Nothing Special   »   [go: up one dir, main page]

Compressive Membrane Action in Bridge Deck Slabs

Download as pdf or txt
Download as pdf or txt
You are on page 1of 283
At a glance
Powered by AI
The document discusses compressive membrane action in bridge deck slabs and how it can enhance strength and serviceability behavior. A relatively simple form of nonlinear finite element analysis is developed to model bridge deck behavior accounting for membrane action.

The thesis examines compressive membrane action in bridge deck slabs through elastic analysis of restrained slab strips. It develops a nonlinear finite element analysis method to model bridge behavior allowing for membrane action.

A relatively simple form of nonlinear finite element analysis is developed which is able to model bridge deck behavior allowing for membrane action. This reduces some of the disadvantages of nonlinear analysis which prevented its use in practice.

COMPRESSIVE

MEMBRANE

ACTION

::ln

BRIDGE DECK

SLABS

by

Paul Austin JACKSON


BSc CEng MICE MIStructE

Submitted to the
Council of National Academic Awards
in partial fulfilment of
the requirements of the degree of:

DOCTOR OF PHILOSOPHY

Sponsoring Establishment:
Polytechnic South West
Department of Civil Engineering
Collaborating Establishment:
British Cement Association

April 1989

COMPRESSIVE MEMBRANE

ACTION

:l.n

BRIDGE DECK SLABS

by

Paul Austin JACKSON

ABSTRACT
An elastic analysis of restrained slab strips shows that membrane action
enhances serviceability behaviour.
However, the enhancement is not as
great as for strength and serviceability is critical when membrane action
is considered in design.

A relatively simple form of non-linear firiite element analysis is developed


which is able to model bridge deck behaviour allowing for membrane action.
This reduces some of the disadvantages of non-linear analysis which have
prevented its use in practice. It uses line elements but, -because-of-novelfeatures of the elements and because it considers all six degrees of
freedom at each node, it is still able to model in-plane forces reasonably
realistically. It gives acceptable predictions for behaviour.
The tension stiffening functions used in non-linear analysis, which are
important to the prediction of restraint, are considered. Explanations are
proposed for several aspects of the behaviour and a new function is
developed.
This gives better results than previous expressions,
particularly for deflections on unloading and reloading.
Tests under full HB load have been performed on two half scale bridges.
These, and the analysis, show that conventional design methods for deck
slab reinforcement are very conservative.
They also show that the
restraint required to develop membrane action is not dependent on
diaphragms; it comes from under-stressed material surrounding the critical
areas. Thus, over much of a bridge's span, there is transverse tension in
the slab and membrane action does not significantly enhance the resistance
to global moments.
It is
Both bridge models failed by a wheel punching through the slab.
shown that these were primarily brittle bending compression failures which
were strongly influenced by global behaviour. This is confirmed both by
the analysis and by the higher wheel load at failure in single wheel tests.

Recommendations are made for using the results in design and assessment.

ACKNOWLEDGEMENTS
The research reported in this thesis was undertaken at the British Cement
Association, formerly the Cement and Concrete Association, whilst I was an
employee of that organisation.
former

directors

of

the

I wish to thank the Council, directors and

Association

for

instigating

the

project,

for

continuing to support it during a difficult period for the Association and


for allowing me to submit this thesis.

I should also like to thank Gifford

and Partners for practical help during the latter phase of the project,
particularly for the use of several hours of main-frame computer time.
Thanks are also due to my two' supervisors, Professor Robert Cope and
Doctor Andrew Beeby, for assistance with different aspects of the study.
In particular, I wish to thank Professor Cope for many helpful discussions
regarding the analysis.
Too many of the staff and former staff of the Association contributed to
the experimental phase of the project to name them all.
like to thank them all.

However, I should

Particular thanks are due to Colin Cook for the

instrumentation and to Daran Morahan and Ron Jewel, whose enthusiasm and
expertise enabled me to complete the testing in what became a very tight
timetable.
Thanks are due to my wife, Sue, for typing some of the text and for proof
reading the document.

Last, but by no means least, I should like to thank

our two sons Andrew and Simon, as well as Sue, for putting up with my
absence or preoccupation whilst working on the thesis.

CONTENTS
vii

LIST OF FIGURES
LIST OF TABLES

xi

1 INTRODUCTION

2 CURRENT DESIGN PRACTICE

2.1 Introduction

2.2 Outline Design

2.2.1 Choice of Form

2.2 .2 To Stress or not to Stress?

2.3 Detailed Design and Codes of Practice

2.3.1 The Importance of Codes of Practice

2.3.2 Sources of Code Clauses

2.3.3 Limit State and Working Stress Codes

2.3.4 BS 5400: Part 4: 1984


2.4 Analysis for Design - British Practice

11
15

2.4.1 Reasons for Linear Analysis

15

2.4.2 Section Properties

15

2.4.3 Global and Local Functions

16

2.5 Analysis for Design - North American Practice

17

2.6 Conclusions and Implications for this Study

18

3 PREVIOUS RESEARCH

20

3.1 Introduction

20

3.2 Reinforced and Plane Slabs

20

3.2.1 Bending Strength

20

3.2.2 Flexural Shear Strength

31

3.2.3 Punching Shear Strength

31

3.2.4 Ductility

44

3.2.5 Serviceability

47

3.2.6 Restraint

49

3.2.7 Global Behaviour

51

3.2.8 Empirical Design Rules

54

- i -

3.3 Prestressed Slabs

57

3.4 Conclusions

60

62

4 ELASTIC ANALYSIS
4. 1 Introduction

62

4.2 Assumptions

62

4.3 Stress

63

.4.4 Comparison with Other Analyses

'63

4.5 Crack Widths

64

4.6 Deflection

65

4c. 7 Effect of Restraint Flexibility

65

4.8 Conclusions

66

5 NON-LINEAR FINITE

El.EMENT ANALYSIS

67

5.1 Introduction

67

5.2 General Approach

67

5.3 Element Type

68

. 5.3.1 Slabs

68

5.3.2 Beams

69
69

5.4 Material Properties

69

5.4.1 Steel

70

-5.4.2 Concrete
5.5 Application to Membrane Action

72

5.6 Use in Design

73

5. 7 Conclusions

74

6 TENSION STIFFENING.

75

6.1 Introduction

75

6.2 Theory

76

6. 2. 1 Mechanisms

76

6.2 .2 Steel Stress

78

6.2 .3 Mesh Dependence

78
-

1i -

6.2.4 Cyclic Loading

79

6.2.5 Unloading

79

6.3 Analysis of Previous Tests

80

6.3.1 Direct Tension Tests

80

6.3.2 Flexural Tests

83
84

6.4 Tests
6.4.1 Design of Specimens

84

6.4.2 Loading Rig

85

6.4.3 Materials

86

6.4.4 Loading

88

6.4.5 Processing of Results

89

6.4.6 Results and Analysis

90
99

6.5 Conclusions
7 A SIMPLER NON-LINEAR ANALYSIS

100

7.1 Introduction

100

7.2 General Approach

101

7.3 Displacement Function

102

7.4 Element Initial Stiffness Calculation

106

7.5 In-Plane Forces

108

7.6 Large Displacements

111

7.7 Material Models

112

7.7.1 Steel

112

7.7.2 Concrete in Compression

113

7.7.3 Concrete in Tension

115

7.8 Stress Integration

119

7.9 Solution Scheme

120

7.9.1 Control

121

7.9.2 Initial Stiffness Method

121

7.9.3 Accelerators

122

7.9.4 Stiffness Recalculation

125

7.9.5 Convergence Criteria

127

7.10 Calibration

128

7.10.1 Duddeck's Slabs

129
- iii -

7.10.2 Taylor and Hayes' Slabs

131

7 .10.3 Batchelor and Tissington's Specimens

135

7.10.4 Kirkpatrick's Model

136

7.11 Conclusions

139

8 MODEL BRIDGE TESfS

140

8.1 Introduction

140

8.2 Design of Models

140

8.2.1 Scheme

140

8.2.2 Beams

143

8.2.3 Diaphragms

145

8.2.4 Slab Reinforcement

145

8.2.5 Bearings

148

148

8.3 Materials

8.3.1 Concrete

148

8.3.2 Reinforcement

154

8.3.3 Prestressing

155

8.4 Construction

155

8.5 Loading

157

8.5.1 Loads Applied

157

8.5.2 Loading Rig

160

8.6 Instrumentation

161

8.7 Tests on First Deck

163

8.7 .1 Global Service Load Tests

163

8.7.2 Global Failure Test

169

8.7.3 Local Failure Tests

177
181

8.8 Tests on Second Deck


8.8.1 Global Service Load Tests

181

8.8.2 Global Failure Test

190

8.8.3 Local Failure Tests

196

8.9 Tests on Single Beam

198

8.10 Discussion and Conclusions

202

8.10.1 Service Load Tests

202

8.10.2 Failure Tests

202
- iv -

9 ANALYSIS OF MODEL BRIDGE TEsTs

205

9.1 Introduction

205

9.2 Conventional Analysis

205

9.2.1 Analysis for Design of Deck Slabs

205

9.2.2 Analysis for Design of Beams

207

9.3 Non-Linear Analysis

213

9.3.1 Single Beam

213

9.3.2 First Deck

214

9.3.3 Second Deck

229

9.4 Conclusions

233

10 USE OF MEMBRANE ACTION IN DESIGN AND ASSESSMENT

235

10.1 Introduction

235

10.2 Use in Design

235

10.2.1 M Beam Type Decks

235

10.2.2 Other Beam and Slab Decks

239

10.2.3 Other Types of .Deck

239

10.3 Use in Assessment

240

11 CONCLUSIONS AND RECOMMENDATIONS

241

11.1 Conclusions

241

11.2 Recommendations

243

11.2.1 Recommendations for Design and Assessment


I

1.2.2 Recommendations For Further Research

243
243
245

REFERENCES
APPENDICES

A. Restrained Slab Strip to Elastic Theory

AI

AI Stresses

AI

A2 Deflection

A3

A3 Effect of Restraint Flexibility on Stress

A6

B. Transverse Shear Deformation of Line Elements

AB

All

C. Large Displacements

Cl Example Showing Effect of Vertical Component of

AB

Axial Force
C2 Effect of Slope on Axial Extension

Al2
A13

D. Notation

- vi -

LIST

OF

FIGURES

3.1

Geometry of restrained slab strip

3.2

Load-displacement relationship of restrained

21

slab strip

22

3.3

Comparison of flow and deformation theory

25

3.4

Elastic-plastic theory (1/h = 10>

27

3.5

Elastic-plastic theory <1/h

= 30)

28

3.6

Kinnunen and Nylander's model

33

3.7

Effect of concrete strength

38

3.8

Seal's Model Two

39

3.9

Kirkpatrick's model

40

3.10

Effect of span

42

3.11

Compressive membrane action to resist global


moments

52

3.12

Transverse stresses in a wide compression flange

53

4.1

Restrained slab strip

63

4.2

Effect of restraint flexibility

66

6.1

Tension stiffening functions

75

6.2

Stresses in concrete as cracks develop

77

6.3

Effect of tension stiffening on analysis'in


direct tension

82

6.4

Analysis of Williams' Specimen

82

6.5

Analysis of Clark's Beam 4

84

6.6

Detail of test specimens

85

6.7

Half size specimen under test

86

6.8

Results of first full size test

92

6.9

Results of first half size test

92

6.10

Results of third half size test

93

6.11

Ideal tension stiffening function

95

6.12

Tension stiffening function adopted

96

7.1

Displacement function

104

7.2

Effect of change to displacement function

106

7.3

Transverse displacements of a line element

108

7.4

Effect of in-plane shear

110

- vii -

7.5

Steel properties

112

7.6

Properties used for concrete in compression

114

7.7

Effect of concrete tensile strength

117

7.8

Initial stiffness method

122

7.9

Modified initial stiffness method

123

7.10

Analysis of Duddeck's slab

129

7.11

Analysis of Duddeck's slab 3

130

7.12

Analysis of Taylor and Hayes' slab 2S4

132

7.13

Analysis of Taylor and Hayes' slab 2R4

134

7.14

Batchelor and Tissington's specimen

135

7.15

Analysis of Batchelor and Tissington's specimen

136

7.16

Analysis of Kirkpatrick's panel C2

137

8.1

Details of first deck

141

8.2

First deck under test

141

8.3

Details of second deck

142

8.4

Second deck under test

143

8.5

Comparison of full size T2 and half size M4 beams

144

8.6

Detail of reinforcement in slab of first deck

146

8.7

Reinforcement in corner of first deck

147

8.8

Stress-strain relationship for reinforcement

155

8.9

First deck under construction

156

8.10

Spreader beam assembly

160

8.11

Load positions for first deck

164

8.12

Transverse soffit strain under wheel 10


<Service load tests)

165

8.13

Deflection under wheel 9 <Service load tests)

166

8.14

Beam deflections

170

8.15

Deflection under wheel 9

170

8.16

Shear cracks in Beam B

172

8.17

Soffit cracks under wheel 4

173

8.18

First deck under 400kN per jack

173

8.19

View across deck as failure approached

174

8.20

Failure; Wheel 4 punched through deck

175

8.21

Failure cone viewed from below

176

8.22

Single wheel test rig

178

8.23

Result of single wheel test

178

8.24

Crack pattern under single wheel

179

- viii -

8.25

Failure cone viewed from below

180

8.26

Load positions for second deck

181

8.27

Transverse strains adjacent to wheel 14


<Service load tests)

183

8.28

Deflection under wheel 14 <Service load tests)

184

8.29

Top cracks on completion of service tests

189

8.30

Transverse strains adjacent to wheel 14

191

8.31

Beam deflections

192

8.32

Top cracks immediately prior to failure

193

8.33

Second deck after failure

195

8.34

Single wheel tests A and C

197

8.35

Local tests B and D

197

8.36

Single beam under test

200

8.37

Load-deflection response of single beam

200

8.38

Shear cracks in single beam

201

8.39

Beam after failure

201

9.1

Beam deflections of first deck (conventional


analysis)

208

9.2

Beam strains in first deck

209

9.3

Beam deflections of second deck <conventional


analysis)

211

9.4

Slab strains over beams in second deck

212

9.5

Beam soffit strains in second deck

213

9.6

Analysis of single beam test

214

9. 7

Beam deflections of first deck <from non-linear


analysis using a coarse element mesh)

9.8

Deflect ion under wheel 9 <from non-linear an lysis


using a fine element mesh)

9.9

215
219

Beam deflect ions of first deck (from non-linear


analysis using a fine element mesh)

220

9.10

Predicted force across centre-line of first deck

221

9.11

Restraint forces predicted under single wheel load

225

9.12

Elastic analysis of stresses around a hole

226

9.13

Beam deflections of second deck <from non-linear

9.14

analysis)

230

Predicted force across deck slab of second deck

232

- ix -

A!

