of each bridge. Šavor

and Bleiziffer (2008) discuss long span arch bridges in Europe and provide a table with the names, coun-

tries, years of completion, spans, deck widths, and structural types of thirty-nine of the spanning arch

bridges in the world. Mondorf (2006) describes the construction of specific concrete arch bridges and

provides dimensioned drawings for many of them. Selected data from these sources are compiled in Table

5-1.

The information presented in Table 5-1 does not inform the reader on how concrete bridges are typically

proportioned, or on their structurally efficiency relative to one another. These kinds of comparative analys-

is require a rigorous database that considers the sectional properties and geometric ratios of each bridge.

Such a database is developed in the next section.

154
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-1. Concrete arch bridge data compiled from reference databases
bridge country year span bridge country year span
in m in m

• Wanxian Yangtze China 1997 420 Kylital Germany 1998 223


• Krk-I Croatia 1980 390 Xuguo China 2001 220
Jiangjiehe China 1995 330 Kashirajima Japan 2003 218
• Colorado River USA 2010 323 Tercer Milenio Spain 2008 216
Yongjiang China 1996 312 Novi Sad Yugoslavia 1961 211

• Gladesville Australia 1964 305 Žeželj Serbia 1961 211


• Rio Parana Paraguay 1965 290 Bregenzerach Austria 1967 210
Chishi Datong China 1997 280 • Galena Creek USA 2011 210
• Infant Henrique Portugal 2003 280 Lingenau Austria 1967 210
• Bloukrans S. Africa 1983 272 Rio Esla Spain 1942 210

Arrábida Portugal 1963 270 Xingduicha China 2007 205


Sanan China 1998 270 • Krka-Scradin Croatia 2004 204
• Sandö Sweden 1943 264 Usagawa Japan 1983 204
Fujikawa Japan 2003 265 Morbihan France 1996 201
Châteaubriand France 1991 260 Fuling China 1989 200

• Tensho Japan 2000 260 Ikeda H. Ohashi Japan 2001 200


Los Tilos Spain 2004 255 Roche Bernard France 1996 200
• Wilde Gera Germany 2000 252 • Maslenica Croatia 1997 200
Svinesund II Norway 2005 247 Pfaffengerg Austria 1969 200
• Sibenik Croatia 1966 246 Van Stadens S. Africa 1971 200

Barelang V Indonesia 1998 245 Zadar Croatia 1997 200


• Krk-II Croatia 1980 244 Weiping China 2001 198
Xiaonanmen China 1990 240 Liuguihe China 2005 195
Beppu-Myouban Japan 1989 235 • Pag Croatia 1968 193
Myoban Japan 1989 235 Huapichong China 1999 180

Fiumarella Italy 1961 231 • Modong China 1999 180


Zaparoze Ukraine 1952 228 • Nosslach Austria 1968 180
Rio Zezere Portugal 1993 224 Shatuo China 2002 180
Kylital Germany 1998 223 Yanxi China 2003 180

• bullets mark bridges that are part of the database that is described in Section 5.2

155
5. Comparative Study of 58 Concrete Arch Bridges

5.2 Database of concrete arch bridges

This section describes how details and properties of concrete arch bridges from references were processed

for use in the database. The main sources of information were journal articles, conference proceedings,

construction drawings, and textbooks. The name, location, year of completion, key dimensions, construc-

tion method, and drawings were collected for each bridge accepted into the database. A succinct integrated

version of the database, from which all figures in this chapter are adapted from, is provided in Appendix A.

5.2.1 Sources of information

Technical engineering journals report on new and innovative developments in the field, and are con-

sidered primary sources of information for the profession because of the high standards of peer review and

evaluation. In order to produce a database of arch bridges that is accurate, reliable, and representative, a

comprehensive search of authoritative English and German structural engineering journals was carried

out. Table 5-2 lists the journals and volumes that were searched and identifies articles that were found to

have information on the design and construction of concrete arch bridges. In addition to keyword searches

of journal indexes, the table of contents of each issue was inspected. Because the author and assistant were

not fluent in German, each German journal volume was visually skimmed through in search of drawings

of concrete arch bridges.

A second approach used for finding information on concrete arch bridge references involved inspecting

reference lists of journal articles, conference papers, and textbooks, especially those mentioned in Section

5.1. These lists often included references to old journal articles, which were in some cases retrievable. An-

other major source of information was the proceedings from several International Conferences on Arch

Bridges, which contained many papers dedicated to recently completed projects around the world. Several

construction drawings of Canadian and American concrete arch bridges were obtained from contacts of

the author.

156
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-2. Journal article search and results

journal language dates searched articles found


Author year: volume(issue)

ASCE. Journal of Bridge Engineering English 1996-2009 Baxter and Balan 2008: 13(5)
Liu et al. 2002: 7(1)
ASCE. Journal of Structural Engineering English 1983-2009 Laffranchi and Marti 1997: 123(10)
ASCE. Journal of Structural Design and Construction English 1996-2009 none
CSCE. Canadian Journal of Civil Engineering English 1974-2009 Ostrander and Oliver 1987: 14(4)
Proceedings of the ice. Bridge Engineering English 2003-2009 Radic et al. 2006: 159(3)
Structural Engineering International English 1991-2009 Tanner and Bellod 2005: 15(3)
Fonseca 2005: 15(2)
Blinkov et al. 2001: 11(3)
Wang and Au 2001: 11(2)
Cheng 1994: 4(4)
Der Bauingenieur German 1985-2009 Hünlein and Ruse 1985: 60(8)
Bautechnik German 1985-1996, 2004-2009 Reintjes 2005: 82(11)
Beton- und Stahlbetonbau German 1985-2009 Freytag et al. 2009: 104(3)
Viet Tue et al. 2005: 100(11)
Eilzer et al. 2005: 100(3)
Zimmermann 2004: 99(4)
Werschnick 2000: 95(2)
von Wölfel 1999: 94(12)
Ewert 1999: 94(9)
Dajun and Weiqing 1999: 94(4)
Standfuß 1990: 85(5)
Schweizer Ingenieur und Architekt German 1985-1996 Koppel and Walser 1991: 109(11)
Herzog 1990: 108(26)

5.2.2 Acceptance criteria

Bridges were accepted into the database based on their structural system and the amount of information

available on them. The following acceptance criteria were used:

• bridge uses reinforced concrete as the primary building material (composite decks also accepted),

• bridge was built in the 20th or 21st century,

• bridge was designed for highway traffic loads,

• bridge deck is supported by the arch from below (half-through and tied arches not accepted), and

• references had scaled drawings with dimensions (elevation and cross-section views).

157
5. Comparative Study of 58 Concrete Arch Bridges

Many of the longest concrete arch bridges in the world have been included in the database. According to

Table 5-1, eighteen of the longest spanning concrete arch bridges in the world have been included in the

database. Each included bridge has been marked with a bullet in the table.

Figure 5-1 shows the distribution of span lengths of bridges accepted into the database. Each bridge is rep-

resented by one mark and is assigned a bridge id (1 being the longest spanning arch and 58 being the

shortest spanning arch in the database). Having reference bridges from all span ranges is important be-

cause observations in Section 5.3 will be made primarily as functions of span length.

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

Figure 5-1. Distribution of span lengths of bridges in database

5.2.3 Drawing database

Drawings of bridges in the database are shown in Figure 5-2, ordered by length of main span. These draw-

ings were traced from their original sources in accordance with all specified dimensions. Elevation views

are shown at 1:15000 scale, while section views are shown at 1:1500 scale (see Appendix A to view these

drawings at larger scale). Bridges are identified by name, year of completion, and bridge id number. Arch

section views are cut perpendicular to the arch axis, and are shown with dimensions averaged from sec-

tions at the crown and springing line. Section views show outlines of spandrel columns along with single

column sections shown on the right. In general, all variable dimensions of sections are drawn using av-

erage values. In some cases, dimensions of spandrel columns were inferred from elevation and section

views. Unless otherwise specified in references, spandrel columns with thicknesses of 1.5 m or greater were

assumed to be hollow with wall thicknesses of 305 mm. Original sources of information for each bridge in

the database are listed in Figure 5-3 in order of bridge id.

158
5. Comparative Study of 58 Concrete Arch Bridges

Bridge drawings bridge ID


Elevations at 1:10000 scale name
Sections at 1:1000 scale year of completion

typical section
column section

Figure 5-2. Drawings of bridges in database.


Scale is 2:3 of original. Original scale version shown in Appendix A.

159
5. Comparative Study of 58 Concrete Arch Bridges

1.! Dajun, D., and L. Weiqing. 1999. Neue Entwicklungen Bei Hochhäusern Und Groβen Brücken Aus Beton in China. Beton- und Stahlbetonbau 94(4): 178-85.
2/11.!Radić, J., J. Bleiziffer, and D. Tkalčić. 2005. Maintaining Safety and Serviceability of Concrete Bridges in Croatia. Bridge Structures 1(3): 327-44.
3.! US Dept. of Transportation FHWA. 2003. AZ-NV HPP 93(3) Hoover Dam Bypass Colorado River Bridge: 60% Submittal. April 4.
4.! Baxter, J. W., A. F. Gee, and H. B. James. 1965. Gladesville Bridge. Proceedings of the Institute of Civil Engineers.! !
5.! Stellmann, W. L. O. 1966. Brücke Über Den Rio Paraná in Foz Do Iguaçú, Brasilien. Beton- und Stahlbetonbau 61(6): 145-49.
6.! Fonseca, A., and F. Mato. 2005. Infant Henrique Bridge Over the River Douro, Porto. Struct. Eng. Int. 15(2): 85-87.
7.! Liebenberg, A.C., and M.G. Latimer. 2001. Bloukrans Bridge. Arch'01: troisième conférence internationale sur les ponts en arc, Sept. 19-21, Paris.
8.! Fritzell, G. 1960. Deflection Measurements on the Sandö Bridge 1942-1958. Sixth Congress of the IABSE.
9.! Kamimura, K., M. Kouno, K. Okada, and T. Tsuka. 2001. Design and Construction of the Tensho Concrete Arch Bridge. Arch'01: Third Int'l Conf. on Arch Bridges, Paris.
10.! Radić, J., Z. Šavor, J. Bleiziffer, and I. Kalafatić. 2008. Ŝibenik Bridge - Design, Construction and Assessment of Present Condition. Chinese-Croatian Joint Colloquium:
Long Arch Bridges, July 10-14, Brijuni Islands, China.
12.! von Wölfel, R. 1999. Die Talbrücke Über Die Wilde Gera. Beton- und Stahlbetonbau 94: 546-51.
13.! State of Nevada Department of Transportation. 2000. B-1948-N/S Galena Creek Bridge. April 19.
14.! Šavor, Z., N. Mujkanović, and G. Hrelja. 2008. Design and Construction of Krka River Arch Bridge. Chinese-Croatian Joint Colloquium Long Arch Bridges, Brijuni Islands.
15.! Radić, J., Z. Šavor, V. Prpić, M. Friedl, and Ž. Žderic. 2008. Design and Construction of the Maslenica Highway Bridge. Chinese-Croatian Joint Colloquium, Brijuni Islands.
16.! Šavor, Z., N. Mujkanović, G. Hrelja, and J. Bleiziffer. 2008. Reconstruction of the Pag Bridge. Chinese-Croatian Joint Colloquium: Long Arch Bridges, Brijuni Islands.
17.! Wang, J. J., and F. T. K. Au. 2001. Modong Hongshui River Bridge, China. Structural Engineering International 11(2): 101-03.
18.! Aigner, F. 1968. Stahlbeton Bogenbrücken Auf Der Österreichischen Brenner Autobahn. Der Bauingenieur.
19.! Igarashi. 1976. Die Hokowazu-Brücke in Japan, Ein Stahlbetonbogen Im Freien Vorbau. Der Bauingenieur (8).
20.! Anonymous. 1978. Communications Techniques Belges 8. Annales des Travaux Publics de Belgique, No. 1-2.
21.! Cheng, K. M. 1994. The Pitan Bridge, Taiwan. Structural Engineering International 4(4): 231-34.
22.! Standfuβ, F. 1990. Nationalbericht: Brücken in Der Bundesrepublik Deutschland. Beton- und Stahlbetonbau (5): 106-13.
23.! Wöβner, K., H. Gebhardt, R. Schnabel, and H. Wörner. 1979. Die Talbrücke Rottweil-Neckarburg Im Zuge Der A81. Beton- und Stahlbetonbau (10-11).
24.! Bouchet. 1964. L'Echangeur La Araña Du Croisement Des Autoroutes Nord-Sud Et Est-Ouest À Caracas (Venezuela). La Technique des Travaux (Belgique).
25.! Hünleim, W., and P. Ruse. 1985. Ein Neues Verfahren Für Den Bau Von Bogenbrücken Dargestelit Am Bau Der Argentobelbrücke Würgau. Der Bauingenieur (12).
26.! Žderic, Ž., A. Runjić, and G. Hrelja. 2008. Design and Construction of Cetina Arch Bridge. Chinese-Croatian Joint Colloquium on Long Arch Bridges, Brijuni Islands.
27.! Koppel, A. J., and R. Walser. 1991. Hundwilertobelbrucke: Ein Bemerkenswerter Neubau. Schweizer Ingenieur und Architekt 109(11): 250-56.
28.! Taiwan Area National Expressway Engineering Bureau, MOTC. 1999. (In Mandarin) The Second Freeway: An Anthology of Particular Bridges.
29.! BC Ministry of Transportation and Highways. 1995. R6-V433-3051 Big Qualicum River Upstream Bridge No. 3051 (Drawings).
30.! Goodyear, D., N. Smit, J. Heacock, and S. Starkey. 2001. Respect for Tradition: The New Crooked River Gorge Bridge. Arch'01: Third Int'l Conf. on Arch Bridges, Paris.
31.! Podolny, W., and J. M. Muller. 1982. Construction and Design of Prestressed Concrete Segmental Bridges. New York: Wiley.
32.! Inaudi, D., A. Rufenacht, B. Arx, H.P. Noher, and S. 2002. Monitoring of a Concrete Arch Bridge During Construction. Proc. SPIE International Society for Optical Eng.
33.! Vogel, T, and P. Marti. 1997. Christian Menn, Brückenbauer. Basel, Switzerland: Birkhäuser Verlag.
34.! State of California Department of Public Works Division of Highways. 1932. Bridge Across Bixby Creek (as Built Plans) Doc. No. 50001001.
35.! Vogel, T, and P. Marti. 1997. Christian Menn, Brückenbauer. Basel, Switzerland: Birkhäuser Verlag.
36.! Eilzer, W., W. Schmidtmann, and R. Jung. 2005. Die Wirrbachtalbrücke. Beton- und Stahlbetonbau 100(3): 236-40.
37/55/57/58. Bill, M., and R. Maillart. 1955. Robert Maillart. Translated by W. P. M. K. Clay. Zürich, Switzerland: Girsberger.
38/50/56. Sinotech Engineering Consultants, Ltd. Personal correspondence, August 2007.
39.! Mörsch, E. 1909. Die Gmündertobel-Brücke Bei Teufen Im Kanton Appenzeil. Schweizerische Bauzeitung 53 (7): 81.
40.! Schwaab, E., and A. Gattner. 1953. Straβenbrücke Über Das Tiefe Tal Bei Rosshaupten, Allgau. Beton- und Stahlbetonbau (11).
41.! Tonello, J., and P. Dal-Palu. 2001. Le Pont Du Triple Saut (the Triple Jump Bridge). Arch'01: troisième conférence internationale sur les ponts en arc, Sept. 19-21, Paris.
42.! State of California Department of Public Works Division of Highways. 1996. Bridge Across Russian Gulch, Earthquake Retrofit Project No. 714.0.
43.! Petrangeli, M.P. 2007. Prefabrication of Medium Span Arch Bridges. Arch'07: 5th Intʼl Conference on Arch Bridges, Sept. 12-14, at Madeira, Portugal.
44.! Zimmermann, W., and N. im Gailtal. 1999. Der Bau Der Stampfgrabenbrücke. Beton- und Stahlbetonbau 99(4): 304-10.
45.! State of California Department of Public Works Division of Highways. 1932. Bridge Across Rocky Creek (as Built Plans) Doc. No. 50001081.
46.! Baxter, D. J., and T. A. Balan. 2008. Design of the Fulton Road Bridge Precast Segmental Concrete Arches. Journal of Bridge Engineering: 476-82.
47.! Blinkov, L. S., E. Cosolo, and S. N. Valiev. 2001. Rehabilitation of a Bridge Over the Matsesta River, Russian Fed. Structural Engineering International 11(3): 181-83.
48.! Ostrander, J. R., and D. C. Oliver. 1987. Construction of the Broadway Bridge At Saskatoon, Saskatchewan, in 1932. Canadian Journal of Civil Engineering 14: 429-38.
49.! State of California Department of Public Works Division of Highways. 1937. Bridge Across Big Creek (as Built Plans) Doc. No, 50001093.
51.! Viet Tue, N., F. Dehn, T. Schliemann, K. H. Reintjes, and F. Tauscher. 2005. Anwendung Von Hochleistungsbetonen Bei Der Bogenbrücke Wölkau Im Zuge Der Bab A17.
Beton- und Stahlbetonbau 100(11): 931-38.
52.! Tanner, P., and J. L. Bellod. 2005. Widening of the Elche De La Sierra Arch Bridge, Spain. Structural Engineering International (3): 148-50.
53.! Tandon, M. 1995. Arch Bridge At Dodan Nallah. Proceedings of the First International Conference on Arch Bridges, Bolton, UK.

Figure 5-3. List of sources ordered by bridge ID

5.2.4 Recorded data

Table 5-3 shows the year of completion, location, construction method, and global geometry of bridges in

the database. Geographical locations, expressed as latitudes and longitudes, were found by searching satell-

ite images online using Google Maps (2009). Measurements of span, rise, height over arch crown, and total

spandrel column length measurements were read from drawing dimensions and textual descriptions, or

scaled from drawings. The height over arch crown is measured from mid-depth of arch crown to the top of

roadway. Total spandrel column length is the sum of all column lengths between springing lines, excluding

those directly above the springing lines. Arch fixity refers to the number of hinges present in the structural

system.

160
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-3. Location and global geometry of arch bridges in database


total spandrel column length
height over arch crown

 
interior spans
construction method
span rise
ID bridge   latitude longitude m m # m m
1 Wanxian-Yangtze 1997 China N 30° 45' 27" E 108° 25' 28" 425.0 85.0  14  11.8 355
2 Krk I 1980 Croatia N 45° 14' 51" E 014° 34' 15" 390.0 60.0  14  4.5 148
3 Colorado 2010 United States N 36° 00' 45" W 114° 44' 29" 323.0 84.4  9  8.3 202
4 Gladesville 1964 Austrailia S 33° 50' 32" E 151° 08' 53" 309.0 39.0 s 10  2.1 36
5 Rio Parana 1965 Brazil S 25° 30' 34" W 054° 36' 04" 290.0 53.0 s 13  0.0 137
6 Infant Henrique 2003 Portugal N 41° 80' 29" W 008° 36' 06" 280.0 25.0  8  3.0 26
7 Bloukrans 1983 South Africa S 33° 58' 00" E 023° 38' 44" 272.0 62.0  15  5.7 289
8 Sandö 1943 Sweden N 62° 53' 00" E 017° 52' 40" 264.0 40.0 t 22  1.4 193
9 Tensho 2000 Japan N 31° 54' ??" E 131° 25' ??" 260.0 32.5  14  2.2 105
10 Sibenik 1966 Croatia N 43° 45' 46" E 015° 50' 56" 246.0 31.0  11  2.9 81
11 Krk II 1980 Croatia N 45° 14' 38" E 014° 33' 55" 244.0 47.0  8  4.2 85
12 Wilde Gera 2000 Germany N 50° 42' 54" E 010° 47' 16" 242.0 68.0  6  6.1 75
13 Galena Creek 2011 United States N 39° 20' 55" W 119° 47' 32" 210.0 42.5 s 3  3.0 G
14 Krka-Skradin 2004 Croatia N 43° 48' 31" E 015° 54' 57" 204.0 52.0  7  5.5 89
15 Maslenica 1997 Croatia N 44° 14' 12" E 015° 31' 20" 200.0 65.0  7  9.1 139
16 Pag 1968 Croatia N 44° 19' 29" E 015° 15' 31" 193.0 28.0  9  0.0 57
17 Modong Hongshui 1999 China N 23° 44' 15" E 109° 11' 54" 180.0 30.0  14  0.0 85
18 Nosslach 1968 Austria N 47° 04' 03" E 011° 28' 52" 180.0 45.0 s 12  4.0 154
19 Hokowazu 1974 Japan N 33° 30' 55" E 129° 50' 53" 170.0 26.5  11  2.4 80
20 Houffalize 1979 Belgium N 50° 08' 40" E 005° 47' 03" 162.0 32.4 s 6  5.6 53
21 Pitan 1993 Taiwan N 24° 57' 31" E 121° 32' 09" 160.0 21.7 s 3  0.0 G
22 Wertachtal ? Germany N 47° 38' 29" E 010° 30' 09" 156.0 34.0  11  1.6 123
23 Neckarburg 1978 Germany N 48° 12' 08" E 008° 36' 47" 154.0 49.5  7  17.0 151
24 Caracas 1953 Venezuela N 10° 33' 11" W 067° 00' 17" 151.8 32.0  12  1.6 66
25 Argentobel 1986 Germany N 47° 38' 30" E 010° 01' 40" 150.0 29.0  10  4.6 85
26 Cetina 2007 Croatia N 43° 37' 02" E 016° 43' 45" 140.3 20.6  7  2.9 38
27 Hundwiler 1991 Switzerland N 47° 20' 11" E 009° 04' 50" 138.4 35.8  13  2.1 122
28 Maling 1999 Taiwan N 25° 06' 42" E 121° 41' 11" 138.0 33.0  7  1.4 50
29 Big Qualicum 1996 Canada N 49° 20' 01" W 124° 32' 07" 132.0 24.0 s 6  1.4 45
30 Crooked River 2000 United States N 44° 23' 35" W 121° 11' 35" 125.0 24.0  8  2.0 42
31 Niesenback ? Germany ? ? 120.0 37.5  6  5.1 53
32 Siggenthal 2000 Switzerland N 47° 28' 18" E 008° 18' 40" 117.0 24.4 s 7  1.2 27
33 Nanin 1967 Switzerland N 46° 24' 45" E 009° 13' 48" 112.0 24.5 t 8  0.0 28
34 Bixby Creek 1932 United States N 36° 22' 17" W 121° 54' 06" 100.6 36.6  11  3.8 111
35 Rhein 1963 Switzerland N 46° 49' 43" E 009° 24' 55" 100.0 20.9 ? 8  0.0 24
36 Wirrbachtal 2002 Germany N 50° 43' 29" E 010° 50' 10" 100.0 25.3 s 7  1.6 40
37 Salginatobel 1930 Switzerland N 46° 58' 54" E 009° 43' 05" 90.0 13.0 t G  0.0 20
38 Wunshuei ? Taiwan N 24° 27' 37" E 120° 52' 34" 85.0 13.5 ? 14  2.2 50
39 Gmundertobel 1909 Switzerland N 47° 23' 16" E 009° 20' 53" 80.0 26.0 t 18  1.0 120
40 Tiefetal 1952 Germany N 47° 38' 38" E 010° 43' 46" 77.7 12.6 t 10  9.8 28
41 Guiers 2000 France N 45° 35' 53" E 005° 37' 52" 73.8 12.5 s 9  1.7 31
42 Russian Gulch 1939 United States N 39° 19' 45" W 123° 48' 17" 73.2 25.9 ? 15  3.6 128
43 Mazzocco 2007 Italy N 42° 29' 16" E 014° 09' 51" 70.0 14.0 s 3  0.0 G
44 Stampfgraben 2003 Austria N 46° 41' 17" E 012° 51' 28" 70.0 21.0  6  1.4 27
45 Rocky Creek 1932 United States N 36° 22' 45" W 121° 54' 09" 68.6 17.1 t 11  3.6 67
46 Fulton Road 1932 United States N 41° 26' 56" W 081° 43' 02" 64.0 12.7 t 5  4.2 18
47 Matsesta 1938 Russia N 43° 33' 16" E 039° 47' 37" 63.6 23.6 ? 12  0.9 55
48 Broadway 1932 Canada N 52° 07' 19" W 106° 39' 35" 54.7 13.9 t 11  2.1 45
49 Big Creek 1937 United States N 36° 04' 12" W 121° 36' 01" 54.1 23.0 t 11  3.1 75
50 Lianlao ? Taiwan N 25° 00' 32" E 121° 38' 55" 50.0 12.5 ? 8  0.5 20
51 Wölkau ? Germany N 50° 57' 31" E 013° 51' 53" 47.5 5.1 s 3  0.0 G
52 Elche de la Sierra 1927 Spain N 38° 24' 38" W 002° 00' 38" 40.0 4.0  20  1.4 19
53 Dodan Nallah 1995 India N 32° 13' ??" E 076° 19' ??" 40.0 14.5 s 7  1.3 26
54 Bronte Creek 1934 Canada N 43° 25' 08" W 079° 48' 38" 39.8 17.1 ? 11  2.4 56
55 Schwandbach 1933 Switzerland N 46° 49' 40" E 007° 23' 53" 37.4 6.0 t 9  0.7 15
56 Nan Ke ? Taiwan N 23° 05' 36" E 120° 15' 56" 25.0 3.3 ? 3  0.4 G
57 Ziggenbach 1924 Switzerland N 47° 04' 39" E 008° 55' 45" 20.0 4.7 t 8  0.2 10
58 Bohlbach 1932 Switzerland N 46° 43' 49" E 007° 51' 59" 14.4 2.7 t 9  0.0 4

G means value not t timber scaffolding / centering 3-hinged arch k=0.54 
applicable for s steel scaffolding / centering 2-hinged arch k=0.50 
comparison  cantilever using cable-stayed IB0/,<.35  
 cantilever using effective truss k=effective arc length factor
? means value  pre-erected reinforcement arch
unknown  rotation method

161
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-4 shows the section types and geometric properties of the decks, arches, and spandrel columns of

bridges in the database. Deck and arch sections were categorized as one of the following: rectangular solid

sections (slabs or solid ribs), voided slab sections, tee or slab-on-girder sections, single or multi-cell box

girder sections, or composite concrete and steel sections. Slab-on-girder sections were assumed to act

compositely, as references did not usually provide enough information to suggest otherwise. Deck widths

were measured from edge to edge of the section, including pedestrian corridors and traffic shoulder. This

was done for uniformity in data, as not all sources marked the location of traffic lanes. Cross sectional

areas and moments of inertia of deck and arch sections were calculated by discretizing sections into sets of

simple rectangular elements. For variable depth arches, the depth used for calculation was the average

between maximum and minimum depths. For transformed moments of inertia of composite sections, a

modular ratio of 8 was assumed. Deck continuity refers to the support condition over the spandrel

columns. Discontinuous deck refers to decks that are simply supported between spandrel columns. In these

bridges, all flexible system moments are carried by the arch. Continuous deck refers to decks that are con-

tinuous over spandrel columns, which allows for their participation in resisting flexible system moments.

Table 5-5 shows calculated values of various span-to-rise ratios, span-to-depth ratios, slenderness ratios,

moment of inertia ratios, equivalent slab thicknesses, and quantities related to the effects of dead and live

load. All these quantities are related to some aspect of arch structural behaviour. Each quantity will be de-

scribed in detail and presented graphically in Section 5.3.

162
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-4. Sectional geometry of arch bridges in database


deck section arch section column section
type width depth area inertia cont. type depth var. area inertia area inertia
ID bridge m m m2 10-3 m4 m m m2 10-3 m4 m2 10-3 m4
1 Wanxian-Yangtze T 21.0 ? ? ? -- 3 7.00 27.9 220000 5.4 4400 deck width
2 Krk I T 10.8 2.00 4.8 1800 -- 3 6.50 21.5 130000 1.7 300
deck depth
3 Colorado  26.9 2.50 @ 8200 —  4.27 14.8 40000 6.9 2900
4 Gladesville T 26.4 1.80 9.8 2900 —  5.60 ± 1.34 25.1 100000 3.0 63
5 Rio Parana T 13.5 2.20 10.5 2100 — 3 4.00 ± 0.80 17.7 44000 0.6 92 column
6 Infant Henrique  20.2 4.50 16.8 57000 — v 1.50 18.1 4400 17.4 2500 section
7 Bloukrans v 15.7 1.50 10.8 2300 — 3 4.60 ± 1.00 17.3 53000 4.7 2500
8 Sandö T 12.4 1.23 5.7 540 — 3 3.95 ± 0.95 13.0 31000 0.9 31
9 Tensho v 8.6 0.80 4.4 280 —  5.00 ± 1.00 17.0 54000 8.1 1500 arch depth
10 Sibenik T 11.1 1.50 2.6 680 -- 3 3.30 ± 0.40 7.7 14000 0.6 3
11 Krk II T 10.8 2.00 5.4 1700 -- 3 4.00 9.0 21000 1.7 300
12 Wilde Gera  26.3 3.74 @ ? — 2 4.40 ± 1.10 11.3 34000 7.6 11000
13 Galena Creek 2 18.9 3.00 12.1 14000 —  3.70 6.1 13000 @ @
14 Krka-Skradin  22.3 1.96 @ 3700 — 2 3.00 10.9 15000 4.1 1700
15 Maslenica T 21.2 2.00 13.3 4800 — 2 4.00 12.6 29000 5.2 3900
16 Pag T 9.0 1.42 1.9 350 -- 3 2.65 ± 0.35 4.0 4500 1.4 77
17 Modong Hongshui  11.9 0.65 8.1 290 --  3.50 5.7 9600 1.3 35
18 Nosslach ? ? ? ? ? — 3 4.00 ± 1.30 12.5 26000 ? ?
19 Hokowazu v 10.1 0.92 4.4 320 — 2 3.00 10.4 13000 3.0 560
20 Houffalize T 16.0 2.00 6.1 3100 -- 2 3.20 13.0 19000 4.5 1900
21 Pitan  15.9 5.00 16.1 59000 —  6.25 17.5 97000 @ @
22 Wertachtal T 12.7 1.42 6.7 1100 — 2 2.80 ± 0.80 8.5 9900 2.1 120
23 Neckarburg  15.5 2.30 9.2 6300 — 2 3.00 7.2 9200 3.8 4800
24 Caracas T 20.8 2.00 5.4 1800 --  3.10 9.4 13000 3.0 110
25 Argentobel T 14.0 1.60 7.7 1500 — 2 2.75 ± 0.75 7.3 8520 2.4 200
26 Cetina T 10.2 1.40 5.4 870 —  2.50 8.0 7500 3.2 1200
27 Hundwiler T 9.6 0.82 4.3 200 —  2.15 ± 0.65 5.6 2200 1.2 36
28 Maling v 16.0 1.10 7.6 1200 — 4 2.60 ± 0.60 20.2 16000 5.9 500
29 Big Qualicum T 24.0 1.73 7.2 2200 —  1.50 9.0 1700 7.2 860
30 Crooked River 4 24.2 1.90 12.5 6400 — T 1.63 ± 0.38 11.0 2500 5.9 910
31 Niesenback T 17.6 2.10 7.6 2300 — 2 2.50 8.9 8100 2.0 170
32 Siggenthal ? ? 1.40 ? ? — ? 1.20 ± 0.20 ? ? ? ?
33 Nanin  9.9 1.00 4.1 400 —  1.20 ± 0.20 5.1 600 2.5 100
34 Bixby Creek T 8.4 0.92 3.4 200 --  2.00 ± 0.50 5.6 1900 1.5 120
35 Rhein  8.5 1.00 2.8 350 —  1.50 ± 0.25 6.1 1100 2.6 110
36 Wirrbachtal T 11.5 1.40 7.6 1300 —  1.10 ± 0.20 8.0 810 1.0 40
37 Salginatobel T 3.5 0.90 1.7 75 —  @ @ @ 0.8 22
38 Wunshuei v 11.0 0.45 4.3 81 — 2 2.00 7.8 4100 2.0 69
39 Gmundertobel T 6.9 0.70 1.9 60 —  1.65 ± 0.45 11.9 2700 1.2 37
40 Tiefetal T 10.0 1.03 3.6 310 —  1.50 ± 0.30 4.8 1400 1.1 25
41 Guiers  9.8 0.75 10.2 480 — T 1.25 ± 0.20 4.0 570 1.7 170
42 Russian Gulch  9.7 0.55 3.5 45 --  1.65 ± 0.35 4.3 970 1.2 64
43 Mazzocco 4 17.2 1.85 7.6 2700 -- T 1.50 6.2 1300 @ @
44 Stampfgraben T 9.3 0.91 5.4 350 —  1.33 ± 0.28 6.0 880 1.3 26
45 Rocky Creek T 8.2 1.07 4.2 210 --  1.77 ± 0.55 4.3 1100 1.0 64
46 Fulton Road T 23.5 1.17 7.7 820 —  1.50 5.7 1300 5.0 650
47 Matsesta T 18.5 0.83 3.7 220 —  1.11 ± 0.21 6.6 680 3.2 120
48 Broadway T 13.7 1.86 6.3 1600 —  1.20 ± 0.30 11.7 1400 2.0 120
49 Big Creek T 8.2 0.76 3.3 110 —  1.20 ± 0.30 2.2 260 1.1 78
50 Lianlao T 20.5 0.60 6.2 120 —  1.00 18.0 1500 1.8 54
51 Wölkau T 4.5 0.60 2.1 59 —  0.53 ± 0.13 1.3 30 @ @
52 Elche de la Sierra T 8.0 0.49 2.8 42 —  1.01 ± 0.07 2.0 172 0.5 3
53 Dodan Nallah  8.3 0.51 2.5 28 —  0.53 ± 0.18 2.3 53 1.7 23
54 Bronte Creek  16.4 0.23 3.8 17 —  0.81 ± 0.05 4.3 241 2.2 32
55 Schwandbach T 4.2 0.85 1.0 31 —  0.20 1.3 4 0.8 2
56 Nan Ke T 16.0 0.80 9.6 500 —  0.45 4.1 70 @ @
57 Ziggenbach T 5.2 0.88 1.5 89 —  0.41 ± 0.15 2.0 29 0.2 1
58 Bohlbach T 4.5 1.16 1.3 160 —  0.16 1.0 2 0.9 2

@ means value not  rectangular section -- discontinuous —average of max and min depth
applicable for v voided slab section over columns — max depth = depth + var. of spandrel columns used to
comparison T tee or slab-on-girder section — continuous — min depth = depth  var. calculate area and moment of inertia
n n-cell box girder section over columns — depth measured  to arch axis
? means value  composite steel box section — depth used to calculate area and moment of inertia
unknown modular ratio of 8 assumed for transformed moment of inertia

163
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-5. Geometric ratios of arch bridges in database


eccentricity of live load : half arch depth
horizontal reaction / (unit weight  deck width)

normalized
equiv. system moment
slab thickness of inertia
by system inertia
equiv. slab thickness by volume
column inertia : system inertia
arch inertia : system inertia

mod. slenderness ratio


slenderness ratio
span : arch + deck depth
span : arch depth

system rad. of gyr.


span2 : rise
span : rise
ID bridge m m % % m m m2 %
1 Wanxian-Yangtze 5.0 2130 2.80 61 ? 58 61 100 2 ? 5.01 ? ?
2 Krk I 6.5 2540 2.46 60 46 59 49 100 <1 2.61 5.25 1020 6
3 Colorado 3.8 1240 1.80 76 48 73 94 83 6 @ 2.78 @ @
4 Gladesville 7.9 2450 2.02 55 42 56 39 97 <1 1.37 3.60 425 8
5 Rio Parana 5.5 1590 1.61 73 47 68 66 95 <1 2.22 3.45 643 12
6 Infant Henrique 11.2 3140 1.84 187 47 54 27 7 4 1.83 3.32 511 1
7 Bloukrans 4.4 1190 1.79 59 45 60 69 96 5 2.25 3.48 611 12
8 Sandö 6.6 1740 1.56 67 51 63 51 98 <1 1.62 3.12 427 15
9 Tensho 8.0 2080 1.79 52 45 53 36 99 3 2.95 4.23 768 6
10 Sibenik 7.9 1950 1.35 75 51 66 46 100 <1 0.98 2.48 240 27
11 Krk II 5.2 1270 1.53 61 41 61 62 100 1 1.46 2.86 357 17
12 Wilde Gera 3.6 860 ? 55 30 ? ? ? ? @ ? @ @
13 Galena Creek 4.9 1040 2.10 57 31 38 40 48 @ 1.00 2.58 209 @
14 Krka-Skradin 3.9 800 1.31 68 41 63 79 80 9 @ 2.16 @ @
15 Maslenica 3.1 620 1.64 50 33 53 79 86 12 1.53 2.67 306 17
16 Pag 6.9 1330 1.06 73 47 67 53 100 2 0.72 1.82 140 45
17 Modong Hongshui 6.0 1080 1.30 51 43 52 46 100 <1 1.24 2.13 224 23
18 Nosslach 4.0 720 ? 45 ? ? ? ? ? ? ? ? ?
19 Hokowazu 6.4 1090 1.13 57 43 80 47 98 4 1.67 2.51 284 @
20 Houffalize 5.0 810 1.21 51 31 52 54 100 10 1.36 2.42 221 23
21 Pitan 7.4 1180 2.99 26 14 20 15 62 @ 2.17 4.90 346 @
22 Wertachtal 4.6 720 1.14 56 37 54 60 90 1 1.40 2.18 219 27
23 Neckarburg 3.1 480 1.47 51 29 45 67 59 31 1.41 2.29 217 21
24 Caracas 4.7 720 1.18 49 30 72 54 100 1 0.82 1.96 125 37
25 Argentobel 5.2 780 1.17 55 34 49 50 85 2 1.22 2.05 183 28
26 Cetina 6.8 960 1.02 56 36 51 40 90 14 1.44 2.14 202 19
27 Hundwiler 3.9 540 0.65 64 47 86 109 92 2 1.23 1.44 170 46
28 Maling 4.2 580 0.92 53 37 59 72 93 3 2.04 2.35 282 20
29 Big Qualicum 5.5 730 0.66 88 41 76 73 44 22 0.81 1.25 107 26
30 Crooked River 5.2 650 0.90 77 35 53 53 28 10 1.09 1.64 137 14
31 Niesenback 3.2 380 1.08 48 26 47 69 78 2 1.10 1.92 132 40
32 Siggenthal 4.8 560 ? 98 45 ? ? ? ? ? ? ? ?
33 Nanin 4.6 510 0.44 93 51 99 111 60 10 1.05 1.07 118 54
34 Bixby Creek 2.8 280 0.58 50 34 78 126 100 6 1.46 1.39 147 67
35 Rhein 4.8 480 0.49 67 40 79 86 76 8 1.20 1.27 120 53
36 Wirrbachtal 4.0 400 0.51 91 40 78 99 38 2 1.50 1.30 150 34
37 Salginatobel 6.9 620 @ @ @ @ @ @ @ @ @ ? @
38 Wunshuei 6.3 540 0.73 43 35 43 37 98 2 1.25 1.66 106 40
39 Gmundertobel 3.1 250 0.48 48 34 72 108 98 1 2.67 1.69 213 49
40 Tiefetal 6.2 480 0.60 52 31 75 42 82 1 0.91 1.27 71 @
41 Guiers 5.9 440 0.51 59 37 54 49 54 16 1.56 1.09 115 25
42 Russian Gulch 2.8 210 0.47 44 33 98 109 100 7 1.15 1.06 84 @
43 Mazzocco 5.0 350 0.46 47 21 59 61 100 @ 0.84 0.97 59 @
44 Stampfgraben 3.3 230 0.45 53 31 65 93 72 2 1.42 1.17 99 59
45 Rocky Creek 4.0 280 0.51 39 24 54 68 100 6 1.23 1.17 84 69
46 Fulton Road 5.0 320 0.61 43 24 40 42 61 31 0.65 1.03 42 50
47 Matsesta 2.7 170 0.37 58 33 78 128 76 13 0.81 0.84 52 126
48 Broadway 3.9 220 0.51 46 18 44 55 47 4 1.57 1.38 86 36
49 Big Creek 2.4 130 0.41 45 28 90 112 70 21 0.96 0.82 52 @
50 Lianlao 4.0 200 0.30 50 31 67 83 93 3 1.35 0.98 67 70
51 Wölkau 9.3 440 0.26 90 42 93 39 34 @ 0.76 0.62 36 @
52 Elche de la Sierra 10.0 400 0.33 40 27 44 24 80 1 0.64 0.68 25 85
53 Dodan Nallah 2.8 110 0.19 76 39 96 155 65 28 0.79 0.49 32 310
54 Bronte Creek 2.3 90 0.24 49 38 78 139 93 12 0.78 0.57 31 252
55 Schwandbach 6.2 230 0.16 187 36 85 73 12 6 0.64 0.47 24 104
56 Nan Ke 7.7 190 0.37 56 20 24 17 12 @ 0.87 0.75 22 @ @ means value not
57 Ziggenbach 4.3 90 0.24 49 16 33 39 25 <1 0.75 0.65 15 103 applicable for comparison
58 Bohlbach 5.3 80 0.40 90 11 14 13 1 1 0.58 0.76 8 17 ? means value unknown

164
5. Comparative Study of 58 Concrete Arch Bridges

5.3 Trends in recorded data and geometric ratios

This section describes trends observed from the recorded data presented in Section 5.2. In the following

analyses, arch span length is treated as the primary independent variable. Bridges in the database are inter-

preted as unique, feasible design solutions that respond to given span length requirements.