Restrained slab strip under line load

Al

Bl

Plan of line element

AB

B2

Unrestrained deformation

A9

Cl

Three element strut

All

C2

Inclined element

A12

X -

LIST

OF

TABLES
87

6.1

Typical mixes

7.1

Stiffness matrix of an off-set beam element

107

8.1

Mixes for in situ concrete

149

8.2

Test results for slab concrete from first deck

150

8.3

Test results for slab concrete from second deck

151

8.4

Test results for other in situ concrete from


second deck

152

8.5

Mix for precast concrete

152

8.6

Test results for precast concrete from first deck

153

8.7

Test results for precast concrete from second deck

154

- xi -

CHAPTER

INTRODUCTION
If in-plane restraint prevents material in the tension region of a beam or
slab from expanding as load is applied, a compressive force is developed.
This

force

can

lead

to

greater

predicted by normal flexural theory.


enhancement

is

relatively

small

strengths

and

stiffnesses

than

are

In a simple steel beam, however, the


and

arises

only

when

the

in-plane

restraint is applied below, that is on the tension side of, mid-depth.

In

concrete, and also in masonry, the low tensile strength and consequent
cracking mean that the effect can arise even when the restraint is applied
at mid-depth.

The enhancement can also be very much greater since the

compressive force enables even unreinforced slabs to support large loads.


This effect, which is known as compressive membrane action, arching action
or dome effect, has been known since the earliest days of reinforced
concrete.

It was described by Westergaard and Slater <1 > in 1921 and as

early as 1909 Turner <2> wrote of his flat slabs "such a slab will act at
first somewhat like a flat dome and slab combined".

Turner built many

flat slabs with reinforcement designed by empirical means.

At the time

there was good reason to use empirical design methods; the theory of flat
plates was not well developed.

However, Turner's contemporaries used more

conservative design methods and Sozen and Siess<3> report that, in 1910,
the weight of_ steel required in the interior panel of a flat slab varied by
a factor of four according to the design method used.

As they put it

"design methods could not be correct if the variation in results was 4-00%".
When an analysis based on simple statics was published in 1914-<4>, it
suggested that Turner's slabs were grossly under-designed; yet they had
behaved well both in service and in load tests.

Lord(5) had even measured

strains in a load test which appeared to support Turner and defy the laws
of statics.

Compressive membrane action was an important reason for these

discrepancies although there were others, including tension stiffening <3>.


Despite the satisfactory behaviour of Turner's slabs, design methods which
can be justified by statics are now preferred and purely empirical methods
have tended to fall out of favour whenever more rational methods have
become available.

Thus even flat slabs are now designed using flexural

theory, although Sozen and Siess <3> report that the charige was gradual
whilst Beeby (6) has shown that it is sUll not complete.
-

1 -

Apart from indirect

(and very limited> use in the Soviet dfsign code <7>,

compressive membrane action seems to have been largely forgotten for many
years.

Thus, in Braestrup's words <8>, "it therefore came as a surprise

when Ockleston <9> tested a real structure in South Africa and recorded
collapse loads that were three or four times the capacities predicted by
yield-line theory".

In fact Guyon<lO> had found similar results slightly

earlier, when he tested a multi-bay continuous slab, but this seems to


have been considered a characteristic of prestressed concrete.
Ockleston's results stimulated research into compressive membrane action
which has continued ever since.

Despite this research, which will be

reviewed in Chapter 3, the effect is still not normally used in design.


Recently, however, new design rules for bridge deck slab reinforcement,
which do allow for the effect, have been developed and incorporated into
the Ontario Highway Bridge Design Code <11>.

These rules lead to major

savings

methods;

compared

with

conventional

design

reduction in main steel plus a saving in design time.

typicslly

70%

More recently still,

similar rules have been adopted in other parts of the World, including
Northern Ireland.

The rules used in Northern Ireland <12 > were proposed by

Kirkpatrick et al<13) for use in the whole of Britain but have not yet
been accepted on the mainland.
One objection to these rules is simply that they are empirical.

Existing

theory shows, as will be seen in Chapter 3, that slabs designed to the


rules will have ample strengths under local wheel loads, provided there is

adequate restraint.

It even suggests that there would still be ample

strength with no reinforcement at all.

This, however, is the limit of the

extent to which the rules are proven theoretically.

There is also an

apparently serious omission from the experimental work on which they are
based.

An extensive series of tests on laboratory specimens, model bridges

and real bridges was undertaken yet none of the tests produced anything
approaching the full design global load on a bridge.

Thus the integrity of

the deck slabs under combined global and local effects is unproven.

Also,

they msy not give the load distribution which is assumed in the design of
the

beams;

particularly

as

global

analysis

based

on

uncracked

slab

properties is recommended <11,1 ~> for use with the rules.


Concrete slab design has come full circle; bridge deck slsb design is now
in the position which flst slab design occupied in
-

2 -

191~.

On the one hand

there is an empirical design method which seems to work and which is very
economical yet which could be considered unproven: on the other there is
the conventional method which is supported by flexural theory but which
seems to be very uneconomical.

Just as in 1910 design methods for flat


des~n

slabs could not be correct when they differed by 400%, so

rules for

bridge deck slab reinforcement cannot be correct now when they differ by
300%.

There is clearly a need for further research.

In recent years the assessment of existing structures has assumed equal


importance to the

des~n

of new construction.

Current design standards

are used in these assessments, but they often suggest that structures
which have given many years of satisfactory service are unsafe.

In many

such cases, compressive membrane action offers the possibility of more


realistic

assessment

reconstruction

which

work.

could

Previous

avoid

expensive

research,

having

strengthening

concentrated

and

on

new

construction, does not enable this potential to be fully used.


Another

problem

which

has

reinforcement corrosion.

become more

important

in recent

years

is

Resistance to this can be greatly improved by

increasing cover or by using epoxy coated, or other special reinforcement.


Both these approaches would become more economical if membrane action
were considered in

des~n.

It has even been suggested that satisfactory

deck slabs could be built without any reinforcement at all, which would
certainly avoid the problem of reinforcement corrosion.
"Localised"

reinforcement

corrosion

is

believed

to

be

particularly

dangerous <15) but an interesting implication of membrane action, which has


not previously been considered, is that this may have no

s~nificant

effect

on the behaviour of slabs.


In the present study the behaviour of bridge decks is
order .to develop and justify a rational

des~n

invest~ated

in

and assessment method which

can be adopted in British practice but which takes as much advantage as


possible of compressive membrane action.

The approaches used in the study

include tests on large scale model bridges and a simple elastic analysis.
However, because model tests alone can produce only empirical results,
whilst the behaviour considered is too complex to analyse in full by hand,
non-linear computer analyses are also used.

3 -

CHAPTER
CURRENT

DESIGN

2
PRACTICE

2.1 INTRODUCTION

In order to direct this study towards those areas which are important in
design, and to ensure that the knowledge gained will be usable in practice,
it is necessary to begin the study with a good understanding of current
bridge deck design practice.

That is, of the way bridges are assumed to

behave for design purposes, of the way they are designed, and of the
criteria and codes of practice they are designed to.
provide such an understanding.

This chapter aims to

There are also more fundamental reasons

for respecting past practice which will be discussed.


Design practice, unlike the real behaviour of bridges, differs significantly
between countries.

It is not practical, or necessary, to review practice

throughout the world.

This study is aimed at improving British practice,

so this chapter will concentrate on British practice.

Much of the most

relevant previous research has, however, been undertaken in North America


against a background of North American design practice.

There are several

important differences between British and North American design practice


which have greatly influenced the research and render its application in
Britain more difficult than might be expected.

In order to appreciate

these problems it is necessary to review the relevant aspects of North


American design practice.

Only conventional design methods, which ignore

membrane action, will be considered here.

The newer empirical design

approach, which allows for membrane action, will be considered in the next
chapter, along with the research from which it was developed.
2.2 OU11.INE DESIGN
2.2.1 Choice of Form

Before the detailed design of a bridge can be started the form of the
bridge has to be decided; for example solid slab, voided slab, beam and
slab, box girder or arch.
experience.

In making this decision, engineers are guided by

For particular ranges of span and sets of circumstances,

certain forms of structure have been found to be most economical.

Over

the years these favoured forms change, usually because of changes in


construction

technology

rather

than

because

of

advances

Construction considerations are always very important <16>.

- 4 -

in

analysis.

The desired

erection method nearly always decides the form of the bridge, rather than
the reverse.

For example, one does not choose to use precast elements in

a bridge because it is a beam and slab bridge, one chooses a beam and slab
bridge because it is convenient to precast.
The few cases when advances in analysis have changed the form of bridges
have arisen when those advances have enabled the analysis of structures
which are physically simpler but analytically more complicated.

An example

of this is the virtual extinction of intermediate diaphragms in beam and


slab bridges since load distribution analysis has been in widespread use.
These diaphragms served not so much to distribute load between beams as
to enable this distribution to be analysed.

With modern analytical methods

they are eliminated, sometimes at the price of doubling the transverse


steel in the deck slab.

In terms of material cost this change may be

uneconomic, but the difficulty of forming diaphragms in the span is such


that eliminating them leads to significant overall savings.

Thus if re-

introducing these diaphragms solved a problem in using membrane action


<and

Chapter

shows

that

this

is

the

case)

it

would still not

be

economical.
The dominance of construction considerations in the choice of the form of
bridges

means

behaviour
bridges.

of

that

study

such

as

this,

which

completed structures,

is

unlikely

considers

to

alter

the

only
form

the
of

It is thus essentially concerned with detailed design rather than

with scheme design.


2.2.2 To Stress or not to Stress?
Another decision which has to be taken in the early stages of a concrete
bridge design is whether or not to prestress and if so whether to pre- or
post-tension.
Again

this

construction.

decision

is

often

dictated

by

practical

considerations

of

It is not possible to build a glued segmental bridge without

post-tensioning and it would be difficult to build any long-span bridge


<except an arch> of ordinary reinforced concrete.

On the other hand, a

small slab is obviously more conveniently reinforced and small precast


beams are more easily pre-tensioned on a long line bed.

Only over narrow

ranges of structures (such as large voided slabs) is the decision marginal,

- 5 -

and

therefore

sensitive

to

small

changes

in

the

relative

costs

or

quantities of steel required.


In Britain, and most of the rest of the world, it has been found that,
because of the extra operations involved, transverse stressing of bridge
deck slabs is rarely economical.

A large number of tendons have to be

fixed, threaded, stressed and grouted, usually with very difficult access.
Much of the cost of these operations is fixed so that, even if the
required force were greatly reduced, transverse stressing would still be
unattractive.

Because of this, the present study assumes that deck slabs

will not be transversely stressed.

Accordingly, the rest of this chapter

concentrates on the design of ordinary reinforced concrete.


however,

In reality,

<unlike in most codes of practice> prestressed and reinforced

concrete are not fundamentally different.


effectively

prestressed

behaviour of the deck.

in

the

Also bridge deck slabs are often

longitudinal

direction

by

the

global

Thus research on stressed slabs can be relevant

and some of it will be considered in Chapter 3.


2.3 Df:I'AILED DESIGN AND CODES OF PRACTICE
2.3.1 The Importance of Codes of Practice
Most major bridge owners, including all of those in Britain, require new
construction to be designed to specified codes of practice.
codes are also frequently specified for use in assessment.

The same
Because of

this, codes have an importance which they owe as much to their contractual
position as to their engineering merit.

This alone justifies the extensive

reference which is made to them throughout this chapter.

It also means

that a new design method, such as one which allows for membrane action,
will be much more easily put into practice if it can be used within
existing codes.
as

this

Despite this, it is arguable that a research thesis such

should

be

concerned

only

with

fundamental

requirements

of

structural behaviour, and not with the sometimes arbitrary provision of


codes of practice.
distinct

from

Codes do, however, have a considerable engineering, as

contractual,

significance

which

arises

from

their

two

different, but overlapping, types of source.


2.3.2 Sources of Code Clauses
The first of these sources is the philosophy, theory and test data on
which codes are based.

The second is the cumulative experience which they


-

6 -

represent.

The latter means that a code can be considered as a set of

arbitrary rules which have been found to produce satisfactory structures


in the past.

Paradoxically, this applies to new codes, as well as to long-

established ones, because they are adjusted to make significant changes to


past designs only where they are known to be at fault.
The two sources of code clauses each have their faults.

Our understanding

of structural behaviour and our stock of test results are too incomplete
to enable them to be used as the sole basis of a code of practice.

On

the

other hand experience, as a source of code clauses, allows no innovation


and shows only where provisions are inadequate, not where they are overconservative or even

unnecessary.

It

has

also

been

pointed out

by

Beeby <17) that experience is an unreliable guide to design practice when,


as is usual with bridges, the design life is long compared with the timescale of change in loading, materials and design methods.
Code clauses owe

their origins

to a complex mixture of

theory,

test

results, experience and the engineering judgement of the code writers.


Theories are fitted to test results and to experience.

New theories and

test results are used to design structures which become part of the stock
of experience.

Experience is reviewed in the light of new theories, whilst

structures which were designed using discarded theories remain in the


stock of experience.

Finally, when experience shows that a subject needs a

code clause but not what the clause should be, and when there is no clearcut theory or evidence to go on, the code committee makes an arbitrary
decision.

By now it is often difficult to tell what specific source, or

even what type of source, any particular code clause is based on.

This

may not matter to the ordinary user of the code, but it is important when
the code comes to be reviewed in the light of new discoveries.
Even when the source of a code clause can be identified it may be a
matter

of

opinion

whether

the

clause

requirement or an arbitrary rule.

is

logical

and

fundamental

A classic example of this is the no-

tension rule in prestressed concrete, which can easily be traced back to


Freyssinet <18).

This rule illustrates how the source of a code clause

<that is, whether it is a fundamental requirement or an arbitrary rule


which has been found to work) affects, or should affect, the way it is
reviewed in the light of new discoveries.
detail.
-

7 -

This will be considered in more

In the 1970's Emerson<19> observed that bridges were subjected to large

temperature differentials with non-linear distributions.


many

bridges designed

tensile stresses.

to

the

no-tension

rule

This implied that

experienced

s~ificant

If the .no-tension rule was a logical and fundamental

requirement this was an alarming discovery indicating that the prestress


in those bridges needed to be increased.

On the other hand, 1f the no-

tension rule was simply an arbitrary design criterion which has been found
to produce satisfactory structures in the past, the discovery that some of
those satisfactory structures do experience tension is no cause for alarm.
If anything, it implies that the remainder of the structures

des~ed

to

the rule, which do not experience tension, have more prestress than they
need.

It

is now widely accepted that

the no-tension rule is largely

arbitrary eg. see Low<20>l but at the time it was treated as though 1t
was a rational and necessary requirement.

The result was that from the

introduction of non-linear temperature distributions into bridge design


practice in 1973<21>, up to the implementation(22> of BS 5400(23) and the
use of a degree of so-called "partial prestressing" in 1983, many bridges
were provided with unjustifiably large amounts of prestress.
Research, by providing new theory and test results, can invalidate code
provisions which are based on theory and test results.

Where new research

provides sufficient understanding of the relevant aspects of behaviour 1t


can also supersede code provisions which are based on practical experience.
des~

criteria for bridges are difficult to

define, let alone check by analysis.

A bridge has to survive a long life

Often, however, the critical

in an adverse environment and to remain serviceable after experiencing a


complex history of loads: environmental, functional and accidental.

When

we cannot fully analyse these things we rely upon experience to fill in


the gaps.
This

inability

consequent

to

fully

analyse

all

dependence upon experience,

conservative.

If code provisions,

aspects
tends

of

behaviour,

and

the

to make bridge engineers

and hence design methods, are based

purely on experience how can we know if it is safe to reduce the steel


area in bridge deck slabs?

One might think that until we can fully

understand all aspects of the behaviour of structures, we have to .keep


using as much steel as we always have.

In truth, however, theory can be

used to extrapolate experience and to use experience of one type of


structure in the design of another.

- 8 -

If we can prove with new theory that the steel in deck slabs, which is

designed for a stress of 345N/mm 2


say, 80N/mm 2

,.

not

know,

actually experiences a stress of only,

it does not prove that it would be safe to design the slab

with the new theory.


do

Until we can understand all aspects of behaviour we

from

theory

alone,

that

it

behaviour of the slab would be satisfactory.


term

or

cyclic

stress

of

over

80N/mm 2

would

be safe or

that

the

Is it possible that a longwould

cause

problems?

Our

experience of slabs does not answer this question because all the slabs we
can observe were designed by the very conservative method which we are
trying to supersede.

Simply supported beams are statically determinate,

however, so we know they experience the stress they are designed for.
Thus we do not need to fully understand all the implications of allowing a
higher stress to know if it is safe; we know it works in beams.

Thus,

even if new theory cannot prove that a slab design will be satisfactory, it
can show that the maximum stresses the slab will experience are less than
.those experienced by beams whose behaviour we know to be satisfactory.
Thus it enables the reinforcement in slabs to be reduced, refining the
safety margins towards, but not below, those already found satisfactory in
beams.
2.3.3 Limit State and Working Stress Codes
The great

majority of bridges built

Department of Transport standards.


Part

2:

1978 <24)

as

in Great Britain are designed

to

The loading standard used is BS 5400:

implemented

<and

significantly

modified)

by

BD

14/82<25) whilst the design standard for concrete bridges is BS 5400: Part
4: 1984 <26) which is implemented by BD 24/84 <27 >.

These are limit state

codes but the Department has only recently changed from using its own
standards <21,28,29> which were based on the working stress approach.

It is

helpful to review what this change in concept means.


The basic idea of a limit state code is that the various ways in which a
structure could exhibit unacceptable behaviour are considered in turn.

structure which is on the limit of acceptable behaviour is said to have


reached a certain "limit state".

Thus a structure which has the maximum

acceptable deflection could be said to have reached the limit state of


deflection.

Checking a design involves checking each limit state in turn.

Partial safety factors and the concept of probabilistic design have been
introduced at the same time as limit state philosophy but they are not
central to the concept or definition of a limit state code.
-

9 -

A working stress code specifies allowable stresses.

Checking a design

involves using elastic theory to calculate the stresses which exist in the
structure under working loads.
allowable stress.

It

These stresses are then compared with the

is the code writer's responsibility

to set

the

allowable stress at a level such that compliance with the limit ensures
satisfactory behaviour of the structure.
At first sight the two approaches seem quite fundamentally different.

It

might also be thought, as some engineers have argued<30>, that the limit
state approach involves the designer in a great deal more calculation than
the working stress approach.
cut.

In practice the difference is far less clear-

This is largely because it has never been possible to develop

reasonable

stress

limits

which

structure in every respect.

ensure

satisfactory

behaviour

of

The result is that so called "working stressA

codes require separate checks on what are really limit states; such as
deflection and crack widths.

Conversely, it has been possible to write

many limit state codes in such a way that compliance with one limit state
<and perhaps some nominal rules as well> ensures compliance with other
limit states.

In CP110 <31) - now BS 8110 <32) - this has been taken to the

point where it is normally only necessary to check one limit state, the
ultimate limit state.
In principle a limit state code needs only to specify the design criteria
for each limit state.

It could leave the designer

method used to check compliance.

free to choose the

In practice limit state codes do give

methods for checking compliance, although these are often optional.

The

important point is that, in principle at least, the design criteria are


fundamental characteristics of structural behaviour <such as strength or
deflection> and are independent of the method used to check compliance.
This differs from the situation in a working stress code where the design
criterion is that the stress, as calculated using elastic theory, should
comply with the limits.

There the design criterion <stress> and the method

for checking compliance <elastic theory) are not independent.

The result

is that the adoption of limit state codes should make the introduction of
new methods of analysis and design into practice much easier than it was
under working stress codes.

It should be simply a case of using the new

method to check compliance with the existing fundamental design criteria.


In practice it is not this straightforward, because the design criteria in
limit state codes are not always truly fundamental or independent of the
- 10 -

methods used to check compliance.


for the

This will be considered in more detail

particular case of BS 5400: Part 4: 1984<26>.

2.3.4 BS 5400: Part 4: 1984

BS 5400 is a limit state code and, as far as reinforced concrete is


concerned,

the

major

limit

states

which

the

des~er

is

required

to

consider are the ultimate limit state and the serviceability limit states
of crack widths and stress limits.
durability,

deflection

and

There are other considerations, such as

reinforcement

normally critical in conventional design.

fatigue,

but

these are

not

The important limit states will

be considered in turn.
a. Ultimate Strength
The need for a check on the ultimate limit state <formerly, and arguably
more correctly, known as the limit state of collapse> is obvious.

The

consequences of failure at this limit state are clearly very serious so the
acceptable probability of failure is very low.

For this reason the partial

safety factors used in BS 5400 for both loads and materials are larger for
this limit state than for the serviceability limit state.
In principle, the design criterion for the ultimate limit state is simply

that the structure should not collapse under the specified loads.
a

This is

fundamental design criterion so, having specified loads and material

strengths, the code is able to give some freedom as to how it is checked.


The usual approach is to analyse the structure using methods which will be
discussed in 2.4 and then to check sections separately for bending and
shear.

The bending strength check is done by assuming that plane sections

remain plane and using the code specified stress-strain relationship for
concrete and reinforcement.

There is an additional proviso that

the

reinforcement should yield at failure which was introduced to ensure a


ductile failure mode.
proved difficult

As the clause is of questionable value, and has

to comply with in some sections, the code allows the

alternative of providing

15~

extra ultimate strength.

b. Crack Widths
The need for the two main serviceability limit states, crack width and
stress, is less obvious and requires some explanation.

It is desirable to

limit crack widths for aesthetic reasons but the restriction in BS 5400 is
unnecessarily severe for this purpose.

This has arisen because it has

- 11 -

...

been

assumed

that

there

reinforcement corrosion.

is

relationship

between crack widths

and

Beeby <33> and others have said that neither the

available test evidence nor the accepted theory of reinforcement corrosion


support such a relationship.

It seems likely, therefore, that the BS 5400

crack width restriction .is unnecessarily severe, although at present it has


to be complied with in design.
In principle a crack width is a fundamental design criterion which is
independent of the method used to calculate it.
available

crack

width

prediction

formulae

In practice, however, the

give

such

widely

different

results [see Beeby <34)] that the criterion and the method for checking
compliance are interdependent.

For this reason BS 5400 explicitly states

that its criterion is that the crack width as calculated using the code
method should not

exceed the specified values.

The particular formula

specified in BS 5400 is based on that given in CP 110.

The background to

this is given by Beeby <34>.


The code only requires crack widths to be checked for functional, not
environmental, loads.
considered,

not

the

It also only requires 25 units of HB load to be

full

design

value

of

up

to

45

units.

Tension

stiffening is not used if more than half of the bending moment in the
section is due to live load.

This is to allow for the effect of repeated

loading and for the possibility that a section could have been pre-loaded
to a higher load than that for which cracking is checked.

This differs

from CP 110 and makes the crack width prediction formula conservative.
Despite this, and unlike under BE1/73 (28>, it is rarely critical in the
design of the main steel for bridge deck slabs.
c. Stress Limits
The provision of stress limits in a limit state code is something of an
anomaly.

It is contrary to the basic concept of a limit state code.

If

the deflections, strengths and crack widths are satisfactory it is hard to


see how a structure can exhibit unacceptable behaviour due to stress.

The

stress

the

limits

author<35).

in BS 5400 have
This showed

that

been

the

their

linear-elastic structural behaviour.

subject

purpose

is

of a study by

to ensure reasonably

This is not a fundamental design

criterion either but it is desirable for two reasons.

Firstly, the methods

given in the code for checking the other serviceability criteria, such as
crack width and deflection, assume linear elastic behaviour.
- 12 -

Thus the

check is needed simply as a


checks.
elastic

Secondly,
range,

chec~.

on an assumption made in the other

in a structure which. went significantly out of the

transient

loads

would

cause

permanent

deformations.

This would mean that a structure could be influenced by the cumulative


effect of all the loads which it had experienced throughout its life.
would be extremely difficult, if not impossible, to assess.

This

It is much

simpler to assume that a structure recovers from transient loads and limit
stress so that this is approximately true.