This study is biased toward concrete arch bridges that were sufficiently detailed in references and accessible

to the author. Because the database draws extensively from journal articles and international conferences

on arch bridges, it can be argued that the collected body of data represents the state-of-the-art of concrete

arch bridge technology, especially in North America and Europe. Generally speaking, most new bridges

are designed with proportions that do not depart appreciably from past completed works. In other words,

new bridges are typically designed in reference to or with knowledge of designs that have been previously

constructed. If a particular bridge is innovative or unique in terms of aesthetics, slenderness, shallowness,

lightness, or method of construction, then it is likely that information on its design would have been publ-

ished, and thus included in this database. Based on these observations, it is assumed that the selective

sample of 58 concrete arch bridges in this study is representative of nearly all concrete arch bridges built in

modern times.

165
5. Comparative Study of 58 Concrete Arch Bridges

5.3.1 Geographic location

The geographic locations of bridges included in the database are shown in Figure 5-4. The number of

bridges in the database by country is tabulated in Table 5-6. Swiss, German, Croatian, and Taiwanese con-

crete arch bridges are well represented in the database, as well as those along the west coast of the United

States. Arch bridges in China are poorly represented in the database. According to a study by Chen (2007),

151 concrete arch bridges with spans greater than or equal to 100 metres were built or under construction

as of March 2006. Detailed information about Chinese arch bridges were inaccessible to the author, except

for two Chinese arches found in international journal articles.

ÇÇ Ç
Ç Ç
Ç Ç Ç
Ç ÇÇ
A
Ç Ç ÇÇ ÇÇÇ
ÇÇÇ Ç ÇÇ ÇÇ ÇÇÇÇ Ç Ç
Ç Ç Ç
Ç B ÇÇ Ç
Ç ÇÇ
Ç Ç
Ç
Ç B
ÇÇ
Ç Ç Ç
Ç
Taiwan Europe

Figure 5-4. World map showing location of bridges in database

Table 5-6. Number of bridges in the database by country


country No. country No.

Switzerland 9 Brazil 1
Germany 8 France 1
United States 8 India 1
Croatia 7 Italy 1
Taiwan 5 Portugal 1
Canada 3 Russia 1
Austria 2 South Africa 1
China 2 Spain 1
Japan 2 Sweden 1
Australia 1 Venezuela 1
Belgium 1

166
5. Comparative Study of 58 Concrete Arch Bridges

5.3.2 Methods of construction


span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

year of completion 2020

2010 s ⩚ Colorado
s ⩚
⦿ ⩚ ☒
s
2000 s s ⩚ ⦿ ⩚ ⩚ ☒
s ⩚ ⦿
s
s
1990 ⦿
Wanxian-Yangtze
Argentobel ⦬

1980 ☒ ☒ Krk I
⩚ s

1970
t s⩚
⩚ s s
? Gladesville
1960

t ⩚
1950
t timber scaffolding / centering t Sandö
s steel scaffolding / centering 1940 ?
⩚ cantilever using cable-stayed t?
?
☒ cantilever using effective truss t t t ⦿
⦿ pre-erected reinforcement arch 1930 tt t
⦬ rotation method ⦿
t Elche de la Sierra
1920

Figure 5-5. Year of completion and method of construction versus span length

Figure 5-5 shows the years of completion and methods of construction of bridges in the database versus

span length. Bridges with the longest span for each method of construction have been labelled. Most con-

crete arches in the first half of the 20th century were built cast-in-place on timber falsework. This falsework

would support timber centering, which is the curved temporary framing that is used to support the form-

work for the concrete arch. This labour intensive construction technique limited the lengths that concrete

arches could span economically. The longest spanning arch in the database using timber scaffolding is the

Sandö bridge in Sweden, with a span of 264 metres. Figure 5-6 shows the complex geometry and large

volume of throwaway materials needed to construct the Sandö arch. This method of construction causes

high impact and disruption to the surrounding environment.

Figure 5-6. Sandö Bridge construction using timber scaffolding. Adapted from Mondorf (2006).

167
5. Comparative Study of 58 Concrete Arch Bridges

According to trends in Figure 5-5, steel replaced timber as the preferred material for scaffolding and cen-

tering after 1960. In 1965, the Gladesville Bridge in Australia was constructed using structural steel to form

temporary piers and centering as shown in Figure 5-7. At the time, this bridge held the world record for

longest concrete arch span ever built. The arch is composed of four arch ribs that were built-up from

precast segments.

Figure 5-7. Gladesville Bridge construction using steel centering. Adapted from Seegers (1963).

Instead of using steel arch ribs solely as temporary centering, they can also be designed to remain in place

as permanent reinforcement for the arch. An early example of a bridge that used this technique is the

Elche de la Sierra Bridge in Spain, built in 1927 (Tanner and Bellod 2005). During construction, a pilot

truss composed of steel members were used to support formwork for the arch as shown in Figure 5-8. In

the final stages of construction, the steel pilot truss was embedded in the concrete, making a composite

arch section. This technique reduced the complexity of falsework and the amount of throwaway material.

A similar technique, albeit at a much larger scale, was used 70 years later to construct the Wanxian-

Yangtze bridge, which is currently the longest spanning concrete arch bridge in the world. Instead of using

falsework to support the steel pilot truss as in the Elche de la Sierra Bridge, towers were built above the

springing lines from which cables were stayed. These stays supported the cantilevered steel trusses as they

were built out from both ends of the bridge.

168
5. Comparative Study of 58 Concrete Arch Bridges

Figure 5-8. Steel pilot truss method of construction.


Elche de la Sierra Bridge (left) and Wanxian-Yangtze Bridge construction (right).
Adapted from Tanner and Bellod (2005) and Yan and Yang (1997).

The cantilevering technique used in the Wanxian-Yangtze Bridge can also be used to erect precast or cast-

in-place concrete arch segments, although the weight of concrete cantilevers are normally heavier than

steel ones. This cantilevering method of construction using towers and stays was recently used in the con-

struction of the Colorado River Bridge as shown in Figure 5-9 (US DOT FHWA 2003). Stays emanating from

the towers support the arch cantilevers at several locations along the arch. The forces in these stays are

equilibrated by reactions from the temporary towers, and back stays that are anchored into the rock. Based

on photos of the actual construction of this bridge, all stays were directed upwards to the top of the tower

as opposed to what is shown in the figure.

Figure 5-9. Colorado River Bridge construction using tower and stay method. Adapted from usdot-fhwa (2003).

169
5. Comparative Study of 58 Concrete Arch Bridges

Another variation of cantilevering in the construction of arches involves creating an effective truss, in

which the uncompleted arch ribs form the compression chord of the truss. This was done in the construc-

tion of the Krk Bridges in Croatia in 1980. For these bridges, steel ties were used to carry tensile forces

along the top chord and diagonals of the truss, as shown in Figure 5-10 (Šram 1982). Also shown in the fig-

ure is an overhead cableway, which supports cranes that can be used to transport materials to the free ends

of the cantilevers.

Figure 5-10. Krk Bridge construction using trussed cantilever method.


Adapted from Šram (1982).

In the construction of the Infant Henrique Bridge in Portugal, a similar trussed cantilever method was

used. The tensile forces for the top chord of the truss, however, were carried by the prestressed deck girder,

which was being built simultaneously with the cantilevered arch (Fonseca and Mato 2005). Compared to

using only steel ties, using the deck as part of the truss improves the stiffness and stability of the overall

system during construction, but also increases the complexity of construction.

170
5. Comparative Study of 58 Concrete Arch Bridges

The last method of construction commonly used for concrete arches is the rotation method. The Argento-

bel Bridge in Germany built in 1986 used this method (Hünleim and Ruse 1985). As shown in Figure 5-11,

each concrete arch half-rib was cast in an upright position using slip-forming. The ribs were then rotated

66.5° downward to their final positions. To achieve the rotation, special temporary hinges were built at the

springing lines, and were later fixed with closure pours. Forming the arch upwards, as opposed to out-

wards over the span, gives the advantage of having a smaller construction footprint and better access to the

advancing end of the cantilever. A variation of this method has the arch half-ribs rotated horizontally

rather than vertically. Where the topography of the site allows, this method allows the arch to be built over

land, over the banks at either side of the crossing. These rotation methods of construction are typically

limited to short and medium span arches because of the high degree of movement and bending moments

carried by the arch during construction.

Figure 5-11. Argentobel Bridge construction using the cantilever rotation method.
Adapted from Hünleim and Ruse (1985).

5.3.3 Structural depths of arches and decks

Figure 5-12 shows the structural depths of arches and decks of bridges in the database versus span length.

Nearly all arches with depths of 2 metres or greater are box girder sections. Of these arches, 19 out of 27

(70%) have box girders with multiple cells. Only 3 out of 57 arches in the database (5%) have tee-shaped

sections. This low quantity is probably because arches must resist moments of similar magnitude in both

positive and negative bending. Tee-shaped sections are typically better suited for systems that are biased

toward positive bending.

171
5. Comparative Study of 58 Concrete Arch Bridges

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

max arch depth in m 8


average arch depth in m
min arch depth in m
7 3⧈

3⧈

6

5 ⧈
3⧈
2⧈

4 3⧈ 2⧈ 3⧈ 3⧈ 3⧈


3⧈
3
⧈ 2⧈
2⧈ 2⧈ 2⧈
2⧈
4⧈ 2⧈ 3⧈
2⧈ ⧈
2⧈ −
2 −
− −− − T

−T ⧈ v

− rectangular solid section − −T − ?


v voided slab section 1 − −− −
T tee or slab-on-girder section −
n⧈ single or n-cell box-girder section
− −
− −
0
− −

deck depth in m 6

5 ⧈


4

3 2⧈


⧈ T
2 T T T T
T ⧆ T
T 4⧈ 4⧈
T
T 2⧈
T T v
− rectangular solid section
T
T ? T T T
T
v voided slab section T TT ⧈ v
T tee or slab-on-girder section 1 ⧈ v
T T T T
n⧈ single or n-cell box-girder section T T TT T − T

v
T T
⧆ composite steel-concrete section −
T v


0

Figure 5-12. Section type and structural depths of arches and decks versus span length

Figure 5-13 shows arch depths and deck depths superimposed on the same plot. The arch depths of the

bridges in the database tend to increase linearly with span length. The two bridges that depart from this

trend are the Pitan and Infant Henrique bridges. The Pitan bridge has no spandrel columns and must resist

large bending moments resulting from the frame action between its fused arch and deck. Thus, deep sec-

172
5. Comparative Study of 58 Concrete Arch Bridges

tions are needed to carry these moments. The Infant Henrique bridge achieves a thin arch by minimizing

arch bending moments through deck-stiffening.

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

max arch depth 8


average arch depth
min arch depth Pitan
6
continuous deck depth
discontinuous deck depth
4

Infant-Henrique
0

Figure 5-13. Structural depth of arch and deck versus span length

The deck depths of bridges in the database are less correlated with span length than arch depth. In most

cases, the deck depth is lower than arch depth. This shows that there is a preference among bridges in the

database toward self-stiffened arch systems. Decks of self-stiffened and discontinuous deck arch bridges

tend not to be deeper than 2.5 metres because they do not participate in resisting global flexible system

moments.

Figure 5-14 shows span-to-depth ratios of bridges in the database versus span length. The ratios calculated

for the top graph only consider the depth of the arch. In this case, 48 out of 57 bridges in the database

(84%) have span-to-arch-depth ratios in the range 39 to 77. Thus, this range constitutes the typical range

for span-to-arch-depth ratios in concrete arch bridges. This range includes all bridges with decks that are

simply supported between spandrel columns. The Schwandbach and Infant Henrique Bridges have span-

to-arch-depth ratios of 187, which are nearly twice that of the next highest ratio in the database. These two

bridges achieve thinness in the arch by having stiff decks that attract most of the flexible system bending

moments.

173
5. Comparative Study of 58 Concrete Arch Bridges

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

span-to-arch-depth ratio 200


with continuous deck Schwandbach Infant-Henrique 187
with discontinuous deck
150

max ratio
average ratio 100
min ratio 77
50
39

span-to-system-depth ratio 80
with continuous deck
Schwandbach Infant-Henrique
with discontinuous deck
60
system = arch + deck
max ratio 48
average ratio 40
min ratio
24
20

Figure 5-14. Span-to-depth ratios versus span length

The ratios calculated for the bottom graph in Figure 5-14 consider the sum of the arch and deck depths, or

system depth, in order to capture overall system behaviour. In this case, 46 out of 55 bridges in the data-

base (84%) have span-to-system-depth ratios in the range 24 to 48. Thus, this range constitutes the typic-

ally range for span-to-system-depth ratios in concrete arch bridges. The total amount of bridges between

the top and bottom graphs is different because the deck depth of two bridges are not known. When the

deck depth is considered, the span-to-depth-ratios for the Schwandbach and Infant Henrique Bridges are

not exceptional compared to the ratios of other bridges in the database. These observations suggest that

span-to-system-depth ratios are better measures of arch bridge proportions than are span-to-arch-depth

ratios.

5.3.4 Moments of inertia of arches, decks, and columns

The moments of inertia of arch, deck, and columns are indicators of flexural rigidity, which is the product

of the elastic modulus and moment of inertia. The flexural rigidities of the arch, deck, and columns de-

termine the distribution of flexible system moments, as previously illustrated in Figure 4-15 on page 106.

Because the elastic modulus of concrete is not known for each bridge in the database, flexural rigidities

174
5. Comparative Study of 58 Concrete Arch Bridges

among bridges cannot be directly compared. Instead, moments of inertia and ratios between moments of

inertia will be used as an approximate means to compare distributions of flexural rigidity of bridges in the

database. Effectively, this is equivalent to assuming that the elastic modulus is more or less equal among

bridges in the database. A typical range for the elastic modulus of conventional concrete is between 25000

and 35000 MPa. This means that the flexural rigidities of sections with equal moments of inertia may in

fact differ by up to 40%.

Figure 5-15 shows the arch and deck moments of inertia of bridges in the database versus span length. For

purposes of comparison, these values are normalized by deck width. Trends seen on the top graph for

these values plotted on a log scale are similar to those in Figure 5-13, which shows structural depths of

arches and decks plotted on a linear scale. The general trend is that arch moments of inertia increase faster

than deck moments of inertia, except for those bridges with stiffening decks.

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

arch inertia / deck width in m3 102


deck inertia / deck width in m3
with continuous deck
with discontinuous deck 101

100

10 1

10 2

10 3

10 4

normalized system 6
moment of inertia in m linear regression for :
Pitan y = 0.0109 x + 0.337
= 3 12 system inertia
r2 = 0.906
deck width Tensho
4
y is normalized system inertia in m
excluded from regression analysis x is span length in m
Colorado r2 LVFRHIÀFLHQWRIGHWHUPLQDWLRQ
2

Figure 5-15. Moment of inertia of arch, deck, and system versus span length

175
5. Comparative Study of 58 Concrete Arch Bridges

System moment of inertia Isys is the sum of arch and deck moments of inertia. A means of normalizing this

quantity is needed to make meaningful comparisons among bridges in the database. The moment of iner-

tia I of a rectangular solid will be used as a model for normalizing Isys.

I = 121 bh 3 Equation 5-1

By setting I=Isys and rearranging, the definition for normalized system moment of inertia hI given by Equa-

tion 5-2 is obtained, where bdeck is width of bridge deck:

12I sys
hI = 3 Equation 5-2
bdeck

Quantities of hI versus span length are plotted in the bottom graph of Figure 5-15. A linear regression ana-

lysis was performed on the data, excluding the Pitan Bridge, which is clearly an outlier. The analysis resul-

ted in the following trend line, where L is span length in metres:

hI = 0.0109 ! L + 0.337 m Equation 5-3

Based on this trend line, the data shows that there is a strong correlation between normalized system iner-

tia and span length. The coefficient of determination R2 was found to be 0.906. The two bridges that depart

most from the trend line are the Tensho and Colorado River Bridges. The former has high span-to-rise ra-

tio and probably requires high flexural rigidity to prevent excessive deflections. The latter has relatively

high concrete strength and also has a composite deck. These factors may hinder direct comparison with

the other concrete bridges in the database with respect to bending stiffness in the system.

176
5. Comparative Study of 58 Concrete Arch Bridges

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

system radius of gyration in m 3 Pitan


linear regression for :
= system inertia
arch area y = 0.0062 x + 0.0768
r2 = 0.913
excluded from regression analysis 2
y is system radius of gyration in m
x is span length in m
r2 LVFRHIÀFLHQWRIGHWHUPLQDWLRQ
1

Figure 5-16. System radius of gyration versus span length

Figure 5-16 shows system radius of gyration of bridges in the database versus span length. The expression

for system radius of gyration rsys, previously given in Equation 4-38 on page 114, is modified as:

I arch + I deck
rsys = Equation 5-4
Aarch

In this modification, modulus of elasticity E is assumed to be constant among deck and arch, and arch

cross sectional area Aarch is taken as an average value rather than the value at the crown Acrown. A linear re-

gression analysis was performed on the data, excluding the Pitan Bridge, which again is clearly an outlier.

The analysis resulted in the trend line given by Equation 5-5, where L is span length in metres:

rsys = 0.0062 ! L + 0.0768 m Equation 5-5

Based on this trend line, the data shows that there is a strong correlation between system radius of gyra-

tion and span length. The coefficient of determination R2 was found to be 0.913.

177
5. Comparative Study of 58 Concrete Arch Bridges

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

arch inertia to system inertia ratio 1


with continuous deck 32 self-
with discontinuous deck stiffened
0.8

0.6
18 partially
deck-
0.4 stiffened

0.2
4 deck-
stiffened
0

Figure 5-17. Ratio of arch inertia to system inertia versus span length

Figure 5-17 shows the ratios of arch moment of inertia to system moment of inertia of bridges in the data-

base versus span length. This ratio is used to estimate how flexible system moments are distributed among

arch and deck. As discussed in Section 4.4.1, flexible system moments are distributed to arch and deck in

proportion to their relative flexural rigidity. When the arch inertia to system inertia ratio equals one, the

arch is effectively unstiffened by the deck. This ratio also equals one if the deck is discontinuous over the

columns. Of the bridges in the database, 32 out of 54 (71%) have arch inertia to system inertia ratios greater

than or equal to 0.8, which implies that most are self-stiffened or have discontinuous decks. When this ra-

tio equals zero, the system is stiffened completed by the deck, which carries all flexible system bending

moments. The Infant Henrique bridge is the only deck-stiffened bridge in the database with a span greater

than 100 metres. This design raises the question: why are more long span bridges not built with slender

arches and stiff decks.

Figure 5-18 shows the ratios of column moment of inertia to system moment of inertia of bridges in the

database versus span length. Column moments of inertia are calculated based on average dimensions

among spandrel columns. This ratio of inertias affects the magnitude of flexible system bending moments

carried by the deck and arch. As shown in Figure 4-15 on page 106, lower column flexural rigidities corres-

pond with higher flexible system moments in the arch and deck. When this ratio is increased to one and

spandrel columns are monolithically connected to the arch and deck, the system behaves as a Vierendeel

truss, reducing bending moments in the arch and deck, but also increasing moments in the columns.

178
5. Comparative Study of 58 Concrete Arch Bridges

When the ratio is low, the columns behave like pin-ended columns regardless of their connection type,

since very little bending moment can be distributed into the columns. In the database, 39 out of 49 bridges

(80%) have column moment of inertia to system moment of inertia ratios of 0.1 or less. In conceptual

design, spandrel columns are often assumed to be pin-ended, even if they will eventually be detailed as

fixed or monolithic connections. The generally low flexural rigidity of the spandrel columns relative to that

of the system makes this preliminary design assumption acceptable in most cases.

span length in m 0 100 200 300 400 500


bridge ID 51 41 31 21 11 1

column inertia to system inertia ratio 1

0.1

0.01

0.001

Figure 5-18. Ratio of average column inertia to system inertia versus span length

5.3.5 Arch slenderness ratios

Arches are subjected to combined compression and bending and are thus susceptible to second-order

effects. The severity of these additional deflections and bending moments depend on the slenderness of the

system. In general, slenderness is a function of unbraced length, support conditions, and radius of gyration

of a member. For arches, system slenderness ratio ( is given by (Galambos 1998):

kS
!= Equation 5-6
rsys

Effective arc length kS has been previously discussed in Section 4.3.1 on page 86. Effective arc length

factors k are determined based on the number of hinges present in the arch, and are given by Equation 4-1.

Arc lengths S is given by Equation 4-3, which assumed that the arch axis forms a parabola. System radius

of gyration rsys is given by Equation 5-4, which includes the effect of deck-stiffening.

179
5. Comparative Study of 58 Concrete Arch Bridges

span length in m 0 100 200 300 400 500


bridge ID 51 41 31 21 11 1

system slenderness ratio 120


Russian
effective arc length factor arc length Gulch
= 100 Nanin 99
system radius of gyration
80 80

60

40 38

20

Figure 5-19. System slenderness ratio versus span length

Figure 5-19 shows system slenderness ratios of bridges in the database versus span length. In the database,

43 out of 54 bridges (81%) have system slenderness ratios in the range of 38 to 80. Thus, this range consti-

tutes a typical range for system slenderness ratios in concrete arch bridges. The Russian Gulch and Nanin

Bridges have the highest system slenderness with ratios of 99. Bridges with high slenderness, such as these,

are susceptible to second order effects, and thus require structural analyses that account for material and

geometric nonlinearity.

5.3.6 Span-to-rise ratios

The choice of span to rise ratio is strongly influenced by the geometric requirements of given bridge sites.

Elevations of roadways are often determined by the start and end points of existing or planned transporta-

tion networks or navigation clearances. The location of arch foundations are constrained by site accessibil-

ity and the ability of the ground material to provide adequate reactions. Along with these considerations,

potential arch bridges should respond naturally to the surrounding landscape from a visual perspective.

Figure 5-20 shows the ground and arch profiles for selected bridges in the database. In most of these cases,

the elevation of the arch crown is chosen to be close to the elevation of the higher, shallower slopes at each

end of the bridge site. The springing lines of the arch are positioned to traverse deep valleys and avoid

steep slopes, such as in the Colorado River bridge site. In some cases, the springing lines are placed adja-

cent to edges of bodies of water, while in others, they are not. In the case of the Neckarburg Bridge, mov-

ing the springing lines closer to the edges of the water surface would result in span-to-rise ratios less than

180
5. Comparative Study of 58 Concrete Arch Bridges

3, which is possible to build but lacks in visual appeal. This aesthetic judgement can be made if the viewer

holds the following values: (1) that longer spans are more visually appealing than shorter spans because

they are more daring and impressive, and (2) that less material is more visually appealing than more ma-

terial. In terms of the second value, the ratio of arc length S to span length L increases rapidly as span-to-

rise ratios are decreased below 3. This vertical elongation of the arch may be perceived as an inefficient use

of material.

Figure 5-20. Ground and arch profiles of selected bridges in the database.
All drawn at 1:20000 scale. Inverted triangles mark the surfaces of water.

To illustrate the visual effect of different span-to-rise ratios, variations of the Colorado River Bridge arch

are shown in Figure 5-21. Variation A has a very deep arch with a starting inclination angle of about 60°.

This angle makes the parabolic form similar to an equilateral triangle, accentuating the vertical rising of

the arch rather than its horizontal spanning. In this case, the arch would be difficult to build because the

left slope is steep and not easily accessed. If this arch was built using cast-in-place construction, the high

angle would require top and bottom forms, which is more costly and more complex than casting with bot-

tom forms only. Variations B and C move away from the triangular form, and accentuate the horizontal

spanning of the arch. According to the aesthetic values previously stated, Variations B and C are superior

to Variation A. Visually, Variations B and C are not that different from the original, and respond reason-

ably well to the landscape. Although Variation D spans much farther than the original and may seem more

structurally daring, its aesthetic quality is compromised by the impractical positioning of the left springing

181
5. Comparative Study of 58 Concrete Arch Bridges

line. Variation D overshoots the natural cliff edge on the left side, which makes the economic premium of

increased span length seem unwarranted.

Figure 5-21. Variations of span-to-rise ratio based on the Colorado River Bridge. All drawn at 1:20000 scale.

Span-to-rise ratios of bridges in the database versus span length are shown in the top graph of Figure 5-22.

55 out of 58 bridges in the database (95%) have span-to-rise ratios in the range of 2.3 to 8.0. As previously

discussed, the low end of the range is governed by awkward aesthetics and material efficiency. The high

end of this range is probably governed by precedence, economy, and complexity. The Infant Henrique,

Elche de la Sierra, and Wölkau Bridges show that span-to-rise ratios greater than 8.0 are possible to build.

Three possible reasons why concrete arch bridges with high span-to-rise ratios are not commonly built are:

(1) there are fewer bridge sites that are geometrically suitable for arches with high span-to-rise ratios than

for conventional span-to-rise ratios, (2) in terms of economy, variable depth girders become increasingly

competitive in spaces suitable for arches with high span-to-rise ratios, (3) arches with high span-to-rise ra-

tios require large horizontal reaction forces and thus foundations are costly or impractical, and (4) con-

struction requires a high standard of care, given the potentially severe consequences of misalignments of

the arch and pressure line.

182
5. Comparative Study of 58 Concrete Arch Bridges

span length in m 0 100 200 300 400 500


bridge ID 51 41 31 21 11 1

span-to-rise ratio 16

12
Elche de la Sierra Infant-Henrique 11.2
Wölkau
8 8.0

4
2.3
Bronte Creek
0

span2 / rise in m 4000


VWDWLFFRHIÀFLHQW 16 =1
2 =8
ise=
or Infant-Henrique
3000 nt
spa

2000 =4

1000

Krk-II Colorado
0

Figure 5-22. Span-to-rise ratios versus span length

The bottom graph of Figure 5-22 shows quantities of span squared divided by span length of bridges in the

database versus span length. This quantity is also referred to as the static coefficient. According to Mondorf

(2006), bridge builders used static coefficients to gauge how demanding the construction of a prospective

arch would be. Higher coefficients indicated higher complexity of construction. This coefficient is present

in the expression for horizontal reaction H, as given by Equation 5-7, where q is uniformly distributed

load, L is span, and f is rise:

qL2
H= Equation 5-7
8f

For the same uniformly distributed load, the Krk II and Colorado River Bridges have about the same hori-

zontal reaction, even though they differ in span length by 79 m. The Infant Henrique Bridge, has a static

coefficient 2.5 times that of the Krk II and Colorado River Bridges, which is large considering that it spans

43 m less than the Colorado Bridge.

183
5. Comparative Study of 58 Concrete Arch Bridges

5.3.7 Modified slenderness ratios

As discussed in Section 4.4.3 on page 110, modified slenderness ratio (f, which is given by Equation 5-8

(Cai et al. 2009), governs the transition of shallow arches between pure arch behaviour and pure beam be-

haviour. A minimum value of 20 was previously identified as a threshold for efficient fixed arch behaviour.

Shallow fixed arches with (f >20 carry more than 90% of permanent loads in compression along the pres-

sure line, and less than 10% in bending. Decreases in (f cause increases in bending moment demand and

decreases in arch compression.

2f
!f = Equation 5-8
rsys

Modified slenderness ratios of bridges in the database versus span length are shown in Figure 5-23. Of the

fixed arches in the database, only 3 out of 51 bridges (6%) have modified slenderness ratios less than 20.

This subset includes the Nan Ke and Pitan Bridges, which do not have spandrel columns and are thus in-

herently more beam-like than other bridges in the database. Most of the other bridges in the database have

ratios much greater than 20, which indicates that they are not shallow and that their crown deflections are

small relative to the initial rise of the arch.

span length in m 0 100 200 300 400 500


bridge ID 51 41 31 21 11 1

PRGLÀHGVOHQGHUQHVVUDWLR 160
2 rise
=
V\VWHPUDGLXVRIJ\UDWLRQ
120
À[HGDUFKHV
KLQJHGDUFKHV
80

IRUÀ[HGDUFKHV
40 increasing arch action
Nan Ke Pitan 90% arch action
0 Bohlbach increasing beam action

Figure 5-23. Modified slenderness ratio versus span length

A strong correlation was previously observed between system radius of gyration and span length in Sec-

tion 5.3.4. Using the linear regression trend line given by Equation 5-5, thresholds for efficient arch beha-

184
5. Comparative Study of 58 Concrete Arch Bridges

viour can be expressed in terms of span-to-rise ratios. By substituting Equation 5-5 into 5-8, and imposing

that (f is greater than 20, the following inequality is obtained, where L is span in metres:

L ! L $
< 16.1 # Equation 5-9
f " L + 12.4 m &%

Equation 5-9 gives threshold values for span-to-rise ratio L/f as a function of span length L and is plotted

in Figure 5-24. The figure shows that arches with spans less than 50 m have particularly low span-to-rise

ratio threshold values. For spans greater than 50 m, the L/f threshold values lie around 12:1 and 14:1. Be-

cause Equation 5-9 is based on average system radius of gyration values, the actual threshold curve should

be lower than what is shown in Figure 5-24.

span-to-rise ratio 16

Threshold based on
12 linear regression trend
line of system radius
of gyration.
8

4
Arches with span-to-rise ratios less
than the threshold curve exhibit
0 greater than 90% arch action.
0 100 200 300 400 500
span length in m

Figure 5-24. Efficiency threshold curve based on average system radius of gyration trend line

The interplay of section capacity and demand compounds the problem of arch shallowness and beam-like

behaviour. For a given span length, arches with high span-to-rise ratios have higher sectional forces than

arches with lower span-to-rise ratios. Thus, the former requires deeper and stiffer sections than the latter.

Choosing stiffer sections will tend to increase system radius of gyration rsys, unless flexural rigidity EI and

axial rigidity EA are increased at the same rate. Increases in system radius of gyration rsys cause decreases

in modified slenderness ratio (f, thus moving the system toward beam behaviour. Increased beam beha-

viour will increase overall bending moments in the arch, which may then require increases in flexural ca-

pacity and flexural rigidity. In other words, proportioning shallow arches to be stronger and stiffer in

bending will inherently make them more beam-like. Based on these reciprocative effects, it is likely that

185
5. Comparative Study of 58 Concrete Arch Bridges

systems with high span-to-rise ratios require system radii of gyration that are greater than those predicted

by the least squares trend line given by Equation 5-5.

A more conservative span-to-rise ratio threshold curve can be obtained by increasing the assumed values

for system radius of gyration. A subjective high estimate trend line, which serves as an approximate upper

envelope of system radii of gyration, was chosen by increasing the vertical intercept in Figure 5-16 on page

177 until all but three data points were below the line, while keeping the slope of the line constant. The se-

lected high estimate trend line is given by the following equation, where L is span in metres:

rsys = 0.0062 ! L + 0.3m Equation 5-10

Using Equation 5-10, the resulting span-to-rise ratio threshold inequality becomes:

L ! L $
< 16.1 # Equation 5-11
f " L + 48.4 m &%

This new threshold curve (dashed line) is plotted in Figure 5-25, along with the less conservative threshold

curve (solid line) based on the linear regression trend line (Equation 5-9). The curves on the diagram show

that it is always inefficient to design fixed concrete arches with span-to-rise ratios greater than 16:1. The

shaded area on the diagram represents arch systems that exhibit greater than 90% arch action, based on

the subjective high estimate. The shaded area shows that fixed arches with spans less than 50 m are partic-

ularly sensitive to moments caused by shallowness effects, even in the normal range of span-to-rise ratios

of 3:1 and 8:1. Thus it is important that systems with high values of rsys are not used in this range of spans.

For fixed arches with spans greater than 100 m, the threshold curve ranges between 12:1 and 14:1. The data

point for the Infant Henrique Bridge, which is among the flattest concrete arches in the world, lies just be-

low the threshold curve. Thus there still remains an opportunity to design even flatter, efficient arches that

carry more than 90% of their dead load in axial compression.

186
5. Comparative Study of 58 Concrete Arch Bridges

Threshold based on
linear regression trend
line of system radius
of gyration.
span-to-rise ratio 16

12 Threshold based on
Infant Henrique Bridge subjective high
estimate trend line.
8

bridges in database
4
Arches with span-to-rise ratios less
than the threshold curve exhibit
0 greater than 90% arch action.
0 100 200 300 400 500
span length in m

Figure 5-25. Efficiency threshold curve based on subjective high estimate of system radius of gyration

It could be argued that span-to-rise ratios greater than the proposed threshold curves are possible, given

that the radius of gyration of the system could sufficiently be reduced. However, this means that the cross

sectional area of the arch would have to be increased faster than the sum of the moments of inertia of the

deck and arch. Effectively, this requires that material be added close to the centroid of the arch, which is

inherently inefficient for carrying bending moments caused by nonuniform live loads and restrained de-

formations. A more precise determination of threshold values for span-to-rise ratios is not possible

without doing detailed calculations of capacity and demand for specific design concepts. Thus these span-

to-rise threshold curves should be considered to be efficiency guidelines for designers, rather than rigid

limitations on the design of shallow concrete arches.

5.3.8 Equivalent slab thickness

Equivalent slab thickness hV based on volume of concrete is defined as the total volume of the system V di-

vided by bridge width bdeck and span L:

V
hV = Equation 5-12
bdeck ! L

Estimates of concrete volume V can be made by multiplying the average cross sectional areas of arches,

decks, and spandrel columns by their respective element lengths. The expression used to calculate concrete

volume V is given by Equation 5-13, where Adeck, Aarch, and Acol are cross sectional areas of the deck, arch,

187
5. Comparative Study of 58 Concrete Arch Bridges

and columns, L is span length, S is arc length, and hi is the height of each spandrel column. Because the

volume of diaphragms, transverse ribs, and stiffeners are not considered, this estimate is a lower-bound es-

timate of concrete volume.

n
V = Adeck ! L + Aarch ! S + Acol ! " hi Equation 5-13
i=1

Figure 5-26 shows equivalent slab thicknesses of bridges in the database versus span length. This quantity

gives some measure of efficiency in terms of concrete consumption. Based on the data, the Gmundertobel

and Tensho Bridges consume more than two times the amount of concrete used in arch bridges of similar

span. The Gmundertobel Bridge was built in 1909 and is likely influenced by the proportions of heavier

masonry arches. The Tensho Bridge uses a single-cell box girder arch rib with thick slabs and webs, result-

ing in a very heavy arch. The assumption that its spandrel walls are solid may also have caused a moderate

overestimate of its true concrete volume.

span length in m 0 100 200 300 400 500


bridge ID 51 41 31 21 11 1

equivalent slab thickness by volume 3 Tensho


= volume / (span deck width) in m
Gmundertobel Krk I
volume = deck area span
+ arch area arc length
+ column area total spandrel 2
column length

Gladesville
1
Sibenik
Fulton Road Pag
Bohlbach
0

Figure 5-26. Equivalent slab thickness by volume versus span length

Bridges with the lowest equivalent slab thicknesses for various span ranges are labelled in Figure 5-26. Data

points from these bridges were used to form an envelope of minimum concrete consumption, representing

the highest level of material efficiency achieved by existing concrete arch technology. This envelope can be

used to evaluate the relative efficiency of new concepts for concrete arch bridges, as will be done in

Chapter 6. In equation form, the minimum equivalent slab thickness hV envelope is given by the following

equation, where L is span in metres:

188
5. Comparative Study of 58 Concrete Arch Bridges

# L
% + 0.569 m 0 ! L ! 193m
% 1276
hV = $ Equation 5-14
% L2 L
" + 2.21m L > 193m
% 22800 m 61.7
&

5.3.9 Historical evolution

Another important aspect to consider is the historical evolution of concrete arches in the database. There

has been a tendency to believe that improvements in the quality and strength of conventional concrete

over the past century have resulted in significant increases in structural efficiency. There has been no exist-

ing evidence, however, to support this claim. Figure 5-27 shows the same data as in Figure 5-26, but adds

the dimension of time. Equivalent slab thickness data points have been labelled with the years of construc-

tion of each bridge. The data shows that there is no apparent relationship between year of construction and

equivalent slab thickness. This suggests that structural efficiency is more a product of the talent and ambi-

tion of a particular designer, rather than the strength of conventional reinforced concrete.

equivalent slab 3
thickness by 2000
volume in m

1909 (year of construction)


1980

1983
1965
1993
2 1999

2003
1974
1932 1943
2000 2002 1997
2007 1980
2003 1932 ?
? 1978 1979 1964
1932 ? 1991
1963 1986 1999
1939 ? 2000
1 1967 2011 1966
1937
1952
? 1995 2007 1996 1953
1924 1934 ? 1938 1968
1933 1927 1932
1932

0
0 100 200 300 400 500

span length in m

Figure 5-27. Equivalent slab thickness, span length, and year of construction

189
5. Comparative Study of 58 Concrete Arch Bridges

5.3.10 Horizontal reactions caused by dead load

Dead load due to the self-weight of concrete can be approximated as a uniformly distributed load q, as giv-

en by the following equation, where V is concrete volume (Equation 5-13), # is unit weight of concrete, and

L is span length:

!V
q= Equation 5-15
L

By substituting Equation 5-15 into Equation 5-7, and normalizing the result by deck width bdeck and unit

weight of concrete #, estimates of horizontal reactions H can made using the following equation:

H V "L
= Equation 5-16
! " bdeck 8 f " bdeck

span length in m 0 100 200 300 400 500


bridge ID 51 41 31 21 11 1

horizontal reaction in m2 10000


(unit weight deck width)
= volume span twice HPSLULFDOÀWIRULQ
8 deck width rise the form: y = aL2 + b
1000
half

100 L2 + 10 m2
y
182.7
y is normalized
10 horizontal reaction

Figure 5-28. Normalized horizontal reaction versus span length

Figure 5-28 shows estimates of normalized horizontal reactions of bridges in the database versus span

length. An overall exponential trend can be observed from the data. A Levenberg-Marquardt algorithm

was used to fit the data into the form: y=mx2+b. The resulting trend found for normalized horizontal reac-

tion as a function of span L in metres is given by:

H L2
in m 2 # + 10 m 2 Equation 5-17
! " bdeck 182.7

190
5. Comparative Study of 58 Concrete Arch Bridges

With this empirical equation, designers can estimate lower bound values for axial forces in the arch using

the deck width and unit weight of concrete specified in their design. These estimates can then be increased

by some factor to account for the additional weight of diaphragms, transverse ribs, and stiffeners.

5.3.11 Effects of combined dead and live loads

The effects of live load on fixed arch bridges in the database will be compared in this section. A simple stat-

ical analysis will be used to estimate the eccentricity of the resultant sectional force at the quarter-point of

the arch. Given that the arch is set directly on the pressure line under dead load, the resultant axial forces

along the arch are assumed to be collinear with the centroidal axis of the arch (see Figure 5-29a). The

effects of elastic shortening of the arch are neglected in this simplified analysis. The addition of nonuni-

form live loads causes both axial forces and bending moments in the arch. Live load axial forces are neg-

lected in this simplified analysis, since they are typically small in comparison to axial forces caused by dead

load. Live load bending moments are expressed as displacements of the resultant axial force away from the

centroid of the arch section. Figure 5-29 illustrates the progression of stress profiles and resultant axial

forces N in the arch as live loads are added to dead loads.

d State 1:
2 ‡GHDGORDGRQO\
S N ‡XQLIRUPFRPSUHVVLYHVWUHVV
‡UHVXOWDQWIRUFHLQOLQHZLWKFHQWURLG

d State 2:
2 N e ‡GHDGORDGSOXVQRQXQLIRUPOLYHORDG
S ‡WUDSH]RLGDOFRPSUHVVLYHVWUHVV
‡UHVXOWDQWIRUFHHFFHQWULFWRFHQWURLG

d N State 3:
2 e ‡GHDGORDGSOXVPRUHOLYHORDG
S ‡QRQOLQHDUFRPSUHVVLYHVWUHVV
‡UHVXOWDQWIRUFHKLJKO\HFFHQWULFWRFHQWURLG

Figure 5-29. Eccentricity of resultant compressive force caused by live load bending moments

The displacement of resultant axial force will be referred to as live load eccentricity elive. Estimates of max-

imum live load eccentricity can be calculated using Equation 5-18, where Mtruck and Mlane are moments at

191
5. Comparative Study of 58 Concrete Arch Bridges

the quarter-point of the arch caused by multilane truck and lane loads, and N is axial force at the quarter-

point of the arch.

max(M truck , M lane )


max(elive ) = Equation 5-18
N

Quarter-point axial force N of fixed arches caused by self-weight can be estimated using the following

equation, where # is unit weight of concrete, V is volume of concrete (Equation 5-13), L is span, and f is

rise:

2
!V " L%
N (x = L) =
1
+4 Equation 5-19
8 $# f '&
4

Quarter-point bending moments Mtruck of fixed arches caused by multilane truck loading can be estimated

using Equation 5-20, where Qlive is the weight of a single heavy truck, n is number of design traffic lanes, m

is multilane modification factor, kdyn is dynamic amplification factor, and Iarch and Isys are moments of iner-

tia of arch and system:

(
M truck (x = 41 L) = 59.3!10 "3 !Qlive ! L ! n ! m ! 1 + kdyn !) I arch
I sys
Equation 5-20

Quarter-point bending moments Mlane of fixed arches caused by multilane lane loading can be estimated

using Equation 5-21, where qlive is the uniformly distributed weight of a single lane of traffic:

( )
M lane (x = 41 L) = 59.3!10 "3 ! 108 Qlive ! L + 9.1!10 "3 ! qlive ! L2 ! n ! m !
I arch
I sys
Equation 5-21

Load parameters #, Qlive, qlive, kdyn, n and m are given values according to the Canadian Highway Bridge

Design Code CAN/CSA S6-06, as listed in Table 5-7. All other quantities are retrieved from the arch

database.