It is only because of this

restriction that BS 5400 is able to ignore some load cases when checking
crack widths.
Because of cracking, the real behaviour of reinforced concrete structures
is not linear-elastic.

To ensure even approximately linear behaviour it

would be necessary to limit the tensile stress to the cracking stress of


the concrete, which is not considered practical.

This means that a precise

analysis of a reinforced concrete structure still requires an assessment


of

the

cumulative

throughout
needed.
that

effect

its life.

of

all

the

As this is not

loadings

which

it

experiences

possible some other approach is

The only rigorously safe approach which is practical is to assume

the cumulative damage is total, and hence to ignore the tensile

strength

of

concrete

completely.

This

is done

in

some calculations,

notably in assessing the ultimate strength of sections in flexure.

It is

also done in BS 5400 when assessing the crack widths which occur in
sections loaded predominantly by live load.

The approach is not, however,

followed rigorously and many of the calculation formulae provided in codes


of practice do depend on the tensile strength of concrete<6>.
The stress limits in BS 5400 are not fundamental design criteria as they
are not

independent of the method used to check compliance.

This is

particularly true of the concrete compressive stress limitation of 0.5fcu


Concrete is significantly non-linear at this stress but the code writers
considered it acceptable to allow some redistribution.

This means that the

actual maximum stress in a section with a calculated maximum of 0.5f cu


would be less than 0.5fcu
the code criterion.

Despite this it still only just complies with

In axial compression, where there is no scope for re-

distribution, the code specifies the much lower limit of 0.38f cu

This

inter-relationship between the code's criterion <the stress limit) and the
method

of

checking

compliance

<elastic

theory>

means

that,

if

an

alternative analytical method is to be used in design, the design criteria


- 13 -

_,

~-

_..

have to be reconsidered, a fact which has not always been appreciated by


non-linear analysts.

For example, see Edwards <36 ).

In routine design to BS 5400 it is not usual to check the stresses in


reinforced concrete.
analysis at

The code allows the check to be avoided provided the

the ultimate limit state is elastic without redistribution.

The writer<35) has shown that, for normal sections in flexure, this rule
results in designs which are similar to those which would be obtained by
checking the stress.

However, sections designed to this rule which are

either heavily reinforced or subjected to axial loads can have calculated


concrete stresses which are significantly above the 0.5f cu limit.

Despite

this, such sections behave satisfactorily.


d. Critical Limit State
It

can be seen

normally

critical

concrete.

This

from

the preceding sections that

in

the

conventional

has

led

some

design

researchers<37)

ultimate strength is

approach
to

the

for

conclusion

research on bridges should concentrate on ultimate strength.


however,

ultimate strength

is

critical

in design only

reinforced
that

In reality,

because of

the

conservative approach <elastic structural analysis) which is used to check


it.

This approach is, in effect, deliberately chosen in order to ensure

that ultimate strength is critical and hence to avoid the need to check
other considerations, such as the stress limits.
In the case of deck slabs, which are subjected to concentrated wheel loads,
elastic theory predicts high moment peaks.
distribute.

In reality these peaks re-

Because of this, a yield-line analysis of a typical deck slab

designed by conventional methods shows that the ultimate strength is twice


what is required.

If, however, the designer opted to use this analysis for

design he would have to check the service stress.

The writer(35> has

shown that, because of this, the maximum saving in steel area which can be
obtained from the use of yield-line analysis is only about 11%.

Thus, if

analysis taking account of compressive membrane action is to result in


significant economies, it must indicate improved serviceability behaviour
as well as strength.

- 14 -

2.4 ANALY515 FOR DESIGN - BRITISH PRACTICE


2.4.1 Reasons for Linear Analysis

Linear elastic analysis is nearly always used in design.


because of the code of practice.
methods at

This is partly

BS 5400 does allow the use of inelastic

the ultimate limit state but, as we have already seen for

reinforced concrete design, there is little to be gained from this.


analysis is also convenient for another reason;

Linear

the principle of super-

imposition applies and, with the great number of load cases which have to
be considered, this is a major advantage.

It also makes linear analysis

much easier to computerise than other methods.


Linear elastic analysis is so widely used in bridge design that designers
tend to forget that the real behaviour of reinforced concrete <particularly
lightly reinforced concrete) can be highly non-linear even at service loads.
However, linear elastic analysis does lead to safe lower-bound solutions
which

is

more

important

in

design

than

realism.

Also,

if

<as

was

suggested in 2.3.2) a major justification for the design criteria in codes


is the experience that they have led to satisfactory structures, the mere
fact that linear analysis was used in the design of those structures is
sufficient justification for using it.

2.4.2 Section Properties

Having opted to use linear elastic analysis to analyse a highly non-linear


material, such as reinforced concrete, it is necessary to make some gross
assumptions

to obtain the section properties.

designer considerable freedom.

Here BS 5400 gives the

It allows the use of the gross concrete

section, the gross concrete section plus reinforcement transformed on the


basis of the modular ratio, or the reinforcement <again transformed) plus
the concrete but ignoring concrete which is subjected to tension.
reinforced

concrete

frame

structure

it

makes

little

difference

In a

which

section properties are used because the relative stiffness of the members
is little changed.

Bridge decks, in contrast, are often prestressed <and

hence uncracked) longitudinally but lightly reinforced transversely.


transverse stiffness may differ by as much as a
methods

whilst

significantly
section

their

affects

properties

longitudinal
the

will

results
lead

to

stiffness

but,
a

theory <38 ).
- 15 -

factor of 8 between

is

unchanged.

fortunately,

safe

design

Their

any

This

assumption

according

to

of

plastic

Gross concrete sections are almost invariably used because this enables
the final structural analysis to be performed before it has been decided
how much reinforcement to provide.

It is not always appreciated that the

alternative

cracked transformed section

behaviour.

Moment is proportional to curvature, but only if any axial

force also varies proportionally.


sections

in

sagging

and

is not

strictly linear in its

Furthermore the cracked transformed

hogging

are

different

in

that,

even if the

reinforcement is symmetrical, the moment reversal moves the neutral axis.


It is this tendency of cracked sections to change their section properties

as moment is applied which causes compressive membrane action.


to

the possibility

This leads

that an elastic analysis using cracked transformed

sections would enable compressive membrane action to be used in design


within the existing code.

This would avoid the need to solve the complex

problem of assessing the cumulative effect of load history on non-linear


structures.

Such an analysis, which is only linear under proportional

loading, will be considered in later chapters.


2.4.3 Global and Local Functions
It is difficult to analyse a whole bridge in sufficient detail to design

the deck slab reinforcement.

It is convenient, therefore, to divide the

behaviour into "global" and Nlocal" functions.

The local function of the

deck slab is to support wheel loads spanning between the beams.


be analysed by a variety of elastic methods.

This can

These are all based on

isotropic plate theory which, as we shall see in later chapters, does not
model slab behaviour well.

The most popular methods are those due to

Westergaard(39) and Pucher<40>.


to

be

fully

fixed-ended

or

In this analysis, the slab may be assumed


simply

supported.

Alternatively,

an

intermediate case is sometimes used.


The global functions of the slab are to distribute load between the beams
and to act as the top flange of the beams.

A variety of elastic methods

have been used for global analysis including methods based on orthotropic
plate theory, such as the Morice Little method (41>, and several computer
methods.

The modern trend is to use computerised grillage analysis almost

exclusively <42 >.

As a bridge deck is not a true grillage, this requires

some approximations, particularly to represent the torsional behaviour, and


advice on these has been published by West <43>.

One fault of grillage

analysis for which it is difficult to correct is that it assumes that the


main beams are connected together only by transverse beams which are in
-

16 -

the same plane as the main beams.

The beams in real beam and slab decks,

in contrast, are connected together by their top flanges and the in-plane
shear stiffness of these tends to even out the stress between the beams.
Ignoring this is conservative and, although the effect on the slab stress
is significant [see Hambly <U.)J, the effect on the beam soffit stresses is
quite small.

As the latter are critical in design the effect of the error

on design is not important.


The

calculated

global

and

local

normally simply added together.

transverse

moments

in

the

slab

are

This is not strictly correct as the end

moments assumed in the local analysis should, theoretically, be applied to


the global analysis.
In the longitudinal direction the global behaviour imposes an axial force
on the slab.
reinforcement.

It is common practice to ignore this in designing the slab

The code explicitly allows this at the ultimate limit state,

apparently because it assumes sufficient redistribution capacity.

Where

the force is always compressive <that is in a simply supported deck) it


can

be

shown

that

it

is

conservative

to

ignore

it,

even

at

the

serviceability limit state.

2.5 ANALYSIS FOR DESIGN - NORTif AMERICAN PRACTICE

Bridge design throughout North America is strongly influenced, although not


always

controlled,

Transportation
<AASHTO>.

by

the

American

Association

Officials Standard Specification

of
for

State

Highway

and

Highway Bridges<45)

This differs from the British Standard in philosophy, detailed

design methods, loading specification and analytical method adopted.


these differences the last is the most significant to this study.

Of
It is

also the least well known so it alone will be considered in detail.


The AASHTO standard does allow global analysis
similar manner to that normally used in Britain.

to be performed in a
However, it is usual to

distribute the wheel loads between beams using a table of distribution


coefficients provided in the Standard and then to use simple beam theory.
This gives a less favourable distribution than the British approach.

If

the beams are not closely spaced it gives a static distribution, which is
certainly conservative.

- 17 -

A table of values is also provided for the local analysis.

This is based

on Westergaard<39), which is one of the methods often used in Britain, so


the results are similar.

The most significant difference from British

practice is that the main steel in the deck slab is designed only for the
local

moment.

The global

transverse

moments

are

not

calculated

or

specifically designed for at all.


Global transverse moments obviously do occur in American bridges so it is
interesting to assess their significance.

Where a static load distribution

is used in designing the main beams, global transverse moments are not
needed to maintain equilibrium.

Thus, according to plastic theory, the

American approach leads to designs with adequate ultimate strength.

This

does not necessarily ensure satisfactory service load behaviour, but the
writer is not aware of any cases of failures in American decks which can
be

attributed

to

global

moments.

This

can

be

explained

conservatism of the method used for local analysis.


designed only for

by

the

The reinforcement

local effects is adequate to resist global moments

because the global moments are smaller than the calculated local moments.
This would not apply to many British "M" beam deck designs
Reference 46 ).

<eg. see

The small close-spaced beams lead to higher global, but

lower local, moments than the larger wider-spaced beams which are used in
North

America.

The

British

HB

load

also gives

much

transverse moments than does the American design loading.

higher

global

Thus it seems

likely that the American design approach would not work for many British
bridges.

The author also understands that problems have been experienced

with some bridge deck slabs in the Middle East, apparently due to a
combination of designing British-style decks to AASHTO rules and very poor
control of vehicle and axle weights.
2.6 cotl.USIONS AND IMPLICATIONS FOR THIS snJDY

The

basic

form

of

bridges

is

largely

dictated

by

considerations, so it is unlikely to be changed by this study.

construction
Accordingly,

the remainder of the study will concentrate on the detailed design of the
forms of bridges in current use.
The conventional methods of analysis and design which have been reviewed
in this chapter assume structural behaviour which is often very different
from the real behaviour of reinforced concrete.

Nevertheless they have

produced structures which have behaved in a satisfactory fashion.


- 18 -

They

should be respected for the wealth of experience which they represent.


This does not mean that complicated structures, such as bridge deck slabs,
which appear to have been over-designed in the past, always have to be
over-designed in the future.

It is possible to refine analytical methods

within existing codes so that effects like compressive membrane action are
allowed for in design.

This amounts to reducing the safety margins in

such structures towards, but never below, the standards already accepted
<and found satisfactory> in simple statically determinate structures.
Even if a more radical approach,

based on first

adopted, this review has important lessons.


strength

is

not

sufficient

condition

principles,

is to be

It shows clearly that ultimate


for

satisfactory

structure.

Serviceability criteria and the effect of the complex load history of a


bridge have to be considered.

It shows too, that a deck slab design needs

to consider global, as well as local, effects.

Finally it shows that many

of the design criteria given in codes of practice are only strictly valid
in conjunction with the methods specified for checking them.

If other

methods are to be used the criteria have to be re-considered.

This

applies particularly to the serviceability criteria, such as crack width and


stress limits, which are less fundamental than ultimate strength.

- 19 -

CHAPTER
PREVIOUS

RESEARCH

3.1 INTRODUCTION
The stock of evidence showing that conventional flexural theory underestimates the strength of restrained slabs is vast.
real

structures <5,9,4-7,48,49>,

model

structures <10,13,50,51,52)

specimens (53,54-,55,56 > under

laboratory

both

been

attributed

particularly
membrane

since

action

theoretical

and

to

1955
has

compressive
when

been

Much of this extra strength

membrane

Ockleston<9>

the

experimental.

all the literature in detail.

action.

As

published his

subject

of

Research

has

countries over a long period of time.

and

concentrated <10,13,51,52,55 >

and uniformly distributed (5,9,50,54,56) loads.


has

It includes tests on

extensive
been

test

result,
results,

research,

undertaken

in

both
many

It is thus not practical to review

This chapter aims only to establish the

present state of knowledge of the subject as it affects, or could affect,


the design of bridge deck slabs.
Much of the experimental work which is most directly relevant to this
study, including most of the Canadian work mentioned in Chapter 1, has
been conducted in the last fifteen to twenty years.

Non-linear finite

element analysis, capable of allowing for compressive membrane action, has


been developed over much the same period.

Despite this, there is almost

no reference to the non-linear analytical work in the experimental studies


so it is convenient to consider finite element studies entirely separately
in Chapter 5.
3.2 REINFORCED AND PLANE Sl.ABS
3.2.1 Bending Strength
As

it

was

the

realisation

that

flexural

theory

under-estimates

the

strength of restrained slabs which promoted the interest in compressive


membrane

action,

bending strength.

it

was

natural

that

research

should

concentrate

on

Many researchers have extended flexural theories to

allow for in-plane forces.

Most have used Johansen's Yield-Line Theory<57>

as their starting point but a variety of approaches have been used.

It is

convenient to illustrate each in turn by considering the simplest possible


case;

symmetrical restrained

slab strip with equal top and bottom

reinforcement.
- 20 -

t'l

"0

I_

.I

Fig ure 3 .1:

Geometry o f r estrained slab strip

a. Rigid Plastic Deformation Theory


The simplest approach is to assume that the slab material is fully plastic
at the yield-lines but rigid elsewhere.

This gives the geometry shown in

Figure 3. 1 and it will be seen that :


=

by symmetry

say

Also, since both the restraint force, F, and the reinforcement area, A., are
the same at all sections:
dc2

de

AB2 + BC2

AC 2

<1/2 )2 + w2

Now;

Hence;
[1/2 + 2 (h/2 -dc).2w/.D2

Neglecting second order terms, this leads to;


d.:

h/2 - w/4.

Equation 3.1

Using a rectangular concrete stress block, considering a unit width of slab


and ignoring the tensile strength of concrete, the force in the concrete is

dc f c:; '

where fc ' is the "plastic" concrete stress; approximately 0.6fc: ....


The tensile force in the rein forcement is
- 21 -

A.fv

where A. is the steel area per unit width and fy is the yield stress.
Ignoring any reinforcement in the compression blocks,

this gives restraint

force, F

The total lever arm for the concrete in the compression blocks, taking the
forces to act horizontally, is

h - 2 <de: /2) - w

h/2 - 3w/4

<substituting for de:


from Equation 3.1)

and the total lever arm for the tensile steel forces

2 (d - h/2) + w

2d - h + w

So the total moment, that is the sum of the support and mid-span moments,

f.:' (h/2-w/4)(h/2-3w/4) + A. f Y <2d-h+w)

By using the virtual work approach, the load-displacement relationship of


the slab strip under any symmetrical load case can now be obtained.

The

result for some typical strips subjected to a single central load is shown
in Figure 3 .2.

Pll h2 30
20

10

p=0.5%

0.0

0.2

0. 4

0.6

0.8

1.0

1. 2

1.4

w/ h
Figure 3 .2:

Load-displacement relationship of restrained slab strip

<Rigid-Plastic Theory, f c ' = 20N/mm2

fy = 460N/mm2

d/h = 0 .8)

The unreinforced slab's maximum load, which occurs at zero displacement, is


as great as that of a simply supported slab with some 2% 1e inforcement.

It is also equivalent to an unrestlained, but still fixed-ended, slab with


about 1~ reinforcement.

The strength of the slab with


- 22 -

0.5~

reinforcement

is greatly enhanced by the restraint, but is only slightly greater than


that of the unreinforced slab.
The

load

on

unreinforced

the
slab

slabs

reduces

will not

greater than 0.67h.

as

support

tensile

any

displacement
load at

increases.

all at

The

displacements

The reinforced slabs reach a minimum load but then

the load starts to pick up again.


by

the

membrane

action;

This is because the slabs start to work

the

load

is

supported

component of the tension in the reinforcement.

by

the

vertical

Eventually the load carried

in this way can exceed the initial "ultimateu load.


Real slabs are not rigid between their yield-lines so they do not reach
their maximum compressive membrane load at zero displacement.

Thus the

real peak load is lower than shown in Figure 3.2, and occurs at

s~ificant

displacement.

However, apart from this, Figure 3.2 gives a good indication

of the behaviour of slabs, subject to certain conditions which will be


discussed in 3.2.2 to 3.2.4..

Researchers, such as Brotchie and Holley <56>,

have performed tests under displacement control and traced the descending
and ascending part of the curve after the ultimate compressive membrane
load is exceeded.
The ability of reinforced concrete slabs to support
tensile

membrane

exceptional

action

accident

may

loads.

occasionally
However,

useful

be

because

of

displacements required, it is of no practical use in the


decks.

load by

s~ificant

for

the

resisting
very

des~

large

of bridge

Slabs with realistic span to depth ratios become unserviceable long

before they enter the tensile membrane range.

In most practical bridge

deck slabs a deflection of 0.05h would be excessive.


Although the basic approach of rigid plastic deformation theory is simple,
the algebra becomes complicated when the yield-line patterns of two-way
spanning slabs are considered.
few cases.

One

Solutions have been published for only a

of the first to be solved was the axi-symmetrical case of

a fully restrained circular slab with isotropic reinforcement.


published by Wood (58>,

This was

who went on to use it to give an approximate

solution for square slabs.

He then compared the predictions of this

theory with the available test data.

Because of the elastic deformation,

the theory over-estimated the strengths.

Wood suggested that this could

be allowed for by multiplying the predicted loads by a reduction factor.


He found that the measured factors varied from 0.4. to 0.8; the smaller
- 23 -

factors occurring in the more lightly reinforced specimens.

This was

because heavily reinforced slabs are less sensitive to restraint.

If the

factor is calculated from the increase in load compared with that given by
yield-line theory, rather than from the total load, the range of observed
values is much smaller and there is no consistent trend with steel area.
Brotchie and Holley <56> used an alternative approach for correcting the
unsafe predictions of rigid-plastic theory.

Instead of multiplying the

load predicted for zero displacement by a reduction factor, they used the
load predicted for the displacement at which rigid-plastic theory gave the
same load as an elastic analysis.

Since both theories give upper-bound

solutions for the load at a given displacement on a structure composed of


elastic-plastic materials, it appears that this should over-estimate slab
strength.

This explatns why "theoretical maximum loads are slightly higher

than the test results for the thinnest slabs".

However, the theory tended

to be conservative for the thickest slabs, which had a span to depth ratio
of only 5.

This was because elastic flexibility has little effect on the

strength of such slabs whilst


greater than in shallow slabs.

the effect

of

triaxial enhancement

is

They attempted to allow for this but their

correction was conservative.


b. Rigid-Plastic Flow Theory
Plastic deformation

theory assumes

that

concrete develops

its plastic

compressive stress whenever it is subjected to compressive strain.

In

reality, not only does the strain have to be significant, it has to be


increasing; the stress reduces rapidly if the strain decreases.

Equation

3.1 predicts that the neutral axis moves closer to the compression face as
the deflection increases.

This implies that some concrete, near to the

neutral axis, experiences a reducing strain and so will not develop its
full

compressive stress.

The resulting error in the analysis can be

avoided by using "flow theory" which assumes that the full stress is
developed whenever the strain is increasing.

The derivation of Equation

3.1 is then replaced by its first differential with respect to displacement


or, more correctly, time.
do:

Braestrup(8) has shown that this leads to:

h/2 - w/2

and the load displacement relationship for the simple strip can then be
calculated in the same way as before.

- 24. -

Braestrup (8) noted that no clear distinction is made in the literature


between

flow

theory and deformation theory.

He suggested that, as a

result, much of the past research, which is based on deformation theory, is


in error.
In Figure 3.3, the results of flow and deformation theory are compared for
the

unreinforced strip which

was

considered

in Figure 3.2 .

At

large

displacements there is a very significant difference but in the practical


range of bridge deck slab deflections the difference is not significant .
Also, deformation theory is conservative because the extra concrete force
it predicts is on the wrong side of the undeflected centre-line and so

develops a couple which acts against the resistance moment .


p l/h 2 20

15
10

Deformation

0.8

0.6

0.4

0.2

0.0

w/h
Figur e 3 .3:

Comparison of flow and deforma tion theory

<p = 0,

f ,::'

= 20

N/mm2

Even in slabs which deflect 0 .5h or more before reaching peak load, the
difference between deformation and flow theory is not as great as Figure
3.3 suggests because the extra deflection is due to elastic deformation
and so does not have the same effect on the neutral axis position.
only

when

becomes

post- ultimate

important .

Since

behaviour

is

post- ultimate

considered
behaviour

that
is

the
of

no

It is

difference
practical

importance to the applications considered in this thesis, it is reasonable


to consider deformation and flow theory as interchangeable.
Mor ley <59) developed rigid-plastic flow theory so that, in principle, it is
general and can be applied to any case.

The algebra becomes complicated,

however, and he gave only a limited number of solutions.


for polygonal slabs.

One of these was

He compared this solution with some test results,

assuming that the true maximum load was that predicted for a deflection
of h/2 .

The choice of this deflection was based on work by Park(60) which

- 25 -

will be considered

later.

For the slabs which Morley considered the

predictions were reasonably good.


c. Elastic-Plastic Theory
Johansen's

Yield-Line

strengths
This

is

of

Theory<57)

unrestrained

possible

displacement.

because

Thus

deflections

slabs

without

it

gives
despite

predicts

elastic

very

predictions

ignoring

loads

deformations

affecting

good

strength.

elastic

which
can
When

are

for

the

deformations.

independent

significantly
membrane

of

increase

forces

are

considered, in contrast, there is a relationship between load and deflection


even in plastic theory.

Thus elastic deformations affect strength and it

is useful to consider them in an analysis.


The analysis is particularly sensitive to elastic shortening of the slab
because, as will be seen from Figure 3.1, small movements have a large
effect

on

the

behaviour,

particularly

at

small

deflections.

Ideally,

however, both in-plane and flexural deformations would be considered.

The

full equations for this have been formulated by Massonnet <61> and have
been applied to rectangular concrete slabs by Moy and May field <62 ).

The

mathematical complexity of the equations is such that hand solutions are


not practical so Moy solved the equations numerically by computer using a
non-linear

finite

reasonably

well,

programs.

difference
it

has

approach.

proved

Although

difficult

this

approach

to develop general

works

computer

Because of this the approach has been largely superseded by the

finite element method, which will be considered in Chapter 5.

It would be

particularly difficult to develop a finite difference program which could


be used by a non-specialist in a wide enough range of circumstances to
make it commercially viable.
far

Because of this the finite element method is

more suitable for direct

use in design and the finite difference

method will not be considered further in this study.


When elastic deformation is included in a hand analysis it is necessary to
make some gross approximations to simplify the mathematics.

The approach

adopted by Park <54), which has been followed by many other studies, was to
ignore the flexural deformation and to assume the axial strain to be
constant along the length of the strip.

Since flexibility in the in-plane

restraint

on

has exactly the same effect

the behaviour as

the axial

flexibility of the strip, it is both useful and convenient to include it in


the analysis.
- 26 -

Considering the same simple strip as before, and using deformation theory,
this leads to;

d.~

h/2 - w/4- - EF-/8w - t 1/4-w

where is the axial strain at mid-depth <which is taken to be constant) t


is the movement of each support, and the other notation is as before.

and,

Now
if

is

calculated

from

d._ f c: I

A.. f y

the

gross

concrete

properties

and

the

restraints are taken to be elastic such that;

Kt

d~

h/2

into

the

this leads to;

Substituting

this

result

+ f ,::' (P/8Erhw - 1/4-Kw)


expression

for

the moment,

which

is

otherwise the same as in the rigid plastic theory, the moment and hence
the load can be obtained.
In Figures 3.4 and 3.5 the result of this calculation is shown for slabs
with 0.5% steel and with span to depth ratios of 10 and 30 respectively.
In order to give an indication of the restraint stiffness required this is
expressed as a multiple of the axial stiffness of the slab strip.
Rigid- Plastic Theory
Pl/h 2 25

20

-I

-- Elastic-Plastic Theory
Unrestrained

K=E, h / 1

15

I
10

Kf= 0. lE.- h / 1

-1 I
I

---

0
0.0

0. 1

0. 2

0. 3

0 . 4-

0. 5
w/h

Figure 3. 4:

Elastic-plastic theory

(1/h = 10, p = 0.25%, f c: ' = 20N/mm.:: , f y = 460N/mm:.<: , d/h = 0.8)

- 27 -

At zero deflection there is no restraint force and the load is as given by


normal yield-line theory.

The load increases with deflection but starts to

reduce again before reaching the value predicted by rigid-plastic theory.


Less well restrained slabs support less load and reach peak load at higher
deflections.

The slab with the larger span to depth ratio is much more

sensitive to flexibility of the restraints but restraints which are far


less

stiff

than

the slab still have

very significant

effect

on its

strength.
Park <54) considered a more genera l case and used a different stress block
and notation but, apart from this, his theory is the same.
gave graphs

equivalent

to Figures

3.4 and

3.5

which

However, he

are significantly

different; for example, the load at zero deflection is not equal to that
given by yield-line theory.