192
5. Comparative Study of 58 Concrete Arch Bridges

Table 5-7. Load parameter values


parameter value reference

unit weight of concrete # 24 kN/m3 S6-06 §3.6


single truck load Qlive 625 kN S6-06 §3.8.3.1
single lane load qlive 9 kN/m S6-06 §3.8.3.2
number of design lanes n =f(bdeck) S6-06 §3.8.2
multilane modification factor m =f(n) S6-06 §3.8.4.2
dynamic amplification factor kdyn 0.25 S6-06 §3.8.4.5

For purposes of comparison, live load eccentricities are normalized by half the arch depth 0.5darch. This

normalizing value is chosen because most bridges in the database have symmetric arch sections, whose

distances between centroid and extreme fibres are equal to 0.5darch. When resultant axial forces have dis-

placed to the extreme fibre at the quarter-point of the arch, normalized live load eccentricities equal one.

span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1

live load eccentricity 1.5


half depth of arch

0.5

Figure 5-30. Normalized live load eccentricity versus span length

Normalized live load eccentricities of fixed arch bridges in the database versus span length are plotted in

Figure 5-30. From the figure, it can be observed that the effects of live load diminish as span length in-

creases. In three cases, fixed arch bridges with spans less than 100 metres have normalized live load eccent-

ricities greater than one. This suggests that significant nonlinear bending stresses in the arch are expected

at serviceability limit states. Bridges in the database with spans greater than 200 metres have small live

load eccentricities, suggesting that their arch stresses are mostly compressive at serviceability limit.

The diminishing live load effects in arches can be interpreted as a progression from the governing of live

loads in short-span arches to the governing of dead loads in long-span arches. As dead loads become large

193
5. Comparative Study of 58 Concrete Arch Bridges

relative to live loads, arches benefit from an increase in their natural pre-compression under dead load. In

other words, arches with large resultant axial forces N (Figure 5-29) require smaller eccentricities elive than

arches with small resultant axial forces, given the same live load bending moment demand. A similar beha-

viour occurs in suspension bridges, where main suspension cables in tension equilibrate nonuniform live

loads through deformation. Tensile forces in the main cables are caused primarily by dead load. As long as

cables remain in the elastic range, greater tensile forces require smaller deflections to equilibrate live loads.

This behaviour is sometimes called the “gravity stiffness” of suspension bridges because live load perturba-

tions of the system are trivial when added to the state of equilibrium caused by gravity load.

Based on the data in Figure 5-30, designers may be misled to think that all bending moments in long-span

arches are diminished. The simplified analysis performed here considers only the effects of traffic live

loads. The effects of restrained deformation, including temperature, creep, and shrinkage tend to increase

with span length. These effects will also cause the resultant axial force of the arch to displace away from the

centroid of the arch. In many cases, these effects will be more severe than those caused by traffic live loads.

Thus, it is important to understand that the observations made in this section are limited to the effects of

dead and live load only.

5.4 Summary of observed trends


The following trends on the proportions and properties of 58 concrete arch bridges have been observed:

• 46 out of 57 bridges in the database (84%) have span-to-arch-depth ratios in the range of 39 to 77.

• 46 out of 55 bridges in the database (84%) have span-to-system-depth ratios in the range of 24 to 48.

Span-to-system-depth ratios are better measures of the overall proportions of arch bridges than are

span-to-arch-depth ratios because they account for the effects of deck-stiffening.

• Normalized system moments of inertia hI and span length L were found to be well-correlated. A lin-

ear regression analysis resulted in the trend: hI = 0.0109 ! L + 0.337 m .

194
5. Comparative Study of 58 Concrete Arch Bridges

• System radius of gyration rsys and span length L were found to be well-correlated. A linear regression

analysis resulted in the trend: rsys = 0.0062 ! L + 0.0768 m .

• 32 out of 54 bridges in the database (59%) have ratios of arch inertia to system inertia Iarch/Isys greater

than or equal to 0.80, implying that most arches are self-stiffened or have decks that are discontinu-

ous over spandrel columns.

• 39 out of 49 bridges in the database (80%) have ratios of column inertia to system inertia Icol/Isys less

than 0.1. Because spandrel columns generally have low bending stiffness relative to the arch and deck,

it is acceptable to assume that they are pin-ended during conceptual design.

• 43 out of 54 bridges in the database (81%) have system slenderness ratios ( in the range of 38 to 80.

This ratio includes the effects of deck-stiffening through the use of system radius of gyration rsys.

• 55 out of 58 bridges in the database (95%) have span-to-rise ratios in the range of 2.3 to 8.0.

• 48 out of 51 fixed are bridges in the database (94%) have modified slenderness ratios greater than 20,

which means that these arches exhibit greater than 90% arch action under permanent load.

• A threshold curve of maximum span-to-rise ratios for efficient arch behaviour based on the correla-

tion between system radius of gyration rsys and span length L was calculated to be:

L ! L $
< 16.1 #
f " L + 12.4 m &%

• A threshold curve of maximum span-to-rise ratios for efficient arch behaviour based on a more con-

servative, subjective high estimate envelope of system radius of gyration rsys with respect to span

length L was calculated to be:

L ! L $
< 16.1 #
f " L + 48.4 m &%

195
5. Comparative Study of 58 Concrete Arch Bridges

• Equivalent slab thicknesses by concrete volume hV were calculated for each bridge in the database.

An envelope of minimum slab thicknesses achieved by existing concrete technology is proposed as a

function of span length L:

# L
% + 0.569 m 0 ! L ! 193m
% 1276
hV = $
% L2 L
" + 2.21m L > 193m
% 22800 m 61.7
&

• There is no apparent relationship between year of construction and equivalent slab thickness, sug-

gesting that increases in structural efficiency are not direct outcomes of historical increases of con-

crete strength over the past century.

• An overall exponential trend was observed from the data of normalized horizontal reaction versus

span length. The resulting empirical trend, which excludes the weight of diaphragms, transverse ribs,

and stiffeners, was found to be:

H L2
in m 2 # + 10 m 2
! " bdeck 182.7

• A simple statical analysis was used to estimate the effects of live load in fixed arch bridges in the data-

base. Comparisons of live load eccentricities, or displacements of the resultant axial force away from

the centroid of the arch, showed that the effects of live load diminish with increasing span length.

196
Chapter 6. New Concepts for Arch Bridges

This chapter describes new preliminary design concepts for arch bridges using ultra high-performance

fibre-reinforced concrete. The stress-strain, creep, and shrinkage models assumed in the design of these

concepts are based on material tests described in Chapter 2. The sectional and member capacities of decks,

spandrel columns, and arches are calculated using the general analysis method described in Chapter 3, and

are compared with the simplified design method. Maximum forces at critical sections in the system are es-

timated using the analysis methods described in Chapter 4.

This chapter begins with a description of the objectives, assumptions, and design strategies of a parametric

design study of arch systems using ultra high-performance fibre-reinforced concrete in Section 6.1. The

three parameters considered are: span length, arch rise, and distribution of flexural rigidity. The resulting

design concepts, each responding to different sets of design parameter values, are described in Section 6.2

in terms of their geometries, proportions, and structural demands. In Section 6.2.3, the preliminary design

for one of the design concepts is further developed. In Section 6.3, comparisons of structural efficiency

among design concepts are discussed. Further comparisons are also made with existing concrete arch

bridges. In the last section, general recommendations regarding the potential opportunities offered by

these structures are given.

197
6. New Concepts for Arch Bridges

6.1 Parametric design study

The primary objectives of this parametric design study are: (1) to determine if it is possible to design arches

using ultra high-performance fibre-reinforced concrete that reduce the consumption of concrete by 50% as

compared with current concrete arch technology, (2) to produce preliminary design concepts that respond

to different span length and arch rise requirements, using different distributions of flexural rigidity among

deck and arch, and (3) to quantify and compare the structural efficiency of these preliminary design

concepts.

The three primary design parameters considered are: span length, arch rise, and distribution of flexural ri-

gidity. Parameters are each discretized into a finite set of values. Eight spans between 50 and 400 metres

are considered, each an increment of 50 metres from the previous value. The longest span considered is not

intended as an upper limit of concrete arches, but is instead chosen to be just beyond the range of spans

that have been previously achieved using current concrete arch technology. Four span-to-rise ratios: 4:1,

8:1, 12:1, and 16:1 are considered. The lower values are representative of the range of span-to-rise ratios

commonly used in concrete arch design, while the higher values are greater than the highest ever built. For

each combination of span and span-to-rise ratio, three classes of fixed arch systems are designed: self-sti-

ffened arch, deck-stiffened arch, and partially deck-stiffened arch, each corresponding to different distribu-

tions of flexural rigidity among deck and arch (see Section 4.4.1 on page 99). The combination of 8 span

lengths, 4 span-to-depth ratios, and 3 arch classes results in a design space of 96 potential design concepts.

Each design concept is given an id number, as given in the Table 6-1:

Table 6-1. Design concept ID numbers and design matrix

50 m 100 m 150 m 200 m 250 m 300 m 350 m 400 m


span-to-rise span span span span span span span span

4:1 1+i 2+i 3+i 4+i 5+i 6+i 7+i 8+i


8:1 9+i 10+i 11+i 12+i 13+i 14+i 15+i 16+i
12:1 17+i 18+i 19+i 20+i 21+i 22+i 23+i 24+i
16:1 25+i 26+i 27+i 28+i 29+i 30+i 31+i 32+i

i=0, self-stiffened arches


i=32, deck-stiffened arches
i=64, partially deck-stiffened arches

198
6. New Concepts for Arch Bridges

In order to keep comparisons among design concepts consistent, other geometrical parameters are kept

constant. The specified transverse width of deck for all concepts is 16 metres, which represents a moder-

ately wide bridge that typically holds up to three highway traffic lanes (this width corresponds to four

design lanes of traffic). The specified number of interior spans for all concepts is nine, which ensures that

localized fixed system moments remain small in comparison to global flexible system moments shared by

deck and arch. Increasing the number of interior spans would further reduce fixed system moments, but

would require an increase in the number of spandrel columns, which may result in increases in dead load.

The assumed material properties are kept constant among all design concepts. All sectional analysis calcu-

lations are based on the nonlinear material models presented in Chapter 2. A design compressive strength

of 120 MPa is used, representing a nominal strength that can readily be achieved by many different ultra

high-performance fibre-reinforced concrete mixes, without the use of specialized heat or steam treatment.

A design tensile strength of 14.6 MPa is used, based on the test data presented in Section 2.1.4. This high

tensile strength requires a high steel fibre content similar to the 5.5% fibre content used in the University of

Toronto mix. Figure 6-1 shows the stress-strain curves used for design. Post-peak tensile stresses, which

occur after the formation of localized cracks, are neglected.

compressive fc
stress in MPa

120

84
tensile ft
stress in
MPa

14.6

2.0 3.7 5.2 !c 0.15 1.5 !t


compressive strain in 10-3 tensile strain in 10-3

Figure 6-1. Design stress-strain curves

Other important material properties used for design are summarizes in Table 6-2. The amounts of creep

and shrinkage strains imposed on the completed indeterminate arch system are determined by the method

and schedule of construction. Creep strains do not occur until elastic stresses are applied to the arch. Mo-
199
6. New Concepts for Arch Bridges

ments caused by these creep strains do not occur until systems are made indeterminate by the closure of

the arch. In order to simplify the design process, it is assumed that flat, hydraulic jacks are used to intro-

duce elastic stresses into the system while it is determinate. It is also assumed that the arch crown is made

continuous, or closed, a minimum of 28 days after the jacks are activated. Based on the progression of

creep with time shown in Figure 2-11 on page 31, only about 50% of the total long-term creep strain will

cause redundant forces in the system after closure, corresponding to a creep coefficient ,(t) design value of

0.4. The ageing coefficient is a function of the long-term creep coefficient and of the age of concrete at ini-

tial loading (Bazant 1972). Because this quantity is not easily calculated during preliminary design, it is

usually reasonable to assume a typical ageing coefficient +(t) design value of 0.8 (Menn 1990). Increasing

+(t) to its maximum value of 1 causes relatively small changes in the overall effects of creep.

Table 6-2. Design values for material properties

property symbol design value reference

creep coefficient ,(t) 50% of 0.8 Interim Recommendations (adgc 2002), and Graybeal (2006)

ageing coefficient +(t) 0.8 typical value assumed (Menn 1990)

shrinkage strain %sh(t) 20% of 550·10-6 Interim Recommendations (adgc 2002), and Habel et al. (2008)

coefficient of thermal expansion !T 0.000015 /°C Interim Recommendations (adgc 2002)

Poisson’s ratio * 0.20 Interim Recommendations (adgc 2002)

modulus of elasticity E 42000 MPa based on cylinder and prism tests from Chapter 2

density of concrete #/g 2600 kg/m3 based on mix design quantities presented in Chapter 2
unit weight of concrete # 25.5 kN/m3 based on mix design quantities presented in Chapter 2

Using the previous assumption that the closure of the arch is delayed by at least 28 days, the minimum age

of concrete in the arch at closure is 28 days. Based on the progression of shrinkage with age after casting

(see Figure 2-10 on page 30), only about 20% of the total long-term shrinkage strain will cause redundant

forces in the system, corresponding to a shrinkage strain %sh(t) design value of 110·10-6.

The delay of arch closure may not be required in all cases. If the effects of restrained deformations are

small relative to the effects of dead and live load, then the delayed closure requirement may be ignored. If

the effects of restrained deformation are large, they can be reduced by further postponing the closure of
200
6. New Concepts for Arch Bridges

the arch, or by using heat- and pressure-treated precast segments. The selected design values of ,(t), +(t),

and %sh(t) represent a low to moderate amount of creep and shrinkage that can be expected in the system

without adding further complexity to the construction of the bridge.

All other design material properties, including coefficient of thermal expansion, Poisson’s ratio, modulus

of elasticity, density, and unit weight of concrete are based on the material tests presented in Chapter 2 and

on design values given by the the French Interim Recommendations on ultra high-performance fibre-rein-

forced concrete (AFGC 2002).

6.1.1 Loads, load factors, and material resistance factors

All concepts are designed according to loads and load factors given in the Canadian Highway Bridge

Design Code CAN/CSA S6-06, abbreviated herein as S6-06.

For each concept, permanent dead loads consisting of the total self-weight of the structure, the weight of

traffic barriers, and the weight of wearing surface are approximated as uniformly distributed loads. The

distributed load corresponding to the self-weight of the structure qdead is approximated using Equation 6-1,

where # is unit weight of concrete, L is span, and V is the total volume of concrete in the system.

V
qdead = ! "
L
Equation 6-1
# % n
(
=! " V A
'& deck " L + A " S + Acol $ hi *
"
)
crown
L i=1

The volume of concrete that participates in the longitudinal behaviour of the system V can be estimated by

multiplying the deck, arch crown, and column cross sectional areas Adeck, Acrown, and Acol by their respective

lengths: span L, arc length S, and column height hi. The additional weight of transverse ribs, diaphragms,

stiffeners, and prestressing steel is accounted for by extra weight factor !V, which is calculated for all con-

cepts individually as the total weight of all structural components divided by the weight of all continuous,

longitudinal concrete V. A sample calculation of !V is given in Calculation D3 on page 319.

201
6. New Concepts for Arch Bridges

Figure 6-2. Superimposed dead load and traffic lanes

The design superimposed dead load qsdead is calculated based on the geometry of a standard Performance

Level 3 (pl-3) traffic barrier wall and a 40 mm layer of concrete pavement, as shown in Figure 6-2. The as-

sumed traffic barriers are intended to be conservative, since they are among the heaviest that would be

needed in typical Canadian bridge projects. Performance level requirements depend on factors such as

speed and volume of traffic, and distance from edge of paved roadway (see S6-06 §12.4.3.2.1). The barriers

are made from conventionally reinforced concrete, which suggests that with future research and develop-

ment it may be possible to reduce the barrier weight by using ultra high-performance fibre-reinforced

concrete.

The road surface is normally composed of an asphalt wearing surface and waterproofing membrane,

whose collective depth is typically about 90 mm in Ontario, Canada. In a superstructure rehabilitation of a

Swiss bridge using ultra high-performance fibre-reinforced concrete, Denarié et al. (2005) used only 40

mm of bituminous concrete pavement without a waterproofing membrane to top the road surface. They

were able to significantly reduce the road surface thickness due to the high resistance of ultra high per-

formance fibre-reinforced concrete against degradation and water penetration. This pavement thickness

was also used by Spasojević (2008) in her design study of thin ultra high-performance fibre-reinforced

concrete deck slabs. The same 40 mm reduced thickness is assumed for all concepts in the parametric

design study. Based on two PL-3 traffic barriers and a 40 mm pavement, the total superimposed dead load

was calculated to be 36.8 kN/m.

202
6. New Concepts for Arch Bridges

Truck and lane loads used for design are based on S6-06 §3.8.3.1 and §3.8.3.2. In the calculation of flexible

system moments, truck loads are idealized as single concentrated loads of 625 kN applied in each design

lane, instead of the five-axle CL-625-ont truck specified in S6-06 §A3.4. The simplification of five axles to

one axle is conservative, but not overly conservative since most of the span lengths considered in the study

are much larger than the 18 metre model truck length. The lane load is taken as a 9 kN/m uniformly dis-

tributed load combined with a 500 kN concentrated load. In accordance to S6-06 §3.8.4, multi-lane live

loads are obtained by multiplying single-lane live loads by the number of design lanes, the multi-lane

modification factor, and one plus the dynamic amplification factor, as given below:

Table 6-3. Design live loads

live load symbol single lane number of multi-lane modi- 1+dynamic ampli- multi-lane
value design lanes fication factor fication factor value

truck (concentrated) Qlive 625 kN 4 0.70 1.25 2190 kN


lane (concentrated) Qlive 500 kN 4 0.70 1.00 1400 kN
lane (distributed) qlive 9 kN/m 4 0.70 1.00 25.2 kN/m

Design values for uniform drops in temperature are based on S6-06 §3.9.4. The minimum effective tem-

perature for Toronto, Ontario is -33°C as given by S6-06 Figure A3.1.2 and S6-06 Table 3.7. Based on an as-

sumed effective construction temperature of 15°C and on the temperature modifications for arch depth d

(given in S6-06 Figure 3.5), the uniform drops in temperature ΔT in °C used for design are given below:

$ "48 0 # d # 0.4 m
&
! T = % 6.25d " 50.5 0.4 m < d # 2.0 m Equation 6-2
& "38 2.0 m < d
&'

Four combinations of load factors are considered: two at serviceability limit states, and two at ultimate lim-

it states. For both limit states, the effects of restrained deformations will first be excluded (young bridge

condition) and then included (old bridge condition). The load factors used for design are shown in Table

6-4. The effects of wind, seismic, and unbalanced multilane loading will not be considered.

203
6. New Concepts for Arch Bridges

Table 6-4. Design load factor combinations

load combinations dead live restrained critical


from s6-06 table 3.1 load !D load !L deformation !K sections

serviceability, sls 1 1.0 0.9 0 quarter-point


serviceability, sls 1 1.0 0.9 0.8 midspan
ultimate, uls 1 1.2 1.7 0 quarter-point
ultimate, uls 2 1.2 1.6 1.15 midspan

Material resistance factors are used to account for the inherent variability of material strength. The materi-

al resistance factor for prestressing steel ,p is taken as 0.95, in accordance with S6-06 §8.4.6. The material

resistance factor for ultra high-performance fibre-reinforced concrete ,c is assumed to be 0.75, which is

the same as the factor for conventional concrete in S6-06. A comprehensive material characterization is

needed to verify the validity of this assumed resistance factor for ultra high performance fibre-reinforced

concrete. Of the cylinders tests described in Table 2-3 on page 22, a maximum sample size of only 3 cylin-

ders were tested under the same age of concrete after casting and curing regime, which is insufficient for

any form of statistical analyses. In her study of composite behaviour of ultra high-performance fibre-rein-

forced concrete and reinforced concrete members, Habel (2004) used the same partial safety factors for

both conventional and ultra high-performance fibre-reinforced concrete. In the French Interim Recom-

mendations (AFGC 2002), §2.2 gives a partial safety factor of 1.3 for ultra high-performance fibre-rein-

forced concrete in tension. The reciprocal of this safety factor is 0.77, which can roughly be taken as an

equivalent resistance factor. Additional strength reduction factors are proposed in Interim Recommenda-

tions to account for localized effects of fibre orientation and methods of placement. These additional

factors are not considered in this study.

Only design quantities relevant to the longitudinal behaviour of the system are considered in the paramet-

ric study. For selected concepts, proportions of transverse structural members, diaphragms, and stiffeners

are estimated, but not rigorously designed. The following three sections will describe the strategies used in

designing and proportioning the three different arch systems considered.

204
6. New Concepts for Arch Bridges

6.1.2 Longitudinal proportioning of self-stiffened arch systems

Concepts 1 through 32 are designed as self-stiffened arch systems. In these systems, arches are designed to

be stiffer than decks so that bending moments caused by nonuniform loads and restrained deformations

are carried primarily by the arch. To investigate the limiting case where arches carry all flexible system

bending moments, deck girders are assumed to be simply supported between spandrel columns and are

therefore isolated from the rest of the system.

Figure 6-3. Proportioning of self-stiffened arch systems.


Note: Stiffeners, ribs, and floorbeams will be shown later.

Deck girders are designed as double-tee sections prestressed with internal, bonded post-tensioning, as

shown in Figure 6-3. As an alternative, external, unbonded post-tensioned deck systems are also possible,

but are considered outside the scope of this thesis. One advantage of this system is that much thinner webs

can be used, resulting in reduced concrete consumption and dead load.

205
6. New Concepts for Arch Bridges

The quantities to be designed for the internal, bonded post-tensioned double-tee deck girder are: (1) depth

of deck, (2) thickness of webs, and (3) amount of prestressing steel. In all cases, the deck slab thickness is

specified to be 50 mm. Transverse ribs are used to carry forces transversely to the longitudinal tee-beams

(not shown in Figure 6-3). As a first estimate, the deck depth is chosen such that the interior span-to-deck-

depth ratio is 20:1. The prestressing steel is given a parabolic profile such that the centroid of prestressing is

100 mm from the bottom fibre at midspan. The flexural capacity MR of the beam at midspan is taken con-

servatively as the effective prestressing force after all losses Fp! times the internal lever arm z between the

centroid of tendons and centroid of top slab, as given in Equation 6-3, where ,p is material resistance

factor of prestressing steel, Ap is total area of prestressing steel, fp! is effective prestress after all losses, fpu is

ultimate stress of prestressing steel, and d is depth of section:

M R = Fp! ! z
(
= "p ! Ap ! fp! ! z ) Equation 6-3

= (" p )
! Ap ! 0.6 fpu ! ( d # 125 mm )

Bending moment demands are calculated based on dead and live loads. Dead load calculations are based

on the weight of the deck girder, traffic barriers, concrete pavement, stiffeners, ribs, and floorbeams. Be-

cause the interior span lengths are smaller or of similar size to the model truck length of 18 m, it is overly

conservative to model truck loads as single concentrated loads. Thus bending moment demands are calcu-

lated using the cl-625 ont truck model (S6-06 §A3.4) with up to five axles.

Arches for these systems are designed as thin-walled single-cell box sections without permanent longitud-

inal prestressing, as shown in Figure 6-3. The quantities to be designed are: (4) depth of arch and (5) thick-

ness of top and bottom slabs. In all cases, web thickness is specified to be 50 mm. As a first estimate, the

arch depth is chosen such that the span-to-arch-depth ratio is 50:1. The thickness of top and bottom slabs

is first taken as 50 mm, and then increased in increments of 50 mm until the axial force-moment N-M*

member capacity interaction diagram of the arch exceeds demands. Member capacity interaction diagrams

for the arches are calculated using the general analysis method in Section 3.2. The model column length is
206
6. New Concepts for Arch Bridges

taken as the effective arc length kS of the bridge. Axial force and bending moment demands are calculated

using the methods described in Section 4.5. If no satisfactory design is found when slab thickness is in-

creased to 500 mm, then the depth of section is increased by decrementing the span-to-arch-depth ratio

from 50:1 to 45:1, and then to 40:1, as shown in Table 6-5. The ratio of 40:1 is chosen because it is near the

lower boundary of span-to-arch-depth ratios observed in existing concrete arch bridges, as shown in Fig-

ure 5-14 on page 174.

Table 6-5. Sequence of trial dimensions used in the design of hollow box sections

span-to- slab thick- span-to- slab thick- span-to- slab thick-


trial depth ratio ness in mm trial depth ratio ness in mm trial depth ratio ness in mm

1 50:1 50 11 45:1 50 21 40:1 50


2 50:1 100 12 45:1 100 22 40:1 100
3 50:1 150 13 45:1 150 23 40:1 150
4 50:1 200 14 45:1 200 24 40:1 200
5 50:1 250 15 45:1 250 25 40:1 250
6 50:1 300 16 45:1 300 26 40:1 300
7 50:1 350 17 45:1 350 27 40:1 350
8 50:1 400 18 45:1 400 28 40:1 400
9 50:1 450 19 45:1 450 29 40:1 450
10 50:1 500 20 45:1 500 30 40:1 500

6.1.3 Longitudinal proportioning of deck-stiffened arch systems

Concepts 33 through 64 are designed as deck-stiffened arch systems. In these systems, decks are designed

to be stiffer than arches so that bending moments caused by nonuniform loads and restrained deforma-

tions are carried primarily by the deck. In order to enforce this type of behaviour, trial designs are required

to have ratios of deck flexural rigidity to system flexural rigidity that are 0.95 or greater. Bending moment

demands in the arch and deck are calculated using the methods described in Section 4.5.

Deck girders are continuous over spandrel columns and are designed as thin-walled single-cell box sec-

tions with external, unbonded post-tensioning, as shown in Figure 6-4. The quantities to be designed are:

(1) depth of deck, (2) thickness of top and bottom slabs, and (3) amount of prestressing steel. Prestressing

tendons are anchored to diaphragms above each springing line, and are arranged inside the cavity of the
207
6. New Concepts for Arch Bridges

box such that they provide concentric prestress. The width of the box girder is taken as half the width of

the deck (8000 mm). Transverse ribs are used to carry forces transversely to the 50 mm webs of the

girders. As a first estimate, the deck depth is chosen such that the span-to-deck-depth ratio is 50:1. Trial di-

mensions for the deck girder follow the sequence given in Table 6-5. Only the portion of the deck slab

between the girders are thickened with each design iteration. The thickness of the cantilever portions of

the deck slab are kept constant at 50 mm. Member capacity interaction diagrams for the arch ribs are cal-

culated using the general analysis method in Section 3.2. The model column length is taken as the effective

arc length kS of the bridge (as opposed to the effective span length kL). The effects of prestressing steel is

neglected in member capacity calculations and are instead represented as concentric, externally applied

axial loads. Thus, the axial force N in the deck is taken to be the effective prestressing force after all losses

Fp!, which was previously given in Equation 6-3.

Figure 6-4. Proportioning of deck-stiffened arch systems.


Note: stiffeners, ribs, and floorbeams will be shown later.

208
6. New Concepts for Arch Bridges

Arches in deck-stiffened concepts are designed as solid or hollow twin ribs. Because they are stiffened by

the deck, the arch ribs can be designed to be very slender, as shown in Figure 6-4. The quantities to be de-

signed are: (4) depth of arch ribs, and (5) total width of arch ribs. If the ribs are made hollow, web and slab

thicknesses are assumed to be 200 mm. As a first estimate, the depth of ribs is chosen such that the span-

to-arch-depth ratio is 200:1. This trial ratio is chosen to be higher than 187, which is the highest ratio

among existing arch bridges in the database (see Figure 5-14 on page 174). The second trial span-to-arch-

depth ratio is 180:1. The total width of ribs is taken as 4 m (or one quarter of the deck width). If 4 m is un-

satisfactory, the total width of ribs is increased to 6 m.

single wave arch buckling mode multiple wave arch buckling mode

no rotation at joints rotation at joints

Figure 6-5. Buckling of slender arches over interior spans

Because these arches are stiffened by the deck, the spandrel columns act as lateral braces for the slender

arches, preventing global buckling. The arches, however, may buckle between the spandrel columns as

shown in Figure 6-5. Both single and multiple wave buckling modes are checked. If the arch section is kept

constant along the entire arch, then the arch legs nearest to the springing lines have the most critical

single-wave buckling modes. This is because these arch legs have the highest unbraced lengths and highest

axial forces among all legs in the arch polygon. Arches, however, are typically designed to be stiffer toward

the springing lines. Thus, single-wave buckling loads are calculated assuming that the most critical arch

legs are those nearest to the quarter-points. The effective length of the arch leg (kL)arch, then, is taken as

effective length factor k times the interior arc length S/n divided by the cosine of the angle of inclination '

at the quarter-point:

209
6. New Concepts for Arch Bridges

S 1
( kL )arch = k ! !
n cos "
2 Equation 6-4
S # f&
= k ! ! 4% ( +1
n $ L'

The effective length factor k for the fixed-fixed arch leg is taken as 0.5 and the number of interior spans n is

taken as 9. With these values, (kL)arch is used as the model column length for calculating member capacity

interaction diagrams for deck-stiffened arch ribs using the general analysis method.

6.1.4 Longitudinal proportioning of partially deck-stiffened arch systems

Concepts 65 through 96 are designed as partially deck-stiffened arch systems. In these systems, decks and

arches are designed to have about the same flexural rigidity, and thus share flexible system moments.

Bending moment demands are calculated for the arch and deck using the methods described in Section

4.5.

Figure 6-6. Proportioning of partially deck-stiffened arch systems


Note: Stiffeners, ribs, and floorbeams will be shown later.

210
6. New Concepts for Arch Bridges

As shown in Figure 6-6, the deck girders for this system are designed as thin-walled box girders in the

same way the deck girders are designed in Section 6.1.3. Accordingly, the quantities to be designed are: (1)

depth of deck, (2) thickness of top and bottom slab, and (3) amount of prestressing steel. The first trial deck

depth is chosen such that the span-to-deck-depth ratio is 100:1. The arch ribs are also designed as thin-

walled box girders and are constrained to be identical to the deck girders, except that they are not perman-

ently prestressed. Thicknesses of top and bottom slabs of both deck and arch are first taken as 50 mm.

Amount of prestressing steel in the deck is increased until bending moment demands are within the axial

force-moment N-M* member capacity interaction diagram, as calculated by the general analysis method

in Section 3.2.

To account for the mutual stiffening of the arch and deck, the model column length is taken as 0.707 times

the effective arc length kS, which has the effect of doubling the flexural rigidity of the deck or arch section,

whichever is being analyzed. If member capacities of the deck or arch are unsatisfactory, even with the aid

of deck prestressing, then slab thickness is increased in increments of 50 mm up to a maximum of 500

mm. If the design is still unsatisfactory, then the depths of deck and arch are increased such that their

span-to-depth ratios are 90:1, then 80:1, and then 70:1. Web thicknesses are kept constant at 50 mm.

211
6. New Concepts for Arch Bridges

The longitudinal proportioning of the deck and arch sections among all concepts are summarized below:

Table 6-6. Summary of quantities to be designed for parametric design study

system class element section type design quantity trial proportions

self-stiffened arch deck double tee (1) depth of deck L/n:d = 20:1
(2) thickness of webs 500 mm, 1000 mm
(3) amount of prestressing steel number of 0.6” strands

arch hollow box (4) depth of arch L:d = 50:1, 45:1, 40:1
(5) thickness of top/bottom slabs 50 mm to 500 mm

deck-stiffened arch deck hollow box (1) depth of deck L:d = 50:1, 45:1, 40:1
(2) thickness of top/bottom slabs 50 mm to 500 mm
(3) amount of prestressing steel number of 0.6” strands

arch solid/hollow box (4) depth of arch L:d = 200:1, 180:1


(5) total width of sections 4000 mm, 6000 mm
(6) hollow or solid webs and slabs thicknesses equal
200 mm for hollow sections

partially deck-stiffened arch deck/arch hollow box (1) depth of deck and arch L:d = 100:1, 90:1, 80:1, 70:1
(2) thickness of top/bottom slabs 50 mm to 500 mm

deck hollow box (3) amount of prestressing steel number of 0.6” strands

6.1.5 Wall slenderness

As webs and slabs are made thinner than previously built in conventionally reinforced concrete structures,

local stability problems become increasingly important. The design of hollow rectangular components is

dealt with in S6-06 §8.8.5.8. In this clause, wall slenderness ratios, which are defined as the wall length b

divided by wall thickness t, are used to determine whether a given wall will buckle locally before it can

reach its maximum compressive stress. According to this standard, wall slenderness ratios greater than 15

are subject to decreases in compressive strength by up to 25%. This clause originates from an analytical and

experimental study done by Taylor et al. (1995) on hollow thin-walled piers made from conventionally re-

inforced concrete. Their analysis treats compression flanges of hollow boxes as a thin, rectangular plates

subjected to compressive uniformly distributed loads along two edges, as shown by the diagrams on the

212
6. New Concepts for Arch Bridges

left in Figure 6-7. The support conditions along the edges parallel to the applied load are assumed to be re-

strained elastically against rotation (fixed). The bottom-left diagram shows the buckled shape of the plate.

According to Timoshenko’s Theory of Elastic Stability (1936), critical buckling stresses fE of thin plates can

be calculated using Equation 6-5, where kE is plate buckling coefficient, E is elastic modulus, * is Poisson’s

ratio, t is plate thickness, and b is width of plate:

! 2E $ t '
2

fE = kE & ) Equation 6-5


12 (1 " # 2 ) % b (

/2E £ t ¥
2 thin-walled ribbed ultra high-
fE  kE reinforced performance fibre-

12 1 < p 2 ¤ b ¦ concrete box reinforced concrete box

longitudinal
stiffener

transverse rib

t b t
b

idealized longitudinal pinned pinned


line support fixed fixed pinned pinned
idealized transverse
line support
pinned pinned

pinned

buckled shape

pinned pinned

uniformly distributed load uniformly distributed load

Figure 6-7. Plate buckling in thin-walled concrete box members

213
6. New Concepts for Arch Bridges

The value of the plate buckling coefficient kE depends on the aspect ratio, or width-to-length ratio, of the

plate and the support conditions along the edges of the plate. The minimum plate buckling coefficient for

plates that are simply supported along the edges perpendicular to the load and rotationally restrained

along the edges parallel to the load is 6.97.

In their study, Taylor et al. (1995) use the tangent modulus Etan=d//d% in place of the initial modulus of

elasticity to account for the material nonlinearities of concrete at high stresses. This relatively simple tan-

gent modulus approach was shown by Swarts and Rosebraugh (1974) to adequately model the buckling be-

haviour of thin concrete plates. By imposing that the critical buckling stress fE of the plate is greater than or

equal to the compressive strength of concrete fc! and rearranging Equation 6-5, an expression for maxim-

um wall slenderness ratio b/t is obtained:

b #2 E
! kE " " tan Equation 6-6
t (
12 1 $ % 2
) fc&

The top-right diagram of Figure 6-7 shows an ultra high-performance fibre-reinforced concrete box with

transverse ribs and longitudinal stiffeners (only a few are shown for clarity). Rotational restraint at the

web-flange corners of the hollow ribbed box is highest near the transverse ribs and lowest half way

between the transverse ribs. To simplify the problem, it is assumed that no rotational restraint is provided

along these corners. Thus, the compression flange of the box can be assumed to be simply supported along

all edges, which corresponds to a minimum plate buckling coefficient kE of 4. Substituting material proper-

ties for ultra high-performance fibre-reinforced concrete: Etan=21100 MPa, fc!=120 MPa, and *=0.20, results

in a maximum wall slenderness ratio of 24.5. This means that as wall thicknesses are reduced down to 50

mm, the maximum width of plate between webs needed to attain the full compressive strength of concrete

is 1230 mm, which is too small for the deck and arch members. Thus, longitudinal stiffeners are needed to

increase the allowable width of these thin plates. The longitudinal stiffeners shown in Figure 6-7 are as-

sumed to provide additional lines of pinned supports, effectively dividing the compression flange into mul-

214
6. New Concepts for Arch Bridges

tiple plates. Maximum plate widths for other plate thicknesses are tabulated in Table 6-7, and are used in

the design of the new concepts.

Table 6-7. Maximum plate widths for thin ultra high performance fibre-reinforced concrete plates

plate max. width plate max. width


thickness of plate thickness of plate
in mm in mm in mm in mm

50 1230 250 6140


100 2460 300 7360
150 3680 350 8590
200 4900 400 9820

6.1.6 Proportioning of spandrel columns

As discussed in Section 4.4.1 and illustrated in Figure 4-15 on page 106, spandrel columns can be assumed

to be pin-connected during preliminary design. According to S6-06 §8.8.5.3(g), compression members

with zero end bending moments must be designed with minimum end eccentricities of load of 15 mm plus

3% of the structural depth of the column. The appropriate analytical model for an eccentrically-loaded

column is shown in Figure 6-8. The general analysis method for calculating ultimate loads Q* of columns,

which is described in Section 3.2.5, was used to develop load capacity design curves for various hollow

cross-sections using ultra high-performance fibre-reinforced concrete. The input parameters for the calcu-

lation procedure were: column cross-section, material stress-strain curve, column length, and eccentricity

of load. The output result was ultimate compressive load Q* of the column. For each cross-section, the cal-

culation was repeated for a wide range of column lengths from zero up to the length corresponding to a

slenderness ratio ( of 200. Design curves for five hollow column sections are shown in Figure 6-8. Because

wall slenderness ratios of these sections are all below 24.5, local buckling problems are avoided.

215
6. New Concepts for Arch Bridges

factored ultimate
load capacity !c·Q* in MN 50

!c = 0.75 E

40
Q

30
D
e

L 20

C
B
10 data points calculated
A by program QULT using
Q general analysis method

structural 0
model 0 20 40 60 80 100 column length L in m

eccentricity 100

position
of load Q centroid
of section
50 100

60
50 100
45 45

30 30

b/t = 8 b/t = 3 b/t = 8 b/t = 13


b/t = 18

500 500 1000 1000 1500

Section A Section B Section C Section D Section E

Figure 6-8. Spandrel column load capacity design curves

Generalized column load capacity curves can be obtained by normalizing ultimate load capacities Q* by

cross sectional area of the column A and by the compressive strength of concrete fc!. Column length L is

normalized by column radius of gyration r to obtain slenderness ratios. As shown in Figure 6-9, normaliz-

ing all five design curves from Figure 6-8 results in very similar generalized design curves. The results

show that ultra high-performance fibre-reinforced concrete columns with slenderness ratios of 60 have

load capacities that are about half the failure load of the section in pure axial compression. These general-

ized column design curve can be used to estimate the load capacities of any hollow, rectangular ultra high-

performance fibre reinforced concrete column with pinned supports, as long as wall slenderness ratios are

kept below 24.5.

216
6. New Concepts for Arch Bridges

Q*
normalized load capacity 1
A·fc!

0.8

0.6

0.4

0.2

0
0 50 100 150 200 slenderness ratio "

Figure 6-9. Generalized column load capacity curves

6.2 Preliminary design concepts

This section reports the results of the parametric design study. Design quantities are tabulated in Section

6.2.1. Member capacity interaction diagrams of concept arches are also shown along with axial force-mo-

ment demands. Preliminary drawings of Concept 80, which is a partially deck-stiffened arch system with

main span of 400 m and arch rise of 50 m, is described in Section 6.2.3. The precast segmental construc-

tion method proposed for this concept is also described.

6.2.1 Parametric design solutions

Design concepts with span-to-rise ratios of 16:1 were found to be inherently inefficient because of the un-

avoidable large bending moment demands that arise due to the beam action caused by the shallowness of

the arch. No practical arch design solutions were found. Because of this, these concepts are absent from the

parametric design solutions that follow. The efficiency and practicality of very shallow arches will be furth-

er discussed in Section 6.3.

The proportions of 72 arch design concepts using ultra high-performance fibre-reinforced concrete are giv-

en in Table 6-8. The first group of columns in the table specify the design parameters that each concept

was designed for. The second group of columns in the table specify the proportions of the deck girders.

Girder widths are not specified for self-stiffened arch concepts because they have double-tee sections

217
6. New Concepts for Arch Bridges

rather than box sections. VSL Multistrand Systems are used as the longitudinal deck prestressing steel.

Steel quantities are specified by the number of required Type 0.6” grade 270 ksi/1860 MPa strands. The

third group of columns in the table specify the proportions of the arch. Web and slab thicknesses are not

specified for concepts with solid arch ribs. For the deck-stiffened concepts, arch width is reported as the

combined width of the two arch ribs. The last group of columns in the table specify the proportions of

spandrel columns. Width values for columns refer to the combined transverse width of the two columns

(i.e. 2.0 m means two columns that are each 1.0 m wide). Cross-sectional areas and moments of inertia are

reported as the sum of quantities of the two columns.

Table 6-9 tabulates geometric ratios and structural demands of the 72 design concepts. The first group of

columns tabulates the total volume of continuous, longitudinal concrete V according to Equation 5-13 on

page 188, and volume distributions among deck, arch, and spandrel columns. The second group of

columns tabulates quantities related to system stiffness, including: span-to-system-depth ratio (system

depth is the sum of deck and arch depths), system moment of inertia (sum of deck and arch moments of

inertia), system radius of gyration (system moment of inertia divided by cross sectional area of arch), and

column-to-system-moment of inertia ratio. The third group of columns tabulates slenderness and modi-

fied slenderness ratios (Equation 5-6 on page 179 and Equation 5-8 on page 184), equivalent slab thickness

by volume (Equation 5-12 on page 187), and normalized system moment of inertia (Equation 5-2 on page

176).

The fourth group of columns in Table 6-9 tabulates quantities related to the design dead load, or self-

weight of the structure. Extra weight factors are ratios between the total weight of all structural elements

(including ribs, diaphragms, stiffeners, and tendons) and the weight of continuous, longitudinal concrete.

A sample calculation of extra weight factor is shown in Calculation D3 on page 319. Based on these extra

weight factors, uniformly distributed dead loads and horizontal reactions due to dead load are calculated

for serviceability limit states (unfactored loads). Arch action values are calculated based on modified slen-

derness (f as given by the second factor in Equation 4-43 on page 115: 1/(1+45·(f-2).

218
6. New Concepts for Arch Bridges

Table 6-8. Proportions of design concepts


deck geometry arch geometry twin column geometry
no. of 0.6" strands
moment of inertia moment of inertia
cross sectional area cross sectional area moment of inertia
slab thickness slab thickness cross sectional area
web thickness web thickness wall thickness
width width max col. height
span: rise span-to-depth span-to-depth width
span rise depth depth depth
ID m m :1 m :1 m mm mm m2 10-3 m4 m :1 m mm mm m2 10-3 m4 m m m mm m2 10-3 m4
1 50 12.5 4 0.28 180 ‡ 1000 50 1.3 8 114 1.00 50 8 50 150 2.5 441 0.5 2.0 8 50 0.28 11
2 100 25.0 4 0.56 180 ‡ 1000 50 1.8 57 114 2.00 50 8 50 100 1.8 1490 0.5 2.0 15 50 0.28 11
3 150 37.5 4 0.83 180 ‡ 500 50 1.6 108 114 3.00 50 8 50 50 1.1 1940 0.5 2.0 23 50 0.28 11
4 200 50.0 4 1.11 180 ‡ 500 50 1.9 240 132 4.00 50 8 50 50 1.2 3610 1.0 2.0 30 50 0.38 57
5 250 62.5 4 1.39 180 ‡ 500 50 2.1 443 132 5.00 50 8 50 50 1.3 5880 1.0 2.0 38 50 0.38 57
6 300 75.0 4 1.67 180 ‡ 500 50 2.4 728 186 6.00 50 8 50 50 1.4 8790 1.0 2.0 45 100 0.72 98
7 350 87.5 4 1.94 180 ‡ 500 50 2.7 1090 186 7.00 50 8 50 50 1.5 12400 1.5 3.0 53 100 1.12 368
self-stiffened arch concepts

8 400 100.0 4 2.22 180 ‡ 500 50 3.0 1570 248 8.00 50 8 50 50 1.6 16700 1.5 3.0 60 100 1.12 368
9 50 6.3 8 0.28 180 ‡ 1000 50 1.3 8 114 1.25 40 8 50 250 4.1 1020 0.5 2.0 4 50 0.28 11
10 100 12.5 8 0.56 180 ‡ 1000 50 1.8 57 114 2.00 50 8 50 100 1.8 1490 0.5 2.0 8 50 0.28 11
11 150 18.8 8 0.83 180 ‡ 500 50 1.6 108 114 3.00 50 8 50 50 1.1 1940 0.5 2.0 11 50 0.28 11
12 200 25.0 8 1.11 180 ‡ 500 50 1.9 240 132 4.00 50 8 50 50 1.2 3610 0.5 2.0 15 100 0.52 17
13 250 31.3 8 1.39 180 ‡ 500 50 2.1 443 132 5.00 50 8 50 100 2.1 10500 1.0 2.0 19 50 0.38 57
14 300 37.5 8 1.67 180 ‡ 500 50 2.4 728 186 6.00 50 8 50 100 2.2 15600 1.0 2.0 23 50 0.38 57
15 350 43.8 8 1.94 180 ‡ 500 50 2.7 1090 186 7.00 50 8 50 100 2.3 21700 1.0 2.0 26 50 0.38 57
16 400 50.0 8 2.22 180 ‡ 500 50 3.0 1570 186 8.00 50 8 50 150 3.2 40800 1.0 2.0 30 50 0.38 57
17 50 4.2 12 0.28 180 ‡ 1000 50 1.3 8 114 1.25 40 8 50 250 4.1 1020 0.5 2.0 3 50 0.28 11
18 100 8.3 12 0.56 180 ‡ 1000 50 1.8 57 114 2.00 50 8 50 100 1.8 1490 0.5 2.0 5 50 0.28 11
19 150 12.5 12 0.83 180 ‡ 500 50 1.6 108 114 3.00 50 8 50 100 1.9 3550 0.5 2.0 8 50 0.28 11
20 200 16.7 12 1.11 180 ‡ 500 50 1.9 240 132 4.00 50 8 50 150 2.8 9320 0.5 2.0 10 50 0.28 11
21 250 20.8 12 1.39 180 ‡ 500 50 2.1 443 132 5.00 50 8 50 200 3.7 19300 0.5 2.0 13 100 0.52 17
22 300 25.0 12 1.67 180 ‡ 500 50 2.4 728 186 6.00 50 8 50 300 5.3 40300 1.0 2.0 15 50 0.38 57
23 350 29.2 12 1.94 180 ‡ 500 50 2.7 1090 186 7.00 50 8 50 400 7.0 71800 1.0 2.0 18 50 0.38 57
24 400 33.3 12 2.22 180 ‡ 500 50 3.0 1570 186 8.00 50 8 50 500 8.7 116000 1.0 2.0 20 50 0.38 57
33 50 12.5 4 1.00 50 8 50 50 1.3 249 160 0.25 200 4 ‡ ‡ 1.0 5 0.5 2.0 8 50 0.28 11
34 100 25.0 4 2.00 50 8 50 50 1.4 1090 160 0.50 200 4 ‡ ‡ 2.0 42 0.5 2.0 15 50 0.28 11
35 150 37.5 4 3.00 50 8 50 100 2.3 4270 160 0.83 180 4 200 200 1.9 171 1.0 2.0 23 50 0.38 57
36 200 50.0 4 4.00 50 8 50 100 2.4 7840 160 1.11 180 4 200 200 2.2 361 1.0 2.0 30 50 0.38 57
37 250 62.5 4 5.00 50 8 50 100 2.5 12600 160 1.39 180 4 200 200 2.4 635 1.0 2.0 38 50 0.38 57
38 300 75.0 4 6.00 50 8 50 150 3.4 25200 160 1.67 180 4 200 200 2.6 1000 1.5 3.0 45 100 1.12 368
39 350 87.5 4 7.00 50 8 50 150 3.5 34900 160 1.94 180 4 200 200 2.8 1470 1.5 3.0 53 100 1.12 368
deck-stiffened arch concepts

40 400 100.0 4 8.00 50 8 50 150 3.6 46400 240 2.22 180 4 200 200 3.1 2040 1.5 3.0 60 100 1.12 368
41 50 6.3 8 1.00 50 8 50 50 1.3 249 200 0.25 200 4 ‡ ‡ 1.0 5 0.5 2.0 4 50 0.28 11
42 100 12.5 8 2.00 50 8 50 50 1.4 1090 200 0.50 200 4 ‡ ‡ 2.0 42 0.5 2.0 8 50 0.28 11
43 150 18.8 8 3.00 50 8 50 100 2.3 4270 200 0.83 180 4 200 200 1.9 171 0.5 2.0 11 50 0.28 11
44 200 25.0 8 4.00 50 8 50 100 2.4 7840 200 1.11 180 4 200 200 2.2 361 0.5 2.0 15 100 0.52 17
45 250 31.3 8 5.00 50 8 50 100 2.5 12600 240 1.39 180 4 200 200 2.4 635 1.0 2.0 19 50 0.38 57
46 300 37.5 8 6.00 50 8 50 150 3.4 25200 320 1.67 180 4 200 200 2.6 1000 1.0 2.0 23 50 0.38 57
47 350 43.8 8 7.00 50 8 50 150 3.5 34900 320 1.94 180 4 200 200 2.8 1470 1.0 2.0 26 50 0.38 57
48 400 50.0 8 8.00 50 8 50 150 3.6 46400 400 2.22 180 4 200 200 3.1 2040 1.0 2.0 30 100 0.72 98
49 50 4.2 12 1.00 50 8 50 50 1.3 249 200 0.25 200 6 ‡ ‡ 1.5 8 0.5 2.0 3 50 0.28 11
50 100 8.3 12 2.00 50 8 50 100 2.2 1800 320 0.50 200 6 ‡ ‡ 3.0 63 0.5 2.0 5 50 0.28 11
51 150 12.5 12 3.00 50 8 50 150 3.1 5800 320 0.75 200 6 ‡ ‡ 4.5 211 0.5 2.0 8 50 0.28 11
52 200 16.7 12 4.44 45 8 50 200 4.0 16700 440 1.00 200 6 ‡ ‡ 6.0 500 0.5 2.0 10 100 0.52 17
53 250 20.8 12 6.25 40 8 50 200 4.2 34400 550 1.25 200 6 ‡ ‡ 7.5 977 0.5 2.0 13 50 0.28 11
54 300 25.0 12 7.50 40 8 50 200 4.3 50700 770 1.50 200 6 ‡ ‡ 9.0 1690 1.0 2.0 15 50 0.38 57
55 350 29.2 12 8.75 40 8 50 200 4.4 70200 770 1.75 200 6 ‡ ‡ 10.5 2680 1.0 2.0 18 50 0.38 57
56 400 33.3 12 10.00 40 8 50 250 5.4 111000 990 2.00 200 6 ‡ ‡ 12.0 4000 1.0 2.0 20 50 0.38 57
65 50 12.5 4 0.50 100 8 50 100 2.0 82 200 0.50 100 8 50 100 1.6 66 0.5 2.0 8 50 0.28 11
66 100 25.0 4 1.00 100 8 50 100 2.1 403 200 1.00 100 8 50 100 1.7 330 0.5 2.0 15 50 0.28 11
67 150 37.5 4 1.50 100 8 50 100 2.1 975 200 1.50 100 8 50 100 1.7 804 1.0 2.0 23 50 0.38 57
68 200 50.0 4 2.00 100 8 50 100 2.2 1800 200 2.00 100 8 50 100 1.8 1490 1.0 2.0 30 50 0.38 57
69 250 62.5 4 2.50 100 8 50 100 2.2 2900 200 2.50 100 8 50 100 1.8 2410 1.0 2.0 38 50 0.38 57
partially deck-stiffened arch concepts

70 300 75.0 4 3.00 100 8 50 100 2.3 4270 200 3.00 100 8 50 100 1.9 3550 1.0 2.0 45 100 0.72 98
71 350 87.5 4 3.50 100 8 50 100 2.3 5910 240 3.50 100 8 50 100 1.9 4920 1.5 3.0 53 100 1.12 368
72 400 100.0 4 4.00 100 8 50 100 2.4 7840 320 4.00 100 8 50 100 2.0 6540 1.5 3.0 60 100 1.12 368
73 50 6.3 8 0.50 100 8 50 100 2.0 82 200 0.50 100 8 50 100 1.6 66 0.5 2.0 4 50 0.28 11
74 100 12.5 8 1.00 100 8 50 100 2.1 403 200 1.00 100 8 50 100 1.7 330 0.5 2.0 8 50 0.28 11
75 150 18.8 8 1.50 100 8 50 100 2.1 975 200 1.50 100 8 50 100 1.7 804 0.5 2.0 11 50 0.28 11
76 200 25.0 8 2.00 100 8 50 100 2.2 1800 200 2.00 100 8 50 100 1.8 1490 0.5 2.0 15 100 0.52 17
77 250 31.3 8 2.50 100 8 50 150 3.0 3930 240 2.50 100 8 50 150 2.6 3410 1.0 2.0 19 50 0.38 57
78 300 37.5 8 3.00 100 8 50 200 3.9 7210 320 3.00 100 8 50 200 3.5 6430 1.0 2.0 23 50 0.38 57
79 350 43.8 8 3.89 90 8 50 150 3.2 10100 440 3.89 90 8 50 150 2.8 8780 1.0 2.0 26 50 0.38 57
80 400 50.0 8 4.44 90 8 50 200 4.0 16700 440 4.44 90 8 50 200 3.6 15000 1.0 2.0 30 100 0.72 98
81 50 4.2 12 0.63 80 8 50 100 2.0 140 200 0.63 80 8 50 100 1.6 112 0.5 2.0 3 50 0.28 11
82 100 8.3 12 1.25 80 8 50 150 2.9 862 200 1.25 80 8 50 150 2.5 738 0.5 2.0 5 50 0.28 11
83 150 12.5 12 1.88 80 8 50 200 3.7 2580 240 1.88 80 8 50 200 3.3 2280 0.5 2.0 8 50 0.28 11
84 200 16.7 12 2.50 80 8 50 300 5.4 6460 440 2.50 80 8 50 300 5.0 5900 0.5 2.0 10 50 0.28 11
85 250 20.8 12 3.57 70 8 50 500 8.7 20400 440 3.57 70 8 50 500 8.3 19200 1.0 2.0 13 100 0.72 98
86 300 25.0 12 4.29 70 8 50 500 8.7 30800 550 4.29 70 8 50 500 8.3 29100 1.0 2.0 15 50 0.38 57
87 350 29.2 12 5.00 70 8 50 500 8.8 43500 660 5.00 70 8 50 500 8.4 41200 1.0 2.0 18 50 0.38 57
88 400 33.3 12 5.71 70 8 50 500 8.9 66900 770 5.71 70 8 50 500 8.5 55400 1.0 2.0 20 50 0.38 57
‡ bullet indicates that decks are double-tee ‡ bullet indicates that arch ribs are solid
sections

219
6. New Concepts for Arch Bridges

Table 6-9. Geometric ratios and structural demands of design concepts


equivVODEWKLFNQHVV normalized system moment of inertia
PRGLÀHGVOHQGHUQHVVUDWLR +HV\VWHPPRPHQW¸0I
slenderness ratio OLYHORDGV\VWHPPRPHQW¸0I
FROXPQLQHUWLDV\VWHPLQHUWLD &67V\VWHPPRPHQW¸0I
system radius of gyration max system moment at ULS, Mf
system moment of inertia PD[GHÁHFWLRQ¸ULVHDW8/6
VSDQDUFKGHFNGHSWK JRYHUQLQJORDGFDVH
FROXPQYROXPH SLS dead H
DUFKYROXPH DUFKDFWLRQ
GHFNYROXPH SLS dead UDL
total volume extra weight
ID m3 % % % :1 10-3 m4 m % m m MN/m % MN % MN·m % % %
1 219 29 65 7 50 441 0.42 2.5 48 59 0.27 0.69 1.52 0.17 99 4 5 0.6 11.2 0 98 2
2 414 44 49 7 50 1490 0.91 0.8 44 55 0.26 1.04 1.56 0.16 99 8 5 0.6 22.9 0 97 3
3 468 51 40 9 50 1940 1.33 0.6 45 56 0.19 1.13 1.74 0.14 99 10 5 0.8 35.8 0 93 7
4 723 51 38 11 50 3610 1.74 1.6 46 57 0.23 1.39 1.68 0.15 99 15 4 0.9 49.1 0 88 12
5 1000 53 37 10 50 5880 2.13 1.0 47 59 0.25 1.64 1.62 0.17 99 20 4 1.0 69.3 0 84 16
6 1420 51 34 15 50 8790 2.51 1.1 48 60 0.30 1.88 1.54 0.19 99 28 4 1.2 94.4 0 80 20
7 1940 49 31 21 50 12400 2.88 3.0 49 61 0.35 2.10 1.48 0.21 99 36 4 1.4 126.0 0 75 25
VHOIVWLIIHQHGDUFKFRQFHSWV

8 2380 50 31 19 50 16700 3.24 2.2 50 62 0.37 2.32 1.46 0.22 99 44 4 1.5 161.0 0 72 28
9 282 22 75 3 40 1020 0.50 1.1 36 25 0.35 0.91 1.42 0.20 93 10 3 1.5 21.3 54 43 3
10 381 48 49 4 50 1490 0.91 0.8 40 27 0.24 1.04 1.59 0.15 94 15 3 2.0 31.2 31 59 10
11 429 55 40 5 50 1940 1.33 0.6 41 28 0.18 1.13 1.81 0.13 95 19 3 2.9 48.4 24 57 19
12 673 55 37 8 50 3610 1.74 0.5 42 29 0.21 1.39 1.70 0.15 95 28 2 3.5 75.3 26 47 27
13 1120 48 48 4 50 10500 2.25 0.5 41 28 0.28 1.99 1.53 0.17 95 41 2 2.7 115.0 34 41 25
14 1460 50 47 4 50 15600 2.68 0.4 41 28 0.30 2.27 1.51 0.19 95 53 2 3.1 165.0 34 37 29
15 1840 51 45 4 50 21700 3.09 0.3 41 28 0.33 2.53 1.48 0.20 95 66 2 3.5 226.0 32 34 34
16 2590 46 51 3 50 40800 3.59 0.1 41 28 0.40 3.13 1.40 0.23 95 87 2 3.1 308.0 37 30 33
17 275 23 75 2 40 1020 0.50 1.1 36 17 0.34 0.91 1.43 0.20 86 13 3 3.1 28.0 60 37 3
18 372 49 49 3 50 1490 0.91 0.8 39 18 0.23 1.04 1.60 0.15 88 20 3 4.4 42.3 37 48 15
19 539 44 53 3 50 3550 1.37 0.3 39 18 0.22 1.39 1.64 0.15 88 30 3 5.1 74.2 38 42 20
20 955 39 59 2 50 9320 1.83 0.1 39 18 0.30 1.91 1.49 0.18 88 48 2 5.0 128.0 45 32 23
21 1510 35 62 3 50 19300 2.30 0.1 39 18 0.38 2.44 1.39 0.21 88 71 2 5.1 207.0 47 27 26
22 2390 30 68 2 50 40300 2.75 0.1 39 18 0.50 3.11 1.31 0.27 88 105 2 4.8 331.0 52 22 26
23 3490 27 72 1 50 71800 3.20 0.1 39 18 0.62 3.78 1.25 0.32 88 147 2 4.8 495.0 54 18 28
24 4780 25 74 1 50 116000 3.65 0.0 39 18 0.75 4.43 1.21 0.37 88 195 2 4.9 708.0 55 16 29
33 136 47 42 10 40 254 0.50 4.4 40 50 0.17 0.58 1.75 0.13 98 3 5 0.7 11 0 98 2
34 397 35 58 7 40 1130 0.75 1.0 53 67 0.25 0.95 1.52 0.18 99 9 5 0.7 23 0 96 4
35 735 47 46 8 39 4440 1.51 1.3 40 50 0.31 1.49 1.44 0.23 98 17 3 0.6 39 28 66 6
36 1050 45 47 7 39 8200 1.94 0.7 41 51 0.33 1.83 1.42 0.24 98 24 3 0.7 54 28 63 9
37 1400 44 49 7 39 13200 2.35 0.4 43 53 0.35 2.15 1.40 0.26 98 32 3 0.7 69 28 61 11
38 2250 45 40 15 39 26200 3.17 1.4 38 47 0.47 2.70 1.30 0.34 98 51 2 0.7 100 32 54 14
39 2750 44 41 15 39 36400 3.58 1.0 39 49 0.49 3.01 1.30 0.36 98 62 2 0.7 126 31 53 16
GHFNVWLIIHQHGDUFKFRQFHSWV

40 3290 43 43 14 39 48400 3.98 0.8 40 50 0.51 3.31 1.29 0.38 98 74 2 0.8 158 30 51 19
41 124 52 42 6 40 254 0.50 4.4 36 25 0.15 0.58 1.82 0.11 93 5 3 2.2 14 28 66 6
42 361 38 58 4 40 1130 0.75 1.0 48 33 0.23 0.95 1.57 0.17 96 16 3 2.4 31 30 58 12
43 667 51 46 3 39 4440 1.51 0.3 36 25 0.28 1.49 1.48 0.20 93 29 3 2.2 59 39 47 14
44 980 49 46 5 39 8200 1.94 0.2 37 26 0.31 1.83 1.44 0.23 94 42 3 2.5 86 39 43 18
45 1290 48 48 4 39 13200 2.35 0.4 39 27 0.32 2.15 1.43 0.24 94 56 2 2.7 122 37 39 24
46 1880 54 43 3 39 26200 3.17 0.2 34 24 0.39 2.70 1.37 0.29 93 80 2 2.7 186 41 34 25
47 2310 53 45 3 39 36400 3.58 0.2 36 24 0.41 3.01 1.36 0.30 93 99 2 2.9 243 38 32 30
48 2850 50 45 5 39 48400 3.98 0.2 37 25 0.45 3.31 1.34 0.33 93 122 2 3.3 321 36 29 35
49 146 44 52 3 40 257 0.41 4.4 43 20 0.18 0.58 1.70 0.13 90 9 3 4.7 18 34 56 10
50 533 41 57 2 40 1860 0.79 0.6 45 21 0.33 1.12 1.39 0.24 91 33 3 4.5 49 45 40 15
51 1160 40 59 1 40 6010 1.16 0.2 46 22 0.48 1.65 1.28 0.36 91 73 3 4.8 97 48 30 22
52 2060 39 59 2 37 17200 1.69 0.1 42 20 0.64 2.35 1.23 0.47 90 127 2 4.7 185 55 21 24
53 2980 35 64 1 33 35400 2.17 0.0 41 19 0.75 2.98 1.21 0.55 89 183 2 4.5 312 59 17 24
54 4080 32 67 1 33 52400 2.41 0.1 44 21 0.85 3.40 1.20 0.62 91 254 2 5.3 430 51 16 33
55 5340 29 70 1 33 72900 2.63 0.1 47 22 0.95 3.80 1.18 0.70 92 337 2 5.9 603 46 14 40
56 7080 30 69 1 33 115000 3.10 0.0 46 22 1.11 4.42 1.16 0.81 91 444 2 6.1 868 46 12 42
65 209 48 45 7 50 147 0.30 7.6 67 83 0.26 0.48 1.55 0.19 99 5 5 0.9 12 0 95 5
66 429 48 45 7 50 732 0.66 1.5 61 76 0.27 0.82 1.55 0.20 99 10 5 0.9 24 0 93 7
67 675 47 44 9 50 1780 1.01 3.2 59 74 0.28 1.10 1.53 0.21 99 15 5 0.9 37 0 90 10
68 922 47 44 8 50 3290 1.36 1.7 59 74 0.29 1.35 1.53 0.21 99 21 4 1.0 51 0 83 17
69 1180 47 45 8 50 5310 1.70 1.1 59 73 0.30 1.59 1.53 0.22 99 27 4 1.2 74 0 77 23
SDUWLDOO\GHFNVWLIIHQHGDUFKFRQFHSWV

70 1550 44 42 14 50 7820 2.04 1.3 59 74 0.32 1.80 1.49 0.24 99 35 4 1.5 105 0 71 29
71 1990 41 39 20 50 10800 2.37 3.4 59 74 0.36 2.01 1.46 0.26 99 45 4 1.9 147 0 63 37
72 2320 41 39 20 50 14400 2.70 2.6 60 74 0.36 2.21 1.47 0.27 99 53 4 2.3 195 0 58 42
73 193 52 44 4 50 147 0.30 7.6 61 42 0.24 0.48 1.58 0.18 97 9 3 3.0 12 17 70 13
74 397 52 44 4 50 732 0.66 1.5 55 38 0.25 0.82 1.57 0.18 97 18 3 3.1 28 19 61 20
75 611 52 44 4 50 1780 1.01 0.6 54 37 0.25 1.10 1.57 0.19 97 27 3 3.6 49 19 54 27
76 859 51 43 6 50 3290 1.36 0.5 54 37 0.27 1.35 1.55 0.20 97 38 2 4.5 78 18 42 40
77 1480 51 46 3 50 7340 1.67 0.8 54 37 0.37 1.77 1.41 0.27 97 66 2 4.4 125 20 35 45
78 2300 50 47 3 50 13600 1.98 0.4 55 38 0.48 2.17 1.33 0.35 97 102 2 4.7 196 21 28 51
79 2180 51 46 3 45 18900 2.62 0.3 49 33 0.39 2.42 1.44 0.29 96 96 2 4.9 252 22 29 49
80 3250 49 46 5 45 31700 2.97 0.3 49 34 0.51 2.88 1.34 0.37 96 144 2 5.2 370 22 24 54
81 190 54 44 2 40 252 0.39 4.5 45 21 0.24 0.57 1.64 0.17 91 12 3 5.0 17 31 56 13
82 553 52 46 2 40 1600 0.80 0.7 45 21 0.35 1.06 1.44 0.25 91 35 3 4.9 46 38 43 19
83 1090 52 47 1 40 4860 1.20 0.2 44 21 0.45 1.54 1.35 0.33 91 68 3 5.2 88 40 33 27
84 2110 51 48 1 40 12400 1.58 0.1 45 21 0.66 2.10 1.25 0.48 91 132 2 5.5 167 43 22 35
85 4330 50 49 1 35 39600 2.19 0.2 41 19 1.08 3.10 1.16 0.80 89 265 2 5.0 353 53 15 32
86 5200 50 49 1 35 59900 2.68 0.1 40 19 1.08 3.55 1.17 0.80 89 317 2 5.6 509 51 14 35
87 6120 50 49 1 35 84700 3.18 0.1 39 18 1.09 3.99 1.17 0.80 88 372 2 6.2 707 48 13 39
88 7050 50 49 1 35 122000 3.79 0.0 38 18 1.10 4.51 1.15 0.81 87 424 1 6.8 1060 46 9 45

220
6. New Concepts for Arch Bridges

The last group of columns tabulate maximum deflections and maximum flexible system bending moments

at ultimate limit states, as caused by the five load combinations shown in Figure 4-21 on page 127. These

maximum moments are then separated into the following components: (1) moments caused by creep,

shrinkage, and temperature, labelled as “CST system moment,” (2) moments caused by live load, and (3) ad-

ditional moments caused by second-order effects, labelled as “H-e system moment.” Comparisons among

all design concepts will be discussed in greater detail in Section 6.3, using the data presented in Tables 6-8

and 6-9.

The adequacy of given trial designs were assessed using N-M* interaction diagrams, as calculated by the

general method described in Section 3.2. Diagrams were calculated using factored material resistances.

Five load combinations at ultimate limit states were used for calculating maximum, second-order sectional

forces using the simplified analysis methods described in Section 4.5. Trial designs were deemed satisfact-

ory if member capacities exceeded demands.

Interaction diagrams for the arch members among all design concepts are presented in Figures 6-10, 6-11,

and 6-12. All axial force N results have been divided by the cross-sectional area of the arch A to obtain axial

stress, and then normalized by the factored compressive strength of concrete ,c fc!:

N 1
n= ! Equation 6-7
A "c fc#

All ultimate moment M* results have been divided by section modulus of the arch 2I÷d to obtain effective

flexural stress at the extreme fibre, and then normalized by half the factored compressive strength of

concrete:

M* d 1
m* = ! ! Equation 6-8
I 2 "c fc#

Two sets of sectional force demands are plotted on each interaction diagram. Solid markers represent sec-

tional forces at ultimate limit state while hollow markers represent sectional forces at serviceability limit

states. Because the interaction diagrams were calculated based on factored material resistance, the section-
221
6. New Concepts for Arch Bridges

al forces at serviceability limit states were also factored down by the material resistance of concrete. This

results in graphically-equivalent representations of unfactored demands and unfactored capacity, which is

normally associated with serviceability limit states. Instead of comparing with member capacity interac-

tion diagrams (solid line), unfactored sectional forces at serviceability limit states are compared with con-

tours of cracking, represented by dashed lines on the diagrams. The contour of cracking is defined as the

set of sectional forces at which initial cracking occurs at the flexural tension face of the section. Decks and

arches among all concepts were designed to remain uncracked under service loads so that deformations of

the structure remain small and in the linear range. This also serves to maintain a margin of safety between

service and ultimate loads.

Two other criteria were used to determine wether given trial solutions were satisfactory or not. The first

criterion is related to admissible axial forces imposed on the section. Axial force demands and axial forces

introduced by prestressing were kept below half the maximum load under uniform compression. This en-

sured that additional deflections and moments due to second order effects were not excessively large. It

also ensures some minimum level of ductility, since flexural failures would be initiated by the progressive

pull-out of steel fibres, rather than the crushing of concrete. The second criterion is related to admissible

deflections at ultimate limit states. Trial designs were dismissed if deflections at ultimate limit states were

greater than 5% of the initial rise of the arch. Like the first criterion, this ensured that second-order effects

were not excessively large. More importantly, this deflection criterion ensured the applicability of the sim-

plified calculation methods used to calculate sectional force demands. When crown deflections are larger

than 5%, the reduction of the rise of the arch imposes additional moments on the system, which are not ac-

counted for by the simplified methods presented in Chapter 4.

222
sample
interaction diagram
superposition of
normalized 1
arch interaction diagrams
axial force:
from all self-stiffened concepts
N 1 1
n= · n
A !c·fc!

member capacity N-M*


interaction diagram
0.5 (factored resistance)

contour of cracking (factored)


0.5
sectional forces caused by
critical load combinations at ULS

sectional forces caused by


critical load combinations at SLS
0 reduced by material factor !c
0 0.5 1 1.5
M·d 1 0
normalized moment: m = ·
2I !c·fc! 0 0.5 1 1.5
m

223
concepts L 50 m 100 m 150 m 200 m 250 m 300 m 350 m 400 m
L:f

1-8 4:1

9-16 8:1

Figure 6-10. Interaction diagrams of all self-stiffened arch concepts


17-24 12:1
6. New Concepts for Arch Bridges
sample
interaction diagram
superposition of
normalized 1
arch interaction diagrams
axial force:
from all deck-stiffened concepts
N 1 1
n= · n
A !c·fc!

member capacity N-M*


interaction diagram
0.5 (factored resistance)

contour of cracking (factored)


0.5
sectional forces caused by
critical load combinations at ULS

sectional forces caused by


critical load combinations at SLS
0 reduced by material factor !c
0 0.5 1 1.5

M·d 1 0
normalized moment: m = ·
2I !c·fc! 0 0.5 1 1.5
m

224
concepts L 50 m 100 m 150 m 200 m 250 m 300 m 350 m 400 m
L:f

33-40 4:1

41-48 8:1

Figure 6-11. Interaction diagrams of all deck-stiffened arch concepts


49-56 12:1
6. New Concepts for Arch Bridges
sample
interaction diagram
superposition of
normalized 1
arch interaction diagrams from all
axial force:
partially deck-stiffened concepts
N 1 1
n= · n
A !c·fc!

member capacity N-M*


interaction diagram
0.5 (factored resistance)

contour of cracking (factored)


0.5
sectional forces caused by
critical load combinations at ULS

sectional forces caused by


critical load combinations at SLS
0 reduced by material factor !c
0 0.5 1 1.5

M·d 1 0
normalized moment: m = ·
2I !c·fc! 0 0.5 1 1.5
m

225
concepts L 50 m 100 m 150 m 200 m 250 m 300 m 350 m 400 m
L:f

65-72 4:1

73-80 8:1

Figure 6-12. Interaction diagrams of all partially deck-stiffened arch concepts


81-88 12:1
6. New Concepts for Arch Bridges
6. New Concepts for Arch Bridges

Figure 6-13. Sections of selected design concepts and existing arch bridges

226
6. New Concepts for Arch Bridges

Figure 6-13 shows typical cross-sections of twelve of the proposed concepts. The last column of diagrams

shows typical cross-sections of existing concrete arch bridges with spans similar to those of the concepts in

the same row of diagrams. The proposed concepts shown are proportioned for span-to-rise ratios of 8:1.

Longitudinal stiffeners are spaced so that local buckling does not occur before the maximum compressive

stress of concrete is reached. Transverse and vertical ribs are proportioned to improve the ability of the

sections to transmit transverse and torsional loads. The cross sectional areas of continuous longitudinal

concrete have been lightly shaded to distinguish them from the transverse and vertical ribs. The shading

allows for visual comparisons of the quantity of concrete used among concepts and existing concrete arch

bridges. Diaphragms, which would normally correspond to sections taken through the centres of the span-

drel columns, are not shown.

6.2.2 Sample design calculations including shear

Manual design calculations for Concept 80 are included in Appendix D. These calculations demonstrate

the use of the simplified design methods for calculating sectional force demands in arch systems and cal-

culating member capacity using the reduced n-m interaction diagram. Calculations for two ultimate limit

state load combinations are considered. In both cases, the resulting second-order sectional forces were

shown to be resisted safely by the given arch and deck sections (see Calculation D12 on page 328). These

calculations can be used as a guideline for designers, and as a model example for implementing calcula-

tions on a spreadsheet.

In this parametric study, it has been assumed that axial forces and bending moments govern the design of

arch systems. Because arches carry loads primarily in axial compression, this assumption is typically valid.

For this reason, the effects of shearing forces in arch systems have not been rigorously considered. In order

to validate the use of very thin 50 mm webs without conventional stirrups, a rudimentary shear design

check is shown in Calculations D16 to D18, starting on page 332. Ultimate limit state load combination 1

from S6-06 is used, in which the lane load is applied over half the span along with a uniformly distributed

dead load. Axial force, bending moment, and shear force diagrams are calculated along the arch span. Us-

227
6. New Concepts for Arch Bridges

ing a linear-elastic approach, axial and shear stresses are calculated using properties of the arch section in

Calculation D16. Using Mohr’s circle of plane stress, all principal stresses in the arch member were found

to be less than the factored cracking stress of the material ,c·ft of 5.25 MPa. The most critical (i.e. the most

positive) principal stresses occur near the springing lines, where sectional forces are highest.

It may be the case that there are other design concepts and other load cases that have principal stresses that

exceed the cracking limit at service loads or the tensile strength limit at ultimate loads. The concepts that

would be most susceptible are: (1) concepts that are shallow, because of the beam action under permanent

loads, and (2) concepts that have low arch prestress (i.e. short spans), since precompression helps to reduce

tensile stresses.

For these cases, shear capacity can be increased by increasing web thickness or by prestressing the member

vertically. Increasing web thickness is undesirable since it detracts from the goal of minimizing concrete

consumption and minimizing weight. If webs are thickened, they can be thickened locally in areas of high

shear, rather than uniformly over the whole span. Using vertical prestressing would be the better option,

since it does not add much additional weight to the structure. The vertical compressive stresses introduced

by prestressing would cause the principal stresses in a member to become more compressive (i.e more neg-

ative in a Mohr’s circle diagram). Prestressing steel could be installed at each transverse rib. This would not

only help prevent shear cracks under service loads, but would also create a well-defined truss-like flow of

forces should the structure be cracked at ultimate limit states. Vertical prestressing steel would carry

tensile forces upward, while the webs would carry diagonal compressive forces along the web panels, simil-

ar to how forces are carried in conventionally reinforced concrete beams. This concludes a brief discussion

on the shear resistance and shear design of the proposed arch systems.

6.2.3 Concept 80

In order to present a more detailed design of one of the proposed concepts, preliminary drawings for

Concept 80 have been developed. Plan, elevation, and sections views of the bridge concept are shown in

Figure 6-14. The arch has a span of 400 m, a span-to-rise-ratio of 8:1, and is partially stiffened by the deck.

228
6. New Concepts for Arch Bridges

The ground profile is fictitious and has been adapted to suit the geometry of the arch. The geometry of the

arch centreline is defined in Figure 6-15.

As shown by Section A-A, the deck girder has a constant depth of 4.44 m and is composed of identical pre-

cast segments, each 3 m long. The arch is also made from 3 m long segments that are nearly identical to the

deck girder segments, except for the cantilevers. The arch rib has a constant depth of 4.44 m, except for the

outermost spans near the springing lines, where the arch is made deeper and thicker to accommodate

higher sectional force demands.

According to S6-06 §8.8.5.3, tensile stress across segmental joints without continuous reinforcing bars is

limited to zero. Most of the of the design concepts developed in this study satisfy this requirement. For the

15 out of 72 concepts that do not, alternative joint types or an increase in longitudinal post-tensioning will

be required.

As shown by Sections B-B and C-C, diaphragms are provided above and below each set of columns. These

diaphragm segments can be constructed using pre-cast or cast-in-place methods. The former allows for

high quality of concrete and higher geometry control, while the latter allows for greater flexibility in ad-

justing for geometrical imperfections and misalignments during construction. The proposed spandrel

columns are hollow and thin, which provides for a high degree of visual transparency. The short columns

near the arch crown will probably cause problems for seismic loading. A viable alternative would be to fuse

the arch and deck at the crown. This provides a direct load path for longitudinal seismic-induced inertial

forces to be transferred from the massive deck to the arch, and then from the arch to the foundations.

Along the transverse and vertical ribs of the deck and arch segments, holes for the main longitudinal post-

tensioning are provided. For the deck, twenty VSL Type 6-22 Multistrand tendons are positioned along the

perimeter of the segment cavity to provide concentric prestress. VSL Type E Anchors for these tendons are

shown in Section E-E in Figure 6-15. Thicker diaphragms are provided above the springing lines to accom-

modate for the disturbed regions near the anchors. Figure 6-15 also shows longitudinal sections of half the

arch span, revealing the arrangement and quantity of segments.

229
6. New Concepts for Arch Bridges

Figure 6-14. General arrangement drawing

230
6. New Concepts for Arch Bridges

Figure 6-15. Longitudinal section and anchorage diaphragm section

231
6. New Concepts for Arch Bridges

The trade-off of using minimum materials in highly efficient systems is that the resulting geometrical

shapes are not easily built. Incorporating transverse ribs and longitudinal stiffeners into the segments

poses potential problems for the precasting process. The problem with the proposed ribs and stiffeners is

that they intersect along three dimensions. Every waffle-like recess requires five abutting surfaces to be

formed. The successful release of these forms is another source of complexity. These challenges require a

new and innovative precasting method. One possible segmental precasting method is presented below.

The proposed precasting method involves two separate casting procedures, one for the end ribs of the seg-

ment, and one for the body of the segment between the two ribs. The first procedure, consisting of four

stages, is illustrated in Figure 6-17. The first stage consists of preparing all the components of the form. A

custom-built steel bulkhead is used to form the face of the principal end rib. Block-outs for shear keys are

built into the bulkhead. Steel angles are fastened to the bulkhead to form the perimeter of the rib shape.

Block-outs for prestressing tendon holes are also fastened to the bulkhead. Two lengths of corrugated steel

duct, which will be used later for the transverse post-tensioning of the deck, are installed into the form.

Deformed steel wire is arranged among the areas of the ribs that will abut against the inside body of the

segment. The quantity of deformed steel wire is chosen such that its effective yield force is equal to or

greater than the longitudinal tensile strength of the surrounding concrete.

Once all these components have been setup, the principal end rib is cast and cured (Stage 2). Once the

principal end rib has gained sufficient strength, it is carefully lifted, turned over, and set upon spacer

blocks (Stage 3). This exposes the underside of the rib, which has the recessed impression of the steel bulk-

head. A second set of steel angles is used to line the perimeter of the principal end rib. A second set of

post-tensioning ducts and deformed wires are also installed. The secondary abutting end rib is then cast

directly against the surface of the principal end rib (Stage 4). This technique, called match-casting, pro-

duces optimal bond quality between the contact surfaces of the two ribs. These perfectly match-catch sur-

faces will later be rejoined on site during the erection of the bridge. Once the match-cast end rib has

gained sufficient strength, the ribs are transported to the area dedicated to the second casting procedure.

232
6. New Concepts for Arch Bridges

Figure 6-16. Proposed match-casting method for abutting end ribs

233
6. New Concepts for Arch Bridges

The procedure for casting the interior of the segment consists of eight stages, and is illustrated in Figure 6-

16. This figure is intended as a schematic representation of the procedure, rather than a detailed drawing.

The outside base form for the 3 metre segment is represented by the gray, U-shaped polygon. This form

along with the inside core form would consist of an array of steel sections, plates and stiffeners. It would be

designed so that several components could be retracted away from the completed segment.

The first stage of the procedure is the assembly of the base form, cantilever form, and core form. In the

second stage, precast end ribs are positioned on either side of the core form with their shear keys facing

outward. The two precast ends are not a match-cast pair, since they will be on opposing sides of a segment.