This is due t o a small algebraic error in the

exa mple used to plot his graphs.


----------Rigid-Plasti c Theory
p l/h2 25
-

-Elastic-Plastic Theory

20

- Unrestrained

K=oo - - /

15

---

/K=E.=h / 1
_
/ / / ..- K= 0 . 3E.=h / l
10

I /........- -- -

K = 0 . 1E" h / 1

0 .0

0. 1

0. 2

Figure 3. 5:

0. 3

0. 5

0. 4

w/ h

Elastic- plastic theory

<I/ h = 30, p = 0. 25%, f c ' = 20N/mm 2

f y = 460N/mm-.' , d/h = 0. 8 )

Roberts<5 3) tested a series of restrained strips.

He compared the results

with theory similar to the above, which he attributed to Wood.

Because of

elastic flexibility in bending, the load at low deflections was less than
given

by

Roberts

the

theory.

attributed

this

The peak load exceeded


to

the

effect

of

the predicted value and

transverse

concrete at the supports which enhanced its strength.


proved

that

it

was possible

restraint

to

the

Supplementary tests

to develop stresses in excess of 0.67f '='-' "

After the peak load was passed the load reduced much more rapidly with
- 28 -

This was due to the

increasing deflection than the theory predicted.

difference between the real behaviour of concrete and the ideal plastic
behaviour assumed in the analysis.
Christiansen (63> developed a theory for restrained beams which is similar
except that he added the elastic bending deflection into the analysis.
calculated this using the uncracked section.

He

In addition to applying the

theory to beams he also applied it to slabs, including two-way spanning


slabs.

In

these

enhancement,

the

varies

deflection,

across

the

and

slab

hence

width.

the

effect

of

Christiansen

membrane

avoided

this

complication by "considering only arching action across the shorter span at


the centre of the longer span."

As expected

(and intended> this gave

conservative answers.
Park(54>

used

his

spanning slabs.

strip

theory

to

estimate

the

strength

of

two-way

He did this by assuming a central deflection and using

the strip theory to obtain the moment to use in the virtual work equations
obtained
deflection

from
of

normal
h/2,

yield-line

which

was

theory.

based on

He
a

chose

study of

to

use

test

central

results.

He

acknowledged that this deflection was conservative for slabs with span to
depth ratios below about 20.

He also acknowledged that it is a greater

deflection than his graphs, based on his strip theory, suggest.

In fact,

because of the error mentioned earlier, his graphs show peak loads which
are slightly lower than they should be and which occur at significantly
higher deflections than they should.

Thus the h/2 used by Park does not

agree with the strip theory but Park suggested that this was justified by
the elastic bending which the analysis ignores.

He showed that the theory

gave good predictions for

the strengths of slabs subjected to uniform

loads.

of

However,

because

the

use

of

deflection

of

h/2,

it

is

conservative for slabs with short span to depth ratios.


The

algebraic

complexity

of

this

elastic-plastic

theory

of

two-way

spanning slabs makes it difficult to use and gives it a false impression


of accuracy.

In fact, it is based on gross assumptions.

It is quite

different from the use of elastic-plastic material properties in non-linear


computer analysis.

It assumes that the whole depth of the slab is plastic

at the critical sections.

Elsewhere it is taken to be elastic for axial

behaviour but rigid in flexure.

The assumption that the axial strain of

the slab strips is constant at mid-depth can easily be shown to be wrong,

- 29 -

for example the neutral axis depth at the yield-line is always less than
h/2 which implies an axial shortening at this section.

Thus the major

justification for the equations developed by Park is not


which they are based
give

reasonable

the theory on

much as the fact that they have been shown to

50

results.

This

is significant

as

implies

it

that

the

approach is essentially empirical and thus may not be valid outside the
range of cases for which it has been tested.
Me Dowell et al<64> developed a different form of elastic-plastic analysis.
Although intended for use with masonry walls, it is equally applicable to
unreinforced

concrete.

It

used

the geometry shown

in Figure 3.1

and

assumed that the strain varied linearly in the span direction, from zero at
the crack to a maximum in the compressed region.

This was acknowledged

to be an arbitrary assumption, and it is easy to prove that it is not


Since the

correct, but it is just as reasonable as Park's assumptions.

total reduction in the slab length at any depth can be calculated from the
geometry shown in Figure 3.1, this enables the strain to be calculated at
any position.

Me Dowell used the strains at the yield-line positions to

calculate the stresses, and hence the bending moments, using an elasticplastic stress distribution.

He assumed that, once the plastic stress had

been reached, a subsequent reduction in strain would reduce the stress to


zero.

This made his approach equivalent to flow theory.

Rankin <65) has successfully applied the approach to unreinforced concrete


slabs.

He also adapted it to reinforced slabs by adding the effect of the

reinforcement.

Skates,

Rankin

and

Long<66)

used

similar

approach

although their method for combining the components of the moment capacity
due

to

arching

acknowledged

and

that

reinforcement

his

was

flexural and

slightly

arching

different.

analyses

Rank in

assumed

different

strain fields and the same is true of the approach used by Skates et al.
The main consequence of this is that the assumption that the reinforcement
yields

could

analysis.

be

inconsistent

with

the strains assumed

in

the arching

Although not stated in the other literature, this is a fault

which is shared with all the analyses considered in this section.


suggested

that

the

resulting

unsafe

predictions

could

be

Rankin

avoided

by

limiting the calculated moment capacity to the "balanced" capacity proposed


by_ Whitney <67) which is approximately 0.27f .,~bd2

This restriction appears

to be

taking d/h as 0.8,

maximum

conservative.
possible

Rankin
arching

pointed
moment

out

that,

capacity

- 30 -

of

an

unreinforced

the
slab

approximates

to

this

However,

capacity.

even

if

reinforcement does increase the moment capacity.


decks, d/h is significantly less than 0.8.

it

does

not

yield,

Also, in some bridge

Another conservative aspect of

Rankin's analysis is that although the reduction in the concrete lever arm
due to deflection is included, the increase in the steel lever arm is not.
_Thus the analysis would be conservative for shallow heavily reinforced
slabs . . Despite these faults, Rankin obtained good results and his approach
will be considered further in 3.2.4-.
3.2.2

F~exural

Shear Strength

The theories considered in 3.2.1


shear

failure:

With

few

assume that

exceptions,

flexural failure precedes

this assumption

literature without any particular justification.

is

made

in

the

It is therefore necessary

to investigate the validity of the assumption and again it is convenient


to consider the simple slab strip shown in Figure 3.1.
If the span to depth ratio is less than about 20, rigid plastic flexural

theory implies a shear force which exceeds the ultimate . shear strength
given

by

BS

54-00.

However,

this

ignores

the

fact

that

an

axial

compressive force enhances the shear strength of a concrete section.

simple correction for this, such as that given in the column clauses of
BS 5400, suggests that shear failures are only possible if the span to
depth ratio is less than about 6.

Since the code rules are conservative,

and shear strength is further enhanced if the shear span to depth ratio is
less than around 2.5

<which is equivalent to a flexural span to depth

ratio of 5 >, this means that shear failures in the type of strip shown in
Figure 3.1 are unlikely.
This argument can be extended to show that shear failures are unlikely in
practical restrained slabs subjected to uniform loads. and explains why no
such failures have been reported.
3.2.3 Punching Shear Strength
Even allowing

for

the limitation on the

load

imposed by the bending

strength of a slab, the shear stress in the-vicinity of a concentrated load


is much higher than under a uniform load.

Because of this, slabs subjected

to concentrated loads are likely to fail by punching and test results


confirm this <10,13,51,52,55 >.

Despite this, restrained slabs are stronger

. than unrestraiiled. slabs and, typically, five times stronger than suggested
- 31 -

by

conventional design rules

which assume

flexural

failures.

Several

attempts have been made to analyse the effect.


Aoki and Seki<68l have modified Moe's <69l equation for punching strength
to allow for membrane forces.

However, the correlation with their test

results was not particularly good and they obtained a better relationship
using a purely empirical formula.

Although this formula worked reasonably

well for their tests the author has found that it gives unsafe predictions
for many other restrained slabs and it will not be considered further.
The realisation, following research by Young (70), that bridge deck slabs
fail by punching at high loads prompted the Department of Highways and
Transportation

in Ontario

to sponsor a

major

research

programme

into

punching.

After largely experimental studies by Tong and Batchelor(51l and

Batchelor

and

Tissington<71l,

Hewitt

and

Batchelor<72l

endeavoured

to

develop a theoretical model by modifying an existing theory for punching


in unrestrained slabs.
They found that the best available theory for punching in unrestrained
slabs was that due to Kinnunen and Nylander<73l.

Kinnunen observed that

the punching failure modes of slabs were approximately axi-symmetrical,


even for rectangular specimens, so he used an axi-symmetrical analysis.

In

this model, which is illustrated in Figure 3.6, outer portions of the slab
bounded by a shear crack and two radial cracks are assumed to rotate as
rigid bodies.

The load is taken by the compressed conical shell above the

shear crack which is assumed to be shaped such that the concrete stress
is constant.

The system is taken to deform linearly with load until a

limiting strain

is reached and

the system

fails.

The stress

in the

compressed shell at failure is calculated allowing for the enhancement due


to the triaxial stress state.

Finally an empirical correction factor of 1.1

is applied to allow for dowel effect in the radial bars which the .analysis
ignores.
Hewitt and Batchelor applied the theory to 137 test results and. obtained
good results.

They said that they were better than Moe<69l obtained using

a purely empirical relationship.

However, since Kinnunen and Nylander used

empirical

strain,

factors

for

limiting

triaxial

enhancement

and

dowel

effect whilst Hew it t and Batchelor increased the factor for dowel effect
from 1.1 to 1.2 to improve correlation, the resulting "theoretical punching
load" is largely empirical.

In effect, the model was used only to give a


- 32 -

qualitative explanation of the behaviour and to give the form of the


equations; the actual values are empirical.
By adding the restraining force and moment acting at the outer edge of
the segment into the equilibrium equations, Hewit t was able to correct the
theory for known restraint forces.

Comparison with the results of tests

in which the known restraint was provided by unbonded tendons suggested


that the approach gave good predictions.

a)

SECTION SHOWING BOUNDARY FORCES

b)

FORCES ON SECTOR ELEMENT

Figure 3.6: Kinnunen and Nylander's model (73)

(as modified by Hewitt and Batchelor<72)J


By adding the restraining force and moment acting at the outer edge of
the segment into the equilibrium equations, Hewitt was able to correct the
theory for known restraint forces.

Comparison with the results of tests

in which the known restraint was provided by unbonded tendons suggested


that the approach gave good predictions.
For most practical slabs, the restraining force and moment are unknown.
Hewitt therefore proposed that the actual boundary forces and moments in
real slabs should

be expressed as a

maximum boundary forces and moments.

"restraint

factor",

R,

times

the

These maximum values were obtained

by using "The idealised geometry of displacement as used by Brotchie and


- 33 -

Holley (56>".

This gives a neutral axis depth at the support of h/2 - w/4

as in Equation 3.1.

However, this assumes that the neutral axis depths at

the support and mid-span are equal which conflicts with Kinnunen and
Nylander's assumptions.

There is also no reason why the actual restraint

force and moment should both be reduced by the same percentage relative
to their respective maximum values.

Thus Hew it t 's approach is, in effect,

largely empirical and he appeared to acknowledge this, saying "It is not


implied that the actual boundary restraint and distribution of stress are
known at the instant of failure".
Hewitt obtained the restraint factor values, R, for real slabs by backcalculation from observed failure loads.

Although he said "It is a fact

that

supported

R varies

from

zero

for

simply

slab

to

unity

idealised restraint" the highest value he observed was only 0.77.


appear to be two reasons for this.

with
There

Firstly Hewitt's analysis with full

restraint invariably gives a depth to root of crack which is greater than


h/2.

This is only geometrically compatible with his assumption that the

neutral axis depth at the support is h/2 if the supports are jacked closer
together.

Secondly, Hewitt assumed that the top steel at the supports

reaches yield which, except with large deflections, is incompatible with


Thus "full restraint" in his theory

the assumed neutral axis position.

appears to represent the ideal restraint forces, that is the forces which
r
\i"

lead to the highest

failure load, and not

<as some of his statements

imply> the forces which arise with ideal . <rigid) restraint; R = 1 could
only be obtained by prestressing.

Another oddity of the model is that it

assumes that a volume of concrete, bounded on one side by a shear crack,


rotates as a rigid body until a shear compression failure occurs; yet all
the descriptions of failures show that the shear crack does not appear
until the failure load is reached.
Clearly,

allhough

claimed

essentially empirical.

to

be

theoretical model,

the

approach

is

Hewitt claimed that it gave acceptable predictions

for the strengths of realistic bridge deck slabs and it has been used to
develop charts for assessing the strength of existing bridges <11 >.