The third stage consists of suspending the core form and clamping it to the end ribs using the holes inten-

ded for prestressing tendons in the final structure. Once the forms are sufficiently sealed, the interior body

of the segment is cast (Stage 4). No vibration is required, since the concrete is self-consolidating. The cold

construction joints between the end ribs and body are effectively bonded through the embedded deformed

wires. Once the segment has gained sufficient strength, the core form is unclamped and released, and re-

moved in parts, starting with the top and bottom slab forms, then the web forms, and last the corner forms

(Stages 5 through 7). The geometry of the segment should be adjusted to ease the removal of forms. The

rectangular stiffeners and right-angled corners shown in the figure should be modified as tapered stiffeners

and sloped corners to reduce the friction between the form and cast segment. The final stage consists of

lifting the completed segment out of the base form and then moving it to a storage area.

The proposed partially deck-stiffened systems, including Concept 80, have arch segments that are nearly

identical to their deck segments. This provides unique cost-savings in terms of the cost of the custom bulk-

head and forms. The repetitive precasting procedure, which is the same for the arch and deck, should res-

ult in increased economies of scale and improved efficiency in production, relative to the other arch sys-

tems considered in this study.

234
6. New Concepts for Arch Bridges

Figure 6-17. Proposed method for precasting segments

235
6. New Concepts for Arch Bridges

Renderings of one assembled precast segment and of multiple match-cast segments for Concept 80 are

shown below. An average height human figure is shown to establish a sense of scale.

Figure 6-18. Renderings of precast segments proposed in Concept 80

236
6. New Concepts for Arch Bridges

6.3 Discussion of design concepts

This section compares the design concepts with respect to the three design parameters: span length, arch

rise, and distribution of flexural rigidity among arch and deck. In order to illustrate the similarities and di-

fferences among concepts, various design quantities will be organized in the form of a comparison matrix

as shown in Table 6-10. The effects of span length can be evaluated by comparing values in the same row.

The effects of arch rise can be evaluated by comparing values in each column. The effects of deck-stiffening

can be evaluated by comparing the triplets of values in each column.

Table 6-10. Sample comparison matrix

values in comparison matrix: concept IDs


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 1 2 3 4 5 6 7 8 •each triplet of values
33 to 40 deck-stiffened 4:1 33 34 35 36 37 38 39 40 shows the effect of
65 to 72 partially deck-stiffened 4:1 65 66 67 68 69 70 71 72 changes in deck-stiffening
9 to 16 self-stiffened 8:1 9 10 11 12 13 14 15 16 •each row shows the effect
41 to 48 deck-stiffened 8:1 41 42 43 44 45 46 47 48 of changes in span length
73 to 80 partially deck-stiffened 8:1 73 74 75 76 77 78 79 80
17 to 24 self-stiffened 12:1 17 18 19 20 21 22 23 24 •each column shows the
49 to 56 deck-stiffened 12:1 49 50 51 52 53 54 55 56 effect of changes in arch rise
81 to 88 partially deck-stiffened 12:1 81 82 83 84 85 86 87 88 unit

6.3.1 Proportions of self-stiffened arch concepts

Span-to-deck-depth ratios of self-stiffened concepts are shown in Table 6-11. All deck girders were sat-

isfactorily designed with span-to-depth ratios of 180:1. Given the constant number of interior spans of 9 as-

sumed for all concepts, this ratio corresponds to interior-span-to-depth ratios of 20:1, which is not uncom-

mon for simply supported girder-type structural systems. Six vsl type 0.6” multistrand tendons (three per

web) with 19 to 31 strands each were found to be satisfactory in resisting bending moments at ultimate lim-

it states. Instead of increasing the amount of strands per tendon, eight tendons were used in the 400 m

span concepts to keep the tendon anchorages small. All tendons are internally bonded within the webs of

the double-tee sections. Webs were made 500 mm or 1000 mm wide to allow space for tendons, an-

chorages, and their cover requirements.

237
6. New Concepts for Arch Bridges

Table 6-11. Proportions of self-stiffened arch concepts


span-to-deck-depth ratios of self-stiffened arch concepts
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 180 180 180 180 180 180 180 180 •180:1 corresponds to
9 to 16 self-stiffened 8:1 180 180 180 180 180 180 180 180 interior span-to-depth
17 to 24 self-stiffened 12:1 180 180 180 180 180 180 180 180 :1 ratios of 20:1

span-to-arch-depth ratios of self-stiffened arch concepts


L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 50 50 50 50 50 50 50 50
9 to 16 self-stiffened 8:1 40 50 50 50 50 50 50 50
17 to 24 self-stiffened 12:1 40 50 50 50 50 50 50 50 :1

slab thicknesses of arch ribs in self-stiffened arch concepts, in mm


L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 150 100 50 50 50 50 50 50 •values associated with
9 to 16 self-stiffened 8:1 250 100 50 50 100 100 100 150 deeper 40:1 arch ribs
17 to 24 self-stiffened 12:1 250 100 100 150 200 300 400 500 mm are in red type

Table 6-11 also shows span-to-arch-depth ratios and slab thicknesses of the self-stiffened arch concepts.

Most arches were satisfactorily designed with span-to-depth L:d ratios of 50:1, except for the 50 m concepts

with span-to-rise ratios of 8:1 and 12:1. These concepts with short spans were found to have relatively low

dead load precompression in the arch. Thus deeper arch sections were needed to compensate for the low

flexural resistance of the arch. Results from these concepts also showed a sensitivity to cracking at service-

ability limit states, which is undesirable.

For the arch section, top and bottom slab thicknesses of 50 mm were satisfactory for most concepts with

span-to-rise ratios L:f of 4:1, while 100 mm thicknesses were satisfactory for most concepts with L:f=8:1.

Slab thicknesses up to 500 mm were required for concepts with L:f=12:1 to keep ultimate deflections below

5% of initial arch rise. These shallow arch concepts were not made deeper than L:d=50:1 to decrease deflec-

tions because their modified slenderness ratios were already below the efficiency threshold of 20.

Given that slab thickness requirements are low among concepts with span-to-rise ratios of 4:1 and 8:1, it is

likely possible that arches in these concepts could be made more slender by increasing their span-to-depth

ratios (i.e. to 60:1). This reduction in lever arm would require increases in compressive force, and thus in-

creases in slab thickness. Deflections may also increase due to the increased flexibility of the system, which

would limit how slender the arches could be made.

238
6. New Concepts for Arch Bridges

6.3.2 Proportions of deck-stiffened arch concepts

Span-to-deck-depth ratios and slab thicknesses of deck-stiffened arch concepts are shown in Table 6-12.

Most deck girders were satisfactorily designed with span-to-depth ratios of 50:1. Among these designs, 50

mm top and bottom slab thicknesses were used for concepts with 50-100 m spans, 100 mm slab thick-

nesses for 150-250 m spans, and 150 mm slab thicknesses for 300-400 m spans. Concepts with span-to-rise

ratios of 12:1 required greater slab thicknesses to compensate for larger moment demands. Of these, con-

cepts with spans of ≥200 m were made deeper to reduce deflections. For some of these concepts, it was not

possible to keep deflections below 5% of initial arch rise, which lessens the validity of the simplified calcu-

lation methods used for design. Performing more complex finite element analyses on these deflection-

sensitive concepts may result in more severe bending moment demands and deflections than those calcu-

lated in this parametric study.

Table 6-12. Proportions of deck-stiffened arch concepts

span-to-deck-depth ratios of deck-stiffened arch concepts


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
33 to 40 deck-stiffened L:f = 4:1 50 50 50 50 50 50 50 50
41 to 48 deck-stiffened 8:1 50 50 50 50 50 50 50 50
49 to 56 deck-stiffened 12:1 50 50 50 45 40 40 40 40 :1

slab thicknesses of deck girders in deck-stiffened arch concepts, in mm


L = 50 100 150 200 250 300 350 400 m
33 to 40 deck-stiffened L:f = 4:1 50 50 100 100 100 150 150 150 •values associated with
41 to 48 deck-stiffened 8:1 50 50 100 100 100 150 150 150 deeper 45:1 and 40:1
49 to 56 deck-stiffened 12:1 50 100 150 200 200 200 200 250 mm decks are in red type

span-to-arch-depth ratios of deck-stiffened arch concepts


L = 50 100 150 200 250 300 350 400 m
33 to 40 deck-stiffened L:f = 4:1 200 200 180 180 180 180 180 180 •180:1 arch ribs are hollow
41 to 48 deck-stiffened 8:1 200 200 180 180 180 180 180 180 •200:1 arch ribs are solid
49 to 56 deck-stiffened 12:1 200 200 200 200 200 200 200 200 :1

Table 6-12 also shows span-to-arch-depth ratios of deck-stiffened arch concepts. Due to the stiffening

provided by the deck girder, slender solid arch ribs with span-to-depth ratios of 200:1 and total rib widths

of 4 m were found to be satisfactory for concepts with low span-to-rise ratios and spans of ≤100 m. For

concepts with spans >100 m, solid arch ribs were found to consume more concrete than necessary. Thus,

arch ribs were made hollow to lighten the dead load and made deeper to maintain the flexural rigidity re-

239
6. New Concepts for Arch Bridges

quired to resist buckling between spandrel columns. Because most of the flexible system moments get dis-

tributed to the deck, all arches in these deck-stiffened systems remain uncracked at ultimate limit states (as

shown in Figure 6-11 on page 224). For concepts with span-to-rise ratios of 12:1, the total width of arch ribs

had to be increased by 50% and were kept solid. Increasing the cross sectional area and axial rigidity of the

arch allowed the modified slenderness of the system to stay above the efficiency threshold of 20. If axial ri-

gidity were not increased, the arches would be too shallow and be subjected to large bending moments

caused by beam action. These arch shallowness requirements increase the material consumption of the

system significantly relative to the concepts with low span-to-rise ratios.

6.3.3 Proportions of partially deck-stiffened arch systems

Span-to-deck-depth ratios and slab thicknesses of deck-stiffened arch concepts are shown in Table 6-13.

Most concepts with span-to-rise ratios of 4:1 and 8:1 were satisfactorily designed with decks and arches

with span-to-depth ratios of 100:1 and with slab thicknesses of 100 to 200 mm. For concepts with span-to-

rise ratios of 12:1, deeper sections and thicker slabs were needed to keep deflections below 5% of initial arch

rise.

Table 6-13. Proportions of partially deck-stiffened arch concepts

span-to-depth ratios of arches and decks in partially deck-stiffened arch concepts


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
65 to 72 partially deck-stiffened L:f = 4:1 100 100 100 100 100 100 100 100 ‡DUFKHVDQGGHFNVDUH
73 to 80 partially deck-stiffened 8:1 100 100 100 100 100 100 90 90 are given the same depth
81 to 88 partially deck-stiffened 12:1 80 80 80 80 70 70 70 70 :1

slab thicknesses of arches and decks in partially deck-stiffened arch concepts, in mm


L = 50 100 150 200 250 300 350 400 m
65 to 72 partially deck-stiffened L:f = 4:1 100 100 100 100 100 100 100 100 ‡DUFKHVDQGGHFNVDUH
73 to 80 partially deck-stiffened 8:1 100 100 100 100 150 200 150 200 are given the same
81 to 88 partially deck-stiffened 12:1 100 150 200 300 500 500 500 500 mm slab thicknesses

240
6. New Concepts for Arch Bridges

6.3.4 Design interaction diagrams

Normalized member capacity n-m* interaction diagrams for all concept arches are collectively shown in

Figure 6-19. These diagrams were calculated using the general method, and then normalized according to

Equations 6-7 and 6-8. Normalized contours of cracking are also shown as dashed lines.

The shapes of the n-m interaction diagrams are very similar for all concepts. The diagrams for the deck-sti-

ffened arches tend to have higher normalized moments than arches of the other two systems. This is a res-

ult of the shape of section used: in the deck-stiffened systems, solid or hollow sections with relatively walls

were used, while in the other two systems, hollow sections with thin walls were used.

In Figure 6-19, reduced n-m design interaction diagrams, calculated using the simplified method in Section

3.5, are superimposed on to the member capacity n-m* interaction diagrams calculated using the general

method. The figure shows that in all cases, the proposed design interaction diagram is both conservative

and accurate relative to the interaction diagrams based on more refined calculations. It should be noted

that the value of normalized axial force at Euler buckling nE is given the default value of 0.8 in the figure.

This value is changed according to the slenderness of the column, effectively reducing the dark gray area

above n=0.5. Calibrating this value would result in reduced interaction diagrams that follow the n-m* dia-

grams more closely than what is shown in the figure.

The figure also shows that very few design solutions had normalized axial force demands that were greater

than 0.5. This is a result of the design process, in which arches with very high axial-force demands were

subjected to large secondary moments. This is reflected in Vianello’s deflection magnifier equation, in

which additional secondary deflections and moments increase exponentially as the load approaches the

buckling resistance of the member. Thus, very few arches with normalized axial force demands greater

than 0.5 were capable of carrying the corresponding large second-order moments.

241
6. New Concepts for Arch Bridges

design interaction diagram


n 1
admissible forces for
serviceability limit states
nE admissible forces for
ultimate limit states

superposition of
arch interaction diagrams
0.5
nt —ncr from all self-stiffened concepts
n 1

design contour of cracking


design interaction diagram

0
ncr 0.5 1 1.5 0.5
nt
m
mn=0
QE
nE = = normalized axial force at Euler buckling, but not > 0.8
A·fc!
f
ncr = cr = cracking stress divided by compressive strength
fc!
f!
nt! = t = tensile strength divided by compressive strength 0
fc!
0 0.5 1 1.5
mn=0 = normalized ultimate moment when N=0
m

superposition of superposition of
arch interaction diagrams arch interaction diagrams from all
from all deck-stiffened concepts partially deck-stiffened concepts
n 1 n 1

0.5 0.5

0 0
0 0.5 1 1.5 0 0.5 1 1.5
m m

Figure 6-19. Comparison of simplified and general method interaction diagrams

242
6. New Concepts for Arch Bridges

6.3.5 Proportions of spandrel columns

The structural depths of spandrel columns among all concepts are shown in Table 6-14. In most cases,

500-1000 mm column depths with wall thicknesses of 50-100 mm were capable of carrying the axial forces

and moments caused by construction tolerances and misalignments of load. Deeper 1.5 m column sections

were needed for concepts with taller spandrel columns, including those with spans >300 m and span-to-

rise ratios of 4:1. Transverse loading such as wind loading and temporary construction loads may increase

the required proportions of these columns.

Table 6-14. Proportions of spandrel columns among all concepts


column depths among all concepts, in m
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 0.5 0.5 0.5 1.0 1.0 1.0 1.5 1.5 ‡FROXPQVDUHKROORZ
33 to 40 deck-stiffened 4:1 0.5 0.5 1.0 1.0 1.0 1.5 1.5 1.5 ZLWKZDOOWKLFNQHVVHV
65 to 72 partially deck-stiffened 4:1 0.5 0.5 1.0 1.0 1.0 1.0 1.5 1.5 of 50 mm or 100 mm
9 to 16 self-stiffened 8:1 0.5 0.5 0.5 0.5 1.0 1.0 1.0 1.0
41 to 48 deck-stiffened 8:1 0.5 0.5 0.5 0.5 1.0 1.0 1.0 1.0
73 to 80 partially deck-stiffened 8:1 0.5 0.5 0.5 0.5 1.0 1.0 1.0 1.0
17 to 24 self-stiffened 12:1 0.5 0.5 0.5 0.5 0.5 1.0 1.0 1.0
49 to 56 deck-stiffened 12:1 0.5 0.5 0.5 0.5 0.5 1.0 1.0 1.0
81 to 88 partially deck-stiffened 12:1 0.5 0.5 0.5 0.5 1.0 1.0 1.0 1.0 m

6.3.6 Material efficiency

Equivalent slab thicknesses based on volume of continuous, longitudinal concrete are shown in Table 6-15.

Minimum values among concepts with the same span and rise are highlighted. Among concepts with

spans of 100 m or less, deck-stiffened arches consume the least amount of concrete. For example, Concept

41 achieves an effective slab thickness of 0.15 m, which is 37% less than the thickness of the other two sys-

tems for a span of 50 m. This implies that deck-stiffened arches are the most materially efficient over short

spans. Given the accuracy of the analysis method and resolution of design iterations, all systems can be

considered to be more or less equally efficient for spans of 100 m. For spans of ≥150 m, the data shows that

self-stiffened arch systems consume the least amount of concrete in almost all cases.

243
6. New Concepts for Arch Bridges

Table 6-15. Equivalent slab thicknesses based on volume of longitudinal concrete among all concepts

equivalent slab thicknesses among all concepts, in m


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 0.27 0.26 0.19 0.23 0.25 0.30 0.35 0.37 ‡PLQLPXPYDOXHVDPRQJ
33 to 40 deck-stiffened 4:1 0.17 0.25 0.31 0.33 0.35 0.47 0.49 0.51 triplets arein red type
65 to 72 partially deck-stiffened 4:1 0.26 0.27 0.28 0.29 0.30 0.32 0.36 0.36
9 to 16 self-stiffened 8:1 0.35 0.24 0.18 0.21 0.28 0.30 0.33 0.40 ‡PLQLPXPYDOXHVLQHDFK
41 to 48 deck-stiffened 8:1 0.15 0.23 0.28 0.31 0.32 0.39 0.41 0.45 column are in bold type
73 to 80 partially deck-stiffened 8:1 0.24 0.25 0.25 0.27 0.37 0.48 0.39 0.51
17 to 24 self-stiffened 12:1 0.34 0.23 0.22 0.30 0.38 0.50 0.62 0.75
49 to 56 deck-stiffened 12:1 0.18 0.33 0.48 0.64 0.75 0.85 0.95 1.11
81 to 88 partially deck-stiffened 12:1 0.24 0.35 0.45 0.66 1.08 1.08 1.09 1.10 m
database various various 0.61 0.65 0.69 0.72 0.90 1.30 1.91 2.74 m ‡PLQLPXPVODEWKLFNQHVVHV
from arch database

equivalent slab thicknesses as percentages of


minimum values achieved by concrete arches in database, in %
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 45 40 28 31 28 23 18 14
33 to 40 deck-stiffened 4:1 28 38 45 45 39 36 26 19
65 to 72 partially deck-stiffened 4:1 43 41 41 40 33 25 19 13
9 to 16 self-stiffened 8:1 58 37 26 29 31 23 17 15 ‡YDOXHVJUHDWHUWKDQ
41 to 48 deck-stiffened 8:1 25 35 40 42 36 30 22 16 50% are in red type
73 to 80 partially deck-stiffened 8:1 40 38 37 37 41 37 20 19
17 to 24 self-stiffened 12:1 57 36 33 41 42 38 33 27
49 to 56 deck-stiffened 12:1 30 51 70 89 83 66 50 40
81 to 88 partially deck-stiffened 12:1 39 53 66 91 120 84 57 40 %

Reducing the concrete consumption and overall weight of arch bridges is the primary objective of this

thesis. Thus, equivalent slab thicknesses of the design concepts are shown again in Table 6-15, this time ex-

pressed as percentages of the envelope of minimum equivalent slab thicknesses observed among arches in

the database from Chapter 5. Results show that reductions of concrete volume by 50% or more can be read-

ily be achieved using any one of the three arch systems considered, as long as span-to-rise ratios are kept in

the range of 4:1-8:1. In this range, arches are not likely to be shallow, or sensitive to axial deformations of

the arch. For shallow arch concepts with span-to-rise ratios of 12:1, most deck-stiffened and partially deck-

stiffened concepts were unable to achieve a reduction of concrete volume by 50%. Thus the cost of building

these concepts are not likely to be much different than arches built using conventional concrete techno-

logy. If designers do consider using shallow arches, equivalent slab thickness results show that self-sti-

ffened arches are likely to be the most economical choice among the three arch systems considered in this

study.

A graphical comparison of the consumption of concrete among arches made from ultra high-performance

fibre-reinforced concrete and arches made from conventional concrete is shown in Figure 6-20. This figure

244
6. New Concepts for Arch Bridges

clearly shows the significant material efficiencies offered by the proposed ultra high-performance fibre-re-

inforced concrete arch systems. In most cases, the proposed concepts represent reductions in concrete

volume much greater than 50% when compared to the heavier concrete arch bridges in the database.

equivalent slab thickness in m 3


= volume / (span × deck width)
volume = deck area × span
+ arch area × arc length
+ column area × total spandrel
" " column length
existing concrete arches 2
arch design concepts minimum envelope of
values for existing
concrete arch bridges

0
0 100 200 300 400 500 span length in m

Figure 6-20. Comparison of equivalent slab thicknesses of new arch concepts and existing arches in database

Material efficiency can also be discussed in terms of quantity of main longitudinal prestressing steel in the

deck girder among concepts. Table 6-16 shows that in most cases, self-stiffened arch systems use the least

amount of prestressing. This is not unexpected since the decks in self-stiffened systems carry the least mo-

ments among the three systems.

Table 6-16. Quantity of longitudinal prestressing strands in deck girder among all concepts

number of required 0.6" steel prestressing strands among all concepts


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 114 114 114 132 132 186 186 248 •minimum values among
33 to 40 deck-stiffened 4:1 160 160 160 160 160 160 160 240 triplets are in red type
65 to 72 partially deck-stiffened 4:1 200 200 200 200 200 200 240 320
9 to 16 self-stiffened 8:1 114 114 114 132 132 186 186 186
41 to 48 deck-stiffened 8:1 200 200 200 200 240 320 320 400
73 to 80 partially deck-stiffened 8:1 200 200 200 200 240 320 440 440
17 to 24 self-stiffened 12:1 114 114 114 132 132 186 186 186
49 to 56 deck-stiffened 12:1 200 320 320 440 550 770 770 990
81 to 88 partially deck-stiffened 12:1 200 200 240 440 440 550 660 770

6.3.7 Distribution of concrete volume

Table 6-17 shows the relative concrete volume of arch and deck among all concepts. One might expect that

the stiffening element in the system would naturally require more material than the flexible element it sti-

245
6. New Concepts for Arch Bridges

ffens. Results show, however, that the arch and deck typically have about the same concrete volume regard-

less of which element stiffens the system. This can be possibly be explained by the additional limitations

imposed on the flexible elements in the systems. For example, flexible deck elements in self-stiffened arch

systems are required to transmit loads in bending over interior spans, and thus require some nominal

amount of concrete to achieve this. In deck-stiffened arch systems, flexible arch elements are required to

resist buckling between spandrel columns.

Table 6-17. Relative volume of arch and deck among all concepts
(volume of arch) / (volume of arch + deck) among all concepts, in %
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m

1 to 8 self-stiffened L:f = 4:1 69 53 44 42 41 40 39 38


33 to 40 deck-stiffened 4:1 47 62 49 51 53 47 48 50
65 to 72 partially deck-stiffened 4:1 48 48 48 48 49 49 49 49

9 to 16 self-stiffened 8:1 77 51 42 40 50 48 47 53
41 to 48 deck-stiffened 8:1 45 60 47 49 50 45 46 47
73 to 80 partially deck-stiffened 8:1 46 46 46 46 47 48 48 48

17 to 24 self-stiffened 12:1 77 50 55 60 64 69 73 75
49 to 56 deck-stiffened 12:1 54 58 60 60 65 68 71 70
81 to 88 partially deck-stiffened 12:1 45 47 48 49 49 49 49 49 %

Table 6-18 shows the relative concrete volume of spandrel columns to the total concrete volume of the sys-

tem. Results show that the weight of columns are typically small in comparison to that of the deck and

arch. This, however, is not be true for concepts with very tall piers, as illustrated by the concepts with spans

of ≥300 m and span-to-rise ratios of 4:1. These volume ratios were calculated based on the sectional geo-

metry required for the tallest spandrel columns within each design concept. Thus these values represent an

upper limit, since the proportions of the shorter spandrel columns can be reduced.

Table 6-18. Relative volume of spandrel columns to total volume


(volume of columns) / (volume of system) among all concepts, in %
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 7 7 9 11 10 15 21 19 •values greater than 10%
33 to 40 deck-stiffened 4:1 10 7 8 7 7 15 15 14 are in red type
65 to 72 partially deck-stiffened 4:1 7 7 9 8 8 14 20 20
9 to 16 self-stiffened 8:1 3 4 5 8 4 4 4 3
41 to 48 deck-stiffened 8:1 6 4 3 5 4 3 3 5
73 to 80 partially deck-stiffened 8:1 4 4 4 6 3 3 3 5
17 to 24 self-stiffened 12:1 2 3 3 2 3 2 1 1
49 to 56 deck-stiffened 12:1 3 2 1 2 1 1 1 1
81 to 88 partially deck-stiffened 12:1 2 2 1 1 1 1 1 1%

246
6. New Concepts for Arch Bridges

6.3.8 System slenderness and shallowness

Span-to-system depth ratios among all concepts are shown in Table 6-19. Most of the proposed concepts

have ratios of 39:1-50:1. Figure 6-21 compares span-to-system-depth ratios of the concepts with those from

the concrete arch database in Chapter 5. The figure shows that the proposed concepts have ratios that are

more or less in the same range as those in the database. Comparing the highest values in the figure, con-

cepts with span-to-system-depth ratios of about 50:1 are on par with the thinnest arch systems ever built.

Table 6-19. Span-to-system-depth ratios among all concepts


span-to-system-depth ratios among all concepts
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m

1 to 8 self-stiffened L:f = 4:1 50 50 50 50 50 50 50 50


33 to 40 deck-stiffened 4:1 40 40 39 39 39 39 39 39
65 to 72 partially deck-stiffened 4:1 50 50 50 50 50 50 50 50

9 to 16 self-stiffened 8:1 40 50 50 50 50 50 50 50
41 to 48 deck-stiffened 8:1 40 40 39 39 39 39 39 39
73 to 80 partially deck-stiffened 8:1 50 50 50 50 50 50 45 45

17 to 24 self-stiffened 12:1 40 50 50 50 50 50 50 50 ‡YDOXHVRXWVLGHRIWKHUDQJH


49 to 56 deck-stiffened 12:1 40 40 40 37 33 33 33 33 WRDUHXQGHUOLQHG
81 to 88 partially deck-stiffened 12:1 40 40 40 40 35 35 35 35 :1

span-to-system-depth ratio 80:1


existing concrete arches
arch design concepts
60:1
system = arch + deck
max ratio
average ratio 40:1
min ratio
20:1

0:1
0 100 200 300 400 500 span length in m

Figure 6-21. Comparison of span-to-system-depth ratios

Normalized system moments of inertia among all concepts (as calculated by Equation 5-2 on page 176) are

shown in Table 6-20. Results show that partially deck-stiffened systems tend to have the lowest system mo-

ments of inertia among concepts with span-to-rise ratios of 4:1 and 8:1. This is not unexpected since the

moment of inertia of a single full-depth member (i.e. the arch in self-stiffened systems) is greater than two

half-depth members (i.e. the deck and arch in partially deck-stiffened systems). Among concepts with

span-to-rise ratios of 12:1, system moments of inertia are generally lowest in self-stiffened systems. This is

because members in the other systems had to be deepened to reduce crown deflections.

247
6. New Concepts for Arch Bridges

Table 6-20. Normalized system moment of inertia among all concepts


normalized system moment of inertia among all concepts, in m
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 0.69 1.04 1.13 1.39 1.64 1.88 2.10 2.32 •minimum values among
33 to 40 deck-stiffened 4:1 0.58 0.95 1.49 1.83 2.15 2.70 3.01 3.31 triplets are in red type
65 to 72 partially deck-stiffened 4:1 0.48 0.82 1.10 1.35 1.59 1.80 2.01 2.21
9 to 16 self-stiffened 8:1 0.91 1.04 1.13 1.39 1.99 2.27 2.53 3.13
41 to 48 deck-stiffened 8:1 0.58 0.95 1.49 1.83 2.15 2.70 3.01 3.31
73 to 80 partially deck-stiffened 8:1 0.48 0.82 1.10 1.35 1.77 2.17 2.42 2.88
17 to 24 self-stiffened 12:1 0.91 1.04 1.39 1.91 2.44 3.11 3.78 4.43
49 to 56 deck-stiffened 12:1 0.58 1.12 1.65 2.35 2.98 3.40 3.80 4.42
81 to 88 partially deck-stiffened 12:1 0.57 1.06 1.54 2.10 3.10 3.55 3.99 4.51 m

System radii of gyration among all concepts are shown in Table 6-21. The last rows of the table give the

range of values that can be expected of these new concepts for arch bridges. Low values of system radii of

gyration are preferred since they have high axial rigidity, and are the least likely to have problems associ-

ated with arch shallowness.

Table 6-21. System radius of gyration among all concepts

system radius of gyration among all arch concepts, in m


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m

1 to 8 self-stiffened L:f = 4:1 0.42 0.91 1.33 1.74 2.13 2.51 2.88 3.24
33 to 40 deck-stiffened 4:1 0.50 0.75 1.51 1.94 2.35 3.17 3.58 3.98
65 to 72 partially deck-stiffened 4:1 0.30 0.66 1.01 1.36 1.70 2.04 2.37 2.70

9 to 16 self-stiffened 8:1 0.50 0.91 1.33 1.74 2.25 2.68 3.09 3.59
41 to 48 deck-stiffened 8:1 0.50 0.75 1.51 1.94 2.35 3.17 3.58 3.98
73 to 80 partially deck-stiffened 8:1 0.30 0.66 1.01 1.36 1.67 1.98 2.62 2.97

17 to 24 self-stiffened 12:1 0.50 0.91 1.37 1.83 2.30 2.75 3.20 3.65
49 to 56 deck-stiffened 12:1 0.41 0.79 1.16 1.69 2.17 2.41 2.63 3.10
81 to 88 partially deck-stiffened 12:1 0.39 0.80 1.20 1.58 2.19 2.68 3.18 3.79 m

min 0.30 0.66 1.01 1.36 1.67 1.98 2.37 2.70 ‡PLQLPXPLQFROXPQ
max 0.50 0.91 1.51 1.94 2.35 3.17 3.58 3.98 ‡PD[LPXPLQFROXPQ

248
6. New Concepts for Arch Bridges

normalized system moment of inertia in m 6

= 3 12 × system inertia
deck width
existing concrete arches 4
arch design concepts

0
0 100 200 300 400 500 span length in m

system radius of gyration in m 4

= system inertia
arch area
3
existing concrete arches
arch design concepts
2

0
0 100 200 300 400 500 span length in m

Figure 6-22. Comparison of normalized system moment of inertia and system radius of gyration

Figure 6-22 displays data from Tables 6-20 and 6-21 along with data from the database of existing concrete

arch bridges. The first plot shows that the system moments of inertia of the proposed concepts are gen-

erally lower than those of bridges in the database. This difference can be partially attributed to the differ-

ences in elastic moduli between conventional concrete and ultra high-performance fibre-reinforced con-

crete. Given that the elastic modulus of the latter is about 1.5 times that of the former, the system flexural

rigidity of the proposed concepts are more or less the same as those of bridges in the database. The second

plot shows that the system radii of gyration of the proposed concepts are about the same compared to

those of bridges in the database for spans <200 m, and are generally higher for spans ≥200 m. This means

that the concepts with spans ≥200 m are more susceptible to shallow arch behaviour than bridges in the

database with spans ≥200 m.

249
6. New Concepts for Arch Bridges

Table 6-22. System slenderness ratios among all concepts

system slenderness ratios among all concepts


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 48 44 45 46 47 48 49 50 •values greater than 50
33 to 40 deck-stiffened 4:1 40 53 40 41 43 38 39 40 are in red type
65 to 72 partially deck-stiffened 4:1 67 61 59 59 59 59 59 60
9 to 16 self-stiffened 8:1 36 40 41 42 41 41 41 41
41 to 48 deck-stiffened 8:1 36 48 36 37 39 34 36 37
73 to 80 partially deck-stiffened 8:1 61 55 54 54 54 55 49 49
17 to 24 self-stiffened 12:1 36 39 39 39 39 39 39 39
49 to 56 deck-stiffened 12:1 43 45 46 42 41 44 47 46
81 to 88 partially deck-stiffened 12:1 45 45 44 45 41 40 39 38

System slenderness ratios among all concepts are shown in Table 6-22. The highest system slendernesses

are achieved by partially deck-stiffened concepts with span-to-rise ratios of 4:1. These concepts have slen-

derness ratios of about 60, which is actually not that high when compared to the column tests described in

Chapter 3 (the shortest of the four columns had a slenderness ratio ( of about 60). Figure 3-6 on page 52

shows ultimate member capacity N-M* interaction diagrams for various slenderness ratios. Based on this

figure, slenderness ratios of 60 do not reduce the capacity of the member by much relative its sectional ca-

pacity. This, however, does not mean that additional second-order deflections can be neglected.

Figure 6-23 displays system slenderness ratios of the proposed concepts and concrete arch bridges in the

database. This figure shows that many existing concrete arches have much higher system slenderness ratios

than the proposed concepts. Slender behaviour should not be confused with the slender appearance of

bridge components. Slender behaviour, which is quantified by (, does not change proportionally with visu-

al slenderness, or thinness, as quantified by span-to-depth ratios (Figure 6-21).

system slenderness ratio 120


effective arc length factor × arc length
= 100
system radius of gyration
existing concrete arches 80
arch design concepts
60

40

20

0
0 100 200 300 400 500 span length in m

Figure 6-23. Comparison of system slenderness ratios

250
6. New Concepts for Arch Bridges

Table 6-23. Modified slenderness ratios among all concepts


modified slenderness ratios among all concepts
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 59 55 56 57 59 60 61 62
33 to 40 deck-stiffened 4:1 50 67 50 51 53 47 49 50
65 to 72 partially deck-stiffened 4:1 83 76 74 74 73 74 74 74
9 to 16 self-stiffened 8:1 25 27 28 29 28 28 28 28
41 to 48 deck-stiffened 8:1 25 33 25 26 27 24 24 25
73 to 80 partially deck-stiffened 8:1 42 38 37 37 37 38 33 34
17 to 24 self-stiffened 12:1 17 18 18 18 18 18 18 18 •values less than 20
49 to 56 deck-stiffened 12:1 20 21 22 20 19 21 22 22 (efficiency threshold value)
81 to 88 partially deck-stiffened 12:1 21 21 21 21 19 19 18 18 are in red type

Modified slenderness ratios among all concepts are shown in Table 6-23. As discussed in Chapter 4, modi-

fied slenderness ratios give measures of arch shallowness. As long as modified slenderness ratios of sys-

tems are kept above the threshold value of 20, then the arch will have greater than 90% arch action. The

results in Table 6-23 show that all concepts are well above this threshold, except those with span-to-rise ra-

tios of 12:1.

modified slenderness ratio 160


2 × rise
=
system radius of gyration
120
existing concrete arches
arch design concepts
80

for fixed arches:


40 increasing arch action
90% arch action
0 increasing beam action
0 100 200 300 400 500 span length in m

Figure 6-24. Comparison of modified slenderness ratios

Modified slenderness ratios of concepts are compared with those of existing concrete arch bridges in Fig-

ure 6-24. The figure shows that the proposed concepts are generally more susceptible to arch shallowness

problems than bridges in the database. This is not totally unexpected, since most bridges in the database

have span-to-rise ratios of 8:1 or less.

For concepts with span-to-rise ratios of 12:1, design calculations were governed by arch shallowness and

large crown deflections. The only way to reduce the effects of arch shallowness is to increase (f. This is

done by adding material such that it increases the cross sectional area of the arch faster than it increases

system moment of inertia. Adding material, however, also increases dead load which then also increases

251
6. New Concepts for Arch Bridges

sectional forces in the system. Using the simplified methods of analysis, design iterations for concepts with

span-to-rise ratios of 12:1 eventually led to seemingly satisfactory solutions. The simplified methods of ana-

lysis, however, assume that changes in the rise of the arch are small relative to the initial rise. When deflec-

tions are large, these methods may grossly underestimate the maximum sectional forces imposed on the

system. Thus, the proposed design concepts with span-to-rise ratios of 12:1 should be used with caution.

More refined structural analyses are needed to further validate these designs.

Table 6-24. Maximum deflections at ultimate limit states among all concepts
maximum deflections at ULS ÷ arch rise among all concepts, in %
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 0.6 0.6 0.8 0.9 1.0 1.2 1.4 1.5
33 to 40 deck-stiffened 4:1 0.7 0.7 0.6 0.7 0.7 0.7 0.7 0.8
65 to 72 partially deck-stiffened 4:1 0.9 0.9 0.9 1.0 1.2 1.5 1.9 2.3
9 to 16 self-stiffened 8:1 1.5 2.0 2.9 3.5 2.7 3.1 3.5 3.1
41 to 48 deck-stiffened 8:1 2.2 2.4 2.2 2.5 2.7 2.7 2.9 3.3
73 to 80 partially deck-stiffened 8:1 3.0 3.1 3.6 4.5 4.4 4.7 4.9 5.2
17 to 24 self-stiffened 12:1 3.1 4.4 5.1 5.0 5.1 4.8 4.8 4.9
49 to 56 deck-stiffened 12:1 4.7 4.5 4.8 4.7 4.5 5.3 5.9 6.1 •values greater than or
81 to 88 partially deck-stiffened 12:1 5.0 4.9 5.2 5.5 5.0 5.6 6.2 6.8 % equal to 5.0% are in red type

Table 6-24 shows the greater of deflections at the quarter-point and deflections at crown of the proposed

concepts at ultimate limit states. These deflections are the sum of second-order deflections caused by live

load, creep, shrinkage, and temperature. Elastic deflections caused by dead load are excluded because they

can be accounted for with camber during construction. Most of the concepts with span-to-rise ratios of 4:1

and 8:1 have relatively small ultimate deflections of about 3% of the initial rise or less. Concepts with span-

to-rise ratios of 12:1 have relatively high maximum deflections of about 4% to 6%. Any further modification

of these concepts tended to lead only to increased deflection.

6.3.9 Structural demands

Estimating dead loads during preliminary design can be difficult because the quantity of materials are not

accurately known. Underestimating dead load results in unconservative estimates of structural demand.

Overestimating dead load results in unconservative estimates of precompression in the arch. Thus it is be-

neficial to have a good estimate of dead load during the early stages of design. The total weight of all struc-

tural components, including the weight of the deck, columns, and arch, transverse ribs, stiffeners, dia-

252
6. New Concepts for Arch Bridges

phragms, and prestressing steel were calculated for each concept. Extra weight factors, which are

calculated as the total weight of all structural components divided by the weight of continuous, longitudin-

al concrete, are shown for each concept in Table 6-25. Extra weight factors of the concepts range from 1.15

to 1.82. Figure 6-25 plots extra weight factors versus span length. Results show that using values between

1.4 and 1.6 results in good first estimates of dead load. The lower end of the range is more appropriate for

longer spans, while the higher end of the range is more appropriate for shorter spans.

Table 6-25. Extra weight factors among all concepts

extra weight factors among all concepts


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m

1 to 8 self-stiffened L:f = 4:1 1.52 1.56 1.74 1.68 1.62 1.54 1.48 1.46
33 to 40 deck-stiffened 4:1 1.75 1.52 1.44 1.42 1.40 1.30 1.30 1.29
65 to 72 partially deck-stiffened 4:1 1.55 1.55 1.53 1.53 1.53 1.49 1.46 1.47

9 to 16 self-stiffened 8:1 1.42 1.59 1.81 1.70 1.53 1.51 1.48 1.40
41 to 48 deck-stiffened 8:1 1.82 1.57 1.48 1.44 1.43 1.37 1.36 1.34
73 to 80 partially deck-stiffened 8:1 1.58 1.57 1.57 1.55 1.41 1.33 1.44 1.34

17 to 24 self-stiffened 12:1 1.43 1.60 1.64 1.49 1.39 1.31 1.25 1.21
49 to 56 deck-stiffened 12:1 1.70 1.39 1.28 1.23 1.21 1.20 1.18 1.16
81 to 88 partially deck-stiffened 12:1 1.64 1.44 1.35 1.25 1.16 1.17 1.17 1.15

min 1.42 1.39 1.28 1.23 1.16 1.17 1.17 1.15 ‡PLQLPXPLQFROXPQ
max 1.82 1.60 1.81 1.70 1.62 1.54 1.48 1.47 ‡PD[LPXPLQFROXPQ

extra weight factor 2


weight of all structural components
=
weight of longitudinal concrete

1.5

1
0 100 200 300 400 500 span length in m

Figure 6-25. Comparison of extra weight factors

Figure 4-21 on page 127 shows the five load combinations used to calculate critical sectional forces at the

quarter-point and midpoint of each design concept. Live load cases 1, 2, and 3 cause maximum sectional

forces at the springing lines and at the crown. These sectional forces are combined with those caused by

dead load, creep, shrinkage, and temperature. Live load cases 4 and 5 cause maximum sectional forces at

the quarter-points. These sectional forces are combined with only those caused by dead load.

253
6. New Concepts for Arch Bridges

Table 6-26 shows which of the five live load cases and accompanying loads cause maximum sectional

forces in each design concept. Results show that self-stiffened and partially deck-stiffened concepts with

span-to-rise ratios of 4:1 were governed by sectional forces at the quarter-points caused by truck or lane

loads. The effects of restrained deformation were not critical for these concepts. Nearly all other concepts

were governed by live load cases 2 or 3. This shows that the effects of restrained deformations become in-

creasingly important at span-to-rise ratios of 8:1 and higher.

Table 6-26. Governing live load cases among all concepts


governing live load cases among all concepts
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 5 5 5 4 4 4 4 4 •load cases where quarter-
33 to 40 deck-stiffened 4:1 5 5 3 3 3 2 2 2 points are critical (4 or 5)
65 to 72 partially deck-stiffened 4:1 5 5 5 4 4 4 4 4 are in red type
9 to 16 self-stiffened 8:1 3 3 3 2 2 2 2 2
41 to 48 deck-stiffened 8:1 3 3 3 3 2 2 2 2
73 to 80 partially deck-stiffened 8:1 3 3 3 2 2 2 2 2
17 to 24 self-stiffened 12:1 3 3 3 2 2 2 2 2
49 to 56 deck-stiffened 12:1 3 3 3 2 2 2 2 2
81 to 88 partially deck-stiffened 12:1 3 3 3 2 2 2 2 1

The second-order flexible system moments associated with the governing load cases at ultimate limit states

are shown in Table 6-27. These moments are taken as the greater of the quarter-point and midspan mo-

ments. In most cases, self-stiffened concepts have the least bending moment demands among the arch sys-

tems considered.

Table 6-27. Maximum flexible system moments at ultimate limit states among all concepts
maximum flexible system moments at ULS among all concepts, in MN·m
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 11 23 36 49 69 94 126 161 •minimum values among
33 to 40 deck-stiffened 4:1 11 23 39 54 69 100 126 158 triplets are in red type
65 to 72 partially deck-stiffened 4:1 12 24 37 51 74 105 147 195
9 to 16 self-stiffened 8:1 21 31 48 75 115 165 226 308
41 to 48 deck-stiffened 8:1 14 31 59 86 122 186 243 321
73 to 80 partially deck-stiffened 8:1 12 28 49 78 125 196 252 370
17 to 24 self-stiffened 12:1 28 42 74 128 207 331 495 708
49 to 56 deck-stiffened 12:1 18 49 97 185 312 430 603 868
81 to 88 partially deck-stiffened 12:1 17 46 88 167 353 509 707 1060 MN·m

254
6. New Concepts for Arch Bridges

L:f = 4:1 L:f = 8:1 L:f = 12:1


normalized governing moment at ULS in kN 250
governing moment at ULS
=
span × span-to-rise ratio 200

150
self-stiffened arch concepts
deck-stiffened arch concepts 100
partially deck-stiffened arch concepts
moments caused by load case 4 or 5 50

0
span in m 0 100 200 300 400 0 100 200 300 400 0 100 200 300 400

Figure 6-26. Comparison of governing moments at ultimate limit states

Figure 6-26 plots the governing bending moment results from Table 6-27. Moments are normalized by

span and span-to-rise ratio and are thus expressed in kN units of force. As shown by the first and second

graphs on the left, concepts with spans-to-rise ratios of 4:1 and 8:1 have about the same normalized mo-

ments over the entire range of spans considered in the study. This suggests that for the span lengths con-

sidered, doubling the span-to-rise ratio from 4:1 to 8:1 results in a doubling of moment demand. The

circled data points in the figure refer to governing moments that are caused by dead and live loads only.

These occur only in the concepts with span-to-rise ratios of 4:1. By comparing the centre and right-hand

graphs in the figure, results show that the effects of restrained deformation becomes increasingly severe as

span-to-rise ratio is increased from 8:1 to 12:1 for spans ≥200 m. As the arch system is made flatter and

longer, moments caused by restrained deformation grow exponentially. Thus, there are practical limits to

how flat arches can be made.

255
6. New Concepts for Arch Bridges

Table 6-28. Portions of governing moments caused by first-order live loads among all concepts

SRUWLRQRIPD[LPXPÁH[LEOHV\VWHPPRPHQWVFDXVHGE\ÀUVWRUGHUOLYHORDGVDPRQJDOOFRQFHSWVLQ
L = 50 100 150 200 250 300 350 400 m

1 to 8 VHOIVWLfIHQHG L:f = 4:1 98 97 93 88 84 80 75 72


33 to 40 GHFNVWLfIHQHG 4:1 98 96 66 63 61 54 53 51
65 to 72 SDUWLDOO\GHFNVWLfIHQHG 4:1 95 93 90 83 77 71 63 58

9 to 16 VHOIVWLfIHQHG 8:1 43 59 57 47 41 37 34 30
41 to 48 GHFNVWLfIHQHG 8:1 66 58 47 43 39 34 32 29
73 to 80 SDUWLDOO\GHFNVWLfIHQHG 8:1 70 61 54 42 35 28 29 24

17 to 24 VHOIVWLfIHQHG 12:1 37 48 42 32 27 22 18 16
49 to 56 GHFNVWLfIHQHG 12:1 56 40 30 21 17 16 14 12
81 to 88 SDUWLDOO\GHFNVWLfIHQHG 12:1 56 43 33 22 15 14 13 9

Table 6-28 shows first-order live load moments expressed as percentages of maximum moments at ulti-

mate limit state among all concepts. This set of data measures the validity of using first-order analysis con-

sidering dead and live loads only. The data is presented graphically in Figure 6-27. The results show that

this simple analysis is only accurate for concepts with spans of ≤100 m and span-to-rise ratios of 4:1. For

concepts with spans of >100 m and span-to-rise ratios of 4:1, second-order analysis is needed to account

for the additional moments caused by deformations of the arch off the pressure line. The effects of re-

strained deformation need not be considered for these concepts. For all other concepts (those with span-

to-rise ratios greater than 4:1), second-order analyses including the effects of restrained deformations are

needed because first-order live loads contribute less than 70% of the total flexible system moment at ulti-

mate limit state. In these cases, bending moment demands would be grossly underestimated if the effects

of creep, shrinkage, temperature, and second-order deformations were not considered.

live load portion of governing moments in % 100


first-order live load moment
total second-order moment
80

self-stiffened arch concepts L:f = 4:1


60
deck-stiffened arch concepts
partially deck-stiffened arch concepts
moments caused by load case 4 or 5 40

L:f = 8:1
20
L:f = 12:1

0
0 100 200 300 400 500 span length in m

Figure 6-27. Comparison of first-order live load portions of governing moments

256
6. New Concepts for Arch Bridges

6.3.10 Summary of results and conclusions

The most common geometrical quantities that were found to be satisfactory among design concepts are

summarized in Table 6-29. Concepts whose demands required proportions greater than those shown in

the table include: (1) some concepts with spans of 50 m due to the relatively low amount of natural

prestress in the arch, and (2) most concepts with span-to-rise ratios of 12:1 due to high sectional force de-

mands and sensitivity to axial deformations.

Table 6-29. Most common geometrical quantities used among design concepts
class of system element section type span-to- slab width of box ÷
depth ratio thickness width of deck

self-stiffened arch deck double tee 180:1 50 mm n/a


arch hollow box 50:1 50-100 mm 1/2

deck-stiffened arch deck hollow box 50:1 50-150 mm 1/4


arch solid box 200:1 n/a 1/4
arch hollow box 180:1 200 mm 1/2

partially deck-stiffened arch deck hollow box 90:1, 100:1 100-200 mm 1/2
arch hollow box 90:1, 100:1 100-200 mm 1/2

Regarding efficiency in terms of the consumption of materials, the following observations were made:

• Among the arch systems considered, deck-stiffened systems consume the least amount of concrete

for spans of 50 m, while self-stiffened systems consumed the least amount of concrete for spans of

150-400 m. For spans of 100 m, all systems consumed approximately the same amount of concrete.

• For span-to-rise ratios of 4:1 and 8:1, all concepts consistently had less than 50% of the minimum

equivalent slab thicknesses of existing concrete arch bridges in the database. Concepts with span-to-

rise ratios of 12:1 were not able to achieve the same reduction in concrete volume.

• Self-stiffened arch systems consumed the least amount of prestressing steel.

• Concrete volume in the deck and in the arch tend to be roughly the same, regardless of which ele-

ment stiffens the system.

257
6. New Concepts for Arch Bridges

Regarding slenderness and shallowness, the following observations were made:

• In terms of visual slenderness, concepts tended to have span-to-system-depth ratios between 39:1 and

50:1, which is on par with the ratios observed among existing concrete arch bridges in the database.

• Concepts tended to have normalized system moments of inertia that were less than or about the

same as those observed among existing concrete arch bridges in the database.

• Concepts tended to have system radii of gyration that were greater than or about the same as those

found among existing concrete arch bridges in the database.

• Concepts tended to have system slenderness ratios between 39 and 60, which is not as high as many

existing concrete arch bridges in the database. In this range of slenderness ratios, additional deflec-

tions and moments caused by second-order effects can be significant. For these systems, reduced

member capacity N-M* interaction diagrams are only slightly less than sectional capacity N-M inter-

action diagrams.

• In terms of arch shallowness, concepts with span-to-rise ratios of 4:1 and 8:1 tended to have modified

slenderness ratios much greater than 20. Permanent loads are thus carried primarily in compression

along the arch. Concepts with span-to-rise ratios of 12:1 tended to have modified slenderness ratios

around 20 or less. Their responses to load are thus affected by significant amounts of beam action.

Regarding loads, deflections and complexity of analysis, the following observations were made:

• Extra weight factors of 1.4 to 1.6 can be used to estimate the total weight of the structure based on the

weight of continuous, longitudinal concrete. Shorter spans of 50 m tended to have higher extra

weight factors while longer spans of 400 m tended to have lower factors.

• Most concepts were successfully designed to have crown deflections that were less than 5% of the ini-

tial rise of the arch at ultimate limit states. This suggested deflection limit ensures that results from

simplified methods analyses are sufficiently accurate for preliminary design.

258
6. New Concepts for Arch Bridges

• For most concepts, the bending moments caused by restrained deformations and second-order

effects are significant, and are of similar magnitude to those caused by live load. As such, second-or-

der analyses including restrained deformation should be used in preliminary design.

6.4 Recommendations

In this section, some of the the most important questions regarding the potential opportunities offered by

the arch concepts proposed in this chapter are addressed. Recommendations are made in response to sev-

eral topics that are introduced in the form of questions.

How long can these new arch concepts span?

The parametric study provided examples of arch concepts that are designed for spans of up to 400 m.

These designs were shown to be much lighter than the longest bridges made using conventional concrete

arch technology.

Figure 6-26 on page 255 shows that moments caused by the effects of restrained deformation and second-

order effects grow exponentially with span length. Thus concepts cannot simply be scaled up from 400 m

to 1000 m and be expected to be satisfactory. This means that slab thicknesses much greater than 200 mm

are needed if span-to-depth ratios are kept more or less the same, resulting in heavy segments and greater

complexity of construction. Arches spanning 600 m are likely possible but would require higher equival-

ent slab thicknesses than those used in the parametric design study. Although these long-span concrete

arches are probably possible, and would use materials efficiently to carry loads in the final system, the cost

of temporary works and throwaway materials would outweigh any cost-savings related to the consumption

of concrete. For such long spans, concrete arches would not be economically competitive against cable-

stayed or suspension-type structural systems, since these efficient systems can be built with essentially zero

throwaway materials. These considerations impose economic barriers that limit the applicability of using

ultra high-performance fibre-reinforced concrete arches for spans greater than 400 m.

259
6. New Concepts for Arch Bridges

How shallow can these arch concepts be made?

As the rise of an arch is decreased, its behaviour transitions from carrying permanent loads in pure axial

compression along the pressure line to carrying permanent loads in bending as in conventional beam-type

structures, as discussed in Section 4.4.3. The parametric design study shows examples of design concepts

with span-to-rise ratios of 12:1, which have modified slenderness ratios that are slightly less than the pro-

posed minimum threshold value of 20. Adding material to these designs tended to increase structural de-

mands faster than they increased member capacities. Thus the maximum span-to-rise ratio that should be

used for these proposed systems is less than 12:1, but higher than 8:1.

Using conventionally reinforced concrete, the Infant Henrique Bridge was able to achieve a span-to-rise

ratio of 11.2:1 (Fonseca and Mato 2005). In order to achieve this, it required a very stiff deck girder and a

very wide voided slab for the arch. A similar design could be made using ultra high-performance fibre-re-

inforced concrete but the savings in material would be minimal, since the axial rigidity of the arch would

govern rather than its compressive strength.

How slender can these new arch concepts be made?

Slenderness can be interpreted in three different ways: (1) visual slenderness as measured by span-to-depth

ratios, (2) sensitivity to second-order effects as measured by slenderness ratio, and (3) thinness of elements

as measured by wall slenderness ratios. Each interpretation will be discussed below.

Regarding visual slenderness, the highest span-to-system depth ratio achieved among concepts was 50:1.

Greater visual slenderness can be achieved by reducing the depth of deck and arch, and increasing their

slab thicknesses. Span-to-system-depth ratios only slightly higher than 50:1 are possible because deflec-

tions and second-order effects become increasingly severe as the flexural rigidity of the system is reduced.

If the arch alone is considered, the highest span-to-arch-depth ratio achieved among concepts was 200:1,

which is higher than any of the existing arches in the database.

The highest system slenderness ratio achieved among concepts was 60, which is not high compared to

those in existing concrete arches. This quantity is deceiving, since higher slenderness ratios in these arches

260
6. New Concepts for Arch Bridges

are a result of using sections that are solid, or at least more solid than the sections proposed for the con-

cepts. For example, if a solid member and a hollow member were given the same effective length, depth

and width, it is the solid member that would have the higher slenderness ratio. In this comparison, differ-

ences in slenderness ratio have no impact on the external visual slenderness of the member.

Minimum thicknesses of concrete components are governed by: (1) geometrical requirements for obtain-

ing random fibre distributions, (2) concrete cover requirements for reinforcing steel, and (3) by local plate

buckling problems. Minimum thicknesses for ensuring a random distribution of fibres is usually taken as

three to five times the fibre length, which is 30-50 mm for the mix design considered in this thesis. Because

no conventional reinforcing steel is used in the concepts, only cover requirements for internal prestressing

steel need to be considered. The proposed segment end ribs in this study are specified to be 100 mm thick,

which allows for just enough cover for transverse prestressing reinforcement. Reinforcement-free mini-an-

chorages, such as those used in the Sun-Yu Bridge in Korea (Huh and Byun 2005), allow for post-ten-

sioned concrete components to be thin even near the anchorage zones.

Concepts for thin ribbed slabs using ultra high-performance fibre-reinforced concrete were investigated by

Spasojević (2008). Using various flange and web thicknesses of 40-60 mm combined with internal, post-

tensioned monostrands, Spasojević found that these light-weight ribbed slabs were capable of transmitting

wheel loads over spans of 12 m between supporting girders. Her findings help validate the use of thin, ultra

high-performance fibre-reinforced concrete slabs for the decks of the concepts proposed is this thesis.

Problems associated with local buckling also limit how thin components can be made, as discussed in Sec-

tion 6.1.5. Based on a rudimentary stability analysis of ultra high-performance fibre-reinforced concrete

plates, it was shown that plate buckling could be avoided if wall slenderness ratios less than 24.5 were used.

Wall slenderness ratios are calculated as the width of plate between line supports divided by plate thick-

ness. According to this limit, 50 mm thick plates with maximum plate widths of 1230 mm are capable of

reaching the peak compressive stress of the material without buckling. The maximum allowable plate

width can be extended though the use of longitudinal stiffeners, as is proposed for the design concepts.

261
6. New Concepts for Arch Bridges

Which arch system is most efficient? Should the deck be used to stiffen the arch?

Based on the results of the parametric design study, self-stiffened arch concepts, in most cases, were found

to consume the least amount of concrete among the arch systems considered (see Table 6-15 on page 244).

This result agrees with the conclusion made at the end of Section 4.7: that it is most efficient to maximize

the use of precompression in the arch for carrying flexible system moments. This also serves to isolate deck

girders from the global system, which minimizes their bending moment demands.

For span lengths of 50 m, self-stiffened arches were not the most efficient because the natural prestress of

the arch caused by dead load was low in comparison to the imposed flexural stresses. In these cases, deck-

stiffened arch systems consumed the least amount of concrete. The resistances of deck girders in these sys-

tems were calibrated with demands by selecting an appropriate level of prestressing.

In spite of these findings, designers should not automatically dismiss partially deck-stiffened systems, since

all concepts in the parametric design study were able to reduce the consumption of concrete by at least

50%, except for those with span-to-rise ratios of 12:1. Because the proposed concepts are all more efficient

than the existing concrete bridges in the database, a greater set of criteria should be used to decide which

arch system should be used, such as ease of construction, consumption of prestressing steel, and visual im-

pact on the surrounding environment. For example, one benefit of the proposed partially deck-stiffened

concepts is that they have nearly identical arch and deck segments. This repetition of geometry may lead to

significant economies of scale, since the same set of forms could be modified to precast both the deck and

arch. As a second example, self-stiffened concepts have the highest arch flexural rigidity among the sys-

tems, which means that they have the highest resistances against buckling during the stages of construc-

tion when the arch is unsupported by temporary works. These examples emphasize the importance of hav-

ing a design-oriented mindset that is able to interpret and adapt to the specific requirements of given

designs. Thus, the proposed concepts should be treated as malleable starting points for design rather than

inflexible model solutions.

262
6. New Concepts for Arch Bridges

Would the addition of rebar improve the efficiency or performance of the proposed systems?

Supplementing ultra high-performance fibre-reinforced concrete with rebar results in moments of resist-

ance that are more or less the same as those attained in conventional reinforced concrete members. This is

because moments of resistance are typically governed by the yield force of steel, rather than by the com-

pressive strength of concrete.

The tensile strength provided by the combination of the fibre-reinforced concrete and steel reinforcing

bars is not equal to the sum of tensile strengths provided by each individual component. This is because

the fibre-reinforced concrete reaches its maximum tensile stress at a strain lower than the yield strain of

steel. Thus, in terms of strength requirements, it is best to design members such that tensile strength is

provided: (1) by the fibre-reinforced concrete only, (2) by external unbonded prestressing tendons, which

can be used to delay initial concrete cracking, or (3) by mild steel reinforcing bars only. The first two op-

tions allow for minimum wall thicknesses to be used since conventional cover requirements are not applic-

able. The last option can be used if required wall thicknesses already exceed minimum cover requirements

for mild steel.

The addition of mild steel reinforcing would be useful in members that have high ductility demands. Al-

though ultra high-performance fibre-reinforced concrete is more ductile than ordinary concrete, it is less

ductile than reinforced concrete. Once localized cracks form in ultra high-performance fibre-reinforced

concrete members, curvatures at the cracked section can only be increased if loads are reduced. Because of

the long yield plateau of rebar, reinforced concrete members can withstand large rotations before their mo-

ments of resistance are reduced. It would thus be prudent to supplement ultra high-performance fibre-re-

inforced concrete members with rebar at critical locations in the structure. For the proposed design con-

cepts, mild steel reinforcing or additional prestressing steel should be provided at the sections of the deck

and arch above the springing lines, and at midspan.

263
6. New Concepts for Arch Bridges

Will these concepts with their very thin concrete components be sensitive to impact loads?

Regarding structural robustness and sensitivity to impact loads, analytical and experimental studies have

done by Habel and Gauvreau (2008) on 50 mm ultra high-performance fibre-reinforced concrete plates.

Specimens were subjected to impact loads using drop weights. Results showed that the strength of the ma-

terial increased with increasing strain rate, for strain rates of up to 2 s-1. At this rate, the material continued

to exhibit strain-hardening characteristics and multiple cracking. The bending resistance increased by

more than 25% relative to static behavior. Given these results, it seems that ultra high-performance fibre-

reinforced concrete has the capacity to withstand accidental or exceptional loads. Future research, how-

ever, is certainly needed to further investigate this mode of failure.

Is ultra high-performance fibre-reinforced concrete a good material for arch bridges?

Based on the results of the parametric design study, ultra high-performance fibre-reinforced concrete is an

excellent material to use in arch bridges. This material allows for significant reductions in the consumption

of concrete and in the total weight of the structure, relative to arches made using conventional concrete

technology. These reductions should result in considerable savings related to the costs of raw materials,

temporary works, and hours of skilled labour. These reductions will also reduce the complexity and dura-

tion of construction, resulting in further cost-savings. The enhanced properties of ultra high-performance

fibre-reinforced concrete were found to positively affect the behaviour and efficiency of arch systems in the

following ways:

• The high compressive strength of the material allowed for high levels of precompression in the arch,

which in turn increased arch flexural capacity. The same applied to the high levels of allowable

prestress in the deck. These increases in compression and bending capacities enabled decks and

arches to be designed with highly efficient cross-sections.

• The high tensile strength of the material allowed for the elimination of most mild reinforcing in the

structure. Because the material has a higher cracking strength than conventional concrete, the per-

formance of the structure at serviceability limit states is improved.

264
6. New Concepts for Arch Bridges

• Creep and shrinkage strains are less than in conventional concrete. This reduces long term-deflec-

tions, as well as bending moment demands caused by restrained deformation.

• The short fibre length of the embedded steel fibres allowed for very thin components to be used.

265
Chapter 7. Conclusions and Summary

In this chapter, the principal findings of this thesis related to the the design and validation of new systems

for arch bridges using ultra high-performance fibre-reinforced concrete are presented. A concise list of

conclusions is given in Section 7.1. Summaries of material behaviour, sectional behaviour, member beha-

viour, and arch system behaviour are given in Sections 7.2 to 7.5. This is followed by summaries of the

comparative study of existing concrete bridges and the parametric design study in the last two sections.

Sections 7.2 to 7.7 are written assuming that the reader has not thoroughly read the main body of the thes-

is. Hence, the most important equations and figures are reprinted and briefly explained. Potential research

topics for future work are also identified and discussed in these sections.

7.1 Conclusions

The primary conclusions drawn from the parametric design study are:

• It is possible to reduce the weight and consumption of concrete in arch bridges by at least 50% by us-

ing ultra high-performance fibre-reinforced concrete without conventional steel reinforcing bars.

This reduction applies to spans of 50-400 m and span-to-rise ratios of 4:1 and 8:1.

• Lightness is achieved primarily by the thinning of structural components (i.e. webs, flanges, and

slabs) and the efficient use of precompression in the arch, and rather than by the decrease of bending

stiffness.

266
7. Conclusions and Summary

• Deck-stiffened arch systems consume the least amount of concrete for spans of 50 m, while self-sti-

ffened arch systems consume the least amount of concrete for spans of 150-400 m. For spans of 100

m, all arch systems consumed about the same amount of concrete.

• The use of lightweight segments and thin components requires that stiffeners and ribs be used to pre-

vent local buckling. These additional components increase the complexity of the segmental precast-

ing process.

The primary conclusions drawn from the comparative study of existing arch bridges are:

• The typical range for span-to-rise ratios in concrete arches is 2.3 to 8.0.

• The typical span-to-system-depth ratio in concrete arches is 24 to 48.

• The envelope of minimum equivalent slab thicknesses hV observed among existing arch bridges in

the database, expressed as a function of span length L, is as follows:

# L
% + 0.569 m 0 ! L ! 193m
% 1276
hV = $
% L2 L
" + 2.21m L > 193m
% 22800 m 61.7
&

The primary conclusions drawn from the study of shallow arches are:

• The transition between fixed arch behaviour and fixed beam behaviour was found to be governed by

the modified slenderness ratio, which is proportional to arch rise and inversely proportional to sys-

tem radius of gyration.

• To ensure that given arch designs behave efficiently, a minimum modified slenderness ratio of 20 is

proposed. This threshold ensures that more than 90% of uniform loads are carried in axial compres-

sion along the axis of the arch.

• Based on an empirical relationship between system radius of gyration and span length, the following

efficiency threshold curve, expressed in terms of span-to-rise ratio, is proposed:

267
7. Conclusions and Summary

L ! L $
< 16.1 #
f " L + 48.4 m &%

• There remains an opportunity to design efficient shallow arches that have span-to-rise ratios greater

than 11.2, which is the highest ratio ever used in a concrete arch bridge.

The primary conclusions drawn from the study of slender members under compression and bending:

• Because of second-order effects, slender ultra high-performance fibre-reinforced concrete members

can reach ultimate load before reaching sectional failure. Thus, a reduced member capacity interac-

tion diagram must be used when designing for ultimate limit states.

• Ultimate loads of 4 slender column tests were overestimated by general method predictions by up to

14.5%. Considering the ±2 mm error in measuring initial eccentricities of load (representing ±16.2%

to ±37.7% of the eccentricity), the general method predictions were fairly good.

• Member capacity interaction diagrams calculated using the simplified design method were shown to

be accurate and conservative relative to those calculated using the more rigorous general analysis

method.

268
7. Conclusions and Summary

7.2 Material behaviour

Sets of ultra high-performance fibre-reinforced concrete cylinders and prisms made from local materials

were tested to obtain stress-strain models suitable for use in design calculations. The concrete specimens

exhibited, on average, a 28-day compressive strength of 106 MPa, a tensile cracking strength of 7 MPa, and

a tensile strength of 14.6 MPa. These high strengths are achieved by the use of water reducers, a low water-

to-cement ratio, no coarse aggregate, and 5.5% steel fibre content by volume. Based on test results, the mul-

tilinear stress-strain curves shown below are proposed for use in design calculations. The proposed materi-

al model is assumed to be valid for compressive strengths between 100 MPa and 140 MPa, and is intended

to be used specifically for the mix design developed at the University of Toronto.

The stress-strain response in compression can be modified according to three parameters: compressive

strength fc!, tangent modulus of elasticity in compression Ec, and secant modulus at peak stress Ec!. The

stress-strain response in tension is based on four parameters: strain and stress at cracking, %cr and fcr, and

strain and stress at peak tensile stress, %t! and ft!. Post-peak tensile stresses are neglected.

Material model Compressive


stress in MPa Set 1 (5 cylinders)
Compressive 200 100 mm ! 200 mm
moist-cured for 27 days
stress
fc! ( fc!÷Ec′, fc′ ) avg. material properties:
150 Ec = 48.1 GPa
Ec! = 37.3 GPa
( 1.4 fc!÷Ec!, 0.7fc! )
( 0.7 fc!÷Ec, 0.7fc! ) fc ! = 106 MPa
100

50
0
0 !c′ Compressive strain cylinder test data
material model
0
0 4·10-3 8·10-3
Compressive strain

Reprint of Figure 2-6. Material model in compression

269
7. Conclusions and Summary

Tensile 3 prisms tested


Material model 150 mm ! 150 mm ! 900 mm
stress in MPa
moist-cured for 7 days
Tensile 30 air-cured for 47 days
stress tested at 54 days
ft! avg. material properties:

Obsolete
Et = 46.7 GPa
20 prism test data only !cr = 0.15·10-3 fcr! and !cr! based on
valid before cracking averages from prism data
fcr = 7.0 MPa
fcr
!t! = 1.5·10-3 !t! estimated using results
ft! = 14.6 MPa from direct tension tests by
10
Habel and Gauvreau (2008)

0 ft! estimated using results


0 !cr !t! prism test data
Tensile material model
from 3 split cylinder tests
strain 0
0 1·10-3 2·10-3

Tensile strain

Reprint of Figure 2-9. Material model in tension

A compressive strength of 120 MPa is used for the parametric design study, representing a nominal

strength that can readily be achieved by many different ultra high-performance fibre-reinforced concrete

mixes, without the use of specialized heat or steam treatment. Although the 28-day cylinder tests exhibited

only 106 MPa at 28 days, this modest increase in assumed strength can be attained by further optimizing

the selection and mix proportions of raw materials and by improving the equipment and procedure used

in mixing and casting.

Tests done by others on similar mixes of ultra high-performance fibre-reinforced concrete resulted in av-

erage long-term shrinkage strains of 550·10-6 and long-term creep strains of 0.8. These quantities, which

are used in the design study, result in long-term strains that are much lower than in conventional concrete.

One possible objective for future research is to improve the ductility response of the material. While ultra

high-performance fibre-reinforced concrete is more ductile than plain concrete, both in compression and

tension, it is less ductile than concrete that is reinforced with mild or prestressing steel. This has implica-

tions on the resilience of the structure and its ability to undergo large rotations and deformations. Rather

than resorting to minimum rebar requirements, it might be adequate to provide reinforcement only at crit-

ical sections of the structure, designed specifically to improve the ductility of the member. These bars

would only have to match the tensile force provided by the ultra-high fibre-reinforced concrete.

270
7. Conclusions and Summary

7.3 Sectional behaviour

Using the proposed stress-strain curve, the axial force-moment-curvature response of a 100 mm by 100

mm square cross-sections was investigated in Section 3.2.2. Contours of equal curvature for this cross-sec-

tion are plotted below in axial force-moment space. These contour plots provide: (1) a visual representation

of the nonlinear behaviour of the section, and (2) a means of calculating moment-curvature diagrams

without using iterative calculation. The most important observations made from this analysis are: (1) the

section behaves linear-elastically up until initial cracking, and (2) limited amounts of moment-curvature

ductility can be expected if axial forces are less than the axial force at the balanced condition.

(a) Axial (b) planes of strain with equal !


force N
–1000 kN one contour of equal corresponding
curvature in rad/km stresses are integrated
! = 10 rad/km ... ...
to get N-M point on
–800
diagram

–600 100 mm
! = 10 20 30 40 50

100 mm
! = 30 rad/km ... ...
–400

balanced point ultra high-


–200
performance
fibre reinforced
! = 50 rad/km ... ... concrete section
0
0 4 8 12 16 kN·m
Bending moment M

(c) Axial (d)


force N Moment M
–1000 kN 16 kN·m 1
EI0
N = -500 kN
–800 N = -400 kN
12
N = -300 kN
C
–600
!=10 20 30 40 50 B
N = -500 kN 8 N = -200 kN
cracking point
–400 N = -400 kN
A B C
N = -300 kN N = -100 kN
4 A
–200 N = -200 kN
N = -100 kN
N = 0 kN
0 0
0 4 8 12 16 kN·m 0 30 60 90 120 rad/km
Moment M Curvature !

Partial reprint of Figure 3-3. Contours of equal curvature and corresponding moment-curvature diagrams

271
7. Conclusions and Summary

Moment-Curvature
in MN·m and rad/km N=-55.2 MN
Axial force- Axial force-
Moment interaction Secant stiffness M
in MN and MN·m in MN and EI0
150
Mu
N N
Mcr
post-peak response
!200 depends on the control
of localized cracking

EIu
!150
0
"cr "u "
!100 EIcr
EI10 EIcr
Mcr
Mu EIu
!50

0
0 0 0.5 1 1.5 "
0 50 100 150 M 0 10
EI EI!
!EI0
Secant stiffness-Curvature
(a) (b) in EI0 and rad/km N=-55.2 MN

(c)

Reprint of Figure 3-23. Sectional analysis of an ultra high-performance fibre-reinforced concrete box

The sectional response of a hollow, ultra high-performance fibre-reinforced concrete box section was

investigated in Section 3.5. Based on the results of the analysis (shown in the figure above), expressions for

secant stiffness at ultimate EIu are proposed as follows, where N is axial force, N! is maximum axial force in

pure compression, and EI0 is gross flexural rigidity of the section:

+ # N &
- 1
EI 0 % +1 0 ) N ) 12 N "
-
2
$ N " ('
EI u ! ,
- # N&
1
EI 0 % 2 * ( 1
N" ) N ) N"
- 2
$ N"'
2
.

These conservative secant stiffness values are used in design calculations to account for the softening of

systems caused by material nonlinearity. These expressions result in effective secant stiffness values

between 0.5EI0 and 0.75EI0, which are higher those exhibited by conventionally reinforced concrete

sections.

272
7. Conclusions and Summary

7.4 Slender column behaviour

The response of slender concrete members under combined compression and bending was investigated in

Chapter 3. The analytical model considered is shown in the figure below. In slender columns, additional

secondary moments are caused by deflections of the column away from the line of action of force. In very

slender columns, these second-order effects can cause the ultimate load of columns Q* to be reached be-

fore sectional failure is reached. Hence, the member capacity of very slender columns is always less than

than the sectional capacity of its critical section.

member capacity N-M* sectional capacity N-M


interaction diagram interaction diagram
Axial
force N ultimate limit states Q=Q*
1000 sectional response at N
mid-length for various e:
Q Q e=0.1mm
0.5mm
750 1mm
!=80 2mm
e 4mm
8mm
w 500

M
N 250
16mm
Q 32mm
0
0 4 8 12 16
Bending moment M
(a)

Partial reprint of Figure 3-6. Member capacity interaction diagram

In order to describe this behaviour analytically, a general analysis method for calculating reduced member

capacity N-M* interaction diagrams is presented in Section 3.2.4. The input parameters are pin-to-pin

length, sectional geometry, and stress-strain curve. If initial eccentricity of load is also given, the general

method can also be used to calculate the load-deflection Q-w response of the column, as presented in Sec-

tion 3.2.5. A sample reduced N-M* interaction diagram is shown in the figure above as a dotted curve. The

dashed curve represents sectional failure. The solid curves represent the various possible sectional re-

sponses of the mid-length section, each corresponding to different initial eccentricities of load. The apexes

of the solid curves all correspond to different ultimate limit states of the column, each with different com-

binations of ultimate load N=Q* and ultimate moment M*.

273
7. Conclusions and Summary

The general method accounts for material nonlinearity by using nonlinear moment-curvature diagrams.

Geometric nonlinearity is accounted for by column deflection curves, which map centreline deflections of

a column relative to a fixed line of action of load. Column deflections curves are calculated by first assum-

ing a mid-length moment, and then solving for the end eccentricity of the column, using a numerical in-

tegration procedure. A family of column deflection curves for a given load Q is generated by changing the

assumed mid-length moment. Among these curves, the one at ultimate limit state Q=Q* is the one with

the greatest end eccentricity. Repeating this procedure for many different values of Q produces the reduced

N-M* interaction diagram. All these calculations are implemented in the computer program QULT, which

is described in detail in Appendix B.


design interaction diagram
n 1
admissible forces for
serviceability limit states
nE admissible forces for
ultimate limit states

0.5
nt —ncr

design contour of cracking


design interaction diagram

0
ncr 0.5 1 1.5
nt
m
mn=0
QE
nE = = normalized axial force at Euler buckling, but not > 0.8
A·fc!
f
ncr = cr = cracking stress divided by compressive strength
fc!
f!
nt! = t = tensile strength divided by compressive strength
fc!
mn=0 = normalized ultimate moment when N=0

Reprint of Figure 3-24. Reduced interaction diagram

To reduce the analytical effort required by the general analysis method, a simplified design method is pro-

posed. This method produces a reduced member capacity diagram N-M* that approximates those given by

the general method. The envelope is defined in terms of normalized axial force and normalized moment,

274
7. Conclusions and Summary

as given in the figure above. This envelope was shown to be conservative and accurate relative to the more

precise general method calculations used for the arch concepts in the parametric design study (see Figure

6-19 on page 242).

Four eccentrically loaded columns made from ultra high-performance fibre-reinforced concrete were tes-

ted to failure at the University of Toronto. These columns had nominal slenderness ratios of 60, 80, 100,

and 120. The longest column was 3466 mm long with a square cross-section of 100 mm by 100 mm. The

following measurements were recorded: initial eccentricity of the load, deflection of the column at mid-

span, strains along the flexural compression face of the column, displacements of the centreline of the

column, and the applied load. From this data, load-deflection and axial force-moment responses from the

experiments were compared to responses predicted by the general method (see Figure 3-17 on page 67).

Observed and predicted ultimate loads were in good agreement, with percent differences of +6.1%, +0.8%,

+14.5%, and +10.1%, which are acceptable given the assumed margins of error. Hence, the general analysis

method is validated as a tool for modelling the behaviour of eccentrically loaded columns.

For future work, these column tests can be repeated with greater quantity to increase the statistical signi-

ficance of the results. Load tests of hollow members would better represent the type of highly efficient

members that are proposed in the design concepts, and ensure the applicability of the general analysis

method to various sectional shapes.

7.5 Arch system behaviour

A large quantity of technical information on the statical analysis of arch bridges was presented in Chapter

4. The most important topics and insights have already been summarized in Section 4.6 on page 140. The

most important topics are reviewed here.

The form of arches is chosen such that dead loads are carried in pure compression along the centroidal

axis of the arch. Nonuniform loads and restrained deformations are carried by both the deck and arch in

275
7. Conclusions and Summary

bending. First-order bending moments diagrams, bending moment influence lines and bending moment

envelopes are shown in Figures 4-6 through 4-10, starting on page 93.

The following three assumptions were used to define a simplified statical system for arches: (1) spandrel

columns are pin-connected to the arch and deck, (2) the flexural rigidity of the arch projected on to the

horizontal axis is constant over the entire span and is equal to the flexural rigidity of the arch at the crown,

and (3) vertical deflections of the deck and arch are equal along the length of the span. The validity and

consequences of these assumptions are illustrated and discussed in detail in Section 4.4.1. The overall beha-

viour that results from these assumptions is that bending moments imposed on the system are shared by

the arch rib and deck girder in proportion to their respective flexural rigidities.

Using the simplified statical system, the force method was used to solve closed form solutions of the axial

forces, bending moments, and deflections at critical points in fixed arch system. First this was done for

non-shallow arches, for which axial deformations are neglected in compatibility equations. Solutions for

this analysis are shown in Table 4-3 on page 109. The force method was then resolved for shallow arches,

for which axial deformations are included in compatibility equations. These shallow arch solutions are

summarized in Figure 4-16 on page 113.

The transition between fixed arch behaviour and fixed beam behaviour was identified by comparing sec-

tional forces in shallow arches and fixed beams. The governing non-dimensional ratio, called the arch-

beam parameter " is defined as follows, where f is arch rise, EIsys is system flexural rigidity and EAcrown is

arch axial rigidity:

45 EI sys
!= "
4 f 2 EAcrown

A related shallowness parameter called the modified slenderness ratio, is defined as follows, where rsys is

system radius of gyration:

276
7. Conclusions and Summary

EAcrown 2 f
!f = 2 f =
EI sys rsys

The effect of these parameters are illustrated in the figure below. As (f is decreased from infinite to zero

(i.e. " is increased from 0 to 1), sectional forces are changed as follows: (1) bending moments of the system

caused by uniform loads are increased from zero to those in a fixed-fixed beam, and (2) horizontal arch re-

actions caused by uniform loads are decreased from those in a non-shallow arch to zero. Based on results,

a minimum modified slenderness ratio value of 20 is proposed. This threshold ensures that greater than

90% of permanent loads are carried in pure compression along the pressure line.

Ratio of moments in M(x)arch 1


fixed arches to M(x)beam
moments in fixed
beams under
uniformly distributed
load 0.5 Fixed arches with !f > 20
Proposed threshold for
efficient arch behaviour: have bending moments
!f > 20 that are less than 10% of
those in fixed beams
with the same span and
uniform loading
0
0 20 40 60
Modified slenderness ratio !f

Ratio of horizontal H(! !f ) 1


reaction in shallow fixed H(! !f = " )
arches to horizontal
reaction in non-shallow
arches under uniformly Fixed arches with !f > 20
distributed load: !f > 20 resist more than 90% of
0.5
uniform loads in axial
compression

0
0 20 40 60
Modified slenderness ratio !f

Reprint of Figure 4-17. Transition between fixed arch and fixed beam

The modified slenderness ratio (f has also been used by others to determine whether antisymmetric bifur-

cation buckling or symmetric snap-through buckling modes occur in shallow arches. The buckling resist-

ance of shallow fixed arches are less than the buckling resistance of non-shallow fixed arches, and are con-

siderably less when (f < 19.8, for which symmetric buckling modes are critical.

277
7. Conclusions and Summary

When fixed arches buckle anti-symmetrically, their mode shape is similar to the second buckling mode

shape of straight fixed-fixed beams. Thus an appropriate model for non-shallow arch buckling is that of an

idealized straight fixed-fixed beam that is laterally supported at midspan. The flexural buckling resistance

HE of non-shallow arches can be calculated as follows, where EIsys is the sum of flexural rigidities of the

deck and arch, and kS is effective arc length:

EI sys
HE = ! 2
( kS )2

This quantity along with modified slenderness ratio (f can be used to calculate the buckling resistance of

shallow arches HE,f:

$ "2 '
H E, f = & 0.6 ± 0.4 1 ! 30.7 2 ) * H E # f > 19.8
% #f (

With these buckling resistances, second-order deflections w and moments of arches can be calculated us-

ing Vianello’s method of successive approximations. First-order deflections w0 calculated using the force

method are multiplied by a deflection magnifier, which depends on the ratio of horizontal reaction H to

buckling resistance HE:

1
w(x) = w0 (x)
H
1!
HE

To account for material nonlinearity, the flexural rigidity of the system EIsys is taken as the secant stiffness

of the system κ·EIsys. At serviceability limits states, secant stiffness factor κ can be taken as unity, represent-

ing the uncracked, gross flexural rigidities EI0 of the deck and arch. At ultimate limit states, κ can conser-

vatively be taken between 0.5 and 0.75, representing the expected material softening of the system. Expres-

sions for secant stiffnesses at ultimate are given in Section 7.3.