In

order to ensure that these are safe, and to avoid the need for separate
charts

for

use

with

steel and

concrete

beams,

they are

based

on a

restraint factor of 0.5 even though tests on concrete bridges suggest that
values as high as 0. 7 give more accurate predictions.

- 34 -

Kirkpatrick, Rankin and Long<13) have developed an alternative analysis of


punching in restrained slabs.

Like Hewitt's approach, this was developed

by modifying a theory for punching in unrestrained slabs.

The theory used

was Long's<74-) "two-phase approach" which gave the strength of a slab


which fails in shear before the steel yields as;
4-(Ctd)d

0, 42(fcyl) 0

(10Qp) 0

26

(O. 75 + 4-cil>
where c is the side of the square loaded area,

f cv 1

is the cylinder

compressive strength and the other notation is as used previously.


Kirkpatrick et al took the denominator

<which is a correction for the

effect of the ratio ell> as constant at 1, arguing that


variation was small.

the effect of

This is reasonably true for the type of specimens

originally considered by Long, but the value of the denominator for some
of Kirkpatrick's slabs was as high as 1.6 so the stated reason for ignoring
this factor is unsatisfactory.
For fully restrained slabs they argued, by reference to test results, that
the effect of reinforcement was small and they took the term (100p)0 26 to
represent the influence of flexural strength on shear strength.

The value

of p which they used was the equivalent steel area p.; the area of steel
which would be required to give an unrestrained slab the same moment
capacity according to normal flexural theory which the fully restrained
slab had according to restrained strip theory.

The particular theory which

they used was that due to Me Dowell et al<64>, although it appears that
any of the methods described in 3.2.1

could be used.

Because of the

fourth root term, the choice of approach has little effect.


Kirkpatrick
empirical.
punching

appears

to

have

accepted

that

his

approach

was

largely

However Rankin<65) has developed a similar approach, to analyse


at

columns

theoretical basis.

in

flat

slabs,

and

he

attempted

to

give

it

He assumed that failure occurred when the compression

zone failed in shear.

Because compressive stress tends to enhance the

shear strength of concrete, he said that the critical position was at the
flexural neutral axis.

He calculated the shear strength of the compression

zone assuming an elastic stress distribution and a critical section at d/2


from

the

transmitted

face

of

the

across

the

loaded
shear

area.
crack

Then,

arguing

by . aggregate

- 35 -

that

shear

interlock and

was
dowel

forces,

he said that

the

total shear capacity was 2 to 5 times the

capacity of the compression zone.


There are many faults with this as a theoretical analysis.

Firstly, the

shear failure criterion at the flexural neutral axis was based on maximum
principal tensile stress.

This is not really a failure criterion for the

slab at all; it merely suggests that the shear force reduces the neutral
axis

depth,

fact

which

is

well

known

from

research

on

beams <75 ),

Secondly, if <as Rankin said and as the observed behaviour suggests> slabs
fail as soon as the shear crack appears, dowel forces cannot contribute
significantly
behaviour.
that

to

the

ultimate

strength;

only

to

the

post-ultimate

Thirdly, the geometry of the failure mode appears to suggest

there is no shear displacement across the shear crack; it merely

opens up.

Thus the aggregate interlock force must be small as Ghana <75 >

has found for beams.

However, in beams the load continues to increase

after a shear crack appears and Ghana found that the dowel effect was
very significant.

Using his approach, it is possible to quantify the force

for a punching failure.

Because <as Rankin noted> the inclination of the

shear crack means that the failure surface is very long at the position of
the reinforcement, the dowel force in slabs with conventional quantities of
reinforcement is large.

The assumption that this force is realised before

failure occurs is hard to reconcile with Kirkpatrick's observation


assumption> that reinforcement has little effect on strength.

<and

Although

Rankin was a co-author of Kirkpatrick's paper(13>, they appear to have


differed on this point.

Rankin(65) took the dowel force to represent 25%

of the shear strength of a reinforced unrestrained slab.

He therefore

assumed that the shear strength of a restrained unreinforced slab with the
same depth of concrete in compression at the critical section would be 25%
lower.

Kirkpatrick, like Skates <66) in a more recent paper, used the full

shear stress even in unreinforced slabs.

Despite this, differences in their

methods for estimating neutral axis depth make Kirkpatrick's formula

more

conservative than Rankin's for typical bridge deck slabs.


Kirkpatrick said that his formulae gave good predictions for test results
and it is informative to compare his approach with Hewitt's.

Both are

essentially empirical so they can only be compared by comparing their


predictions.

However, since they were calibrated using sets of data which

are not only very similar but which overlap, the .absolute value of their
predictions give little idea of the relative merits of the appro'aches.
- 36 -

As

might be expected, both give reasonably good predictions for typical slab
test results.

A better indication of their relative merits is given by the

predicted relationship between failure load and the important variables


which affect it.

These will now be considered in turn.

a. Loaded Area
Since Hewitt considered a critical section at the face of the loaded patch,
whilst Kirkpatrick considered a critical section at d/2 from the face, it
might be thought

that Hewitt's predictions would be more sensitive to

patch size than Kirkpatrick's.


corrections for

However, Kinnunen and Nylander's empirical

limiting strain and for triaxial enhancement more than

compensate for this.


Taylor and Hayes'<55> results enable the effect of patch size to be clearly
identified.

They suggest that Kirkpatrick's approach is remarkably good in

this respect.

However, including Long's original correction for ell makes

1t significantly worse, suggesting the factor was removed to improve the

results.

Hewit t 's analysis exaggerates the effect of patch size but it is

only with Taylor's smallest patch size

<2h/3) that

the error is really

significant and this is outside the range of c/h ratios which normally
occur in bridge deck slabs.
b. Concrete Strength
Because Hewitt and Batchelor's theory assumes a shear compression failure,
whilst Kirkpatrick et al's implies a shear tension failure,
significantly in their predictions for
Long's

two

phase

approach

gave

they differ

the effect of concrete strength.


square

root

relationship

<and

he

suggested that a coefficient of 0.4 was slightly better) but Kirkpatrick's


method of calculating p. increased this up to fc.., 0
to depth ratios.
theory itself.

75

for very short span

However, it is not clear if this is justified by the

It is generally accepted that such shear failure loads are

proportional to something between fc.., 0


the tensile strength of concrete

(as in BS 8110 and BS 5400) and

<approximately proportional to f c ... 0

Also, although Long's original paper implied that the term p0

25

5 ).

was purely

empirical, Rank in <65) suggested that 1t was used because, for the relevant
reinforcement ratios, the neutral axis depth is approximately proportional
to p0

25

If so, it would be more logical to use the neutrdl axis depth

given by the arching theory <as Rankin did), rather than going indirectly
to an approximate value via a hypothetical equivalent reinforcement area.
- 37 -

This would reduce the predicted sensitivity to concrete strength t o f c:u 0

for s hort s pan to depth ratios and less for longer span to depth ratios.
The shear compression failure mode consi dered by Hewit t and Bachelor might
be

expected

strength.
t his

to give

However,

failure

loads

which

are

proportional

the empirical expression for

sensit ivity.

Despite

this,

the

approach

to concrete

limiting strain reduces


predicts

s ignificantly

greater sensitivity to concrete strength than Kirkpatrick's as will be seen


from Figure 3 .7.
has

nol

Unfortunately, in the publis hed studies concrete strength

been varieJ sufficiently widely or

which is more realistic.

1000
Failure
Load <kN )
900

systematically

to determine

Hewitt

& Batchelor, R = 0.

Hewitt

& Bat chelor, R

800
700

= 0.5

Kirkpatrick et al

600
500
400
300

25

4.0

35

30

Figure 3 . 7:

4.5
50
f .. .... <N!Jnlli2)

Effect of concrete strength

<1 = 1.5m, h = 160mm, c = 320mm, p = O>

c. Reinforcement Area
The most obvious difference between Kirkpatrick's approach and Hewitt and
Batchelor's approach is that the former ignore s reinforcement whilst the
latter considers it .
aspect

of

However, although Batchelor <76) has criticised this

Kirkpatrick's

approach,

saying

that

reinforcement

is

an

"important consideration", his own theory predicts only a small effect for
well

rest rained

For

slabs.

typical

M beam

slab,

1% reinforcement

increases the predic ted strength by around 15%.


A

c urious

feature

reinforcement

is

of

Kirkpatrick's

ignored,

the

approach

prediction

- 38 -

is

is

that ,

affected

by

although
the

the

assumed

effective depth.

Since this is clearly illogical, the author has used a

hypothetical d of 0.75h for all calculations with the approach.


Analysis of the results of tests on bridge deck slabs appears to give
conflicting

evidence

for

the

significance

of

reinforcement

area.

Kirkpatrick <13) obtained virtually identical failure loads with 0.25%, 0.5%,
1.25% and

However both Beal<77) and Ba tchelor et

1.68% reinforcement.

al's <78> results suggest that Hewitt 's approach under-estimates the effect
of reinforcement.

Beal obtained average failure loads, for his model 2,

which is illustrated in Figure 3.8, of 11, 26.6 and 31.1 kN with 0, 0.23 and
0.35% reinforcement respectively.

He also obtained an average failure load

of 26.7kN with 0.35% bottom steel and no top steel.


the

apparent

contradiction

between

Beal

and

The possibility that

Kirkpatrick's

results

was

because small steel areas have a significant effect, whilst increases above
some critical area have no effect, can readily be eliminated by reference
to other tests such as Taylor and Hayes' (55 ).

They obtained a barely

s ignificant difference with 0, 0.9 and 1.8% steel.


Al e
114

82

C2

CJ

C5

D8

DlO

E8

mo. 35%

[1 0. 23% Ub

0 Unreinforced

0.35% t&b

Figure 3.8:

Beal's Model Two(77>

As Beal and Kirkpatrick's results appear to be so contradictory it is worth


considering them further.
bridges,

which

are

Accordingly the author has analysed the two

illustrated

in

Figures

3.8

and

3.9,

using

both

Kirkpatrick's and Hewitt 's approaches.


A major difficulty

in

interpreting

the results

is

that

Kirkpatrick varied the steel area within the same deck.


- 39 -

both Beal and


There are two

objections to this approach.

Firstly the restraint may be different in

different areas of the structure.

This is confirmed by Bea l's results, a s

tests conducted near the centre of t he bridge gave consi stently highe r
results t han t hose conducted near t he e dge.
36~

higher load than E2.

contributes

to

the

For example, 04 f a iled at a

The second objection is that

restraint ,

so

reinforcement

in

reinforcement

adjacent

contribute to the restraint available to the area under test.

bays

may

Again Beal's

tests show evidence of this as test 08, in the unreinforced a rea, was
stopped when the cracking extended into the adjacent reinforced bay, by
2~

which time the load was already

times that at which 010 later failed.

- --

~----..1..-- ----L------~.-!
~----T---- - -r------r----~

A3

C3

03

83

- --~- --- --- ~-----------"l


C----~-------~-----,----1 A5
I
C5
I
05
85 I

- - --j---- ----c ____


- ------1
--- -- --~-----j
------- 1
1 A2
C2
I
02
I
8'2 :
I ----+ - --- - ~--- -- -J - - - - - 1
~----T-----~- -- - -- T ----A4
I
C4
I
04
I
84 I

r -- ---- -r-----I- - - - I A,
I
C1
~--

t--- --

l.

-'"1---- ---- -__- ,_-- - - ..J


---- '

~ --

01

81

-l- - - - - -- .J---- -- -t------ -l

r - - - - ., - -- -- ---=t - - - - - -

A 168"1.

-_1----- -__]_

c- 0 49"z; o0 -25"1. B1 191~

0
Plan

3COO

r-

u6660 0666uSOJu666u
I

9X)

.,.

Section AA
Figure 3. 9:

Kirkpa trick's model ( 13)

Kirkpatrick's a rrangement of bays might be considered unfortunate as the


two lightly reinforced bays were near mid-span, where Beal's tests gave
higher failure loads, and they were surrounded by more heovily reinforced
areas.

Also

the bays were rather narrow relative to

However, by comparing the results for t he


separately

comparing

eliminated but

the

1.68Z and

0.25~

and

1.19% panels,

there is still no trend.


- 40 -

0.49~

the slab s pan.


panels and t hen

both effects can

be

Thus 1t appears that varying

reinforcement genuinely had no significant effect on Kirkpatrick's results.


They also show no sign of the difference between the centre bay (bay 2>
and the edge

(bays 1 and 3) whereas this effect is very significant in

Seal's results.
These differences between Kirkpatrick's and Seal's results could be because
Kirkpatrick's

stiff

concrete

beams,

diaphragms

and

parapet

upstands

provided adequate restraint whilst Seal's deck, with its flexible steelwork
and no upstands, was more dependent on the slab and its reinforcement for
restraint.

However,

restraint

factors

back-calculated

approach are little different for the two decks.

using

Hewitt's

Those for Kirkpatrick's

deck are in the range 0.5 to 0. 7 whilst those for Seal's reinforced panels
are in the range 0.45 to 0. 75.

Thus the greater effect of reinforcement

on Seal's results cannot be explained by lack of restraint, although it


seems likely that the steel girders in Seal's deck did provide less good
restraint.