278
7. Conclusions and Summary

The eccentricity e of the deformed arch member and the line of action of load is calculated by multiplying

second-order deflections by the eccentricity ratio &, which is 0.733 or 0.5 depending on the type of loading

applied, as shown in the figure below.

distributed loading over distributed loading over


left half of arch span middle third of arch span

q q

H H H H
Z = second-order deflection
Z Z at critical section
H
H H H H undeformed line of action
H H H deformed line of action
e=ƦÃZ e = second-order eccentricity
e=ƦÃZ
e=0.733ÃZ of load at critical section
e=0.5ÃZ

Reprint of Figure 4-20. Analytical model for second-order analysis of arches

Applying these calculation methods can get complicated when considering factored load combinations of

dead load, live load, creep, shrinkage, and temperature changes. Long-term deflections have to be calcu-

lated separately and then expressed as an equivalent first-order short-term deflection before they are added

to actual first-order deflections. A detailed list of calculations for total sectional force demands are sum-

marized in Table 4-4 on page 128. Sample calculations of maximum sectional forces at the crown and

quarter-point of arches are shown in Appendix D.

Many of the analytical methods for calculating forces in arches have been verified experimentally in the

past on full-scale arch bridges and on small-scale linear-elastic models. Load tests on fixed, shallow arches

made from ultra high-performance fibre-reinforced concrete would further substantiate the validity of the

simplified methods presented in this thesis. Other important analytical studies that can be considered for

future work include: (1) the dynamic response of arch systems to seismic and wind loading, and (2) closed

form equations for shear forces in shallow arches. Inertial forces caused by earthquakes are usually mitig-

ated in structures by providing sufficient ductility or damping at critical locations in the system. The added

complication in arch bridges is that deformations of the arch off the pressure line will cause potentially

severe second-order moments. With regard to the second topic, shear calculations can be derived from the
279
7. Conclusions and Summary

force method as done for axial force and moment. Although shear requirements are typically non-critical

in conventionally reinforce concrete arches, they might be critical as more efficient box sections are used.

7.6 Trends observed from existing concrete arch bridges

Articles and drawings of fifty-eight existing concrete arch bridges built over the last century were collected

and processed into a comprehensive database. Because a majority of the reference bridges were collected

from authoritative engineering journals, the database is considered to be representative of the current

state-state-of-the-art of concrete arches bridges. The database serves as a reference guide for bridge design-

ers, as well as a tool for evaluating the material efficiency and feasibility of new design concepts. The com-

plete database is succinctly presented on three 11-inch by 17-inch sheets in Appendix A.

The following sets of data were collected for each bridge: name, location, method of construction, date of

completion, span, rise, number of spandrel columns, arch fixity, and sectional properties of arch, deck, and

spandrel columns. From these quantities, various geometric ratios and systems properties relevant to the

slenderness, shallowness, and degree of deck-stiffening of arch systems are calculated and compared. A

summary of all the trends observed from the database is given in Section 5.4 on page 194. Some of the

trends most relevant to the design study are discussed below.

Advances in arch construction have allowed for concrete arch bridges to span up to 420 m, as in the

Wanxian Yangtze Bridge. Modern cantilever methods of construction have reduced the amount of skilled

labour and throwaway materials through the increased mechanization of work. The construction methods

for most of the longest spanning concrete arch bridges can be categorized as follows: (1) cantilevering by

using an effective truss, where ties at the level of the deck and along the diagonals form the tension mem-

bers, and the arch and spandrel columns form the compression members, (2) cantilevering using stays em-

anating from temporary pylons above the springing lines, (3) cantilevering a steel pilot truss that is used

temporarily to support formwork and used permanently to reinforce the arch, and (4) cantilevering by ro-

tation, where the two halves of the arch are constructed near the springing lines and then rotated vertically

280
7. Conclusions and Summary

or horizontally to their final positions. All these construction methods would also be suitable for building

arches made from ultra high-performance fibre-reinforced concrete.

Of the bridges in the database, 46 out of 55 bridges (84%) have span-to-system-depth ratios in the range of

24 to 48. Span-to-system-depth ratios are better measures of the overall proportions of arch bridges than

are span-to-arch-depth ratios because they account for the effects of deck-stiffening. This ratio largely

effects the visual slenderness and transparency of arch bridges, as well as the internal lever arm of the arch

and deck girder. New design concepts in ultra high-performance fibre-reinforced concrete should aim to

have span-to-system-depth ratios toward the higher end of the range to maintain high transparency and a

slender profile.

Equivalent slab thicknesses by concrete volume hV were calculated for each bridge in the database. An

envelope of minimum slab thicknesses achieved by existing concrete technology is given as a function of

span length L:

# L
% + 0.569 m 0 ! L ! 193m
% 1276
hV = $
% L2 L
" + 2.21m L > 193m
% 22800 m 61.7
&

This minimum equivalent slab thickness envelope serves as a benchmark for the design study. If satisfact-

ory design concepts can achieve slab thicknesses significantly less than this envelope, then significant con-

struction cost-savings can be expected.

A simple statical analysis was used to estimate the effects of live load in fixed arch bridges in the database.

Comparisons of live load eccentricities, or displacements of resultant axial forces away from the centroid

of the arch, showed that the effects of live load diminish with increasing span length. This decreasing sens-

itivity to live load is analogous to the trivial effects of live loads in long-span suspension bridges. The

effects of restrained deformations, however, increase with increasing span length, and thus tend to govern

over the effects of live loads.

281
7. Conclusions and Summary

Because the observed trends had relatively little scatter considering the empirical nature of the study, in-

creasing the sample size of the database would not likely result in many new insights into the typical pro-

portions of concrete arch bridges. Thus, the perpetual increase of the database is not a worthwhile or prac-

tical goal for further research, unless it can be implemented in a more dynamic and interactive form, such

as through a website. Additions from bridges from certain under-represented countries such as Japan and

China would improve the representativeness of the database to those countries.

7.7 Innovative systems for concrete arches

Seventy-two concepts for arches using ultra high-performance fibre-reinforced concrete were designed.

Concepts differ by the three main parameters of the design study: span length (50 m to 400 m), span-to-

rise ratio (4:1 to 12:1), and degree of deck-stiffening (full, partial, or none). Concepts are designed primarily

for longitudinal axial force and bending demands caused by dead load, live load, creep, shrinkage, and

uniform changes in temperature, calculated according to the simplified methods described in Chapter 4.

The adequacy of given trial design were evaluated by comparing sectional force demands with member ca-

pacity N-M* interaction diagrams of the arch rib and deck girder, calculated using the general method de-

scribed in Section 3.2. If given designs were found to be unsatisfactory, slab thicknesses and section depths

were systematically increased until capacities exceeded demands. Sample calculations for Concept 80 are

shown in Appendix D, including a rudimentary shear design check.

The most common geometrical properties that were found to be satisfactory among concepts are summar-

ized in Table 6-29. Concepts whose demands required proportions greater than those shown in the table

include: (1) some concepts with spans of 50 m due to the low amount of natural prestress in the arch, and

(2) most concepts with span-to-rise ratios of 12:1 due to the high sectional force demands and high sensit-

ivity to crown deflections.

282
7. Conclusions and Summary

Reprint of Table 6-29. Most common geometrical quantities used among design concepts

class of system element section type span-to- slab width of box ÷


depth ratio thickness width of deck

self-stiffened arch deck double tee 180:1 50 mm n/a


arch hollow box 50:1 50-100 mm 1/2

deck-stiffened arch deck hollow box 50:1 50-150 mm 1/4


arch solid box 200:1 n/a 1/4
arch hollow box 180:1 200 mm 1/2

partially deck-stiffened arch deck hollow box 90:1, 100:1 100-200 mm 1/2
arch hollow box 90:1, 100:1 100-200 mm 1/2

Wall thicknesses as small as 50 mm are specified. Such thinness is made possible by: (1) the elimination of

conventional concrete cover requirements to depths of reinforcing, (2) the use of short 10 mm steel fibres,

and (3) the use of longitudinal stiffeners to prevent local stability problems. Based on a simple analytical

study on plate buckling, a maximum wall slenderness ratio of 24.5 was determined for ultra high-perform-

ance fibre-reinforced concrete. This upper limit ensures that thin concrete slabs are able to attain full com-

pressive strength without prematurely buckling. This corresponds to a maximum stiffener spacing of 1225

mm for 50 mm slabs or webs.

The need for both transverse ribs and longitudinal stiffeners poses a challenge for forming and casting the

hollow box members. A new segmental precasting method is proposed in Section 6.2.3. This involves cast-

ing the 100 mm end ribs separately from the main body of the segment. End ribs are formed with shear

keys and holes for prestressing against a precision bulkhead (see Figure below). Prior to casting, deformed

steel wire is installed along the edges that will eventually abut against the interior body of the segment.

Once the principal end rib is cast and cured, it is reversed and is used to form a similar match-cast end rib.

283
7. Conclusions and Summary

Partial reprint of Figure 6-16. Proposed match-casting method for abutting end ribs

The main body of the segment is cast using an exterior U-shaped base form and an interior core form.

After the end ribs are positioned 3 metres apart in the base form, with shear keys facing outward, the core

form is clamped to the two end ribs. The interior body of the segment including longitudinal stiffeners are

then cast. Once cured, the interior core form is released in parts and lifted out through the cavity of the

segment. The completed segment can then be lifted off the base form and transported to storage.

The proportions of the parametric design concepts were compared in terms of material efficiency, distribu-

tion of weight, slenderness, shallowness, and structural demands. A summary of all conclusions is given in

Section 6.3.10 on page 257. Some of the most important conclusions are discussed below.

284
7. Conclusions and Summary

Reprint of Table 6-15. Equivalent slab thicknesses based on volume of longitudinal concrete among all concepts.

equivalent slab thicknesses among all concepts, in m


concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 0.27 0.26 0.19 0.23 0.25 0.30 0.35 0.37 ‡PLQLPXPYDOXHVDPRQJ
33 to 40 deck-stiffened 4:1 0.17 0.25 0.31 0.33 0.35 0.47 0.49 0.51 triplets arein red type
65 to 72 partially deck-stiffened 4:1 0.26 0.27 0.28 0.29 0.30 0.32 0.36 0.36
9 to 16 self-stiffened 8:1 0.35 0.24 0.18 0.21 0.28 0.30 0.33 0.40 ‡PLQLPXPYDOXHVLQHDFK
41 to 48 deck-stiffened 8:1 0.15 0.23 0.28 0.31 0.32 0.39 0.41 0.45 column are in bold type
73 to 80 partially deck-stiffened 8:1 0.24 0.25 0.25 0.27 0.37 0.48 0.39 0.51
17 to 24 self-stiffened 12:1 0.34 0.23 0.22 0.30 0.38 0.50 0.62 0.75
49 to 56 deck-stiffened 12:1 0.18 0.33 0.48 0.64 0.75 0.85 0.95 1.11
81 to 88 partially deck-stiffened 12:1 0.24 0.35 0.45 0.66 1.08 1.08 1.09 1.10 m
database various various 0.61 0.65 0.69 0.72 0.90 1.30 1.91 2.74 m ‡PLQLPXPVODEWKLFNQHVVHV
from arch database

equivalent slab thicknesses as percentages of


minimum values achieved by concrete arches in database, in %
concept ID class of arch system L = 50 100 150 200 250 300 350 400 m
1 to 8 self-stiffened L:f = 4:1 45 40 28 31 28 23 18 14
33 to 40 deck-stiffened 4:1 28 38 45 45 39 36 26 19
65 to 72 partially deck-stiffened 4:1 43 41 41 40 33 25 19 13
9 to 16 self-stiffened 8:1 58 37 26 29 31 23 17 15 ‡YDOXHVJUHDWHUWKDQ
41 to 48 deck-stiffened 8:1 25 35 40 42 36 30 22 16 50% are in red type
73 to 80 partially deck-stiffened 8:1 40 38 37 37 41 37 20 19
17 to 24 self-stiffened 12:1 57 36 33 41 42 38 33 27
49 to 56 deck-stiffened 12:1 30 51 70 89 83 66 50 40
81 to 88 partially deck-stiffened 12:1 39 53 66 91 120 84 57 40 %

Concepts with span-to-rise ratios of 4:1 and 8:1 consistently had less than 50% of the minimum equivalent

slab thicknesses of existing concrete arch bridges in the database (see Table above). This suggests that the

costs associated with constructing concrete arch bridge has the potential to be significantly reduced. Con-

cepts with span-to-rise ratios of 12:1 were not able to achieve the same reduction in concrete volume.

Hence, there are cost premiums associated with using arches with span-to-rise ratios in excess of 8:1.

Between self-stiffened, deck-stiffened, and partially deck-stiffened systems, deck-stiffened systems con-

sumed the least amount of concrete for spans of 50 m, while self-stiffened systems consumed the least

amount of concrete for spans between 150 m to 400 m. For spans of 100 m, all systems consumed approx-

imately the same amount of concrete. These results agree with the conclusions made in the modified arch

stress analysis presented in Section 4.7. It was concluded that self-stiffened arch systems are most efficient

among systems because they: (1) maximize the use of increased flexural capacity provided by the natural

precompression in the arch, and (2) isolate deck girders from the the global system, which minimizes their

bending moment demands. For the shorter 50 m span, deck-stiffened arch systems were found to be the

285
7. Conclusions and Summary

most efficient because the natural prestress of the arch caused by dead load was low in comparison to the

imposed flexural stresses.

Most concepts tended to have span-to-system-depth ratios between 39:1 and 50:1, which is on par with the

ratios found among existing concrete arch bridges in the database. Thus the degree of visual slenderness

and transparency that can be achieved by the arch concepts are about the same as in conventional concrete

arches. The savings of concrete are thus related to the thinning of webs and slabs rather than the reduction

of overall depth.

In terms of sensitivity to second-order effects, most concepts had system slenderness ratios between 39 and

60. In this range, additional deflections and moments caused by second-order effects can be significant.

These systems are, however, are not slender enough that ultimate loads Q* are reached before sectional

failure is reached. This means that capacities of the member (N-M* interaction diagram) are only slightly

less than sectional capacities (N-M interaction diagram).

In terms of arch shallowness, concepts with span-to-rise ratios of 4:1 and 8:1 tended to have modified slen-

derness ratios much greater than the efficiency threshold value of 20. Their responses to load are thus

caused primarily by arch action. Concepts with span-to-rise ratios of 12:1 tended to have modified slender-

ness ratios around 20 or less. These concepts are thus affected by significant amounts of bending moment

under uniform load and are sensitive to axial deformations and crown deflections.

The scope of this thesis has been limited to the design of concrete arch bridges for axial forces and bending

moments caused by dead load, live load, creep, shrinkage, and temperature. For conventional concrete

arches, these loads and sectional forces tend to govern over all other design considerations. Although this

is likely true for the concepts proposed in this thesis, there remains a number of design checks and details

that have to be investigated. Design checks for wind, seismic, and construction loads need to be done, as

well as for unbalanced lane loading. The transverse design of the structural system, including the trans-

verse ribs and stiffeners, needs further development and validation.

286
7. Conclusions and Summary

The use of thin 50 mm webs has been validated with a rudimentary shear design check for Concept 80.

Since this concept has one of the longer spans and higher span-to-rise ratios, it is anticipated that the thin

webs specified in most of the parametric design concepts are adequate. The use of thin, ribbed slabs in the

proposed concepts requires further validation with respect to the transmission of local wheel loads to the

primary structural elements of the system. Analytical studies by Spasojević (2008) have shown that thin

post-tensioned ribbed ultra high-performance fibre-reinforced concrete slabs with flange and web thick-

nesses of 40-60 mm were capable of transmitting truck wheel loads over spans of 12 m between supporting

girders. For the concepts proposed in this thesis, similar thin slabs would only have to span 3 m or less

between end ribs of the segment. Much deeper rib depths of 400 mm to 500 mm are specified, which

serves to increase capacity relative to Spasojević’s ribbed deck designs. Based on these reference comparis-

ons, it is anticipated that the proposed thin slabs are satisfactory, or at least satisfactory with very minor

increases in concrete volume.

Other topics for the future development of these systems could include: (1) details and truss models related

to the spread of forces at the anchorage zones of the main prestressing tendons need to be developed, (2)

the transfer of shear forces across the match-cast end ribs of abutting segments, and (3) the transfer of

shear and tension across the spliced joint between the end ribs and interior body of the segment.

Based on the findings in this thesis, ultra high-performance fibre-reinforced concrete is considered to be

an excellent material for arch bridges. Its enhanced properties of increased compressive and tensile

strength, reduced creep and shrinkage strains, and short fibre lengths, all serve to increase member capa-

cities and reduce the dead load demand of the structure. The resulting systems for concrete bridges are

lighter, more efficient, and more economical to build than arches made using convectional concrete

technology.

287
References
Association française du génie civil (AFGC). 2002. Ultra High Performance Fibre-Reinforced Concretes: Interim

Recommendations. France.

Aitcin, P. C., M. Lachemi, R. Adeline, and P. Richard. 1998. The Sherbrooke reactive powder concrete footbridge.

Structural Engineering International. 8(2): 140-144.

Asplund, S. O. 1963. Deflection theory of arches. ASCE Transactions. 128: 307-341.

Austin, W. J. 1971. In-plane bending and buckling of arches. ASCE Journal of the Structural Division. 97(5): 1575-1592.

Bazant, Z. P. 1972. Prediction of concrete creep effects using age-adjusted effective modulus method. ACI Materials

Journal. 69(4): 212-17.

Benaim, R. 2007. The Design of Prestressed Concrete Bridges : Concepts and Principles. London: Taylor & Francis.

Bill, M., and R. Maillart. 1955. Robert Maillart. Translated by W. P. M. K. Clay. Zürich, Switzerland: Girsberger.

Billington, D. P. 1973. Deck-stiffened arch bridges of Robert Maillart. Journal of the Structural Division. 1527-1539.

Billington, D. P. 1979. Robert Maillart's Bridges: The Art of Engineering. Princeton: Princeton University Press.

Billington, D. P.. 2003. The Art of Structural Design: A Swiss Legacy. Princeton, New Jersey: Princeton University Art

Museum.

Bradford, M. A., T. Wang, Y. L. Pi, and R. I. Gilbert. 2007. In-plane stability of parabolic arches with horizontal spring

supports, I: Theory. Journal of Structural Engineering. 133(8): 1130-1137.

Cai, J. G., J. Feng, Y. Chen, and L. F. Huang. 2009. In-plane elastic stability of fixed parabolic shallow arches. Science

in China Series E: Technological Sciences. 52(3): 596-602.

Cai, J., and J. Feng. 2009. Erratum to “In-plane elastic stability of fixed parabolic shallow arches.” Science in China

Series E: Technological Sciences. 52(3).

Candrlic, V., J. Bleiziffer, and A. Mandic. 2001. Bakar Bridge designed in reactive powder concrete. Arch'01: troisième

conférence internationale sur les ponts en arc (Third international arch bridges conference). Paris, France.

Culmann, C., and W. Ritter. 1906. Der Bogen (Arches). ETH Zurich.

288
Chen, B. 2007. An overview of concrete and cfst arch bridges in China. Arch'07: 5th International Conference on

Arch Bridges. Sept. 12-14, Madeira, Portugal.

Cross, H. 1932. Analysis of continuous frames by distributing fixed-end moments. Transactions of the American

Society of Civil Engineers. 96: 1-79.

Denarié, E., E. Brühwiler, A. Znidaric, Y. Houst, and R. Rohleder. 2005. Full scale application of Uhpfrc for the

rehabilitation of bridges: From the lab to the field. Sustainable and Advanced Materials for Road Infrastructure,

Deliverable D22. Accessed online: http://www.fehrl.org, April 2010.

Dickie, J. F., and P. Broughton. 1971. Stability criteria for shallow arches. ASCE Journal of the Engineering Mechanics

Division. 97(3): 951-965.

Dinnik, A. N. 1955. Buckling and Torsion. Moscow.

El-Arousy, A. 1942. Studien über das elastische Verhalten von Brückengewöolben einschliesslich des

Zusammenwirkens mit dem Aufbau. Dissertation. Institut für Baustatik, ETH Zurich.

Ewert, S. 1999. Betonbogenbrücken: Mehr als 200 m Spannweite. Beton- und Stahlbetonbau. 94(9): 377-388.

Favre, R., and J. de Castro San Román. 2001. The arch: enduring and endearing. Structural Concrete. 2(4): 187-200.

Fonseca, A., and F. Mato. 2005. Infant Henrique Bridge over the River Douro, Porto. Structural Engineering

International. 15(2): 85-87.

Freytag, B., M. Reichel, L. Sparowitz, G. Santner, and G. Heinzle. 2009. Groβversuch WILD-Brücke: versuchsgestützte

Bemessung einer UHPC-Bogenbrücke. Beton- und Stahlbetonbau. 104(3): 134-144.

Gaber, E. 1934. Über die Knicksicherheit vollwandiger Bogen (On the bucking strength of solid arches). Bautechnik.

646-651.

Galambos, T. V. 1998. Guide to Stability Design Criteria for Metal Structures. John Wiley.

Gere, J. M. 1963. Moment Distribution. Princeton, New Jersey: Van Nostrand.

Gjelsvik, A., and S. R. Bodner. 1962. The energy criterion and snap buckling of arches. ASCE Journal of the

Engineering Mechanics Division. 88: 87-134.

Google. 2009. Google Maps. Accessed online: http://maps.google.com.

289
Graybeal, B. A. 2006. Material property characterization of Uhpc. No. Fhwa-Hrt-06-103. US Department of

Transportation, Federal Highway Administration.

Graybeal, B. A. 2007. Compressive behavior of ultra-high-performance fiber-reinforced concrete. ACI Materials

Journal. 104(2): 146-152.

Graybeal, B. A., and J. L. Hartmann. 2005. Construction of an optimized Uhpc vehicle bridge. ACI Special

Publication. 228: 1109-1118.

Gruber, L., and C. Menn. 1978. Berechnung und Bemessung schlanker Stahlbetonstützen (Analysis and design of

slender reinforced concrete columns). Report No. 84. Institut für Baustatik und Konstruktion, ETH Zürich,

Zürich, Switzerland.

Habel, K. 2004. Structural behaviour of elements combining ultra-high performance fibre reinforced concretes and

reinforced concrete. Doctoral dissertation. Génie Civil, École Polytechnique Fédérale de Lausanne, Switzerland.

Habel, K., J. P. Charron, S. Braike, R. D. Hooton, P. Gauvreau, and B. Massicotte. 2008. Ultra-high performance fibre

reinforced concrete mix design in central Canada. Canadian Journal of Civil Engineering. 35(2): 217-224.

Habel, K., and P. Gauvreau. 2008. Response of ultra-high performance fiber reinforced concrete (UHPFRC) to impact

and static loading. Cement and Concrete Composites. 30(10): 938-946.

Huh, S. B., and Y. J. Byun. 2005. Sun-Yu Pedestrian Arch Bridge, Seoul, Korea. Structural Engineering International.

15(1): 32-35.

Hünleim, W., and P. Ruse. 1985. Ein neues Verfahren für den Bau von Bogenbrücken dargestelit am Bau der

Argentobelbrücke Würgau. Der Bauingenieur. 60(12): 487-493.

Johnston, B. G. 1976. Guide to Stability Design Criteria for Metal Structures. John Wiley & Sons.

Kennedy, J. S., and A. S. Aggarwal. 1971. Effect of weight on large deflections of arches. ASCE Journal of the

Engineering Mechanics Division. 97(3): 637-644.

Ketchum, M. S. 1934. Thin-section concrete arches as built in Switzerland. Engineering News Record. Jan. 11: 44-45.

Kollbrunner, C. F. 1936. Versuche über die Knicksicherheit und die Grundschwingungszahl vollwandiger Bogen

(Experiments on the buckling strength and fundamental frequency of solid arches). Bautechnik. March:

186-188.

290
Kremer, I. 2008. Hydraulic crown jacking procedure shapes the concrete arch of the Third Millenium Bridge.

InfraStructures. November. Accessed online: http://www.infrastructures.com/0808/enerpac2.htm, April 2010.

Laffranchi, M., and P. Marti. 1997. Robert Maillart's curved concrete arch bridges. Journal of Structural Engineering.

123(10): 1280-1286

Lin, Y., Z. Zhang, B. Ma, and L. Zhou. 2004. Lupu Arch Bridge, Shanghai. Structural Engineering International. 14(1):

24-26.

Ma, J., and M. Orgass. 2004. Comparative investigations on ultra-high performance concrete with and without coarse

aggregates. Leipzig Annual Civil Engineering Report (lacer) No. 9.

Maillart, R. 1934. Gekrümmte Eisenbeton-Bogenbrücken (Curved reinforced concrete arch bridges). Schweizerische

Bauzeitung. 52(11): 132-133.

Melan, J., and D. B. Steinman. 1913. Theory of Arches and Suspension Bridges. Translated by D. B. Steinman. Chicago:

Clark Pub. Co.

Menn, C. 1990. Prestressed Concrete Bridges. Translated and edited by P. Gauvreau. Basel: Birkhäuser Verlag.

Mondorf, P. E. 2006. Concrete Bridges. New York: Taylor & Francis.

Nathan, N. D. 1985. Rational analysis and design of prestressed concrete beam columns and wall panels. Prestressed

Concrete Institute Journal. 30(3): 82-133.

Neville, A. M. 1995. Properties of Concrete. Fourth Edition. Toronto: Prentice Hall.

Ordóñez, F., and J. Antonio. 1979. Eugène Freyssinet. Barcelona: Coop. Industrial Trabajo Association.

Ostrander, J. R., and D. C. Oliver. 1987. Construction of the Broadway Bridge at Saskatoon, Saskatchewan, in 1932.

Canadian Journal of Civil Engineering. 14: 429-438.

Radić, J., Z. Šavor, and I. Gukov. 2005. New contribution to concrete arch bridge construction. Transportation

Research Record: Journal of the Transportation Research Board. 11(1): 235-342.

Rambøll, B. J. 1944. Om buers stabilitet (About arch stability). Bygningsstatiske Meddelelser (2).

Ritter, W. 1877. Versteifungsfachwerke bei Bogen- und Hängebrücken (Stiffening trusses in arch and suspension

bridges). Zeitschrift für Bauwesen. 27: 189-207.

291
Šavor, Z., and J. Bleiziffer. 2008. Long span concrete arch bridges of Europe. Chinese-Croatian Joint Colloguium Long

Arch Bridges, July 10-14, Brijuni Islands, China.

Schreyer, H. L., and E. F. Masur. 1966. Buckling of shallow arches. ASCE Journal of the Engineering Mechanics

Division. 92: 1-19.

Seegers, K. H. 1963. Beton-Bogenbrücke über den Parametta-Fluβ in Sydney. der Bauingenieur 38(12): 477-478.

Spasojević, A. 2008. Structural implications of ultra-high performance fibre-reinforced concrete in bridge design.

Ph.D. dissertation No. 4051. École ploytechnique fédérale de Lausanne, Switzerland.

Spiegel, M. R., and J. Liu. 1999. Mathematical Handbook of Formulas and Tables, Second Edition. Toronto: Schaum's

Outline Series.

Šram, S. 1982. Ausführung der Brücken vom jugoslawischen Festland zu den Adriainseln St. Marko und Krk. Beton-

und Stahlbetonbau. 77(8): 197-203.

Stellmann, W. L. O. 1966. Brücke über den Rio Paraná in Foz Do Iguaçú, Brasilien. Beton- und Stahlbetonbau. 61(6):

145-149.

Stussi, F. 1935. Aktuelle baustatische Probleme der Konstruktionspraxis (Current problems of construction design

practice). Schweizerische Bauzeitung. 106(12): 132-36.

Swartz, S. E., and V. H. Rosebraugh. 1974. Buckling of reinforced concrete plates. ASCE Journal of the Structural

Division. 100(1): 195-208.

Tanner, P., and J. L. Bellod. 2005. Widening of the Elche De La Sierra Arch Bridge, Spain. Structural Engineering

International. 15(3): 148-150.

Taylor, A. W., R. B. Rowell, and J. E. Breen. 1995. Behaviour of thin-walled box piers. ACI Structural Journal. 92(3):

319-333.

Timoshenko, S. 1936. Theory of Elastic Stability. New York: McGraw-Hill Book Company, Inc.

Trost, H. 1967. Auswirkungen des Superpositionsprinzips auf Kriech- und Relaxationsprobleme bei Beton und

Spannbeton (Implications of the superposition principle in creep and relaxation problems for concrete and

prestressed concrete. Beton und Stahlbetonbau (10): 230-238.

292
US Department of Transportation Federal Highway Administration (US DOT FHWA). 2003. Az-Nv Hpp 93(3): Hoover

Dam Bypass Colorado River Bridge: 60% Plan Submittal. April 4.

Vianello. 1898. Graphische Untersuchung der Knickfestigkeit gerader Stäbe. Zeitschrift des Vereins Deutscher

Ingenieure. 1436-1443.

Vogel, T, and P. Marti. 1997. Christian Menn, Brückenbauer. Basel, Switzerland: Birkhäuser Verlag.

Walraven, J. 2008. On the way to design recommendations for Uhpfrc. Second International Symposium on Ultra

High Performance Concrete, Kassel, Germany.

Wang, T., M. A. Bradford, and R. I. Gilbert. 2006. Creep buckling of shallow parabolic concrete arches. Journal of

Structural Engineering. 132: 1641-1648

Yan, G., and Z. H. Yang. 1997. Wanxian Yangtze Bridge, China. Structural Engineering International. 7(3): 164-166.

293
Appendices

294
Appendix A: Integrated Database of Concrete Arch Bridges

This appendix contains the original integrated database of existing concrete arch bridges from which most

figures in Chapter 5 have been adapted from. The complete database is presented on three 11” by 17” sheets.

The first sheet shows the elevation and section views of each bridge, a world map with the location of each

bridge, and a list of references. Each bridge is assigned a bridge id, in order of span length, with id 1 being

the longest spanning bridge in the database. The second sheet tabulates the section properties, geometrical

ratios, structural type, country, type of construction, and year of completion of each bridge. This quantitat-

ive data is presented in graphical form in the second and third sheets. Most plots are accompanied by an-

notations that propose trends that can be observed among bridges in the database.

This database can be used as references and design tools for designers. It allows for the evaluation of new

bridge concepts in terms of structural efficiency and feasibility, relative to a large set of concrete arch

bridges. The integrated form of this database allows for further study of the relationships between data

sets, from which the reader can draw personal insights that are specific to the span range or arch types that

he or she may be interested in. For further reference, readers can consult the primary sources listed on the

first sheet, which contain further information on each bridge in the database.

295
Location of arch bridges
Bridge drawings bridge ID
A
Elevations at 1:10000 scale name
Sections at 1:1000 scale year of completion

ÇÇ Ç
Ç Ç
Ç Ç Ç
Ç ÇÇ
A
Ç Ç ÇÇÇ ÇÇÇ Ç
ÇÇÇ Ç ÇÇ ÇÇ ÇÇÇ
typical section
Ç Ç Ç Ç
ÇÇ
column section
Ç B
Ç
Ç ÇÇ Ç
Ç
Ç
Ç B
ÇÇ
Ç Ç Ç
Ç
Taiwan Europe

References
Organized by bridge ID
1. Dajun, D., and L. Weiqing. 1999. Neue Entwicklungen Bei Hochhäusern Und Gro en Brücken Aus Beton in China. Beton- und Stahlbetonbau 94(4): 178-85.
2/11.Radi
, J., J. Bleiziffer, and D. Tkal i
. 2005. Maintaining Safety and Serviceability of Concrete Bridges in Croatia. Bridge Structures 1(3): 327-44.
3. US Dept. of Transportation FHWA. 2003. AZ-NV HPP 93(3) Hoover Dam Bypass Colorado River Bridge: 60% Submittal. April 4.
4. Baxter, J. W., A. F. Gee, and H. B. James. 1965. Gladesville Bridge. Proceedings of the Institute of Civil Engineers. 
5. Stellmann, W. L. O. 1966. Brücke Über Den Rio Paraná in Foz Do Iguaçú, Brasilien. Beton- und Stahlbetonbau 61(6): 145-49.
6. Fonseca, A., and F. Mato. 2005. Infant Henrique Bridge Over the River Douro, Porto. Struct. Eng. Int. 15(2): 85-87.
7. Liebenberg, A.C., and M.G. Latimer. 2001. Bloukrans Bridge. Arch'01: troisième conférence internationale sur les ponts en arc, Sept. 19-21, Paris.
8. @7BH3:: 
3X31B7=<!3/AC@3;3<BA=<B63'/<2Q@7253   '7FB6=<5@3AA=4B63'
9. Kamimura, K., M. Kouno, K. Okada, and T. Tsuka. 2001. Design and Construction of the Tensho Concrete Arch Bridge. Arch'01: Third Int'l Conf. on Arch Bridges, Paris.
10. Radi
, J., Z. avor, J. Bleiziffer, and I. Kalafati
. 2008. ibenik Bridge - Design, Construction and Assessment of Present Condition. Chinese-Croatian Joint Colloquium:
Long Arch Bridges, July 10-14, Brijuni Islands, China.
12. D=<+Q:43:& 73(/:0@S193J03@73+7:233@/Beton- und Stahlbetonbau 94: 546-51.
13. State of Nevada Department of Transportation. 2000. B-1948-N/S Galena Creek Bridge. April 19.
14. avor, Z., N. Mujkanovi
, and G. Hrelja. 2008. Design and Construction of Krka River Arch Bridge. Chinese-Croatian Joint Colloquium Long Arch Bridges, Brijuni Islands.
15. Radi
, J., Z. avor, V. Prpi
, M. Friedl, and . deric. 2008. Design and Construction of the Maslenica Highway Bridge. Chinese-Croatian Joint Colloquium, Brijuni Islands.
16. avor, Z., N. Mujkanovi
, G. Hrelja, and J. Bleiziffer. 2008. Reconstruction of the Pag Bridge. Chinese-Croatian Joint Colloquium: Long Arch Bridges, Brijuni Islands.
17. Wang, J. J., and F. T. K. Au. 2001. Modong Hongshui River Bridge, China. Structural Engineering International 11(2): 101-03.
18. Aigner, F. 1968. Stahlbeton Bogenbrücken Auf Der Österreichischen Brenner Autobahn. Der Bauingenieur.
19. Igarashi. 1976. Die Hokowazu-Brücke in Japan, Ein Stahlbetonbogen Im Freien Vorbau. Der Bauingenieur (8).
20. Anonymous. 1978. Communications Techniques Belges 8. Annales des Travaux Publics de Belgique, No. 1-2.
21. Cheng, K. M. 1994. The Pitan Bridge, Taiwan. Structural Engineering International 4(4): 231-34.
22. Standfu , F. 1990. Nationalbericht: Brücken in Der Bundesrepublik Deutschland. Beton- und Stahlbetonbau (5): 106-13.
23. +Q <3@306/@2B&'16</03:/<2+Q@<3@ 73(/:0@S193&=BBE37:"319/@0C@5;-C533@ Beton- und Stahlbetonbau (10-11).
24. Bouchet. 1964. L'Echangeur La Araña Du Croisement Des Autoroutes Nord-Sud Et Est-Ouest À Caracas (Venezuela). La Technique des Travaux (Belgique).
25. Hünleim, W., and P. Ruse. 1985. Ein Neues Verfahren Für Den Bau Von Bogenbrücken Dargestelit Am Bau Der Argentobelbrücke Würgau. Der Bauingenieur (12).
26. deric, ., A. Runji
, and G. Hrelja. 2008. Design and Construction of Cetina Arch Bridge. Chinese-Croatian Joint Colloquium on Long Arch Bridges, Brijuni Islands.
27. Koppel, A. J., and R. Walser. 1991. Hundwilertobelbrucke: Ein Bemerkenswerter Neubau. Schweizer Ingenieur und Architekt 109(11): 250-56.
28. Taiwan Area National Expressway Engineering Bureau, MOTC. 1999. (In Mandarin) The Second Freeway: An Anthology of Particular Bridges.
29. BC Ministry of Transportation and Highways. 1995. R6-V433-3051 Big Qualicum River Upstream Bridge No. 3051 (Drawings).
30. Goodyear, D., N. Smit, J. Heacock, and S. Starkey. 2001. Respect for Tradition: The New Crooked River Gorge Bridge. Arch'01: Third Int'l Conf. on Arch Bridges, Paris.
31. Podolny, W., and J. M. Muller. 1982. Construction and Design of Prestressed Concrete Segmental Bridges. New York: Wiley.
32. Inaudi, D., A. Rufenacht, B. Arx, H.P. Noher, and S. 2002. Monitoring of a Concrete Arch Bridge During Construction. Proc. SPIE International Society for Optical Eng.
33. Vogel, T, and P. Marti. 1997. Christian Menn, Brückenbauer. Basel, Switzerland: Birkhäuser Verlag.
34. State of California Department of Public Works Division of Highways. 1932. Bridge Across Bixby Creek (as Built Plans) Doc. No. 50001001.
35. Vogel, T, and P. Marti. 1997. Christian Menn, Brückenbauer. Basel, Switzerland: Birkhäuser Verlag.
36. Eilzer, W., W. Schmidtmann, and R. Jung. 2005. Die Wirrbachtalbrücke. Beton- und Stahlbetonbau 100(3): 236-40.
37/55/57/58. Bill, M., and R. Maillart. 1955. Robert Maillart. Translated by W. P. M. K. Clay. Zürich, Switzerland: Girsberger.
38/50/56. Sinotech Engineering Consultants, Ltd. Personal correspondence, August 2007.
39. !Q@A16 
73;S<23@B=03:@S19337(3C43<;/<B=<>>3<H37:Schweizerische Bauzeitung 53 (7): 81.
40. Schwaab, E., and A. Gattner. 1953. Stra enbrücke Über Das Tiefe Tal Bei Rosshaupten, Allgau. Beton- und Stahlbetonbau (11).
41. Tonello, J., and P. Dal-Palu. 2001. Le Pont Du Triple Saut (the Triple Jump Bridge). Arch'01: troisième conférence internationale sur les ponts en arc, Sept. 19-21, Paris.
42. 'B/B3=4/:74=@<7/3>/@B;3<B=4$C0:71+=@9A7D7A7=<=4756E/GA @72531@=AA&CAA7/<C:16/@B6?C/93&3B@=WB$@=831B"= 

43. Petrangeli, M.P. 2007. Prefabrication of Medium Span Arch Bridges. Arch'07: 5th Intl Conference on Arch Bridges, Sept. 12-14, at Madeira, Portugal.
44. Zimmermann, W., and N. im Gailtal. 1999. Der Bau Der Stampfgrabenbrücke. Beton- und Stahlbetonbau 99(4): 304-10.
45. State of California Department of Public Works Division of Highways. 1932. Bridge Across Rocky Creek (as Built Plans) Doc. No. 50001081.
46. Baxter, D. J., and T. A. Balan. 2008. Design of the Fulton Road Bridge Precast Segmental Concrete Arches. Journal of Bridge Engineering: 476-82.
47. Blinkov, L. S., E. Cosolo, and S. N. Valiev. 2001. Rehabilitation of a Bridge Over the Matsesta River, Russian Fed. Structural Engineering International 11(3): 181-83.
48. Ostrander, J. R., and D. C. Oliver. 1987. Construction of the Broadway Bridge At Saskatoon, Saskatchewan, in 1932. Canadian Journal of Civil Engineering 14: 429-38.
49. State of California Department of Public Works Division of Highways. 1937. Bridge Across Big Creek (as Built Plans) Doc. No, 50001093.
51. *73B(C3"36<('16:73;/<<&37<B83A/<2(/CA163@

<E3<2C<5*=<=16:37ABC<5A03B=<3<373@=53<0@S193+Q:9/C;-C533@/0 
Beton- und Stahlbetonbau 100(11): 931-38.
52. Tanner, P., and J. L. Bellod. 2005. Widening of the Elche De La Sierra Arch Bridge, Spain. Structural Engineering International (3): 148-50.
53. Tandon, M. 1995. Arch Bridge At Dodan Nallah. Proceedings of the First International Conference on Arch Bridges, Bolton, UK.

A Study of 58 Concrete Arch Bridges


—Drawings, Map, and References
prepared by Jason Salonga
assisted by Catherine Chen
Department of Civil Engineering
University of Toronto
296
span length in m 0 100 200 300 400 500

bridge ID 51 41 31 21 11 1
volume = deck area span eccentricity of live load : half arch depth
+ arch area arch length horizontal reaction / (unit weight deck width)
+ column area total spandrel column length max arch depth in m 8
Variable definitions UVYTHSPaLKZ`Z[LTTVTLU[VMPULY[PH
modified system moment of inertia average arch depth in m
interior deck width system inertia = deck inertia + arch inertia equiv. slab thickness by volume min arch depth in m
span spandrel
0 if deck is discontinuous column inertia : system inertia
column 7 3
deck depth arch inertia : system inertia
column PRGLÀHGVOHQGHUQHVVUDWLR 2 rise 3
rise section system radius of gyration mod. slenderness ratio 
arc length 6
slenderness ratio
slenderness ratio = effective arc length factor arc length span : arch + deck depth 
span system radius of gyration span : arch depth
arch depth 5 
system radius of gyration = system inertia system radius of gyr. 3
Bridge geometry and ratios arch area span2 : rise 2

‡ means value not DUFKÀ[LW\ span : rise 4 3 2 3 3 3
applicable for comparison interior spans 
? means value unknown construction method deck section arch section column section 
3
span rise type width depth area inertia cont. type depth var. area inertia area inertia 3
 2
2 2 2
ID bridge year country m m # m m m2 10-3 m4 m m m2 10-3 m4 m2 10-3 m4 m m % % m m m2 % 2
4 2 3
2 
1 Wanxian-Yangtze 1997 China 425.0 85.0 14 T 21.0 ? ? ? -- 3 7.00 27.9 220000 5.4 4400 5.0 2130 2.80 61 ? 58 61 100 2 ? 5.01 ? ?
2 Krk I 1980 Croatia 390.0 60.0 14 T 10.8 2.00 4.8 1800 -- 3 6.50 21.5 130000 1.7 300 6.5 2540 2.46 60 46 59 49 100 <1 2.61 5.25 828 6 2 
2 
3 Colorado 2010 United States 323.0 84.4 9 26.9 2.50 ‡ 8200 — 4.27 14.8 40000 6.9 2900 3.8 1240 1.80 76 48 73 94 83 6 ‡ 2.78 ‡ ‡ 
4 Gladesville 1964 Austrailia 309.0 39.0 s 10 T 26.4 1.80 9.8 2900 — 5.60 ± 1.34 25.1 100000 3.0 63 7.9 2450 2.02 55 42 56 39 97 <1 1.37 3.60 421 8    T

T  v
5 Rio Parana 1965 Brazil 290.0 53.0 s 13 T 13.5 2.20 10.5 2100 — 3 4.00 ± 0.80 17.7 44000 0.6 92 5.5 1590 1.61 73 47 68 66 95 <1 2.22 3.45 440 12
 rectangular solid section  T  ?
6 Infant Henrique 2003 Portugal 280.0 25.0 8 20.2 4.50 16.8 57000 — v 1.50 18.1 4400 17.4 2500 11.2 3140 1.84 187 47 54 27 7 4 1.83 3.32 716 1 v voided slab section 1   
7 Bloukrans 1983 South Africa 272.0 62.0 15 v 15.7 1.50 10.8 2300 — 3 4.60 ± 1.00 17.3 53000 4.7 2500 4.4 1190 1.79 59 45 60 69 96 5 2.25 3.48 335 12 T tee or slab-on-girder section 
8 Sandö 1943 Sweden 264.0 40.0 t 22 T 12.4 1.23 5.7 540 — 3 3.95 ± 0.95 13.0 31000 0.9 31 6.6 1740 1.56 67 51 63 51 98 <1 1.62 3.12 352 15 n single or n-cell box-girder section
 
 
9 Tensho 2000 Japan 260.0 32.5 14 v 8.6 0.80 4.4 280 — 5.00 ± 1.00 17.0 54000 8.1 1500 8.0 2080 1.79 52 45 53 36 99 3 2.95 4.23 768 6
0
 
10 Sibenik 1966 Croatia 246.0 31.0 11 T 11.1 1.50 2.6 680 -- 3 3.30 ± 0.40 7.7 14000 0.6 3 7.9 1950 1.35 75 51 66 46 100 <1 0.98 2.48 239 27
11 Krk II 1980 Croatia 244.0 47.0 8 T 10.8 2.00 5.4 1700 -- 3 4.00 9.0 21000 1.7 300 5.2 1270 1.53 61 41 61 62 100 1 1.46 2.86 232 17 deck depth in m 6
12 Wilde Gera 2000 Germany 242.0 68.0 6 26.3 3.74 ‡ ? — 2 4.40 ± 1.10 11.3 34000 7.6 11000 3.6 860 ? 55 30 ? ? ? ? ‡ ? ‡ ‡
13 Galena Creek 2011 United States 210.0 42.5 s 3 2 18.9 3.00 12.1 14000 — 3.70 6.1 13000 ‡ ‡ 4.9 1040 2.10 57 31 38 40 48 ‡ 1.00 2.58 129 ‡
14 Krka-Skradin 2004 Croatia 204.0 52.0 7 22.3 1.96 ‡ 3700 — 2 3.00 10.9 15000 4.1 1700 3.9 800 1.31 68 41 63 79 80 9 ‡ 2.16 ‡ ‡
5 
15 Maslenica 1997 Croatia 200.0 65.0 7 T 21.2 2.00 13.3 4800 — 2 4.00 12.6 29000 5.2 3900 3.1 620 1.64 50 33 53 79 86 12 1.53 2.67 118 17
16 Pag 1968 Croatia 193.0 28.0 9 T 9.0 1.42 1.9 350 -- 3 2.65 ± 0.35 4.0 4500 1.4 77 6.9 1330 1.06 73 47 67 53 100 2 0.72 1.82 121 45

17 Modong Hongshui 1999 China 180.0 30.0 14 11.9 0.65 8.1 290 -- 3.50 5.7 9600 1.3 35 6.0 1080 1.30 51 43 52 46 100 <1 1.24 2.13 168 23
18 Nosslach 1968 Austria 180.0 45.0 s 12 ? ? ? ? ? — 3 4.00 ± 1.30 12.5 26000 ? ? 4.0 720 ? 45 ? ? ? ? ? ? ? ? ? 4
19 Hokowazu 1974 Japan 170.0 26.5 11 v 10.1 0.92 4.4 320 — 2 3.00 10.4 13000 3.0 560 6.4 1090 1.13 57 43 80 47 98 4 1.67 2.51 228 ‡ 
20 Houffalize 1979 Belgium 162.0 32.4 s 6 T 16.0 2.00 6.1 3100 -- 2 3.20 13.0 19000 4.5 1900 5.0 810 1.21 51 31 52 54 100 10 1.36 2.42 138 23
21 Pitan 1993 Taiwan 160.0 21.7 s 3 15.9 5.00 16.1 59000 — 6.25 17.5 97000 ‡ ‡ 7.4 1180 2.99 26 14 20 15 62 ‡ 2.17 4.90 319 ‡ 3 2
22 Wertachtal ? Germany 156.0 34.0 11 T 12.7 1.42 6.7 1100 — 2 2.80 ± 0.80 8.5 9900 2.1 120 4.6 720 1.14 56 37 54 60 90 1 1.40 2.18 126 27
23 Neckarburg 1978 Germany 154.0 49.5 7 15.5 2.30 9.2 6300 — 2 3.00 7.2 9200 3.8 4800 3.1 480 1.47 51 29 45 67 59 31 1.41 2.29 84 21 
24 Caracas 1953 Venezuela 151.8 32.0 12 T 20.8 2.00 5.4 1800 -- 3.10 9.4 13000 3.0 110 4.7 720 1.18 49 30 72 54 100 1 0.82 1.96 74 37  T
2 T T T T
25 Argentobel 1986 Germany 150.0 29.0 10 T 14.0 1.60 7.7 1500 — 2 2.75 ± 0.75 7.3 8520 2.4 200 5.2 780 1.17 55 34 49 50 85 2 1.22 2.05 118 28 T  T
T 4 4
T
26 Cetina 2007 Croatia 140.3 20.6 7 T 10.2 1.40 5.4 870 — 2.50 8.0 7500 3.2 1200 6.8 960 1.02 56 36 51 40 90 14 1.44 2.14 172 19 T 2
T T v
27 Hundwiler 1991 Switzerland 138.4 35.8 13 T 9.6 0.82 4.3 200 — 2.15 ± 0.65 5.6 2200 1.2 36 3.9 540 0.65 64 47 86 109 92 2 1.23 1.44 82 46  rectangular solid section T ? T T T
v voided slab section T T
28 Maling 1999 Taiwan 138.0 33.0 7 v 16.0 1.10 7.6 1200 — 4 2.60 ± 0.60 20.2 16000 5.9 500 4.2 580 0.92 53 37 59 72 93 3 2.04 2.35 147 20 T TT  v
T tee or slab-on-girder section 1  v
29 Big Qualicum 1996 Canada 132.0 24.0 s 6 T 24.0 1.73 7.2 2200 — 1.50 9.0 1700 7.2 860 5.5 730 0.66 88 41 76 73 44 22 0.81 1.25 73 26 T T T T
n single or n-cell box-girder section T T TT T  T

v
T T
30 Crooked River 2000 United States 125.0 24.0 8 4 24.2 1.90 12.5 6400 — T 1.63 ± 0.38 11.0 2500 5.9 910 5.2 650 0.90 77 35 53 53 28 10 1.09 1.64 89 14  composite steel-concrete section 
T v

31 Niesenback ? Germany 120.0 37.5 6 T 17.6 2.10 7.6 2300 — 2 2.50 8.9 8100 2.0 170 3.2 380 1.08 48 26 47 69 78 2 1.10 1.92 53 40 
0
32 Siggenthal 2000 Switzerland 117.0 24.4 s 7 ? ? 1.40 ? ? — ? 1.20 ± 0.20 ? ? ? ? 4.8 560 ? 98 45 ? ? ? ? ? ? ? ?
33 Nanin 1967 Switzerland 112.0 24.5 t 8 9.9 1.00 4.1 400 — 1.20 ± 0.20 5.1 600 2.5 100 4.6 510 0.44 93 51 99 111 60 10 1.05 1.07 67 54
34 Bixby Creek 1932 United States 100.6 36.6 11 T 8.4 0.92 3.4 200 -- 2.00 ± 0.50 5.6 1900 1.5 120 2.8 280 0.58 50 34 78 126 100 6 1.46 1.39 50 67 year of completion 2020
35 Rhein 1963 Switzerland 100.0 20.9 ? 8 8.5 1.00 2.8 350 — 1.50 ± 0.25 6.1 1100 2.6 110 4.8 480 0.49 67 40 79 86 76 8 1.20 1.27 72 53
36 Wirrbachtal 2002 Germany 100.0 25.3 s 7 T 11.5 1.40 7.6 1300 — 1.10 ± 0.20 8.0 810 1.0 40 4.0 400 0.51 91 40 78 99 38 2 1.50 1.30 74 34 2010 s 
37 Salginatobel 1930 Switzerland 90.0 13.0 t ‡ T 3.5 0.90 1.7 75 — ‡ ‡ ‡ 0.8 22 6.9 620 ‡ ‡ ‡ ‡ ‡ ‡ ‡ ‡ ‡ ‡ ‡ s 
  
38 Wunshuei ? Taiwan 85.0 13.5 ? 14 v 11.0 0.45 4.3 81 — 2 2.00 7.8 4100 2.0 69 6.3 540 0.73 43 35 43 37 98 2 1.25 1.66 84 40 2000 s
s s     
39 Gmundertobel 1909 Switzerland 80.0 26.0 t 18 T 6.9 0.70 1.9 60 — 1.65 ± 0.45 11.9 2700 1.2 37 3.1 250 0.48 48 34 72 108 98 1 2.67 1.69 82 49 s  
s
40 Tiefetal 1952 Germany 77.7 12.6 t 10 T 10.0 1.03 3.6 310 — 1.50 ± 0.30 4.8 1400 1.1 25 6.2 480 0.60 52 31 75 42 82 1 0.91 1.27 55 ‡ s
1990 
41 Guiers 2000 France 73.8 12.5 s 9 9.8 0.75 10.2 480 — T 1.25 ± 0.20 4.0 570 1.7 170 5.9 440 0.51 59 37 54 49 54 16 1.56 1.09 85 25 

42 Russian Gulch 1939 United States 73.2 25.9 ? 15 9.7 0.55 3.5 45 -- 1.65 ± 0.35 4.3 970 1.2 64 2.8 210 0.47 44 33 98 109 100 7 1.15 1.06 30 ‡ 1980  
 s
43 Mazzocco 2007 Italy 70.0 14.0 s 3 4 17.2 1.85 7.6 2700 -- T 1.50 6.2 1300 ‡ ‡ 5.0 350 0.46 47 21 59 61 100 ‡ 0.84 0.97 37 ‡ 
44 Stampfgraben 2003 Austria 70.0 21.0 6 T 9.3 0.91 5.4 350 — 1.33 ± 0.28 6.0 880 1.3 26 3.3 230 0.45 53 31 65 93 72 2 1.42 1.17 41 59 1970
t s
45 Rocky Creek 1932 United States 68.6 17.1 t 11 T 8.2 1.07 4.2 210 -- 1.77 ± 0.55 4.3 1100 1.0 64 4.0 280 0.51 39 24 54 68 100 6 1.23 1.17 42 69  s s
46 Fulton Road 1932 United States 64.0 12.7 t 5 T 23.5 1.17 7.7 820 — 1.50 5.7 1300 5.0 650 5.0 320 0.61 43 24 40 42 61 31 0.65 1.03 26 50 ?
1960
47 Matsesta 1938 Russia 63.6 23.6 ? 12 T 18.5 0.83 3.7 220 — 1.11 ± 0.21 6.6 680 3.2 120 2.7 170 0.37 58 33 78 128 76 13 0.81 0.84 17 126
48 Broadway 1932 Canada 54.7 13.9 t 11 T 13.7 1.86 6.3 1600 — 1.20 ± 0.30 11.7 1400 2.0 120 3.9 220 0.51 46 18 44 55 47 4 1.57 1.38 42 36 t 
1950
49 Big Creek 1937 United States 54.1 23.0 t 11 T 8.2 0.76 3.3 110 — 1.20 ± 0.30 2.2 260 1.1 78 2.4 130 0.41 45 28 90 112 70 21 0.96 0.82 15 ‡
50 Lianlao ? Taiwan 50.0 12.5 ? 8 T 20.5 0.60 6.2 120 — 1.00 18.0 1500 1.8 54 4.0 200 0.30 50 31 67 83 93 3 1.35 0.98 34 70 t timber scaffolding / centering t
s steel scaffolding / centering 1940 ?
51 Wölkau ? Germany 47.5 5.1 s 3 T 4.5 0.60 2.1 59 — 0.53 ± 0.13 1.3 30 ‡ ‡ 9.3 440 0.26 90 42 93 39 34 ‡ 0.76 0.62 42 ‡  cantilever using cable-stayed t?
?
52 Elche de la Sierra 1927 Spain 40.0 4.0 20 T 8.0 0.49 2.8 42 — 1.01 ± 0.07 2.0 172 0.5 3 10.0 400 0.33 40 27 44 24 80 1 0.64 0.68 32 85  cantilever using effective truss t t t 
 pre-erected reinforcement arch 1930 tt t
53 Dodan Nallah 1995 India 40.0 14.5 s 7 8.3 0.51 2.5 28 — 0.53 ± 0.18 2.3 53 1.7 23 2.8 110 0.19 76 39 96 155 65 28 0.79 0.49 11 310  rotation method 
54 Bronte Creek 1934 Canada 39.8 17.1 ? 11 16.4 0.23 3.8 17 — 0.81 ± 0.05 4.3 241 2.2 32 2.3 90 0.24 49 38 78 139 93 12 0.78 0.57 9 252 t
1920
55 Schwandbach 1933 Switzerland 37.4 6.0 t 9 T 4.2 0.85 1.0 31 — 0.20 1.3 4 0.8 2 6.2 230 0.16 187 36 85 73 12 6 0.64 0.47 19 104
56 Nan Ke ? Taiwan 25.0 3.3 ? 3 T 16.0 0.80 9.6 500 — 0.45 4.1 70 ‡ ‡ 7.7 190 0.37 56 20 24 17 12 ‡ 0.87 0.75 21 ‡
57 Ziggenbach 1924 Switzerland 20.0 4.7 t 8 T 5.2 0.88 1.5 89 — 0.41 ± 0.15 2.0 29 0.2 1 4.3 90 0.24 49 16 33 39 25 <1 0.75 0.65 8 103
58 Bohlbach 1932 Switzerland 14.4 2.7 t 9 T 4.5 1.16 1.3 160 — 0.16 1.0 2 0.9 2 5.3 80 0.40 90 11 14 13 1 1 0.58 0.76 6 17

A Study of 58 Concrete Arch Bridges


t timber scaffolding / centering 3-hinged arch k=0.54 rectangular section - - discontinuous —average of max and min depth —Geometry and Ratios
s steel scaffolding / centering 2-hinged arch k=0.50 v voided slab section over columns — max depth = depth + var. of spandrel columns used to prepared by Jason Salonga
cantilever using cable-stayed À[HGDUFKN  T tee or slab-on-girder section — continuous — min depth = depth var. calculate area and moment of inertia
cantilever using effective truss k = eff. arc length factor n-cell box girder section over columns — depth measured to arch axis
assisted by Catherine Chen
n
pre-erected reinforcement arch composite steel box section — depth used to calculate area and moment of inertia Department of Civil Engineering
rotation method modular ratio of 8 assumed for transformed inertia University of Toronto
297
span length in m 0 100 200 300 400 500 span length in m 0 100 200 300 400 500
bridge ID 51 41 31 21 11 1
bridge ID 51 41 31 21 11 1
The arch depths of the bridges in the
database tend to increase linearly with
span length. The two bridges that depart The Infant-Henrique bridge has a span to
max arch depth 8 from this trend are the Pitan and Infant- span-to-rise ratio 16 rise ratio of 11.2, which is much higher
average arch depth Henrique Bridges. The former must resist than the rest of the bridges in the
large moments because of the lack of database. This shows that it is feasible to
min arch depth Pitan spandrel columns. The latter achieves build concrete arches with span-to-rise
6 12
continuous deck depth arch slenderness by minimizing arch Infant-Henrique 11.2 ratios greater than 8.
discontinuous deck depth moments through the stiffening effect of
the deck.
4 8 8.0
Decks of self-stiffened and discontinuous 55 out of 58 bridges in the database
deck arch bridges do not tend to be (95%) have ratios in the range 2.3 to 8.0.
2 deeper than 2.5 m because they do not 4
participate in resisting flexible system
moments. 2.3 The Bronte Creek bridge has the lowest
Infant-Henrique Bronte Creek span-to-rise ratio of 2.3.
0 0
The Schwandbach and Infant-Henrique
span-to-arch-depth ratio 200 bridges achieve visual arch slenderness span2 / rise in m 4000 According to Mondorf (2006), bridge
with continuous deck Schwandbach 187 by having stiff decks that attract most of (static coefficient) 16 =1
2 =8 builders used this geometric ratio, also
the bending moments in the system. = known as the static coefficient, to gauge
with discontinuous deck rise
150 3000 to Infant-Henrique how demanding the construction of the
an arch would be. This ratio also gives an
sp estimate of the horizontal reaction at the
max ratio springing lines caused by uniformly
100 2000 =4 distributed loads:
average ratio
min ratio 77 UDL × (span)2
48 out of 57 bridges in the database horizontal reaction =
8 × rise
50 (84%) have span-to-arch-depth-ratios in 1000
39 the range 39 to 77. All discontinuous deck For the same UDL, The Colorado and Krk-
arch bridges in this study are in this range. II bridges would have about the same
Krk-II Colorado
horizontal reaction.
0 0

span-to-system-depth ratio 80 equivalent slab thickness by volume 3 Tensho


The Schwandbach and Infant-Henrique = volume / (span × deck width) in m Concrete volume calculated based on
with continuous deck Gmundertobel average sectional areas.
Schwandbach Infant-Henrique bridges are no longer outliers when the Krk I
with discontinuous deck deck depth is considered in the ratio. volume = deck area × span
60 + arch area × arc length minimum slab thickness envelope:
system = arch + deck 2
+ column area × total spandrel
48 " " column length L
max ratio 46 out of 55 bridges in the database y= + 0.569 m 0 m ≤ L ≤ 193 m
40 (84%) have span-to-system-depth ratios in 1276
average ratio Gladesville
the range 24 to 48. The continuity of the L2 L
min ratio deck does not appear to affect this ratio. y= − + 2.21 m L ≥ 193 m
1 22800 m 61.7
24 Sibenik
20
Fulton Road Pag y is envelope of equivalent slab thickness in m
Bohlbach L is span length in m
0 0

The moments of inertia of the arch and horizontal reaction in m2


arch inertia / deck width in m3 102 (unit weight × deck width)
10000
deck are indicators of flexural rigidity.
Trends seen using a log scale for the volume × span twice emp. fit empirical fit for in the form: y = aL2 + b
deck inertia / deck width in m3 101 normalized arch and deck moments of =
8 × deck width × rise 2
with continuous deck inertia are similar to those seen for arch 1000 horiz. reaction in m2 ≈ L + 10 m2
half emp. fit
with discontinuous deck 100 and deck depths on a linear scale. The uw × deck width 182.7
general trend is that arch moments of
inertia increase faster than deck moments Horizontal reactions are estimated based
10−1 100 only on the weight of concrete that is
of inertia with respect to span length,
except for those bridges with stiffening continuous in the longitudinal direction.
10−2 decks. Hence, these horizontal reactions should
10 be interpreted as lower-bound values.
Considering the sum of arch and deck
10−3 moments of inertia, or system moment of Traffic loads, superimposed dead load,
inertia, reduces the vertical scatter of the and weight of diaphragms, ribs, and
10−4 overall trend. 1 stiffeners will increase the horizontal
reaction of the arch.
normalized system moment of 6 column inertia to system inertia ratio 1
inertia in m linear regression for :
The column moment of inertia to system
Pitan y = 0.0109 L + 0.337 m moment of inertia ratio gives an estimate of
= 3 12 × system inertia r2 = 0.906 the distribution of flexural rigidity in the
deck width Tensho
4 0.1 system. 39 out of 49 bridges in the database
y is normalized system inertia in m
excluded from regression analysis (80%) have ratios of 0.1 or less.
L is span length in m
Colorado r2 is coefficient of determination During conceptual design, spandrel
columns are often assumed to be pin-
2 Normalized system moment of inertia gives a 0.01 connected, even if they will eventually be
measure of how system stiffness changes with built with fixed connections. The typically
span length. Bridges with values lower than the low flexural rigidity of the spandrel
trend line are more likely to be sensitive to columns relative to that of the system
deflection and stability problems. makes this assumption acceptable.
0 0.001

system radius of gyration in m 3 system slenderness ratio 120


linear regression for : Russian
system inertia effective arc length factor × arc length Gulch The Russian Gulch and Nanin bridges
= y = 0.0062 L + 0.0768 m = 100 Nanin have the highest system slenderness.
arch area system radius of gyration 99
r2 = 0.913
excluded from regression analysis 2 80 80 43 out of 54 bridges in the database
y is system radius of gyration in m (81%) have system slenderness ratios in
L is span length in m the range 38 to 80.
60
r2 is coefficient of determination Deck-stiffening is accounted for in the ratio
by using the system radius of gyration
1 Based on this regression analysis, system 40 38 rather than the arch radius of gyration.
radius of gyration and span length are well- Depending on the construction method,
correlated, despite the differences among 20 arches may have higher slenderness
bridges in the database. For shallow arches, ratios during construction while the deck
it is desirable to have lower system radius of girder is not providing stiffness.
0 gyration to reduce beam action. 0

arch inertia to system inertia ratio 1 modified slenderness ratio 160


with continuous deck 32 self- 2 × rise
with discontinuous deck stiffened 32 out of 54 bridges in the database =
0.8 (59%) have arch inertia to system system radius of gyration
120
inertia ratios greater than or equal
fixed arches
0.6 to 0.80. This implies that most
hinged arches A Study of 58 Concrete Arch Bridges
18 partially bridges in this study are self-
deck- stiffened. In these bridges, flexible 80 —Geometrical Trends
0.4 stiffened system bending moments are prepared by Jason Salonga
carried primarily by the arch. The for fixed arches:
Infant Henrique bridge is the only 40 assisted by Catherine Chen
0.2 Infant-Henrique long span deck-stiffened arch in the increasing arch action
4 deck- database. Department of Civil Engineering
90% arch action Nan Ke Pitan
stiffened University of Toronto
0 increasing beam action 0 Bohlbach
298
Appendix B: Details of Computer Program QULT

This appendix gives additional details on the functions used in the computer program qult. This program,

written by the author, models the behaviour of ultra high-performance fibre-reinforced concrete slender

members under combined compression and bending (see Section 3.3 for an overview of qult). qult is or-

ganized as a hierarchy of functions that implement the general analysis method described in Section 3.2.

qult is written in matlab r2007a and is included with this thesis digitally as a set of matlab m-files.

Table B-1 lists all the major functions used in qult and describes their input parameters, calculation steps,

and output data. Most functions have an option to plot results in matlab. These plots are activated by sett-

ing the variable option to 1 near the beginning of most m-files. These plots allow the user to browse and

check results at each calculation stage. Samples of these output plots are shown in the bottom half of Fig-

ure 3-7, which has been reprinted on the next page.

299
B. Details of Computer Program QULT

sectional
member
geometry & sectional member
response
material analysis deflection
& capacity
response

input parameters stress-strain curve stress-strain curve stress-strain curve stress-strain curve
sectional geometry sectional geometry sectional geometry sectional geometry
member length member length
applied load applied load
eccentricity of load
FUNCTIONS GEOM
superscripts refer CENTROID1
to the sample output SECTPROP
diagrams below
MATMODEL2 SECTFORCE3
GETSTRESS NMBOUND
MFACTOR CONTOURCURV4 MOMCURV5 RESPONSE7
ULTIMATE
DEFLECTION6 CAPACITY8
depth
ENVELOPES

plane stress
width of strain  

element strip
moment

axial moment
force
1. CENTROID
curvature
axial load
3. SECTFORCE ultimate load

stress
5. MOMCURV
 
contour of
axial equal curvature
force   7. RESPONSE
strain + eccentricity
column
 

2. MATMODEL sectional failure


moment
mid-length moment limit
axial
load end moment limit
force
4. CONTOURCURV
half length of member

6. DEFLECTION
moment

8. CAPACITY

Reprint of Figure 3-7. Overview of program QULT

300
B. Details of Computer Program QULT

Table B-1. Description of program functions and calculation steps

function(parameters) description of calculations

matmodel 1. Input concrete strength in compression in MPa (defined within function).


2. Calculate elastic modulus and secant modulus at peak stress
3. Output n×2 matrix. Each row corresponds to one strain-stress ordered
pair. Column 1 is strain, Column 2 is stress in MPa.
4. Override strain-stress matrix with alternate material model (optional).
5. Plot stress-strain material model (optional).

mfactor(index) 1. Input index for material factor.


2. Output factor (index, factor): unfactored (0, 1), concrete (1, 0.75), steel
bars (2, 0.90), prestressing strand (3, 0.95).

getstress(strain) 1. Input strain.


2. Call matmodel to get stress-strain model.
3. Find next and previous strain values from model.
4. Call interpolate to linearly interpolate the stress value.
5. Output stress in MPa.

geom(up) 1. Input up=1 for upside geometry or up=-1 for upside down geometry.
2. Input sectional geometry (defined within function) by entering depth in
mm, width in mm, and number of strips per element. Up to four stacked
rectangular elements can be defined.
3. Calculate lever arms of strips with respect to bottom of section.
Calculate heights and widths of strips.
4. Reverse geometry if up=-1.
5. Output n×3 matrix. Row i is for strip i of the section. Column 1 is for lever
arm from the bottom of the section in mm, Column 2 is for strip height
in mm, Column 3 is for strip width in mm.

centroid(up) 1. Input up=1 for upside geometry or up=-1 for upside down geometry.
2. Call geom(up) to get geometry data.
3. Calculate strip area and strip moment area.
4. Calculate centroid of section.
5. Calculate strip lever arm with respect to centroid
6. Output n×2 matrix. Row i is for strip i of the section. Column 1 is for lever
arm from the centroid of the section in mm, Column 2 is for strip area in
square mm.

7. Plot sectional geometry as strip elements.

301
B. Details of Computer Program QULT

function(parameters) description of calculations

sectprop(up) 1. Input geometry by calling geom(up).


2. Calculate strip area, strip moment area.
3. Calculate centroid of the section.
4. Calculate strip lever arm with respect to centroid.
5. Calculate strip moment area about centroid.
6. Output 1×5 matrix with total area, total moment of inertia, radius of
gyration, distance from centroid to top fibre, and distance from centroid to
bottom fibre.

sectforce(curv, etop, up) 1. Input curvature curv, top strain etop and up=1 for upside geometry
or up=-1 for upside down geometry.
2. Call geom(up) to get geometry.
3. Call centroid(up) to get lever arm and area data.
4. Call mfactor to get material factor.
5. Calculate strain, stress, force, and moment about centroid of each
strip in the section.
6. Calculate axial force by summing forces, and moment by summing
moments about centroid.
7. Output 1×2 matrix: axial force in MN and moment in MN·m.
8. Plot plane of strain, stress profile, and force profile (optional).

nmbound(up) 1. Input up=1 for upside geometry or up=-1 for upside down geometry.
2. Find limiting compressive strain from matmodel.
3. Find distance from centroid to top fibre.
4. Find curvatures corresponding to planes of strain defined by top strain
at limiting compressive strain and the neutral axis crossing at the centre
of each strip.
5. Call sectforce to find axial forces and moments of each plane of strain.
6. Output n×3 matrix: axial force in MN, moment in MN·m, and curvature.

contourcurv(up) 1. Input up=1 for upside geometry or up=-1 for upside down geometry.
2. Find maximum curvature and curvature due to balanced point, using
nmbound, geom, centroid, and matmodel.
3. Define increments of contours of equal curvature to calculate.
4. Calculate sets of axial force and moment for each contour using
sectforce.
5. Output n×3×k matrix to global variable. Each row corresponds to one data
point of axial force in MN, moment in MN·m, and curvature. Each matrix
corresponds to one contour of equal curvature.
6. Plot contours of equal curvature in axial force-moment space (optional).

302
B. Details of Computer Program QULT

function(parameters) description of calculations

momcurv(N) 1. Input axial force N in MN.


2. Call contourcurv to get global variables contour_up and
contour_down.
3. Interpolate values of moment from each contour of equal curvature.
4. Clip descending part of resulting moment-curvature diagrams.
5. Combine positive and negative moment-curvature diagrams.
6. Output n×2 matrix. Column 1 is moment in MN·m, Column 2 is
curvature in rad/mm.
7. Plot moment-curvature diagram (optional).

deflection(Q, L) 1. Input load Q in MN and column length L in mm.


2. Call momcurv to get moment-curvature diagram.
3. Find peak moment and set mid-length moment intervals.
4. Construct column deflection curves.
5. Record end eccentricities.
6. Determine whether sectional failure or member stability failure occurred.
7. Output axial load in MN, maximum mid-length moment in MNm, end
eccentricity in mm, and failure mode.
8. Plot column deflection curves (optional).

capacity(Qmax, L) 1. Input maximum compressive load Qmax in MN and column length L


in mm.
2. Set intervals of load Q to be calculated.
3. Calculate ultimate moments using deflection.
4. Output ultimate load in MN, ultimate moment at mid-length MNm,
and end moment in MNm.
5. Plot member capacity envelope N=Qult, M=Mult.

envelopes(Qmax, Lmax, n) 1. Input maximum compressive load Qmax in MN, maximum length of
member Lmax in mm, and number of shorter columns to calculate n.
2. Set column lengths to be calculated.
3. Call capacity to get for ultimate load and ultimate moment for each
column length.
4. Output m×3×n matrix. Row m corresponds to one data point of ultimate
load in MN, ultimate mid-length moment in MN·m, and end moment in
MNm. Matrix n holds data for one column length.
5. Plot member capacity envelopes for each column length considered
(optional).

303
B. Details of Computer Program QULT

function(parameters) description of calculations

response(Qmax, L, eccen) 1. Input maximum compressive load Qmax in MN, maximum length
of member L in mm, and eccentricity of load in mm.
2. Define intervals of load for calculating column deflection curves.
3. Call momcurv to get moment-curvature diagram for current Q.
4. Find maximum moment and set mid-length deflection conditions for
each column deflection curve.
5. Calculate column deflection curves for current load Q.
6. Record list of mid-length moment and end eccentricity for current load Q.
7. Interpolate mid-length moment for given eccentricity of load eccen.
Repeat for all load Q.
8. Reorder list of load Q and mid-length moment M such that moments are
from lowest to highest.
9. Calculate deflection from load Q, moment M, and eccentricity eccen.
10. Output n×3 matrix. Each row corresponds to a load stage in a deflection
controlled test. Column 1 is load Q in MN, Column 2 is deflection w in
mm, and Column 3 is mid-length moment M in MNm.
11. Plot load-deflection response of column (optional).

ultimate(Qmax, L, eccen) 1. Input maximum compressive load Qmax in MN, maximum length
of member L in mm, and eccentricity of load in mm.
2. Call response to get load-deflection-moment response of column.
3. Find first and last instance of ultimate load and their indices.
4. Average the moments from the first and last instance.
5. Output ultimate load in MN, ultimate deflection in mm, and ultimate
moment in MNm.

304
Appendix C: Slender Column Drawings

This appendix contains drawings and photographs of the slender column load tests described in Section

3.4. Specimens made from high and ultra high-performance fibre-reinforced concrete were prepared and

tested according to these set of drawings and tested at the University of Toronto between September 2007

and December 2007. Specimens E and F, which were made from high-performance fibre-reinforced con-

crete are not specifically dealt with in the body of thesis, but are included here for reference.

Figure C-1 shows the general arrangement of the six specimens and their connection to the mts testing

machine, as well as the list of materials and fabrication. Figure C-2 shows the detail of the steel end plates

that were used to form the ends of each column. Studs were welded at the centre of each plate to provide

shear resistance between the concrete columns and their end plates. Figure C-3 shows the double tongue

assembly that receives the pin at each end of the column. It connects to the steel end plates of the columns

with four nuts and bolts. The larger Specimens E and F use threaded holes instead of nuts because of the

space requirements of the column. The eccentricity of load is created by offsetting the centreline of the pins

from the centreline of the column. Figure C-4 shows the plate gripped by the mts machine head that is

pinned to the top double tongue assembly. The drawing also shows the welded plate that is pinned to the

bottom double tongue assembly. This welded plate is bolted to the circular base plate, which is fixed to the

reaction frame. Figure C-5 shows the general arrangement of the plywood forms used to cast the six speci-

mens. Forms were designed to support the concrete specimens during erection and connection into the

305
C. Slender Column Drawings

test frame. Figure C-6 shows typical sections of the plywood forms and how they were nailed together. The

drawing also shows the positioning of the steel end plates relative to the ends of the forms. End plates were

clamped and secured with additional pieces of plywood to hold them in the correct position. Holes in the

end plates were temporarily plugged to prevent the concrete from spilling out during casting. Figure C-7

shows the types of instrumentation used during the load test. Surface strain gauges with lengths of 60 mm

were glued to the flexural compression face of each column. Tooling balls with 6 mm diameters were

glued along the centerline of the adjacent column face and used as targets for the cohesive laser radar

scanner. The horizontal displacement at mid-length of each column was measured by a linear variable di-

fferential transformer. Figure C-8 shows all the as-built measurements of the specimens including average

compressive strengths (from cylinder tests), eccentricities of load, and lengths, depths, and widths of

columns. These measurements are used as input parameters for the computer program qult (see Ap-

pendix B), which calculates the load-deflection response of the column according to the general method

described in Section 3.2

306
C. Slender Column Drawings

Figure C-1. Slender column tests. Scale reduced to 80% of original.

307
C. Slender Column Drawings

Figure C-2. End plate detail. Scale reduced to 80% of original.

308
C. Slender Column Drawings

Figure C-3. Pin connection detail. Scale reduced to 80% of original.

309
C. Slender Column Drawings

Figure C-4. Top and bottom tongues. Scale reduced to 80% of original.

310
C. Slender Column Drawings

Figure C-5. Formwork. Scale reduced to 80% of original.

311
C. Slender Column Drawings

Figure C-6. Formwork sections. Scale reduced to 80% of original.

312
C. Slender Column Drawings

Figure C-7. Instrumentation. Scale reduced to 80% of original.

313
C. Slender Column Drawings

Figure C-8. As-built measurements. Scale reduced to 80% of original.

314
Appendix D: Design Calculations for Concept 80

This appendix contains a set of manual design calculations pertaining to Concept 80. Concept 80 is one of

the seventy-two ultra high-performance fibre-reinforced concrete arch bridge concepts proposed proposed

in Chapter 6. Concept 80 is described in greater detail in Section 6.2.3. The purpose of these calculations

are: (1) to demonstrate the use of the simplified arch analysis method presented in Chapter 4, which is used

for calculating maximum sectional forces in arch systems, (2) to demonstrate the use of the simplified

design method presented in Section 3.5, which is used for calculating capacities of ultra high-performance

fibre-reinforced concrete members in compression and bending, (3) to demonstrate how each of the sev-

enty-two design concepts presented in Section 6.2.1 were designed, and (4) to serve as a means of checking

the accuracy of spreadsheet calculations.

There are a total of eighteen pages of calculations, each labelled from D1 to D18. Page D1 shows the trial

cross-sections of the arch and deck girder, along with the given material properties and global geometry.

Page D2 lists system properties that are used frequently in design calculations. Page D3 tabulates dead load

quantities from all primary and secondary structural components of the bridges. Page D4 shows calcula-

tions for all load quantities, including dead load, live load, shrinkage strain, and uniform drop in temper-

ature. Pages D5-D12 shows sample design calculations for ultimate limit state load combination 2, for

which sectional forces at midspan are critical. System properties are calculated on pages D5-D6. Long-

term sectional forces, short-term sectional forces, and second-order deflections and moments are calcu-

315
D. Design Calculations for Concept 80

lated on pages D7-D10. Page D11 shows calculations of fixed system moments. Pages D13-D15 shows

sample design calculations for ultimate limit state load combination 1, for which sectional forces at the

quarter-point are critical. Reduced, member capacity n-m interaction diagrams for the arch and deck

cross-sections are shown on page D12. Critical sectional force demands for ultimate limit state load com-

binations 1 and 2 are plotted on to the diagrams. Because these points lie within the envelopes, the pro-

posed cross-sections are satisfactory.

Rudimentary shear force design calculations are shown on pages D16-D18. For simplicity, an equivalent

distributed load applied over half the span (calibrated to cause the same second-order moment shown on

page D14) is assumed. From this, bending moment, shear force, and axial force diagrams are calculated.

On page D17, Mohr’s circle of stress calculations are done for the arch cross-section at x=0.1L. The result-

ing principal stress is less than the factored cracking stress of the material. Shear stress calculations for di-

fferent values of x are done on a spreadsheet. Resulting critical principal stresses along the arch member

are plotted on page D18 (solid line). Principal stresses are most critical within the first 10% of the span, but

are still below the factored cracking limit. A secondary (dashed) curve is shown on the diagram, corres-

ponding to results for a section that has twice the web thickness and twice the top and bottom slab thick-

nesses. This is done because increased cross-section proportions are expected near the springing lines. The

secondary curve shows a significant reduction in critical principal stresses, as compared to the primary

curve.

316
D. Design Calculations for Concept 80

317
D. Design Calculations for Concept 80

318
D. Design Calculations for Concept 80

319
D. Design Calculations for Concept 80

320
D. Design Calculations for Concept 80

321
D. Design Calculations for Concept 80

322
D. Design Calculations for Concept 80

323
D. Design Calculations for Concept 80

324
D. Design Calculations for Concept 80

325
D. Design Calculations for Concept 80

326
D. Design Calculations for Concept 80

327
D. Design Calculations for Concept 80

328
D. Design Calculations for Concept 80

329
D. Design Calculations for Concept 80

330
D. Design Calculations for Concept 80

331
D. Design Calculations for Concept 80

332
D. Design Calculations for Concept 80

333
D. Design Calculations for Concept 80

334
Curriculum Vitae
Jason Angeles Salonga jason.salonga@utoronto.ca

Education
B.E.Sc. in Civil Engineering 2004 University of Western Ontario, London, Ontario, Canada

B.A. in Visual Arts 2004 University of Western Ontario, London, Ontario, Canada

Selected Awards
nserc Canada Graduate Scholarship (Doctoral) May 2007 to April 2009

nserc Summer Program in Taiwan July 2007 to August 2007

Ontario Graduate Scholarship in Science and Technology January 2007 to April 2007

nserc Canada Graduate Scholarship (Master’s) January 2005 to December 2006

Refereed journal articles


Akter, T., S. P. Simonovic, and J. Salonga. 2004. Aggregation of inputs from stakeholders for flood management deci-

sion-making in the Red River Basin. Canadian Water Resources Journal. 29(4): 251-266.

Refereed conference papers


Salonga, J. and P. Gauvreau. 2010. Span-to-rise ratios in concrete arches: threshold values for efficient behaviour. 6th

International Conference on Arch Bridges in Fuzhou, China, October 11-13.

Salonga, J. A. 2008. Rational analysis of slender ultra high-performance fibre-reinforced concrete columns. 7th fib

Ph.D. Symposium in Stuttgart, Germany, September 10-13.

Salonga, J. 2007. Parallels in teaching visual arts and engineering design. Engineering Teaching and Learning Prac-

tices, ASEE St. Lawrence Section Conference in Toronto, Canada, October 19-20.

Non-referred reports
Salonga, J. 2007. The state of modern bridge design and construction in Taiwan. Report for nserc Summer Program

in Taiwan. Host professor: Chung-Yue Wang, National Central University, Taiwan.

Salonga, J. 2004. Aggregation methods for multi-objective flood management decision making: Red River Basin case

study. Undergraduate thesis, University of Western Ontario, Canada.


335

You might also like