The high restraint factors observed near the centre of his deck

appear to be the result of global effects which gave the centre portion of
the slab a significant biaxial compression.
Seal

noted

that

recommended

by

Hew it t 's
the

theory,

Ontario

with

code,

gave

restraint

factor

conservative

of

0.5

results.

as
The

predictions approximated closely to his results for the outer portions of


the deck.
but

the

strengths

He does not appear to have analysed the unreinforced sections


author
by

has
factor

found
of

that
up

Hewitt's

to

just

theory

over

2.

over-estimates
This

is

their

better

than

Kirkpatrick's predictions which are unsafe by a factor of up to nearly 3.


Kirkpatrick's predictions for the reinforced areas are slightly higher than
Hewitt's and are thus closer to the averageobserved values.
Clearly the effect of reinforcement on the strength of Seal's slabs was
greater than Hew it t 's theory predicts and much greater than was observed
by Kirkpatrick.

The reasons for this will be considered later.

Seal said that Hew it t 's theory ignores "compression steel" so the top steel
has no effect

on predicted strength.

compression steel but

he did consider

It

is

true that

top steel at

Hewitt

ignored

the support.

The

reason Seal's had little effect on the predicted strength was that it was
very close to mid-depth.

According to Hew it t 's model, mid-depth steel

should have no effect on the strength of a fully restrained slab.

However,

the fact that his theory still predicts no effect in partially restrained
- 41 -

slabs is merely a consequence of the assumption that the actual restraint


forces and moments are proportional to their respective maximum values.

d. Span
The two theories differ significantly in their prediction for the effect of
span to depth ratio.
beam slab.

This is illustrated in Figure 3.10 for a typical M

If Long's original expression for the effect of cl 1 is included

Kirkpatrick's analysis suggests that strength increases with span, which


seems improbable.

This suggests that Kirkpatrick removed it to improve

correlation rather than because the effect is small.

As the expression

was purely empirical this was a perfectly reasonable thing to do.


------ Hewitt

= 0. 7

Failure
1000
Load <kN )
900

& Batchelor

--Kirkpatrick et al

800
700
600
500
1. 0

Figure 3.10:

the

term,

smaller than Hewitt 's.

3. 0
Span (m)

Effect of span

<h = 160mm, f c '"' = 40N/mm 2


Even without

2.5

2.0

1.5

Kirkpatrick's

c = 320mm, p = 0)

predicted effect

of span

is much

However there is a fundamental difference between

the approaches in that Hewitt's includes a check on the moment equilibrium


of the system whilst Kirkpatrick's does not.
two-phase approach <74),

Within the logic of Long's

on which Kirkpatrick's analysis

is based,

it

is

clear that a separate check on bending strength should be made and this
would be more likely to be critical with longer spans.

Kirkpatrick did not

detail this check because he considered it would not be critical in normal


deck slabs.

However, Rankin <65) did detail such a check and this will be

considered in 3.2.4.
Batchelor also implied that his predictions were not
failures.

He

said

that

these

occur

- 42 -

with

low

valid for bending

reinforcement

and

poor

restraint but, like Kirkpatrick, he assumed that they would not occur in
realistic bridge deck slabs.
It is difficult to clearly identify the effect of span from test results.

As with reinforcement area, tests on model bridges appear contradictory.


Kirkpatrick obtained virtually identical results for his two span lengths
<1/h of 9.4 and 12.5 or 7.2 and 10.2 if only the clear span between the
stiff

beams

is

considered)

but

Batchelor<78) obtained

higher load with an 1/h of 13.7 than with 20.7.


varied

the

span within

single model,

an

average

43%

However, since Kirkpatrick

whilst

Batchelor

varied

it

by

testing three and four-beam models of the same width, it seems likely that
Kirkpatrick's longer spans were better restrained than his shorter spans
whilst Batchelor had the reverse situation.
bays,

Kirkpatrick's

analysis

longer spans by some 30%.

Even ignoring the unreinforced

over-estimates

the

strength

of

Batchelor's

It gives better predictions for the shorter

spans although it is still slightly (18%) optimistic for the unreinforced


bays.
Despite the differences in the restraint, and consequent difficulties of
interpretation, a trend can be detected from the analysis of Kirkpatrick,
Batchelor and Beal's results: increasing the span reduces the strength of
unreinforced

slabs

by

more

than

either

increases the effect of reinforcement.

theory

suggests

but

it

also

This could easily be explained if

the failures were flexural rather than shear failures.

In 3.2.1 we saw

that the greater deflections associated with longer spans reduce the area
of concrete in compression and reduce the lever arm at which it acts,
whilst increasing the lever arm at which the steel acts.

This effect is

allowed for by Hewitt's analysis but it is greatly under-estimated because


the

deflection

is

under-estimated.

Hewitt's

analysis

under-estimated

Kirkpatrick's small deflection at failure by a factor of 2 and Beal's, which


were of the order of h/2, by a factor of up to 10.

Kirkpatrick's analysis

does allow for the reduced concrete contribution with longer spans but,
because of the fourth root term, it. under-estimates the effect.

Neither

Hewitt's nor Kirkpatrick's analyses are capable of allowing for another


effect of deflection; it increases steel strain and hence, if the steel has
not reached yield, steel force.
It appears that both theories would become unsafe if they were applied to
slabs,

particularly

unreinforced

slabs,
- 43 -

with

very

large

span

to

depth

ratios.

However, neither are recommended by their originators for use

above a span to depth ratio of 18.


reasonably safe.

Within this restriction they are

Although Seal's slab was well within this llh ratio the

results which fell below the predictions were for panels which had neither
the nominal steel nor the edge stiffening recommended by both Kirkpatrick
and Batchelor.

It

should also be noted

that

both theories correctly

predict that the failure loads of such slabs are so

h~h

that their precise

values have no practical significance; the safety factors suggested by the


tests were in the range 5 to 30.
e. Multiple Loads
Another effect of increasing the slab span is to increase the effect of
the other

wheels of

the HB

load.

Neither

analysis enables this effect to be assessed.

Kirkpatrick's

nor Hewitt's

Kirkpatrick's choice of a

critical shear perimeter at d/2 from the loaded area implies that wheels
spaced by more than 2c
other.

d centre to centre should have no effect on each

However, the empirical nature of the approach makes this dubious

and Kirkpatrick's own tests confirm this: for the longer spans, two wheels
spaced by over twice this distance failed at only 40'4 more total load than
single wheels.

Hewitt 's analysis implies that wheels spaced by less than 1

could affect each other.

This is confirmed by Kirkpatrick's tests.

For his

shorter spans the HB wheel spacingcorresponded to 1.2 land there was no


effect.

For the longer spans the same spacing corresponded to 0.91 and

the effect was very

s~nificant.

However, since the presence of the second

wheel violates the assumption that the system is axi-symmetrical, Hewitt's


approach does not enable the effect to be quantified.
3.2.4 Ductility
Most of the membrane flexural theories considered in 3.2.1 are based on
plastic

theories,

such

as

Johansen's

yield-line

theory.

An

important

assumption of these theories is that the behaviour is ductile.


Reinforcement
although

is ductile,

in reality

there

whilst
is a

concrete is relatively brittle.


continuous

transition

Thus,

from ductile

to

brittle behaviour, the assumption of ductility is normally considered valid


provided that the tension reinforcement yields before the concrete crushes.
This means that sections are considered ductile provided the ratio dc/d
under ultimate moment is less than some critical value.
varies

sl~htly

according to the material properties.


- 44 -

The critical ratio

In the absence of axial forces, the neutral axis depth ratio is a function
of the reinforcement

percentage.

The requirement

for

ductility

reduces to a critical reinforcement ratio which is around 1.2% <79).

thus
This

includes most bridge slabs and nearly all building slabs.


In contrast. to the situation in unrestrained slabs, the theory considered
in 3.2.1 suggests that the neutral axis depth in a restrained slab is a
function of the in-plane restraint.

It

is also fixed relative to the

overall, rather than the effective, depth.

Simple calculations show that

realistic

bridge

requirement.

deck

slabs

almost

never

comply

with

the

ductility

In many cases, calculations suggest that the steel stresses

should still be quite low when the concrete crushes.

This is confirmed

by researchers who have found that such slabs fail in a brittle fashion
before the reinforcement, often even in the critical areas, has reached
yield.

Thus it appears that few of the theories considered in 3.2.1 are

valid in bridge deck slabs.


Building slabs tend to have larger span to depth ratios, relatively poor
restraint and higher effective depth to overall depth ratios.

Thus they

are more ductile than bridge deck slabs and hence their behaviour is
better predicted by plastic theories.

Despite this, calculations suggest

that the behaviour of some of the slabs which have been tested should be
brittle.

This was often supported by the behaviour at failure.

Yield-

line based theories did, however, agree reasonably well with failure loads.
To some extent this was mere coincidence; the theory under-estimates the
strength of strips so there is some margin for inability to re-distribute
the moments.
loaded

by

However, it is significant that all the test specimens were


uniform

loads.

Under

such

loads

the

yield-line

moment

distribution does not differ greatly from the elastic moment distribution
so plastic theories do not make great demande on rotation capacity.

Under

concentrated loads, in contrast, elastic theory predicts local moment peaks


so plastic theory depends on very high rotation capacity.

Because of this,

the theories considered in 3.2.2 tend to over-estimate strengths under


concentrated loads.

This is presumably why uniform loads were chosen to

test most of the theories, although this was not acknowledged.


Amongst

the few studies to acknowledge that, because of this lack of

ductility, yield-line based analyses may not be valid even in uniformly


loaded slabs, are those due to Skates, Rankin and Long<66> and also Niblock
- 45 -

and Long (80).


problem.

They developed a semi-empirical approach to overcome the

Jn this, the moment capacity of the critical section is calculated

as in their analyses considered in 3.2.1.

The relationship between the

failure load and the moment at the critical section is then calculated
using both elastic and plastic theory.

Only if the moment capacity is

zero, is the plastic relationship considered to be directly applicable.

If

the moment capacity is equal to that of a plastically balanced section the


elastic relationship is used.
these

two

extremes,

interpolation

an

according

balanced moment

For all realistic cases, which are between

intermediate
to

the

capacity.

solution

ratio

of

As might

the

is

obtained

moment

by

capacity

linear
to

the

be expected, since this approach

implies that yield-line theory is only valid in unrestrained slabs with


negligible steel areas, the result tends to be slightly conservative.
Skates, Rank in and Long (66) have applied this analysis to slabs subjected
to concentrated loads whilst Rankin(65) used it for flat slabs subjected to
uniform loads.

Because of the great difference between the elastic and

yield-line moment distributions for such cases, they are a severe test of
a simple linear interpolation.

The use of a strip-based method to obtain

the moment capacity is also questionable as there is no reason why the


distribution of membrane forces
throughout

the span.

across a section should be the same

Also, as Rankin acknowledged,

the slab analysis

implies a different support moment from the strip analysis.


perhaps surprising that

It is thus

they obtained a mean ratio of test result to

prediction of 1.16 and a standard deviation of only

10~.

However, to

achieve this, they used Kirkpatrick's approach as an upper limit imposed by


"shear mode failures".
Whilst many brittle bending failures are reported in the literature for
uniformly loaded slabs, few such failures are reported under concentrated
loads.

As there are theoretical reasons for thinking that slabs are more

likely

to

fail

in

bending

before

reaching

their

yield-line

moment

distribution under concentrated than under uniform loads, this may appear
surprising.

It is instructive to consider what such a failure would look

like.
Elastic theory predicts high moment peaks under the concentrated load.
Thus the highest concrete stress occurs in this region,

but here the

crushing stress is enhanced by the triaxial stress state so the first


- 46 -

- - - - - - - - - - - - - - - - - - -

-----

crushing is likely to occur around the edges of the loaded area.

It may

then extend along the potential yield-lines, 1n which case the failure will
However, the area where the concrete first

be described as "flexural".

reaches its crushing stress is subjected to a high shear stress.

Thus, as

it approaches its crushing stress, 1t is liable to fail suddenly under the


combined effect of shear and compression, in which case the load will
punch through the deck.

Thus there is no clear distinction between a

punching shear failure and a brittle bending compression failure so an


alternative interpretation of the failures considered in 3.2.3 is that they
are essentially flexural failures with shear playing a comparatively minor
role.

It has already been noted that some aspects of the test results can

be explained by flexural theory.

It is also clear from the descriptions of

failure that the characteristic conical shear cracks do not appear until
failure.

Thus this interpretation is worth further investigation and it

will be considered in later chapters.


If the failures considered in 3.2.3 were primarily brittle bending failures,

it

provides

another

explanation

for

the

reinforcement on Kirkpatrick's failure loads.

small

effect

of

varying

Unlike the other researchers,

he used the same secondary steel throughout; he varied only the main
steel.

Increasing this did significantly reduce the deflection at failure,

apparently
sections.

due

to

the

reduced

ductility

of

more

heavily

reinforced

With constant secondary steel this implies that the moments in

the secondary direction at failure must have been greater in the more
lightly reinforced panels.

This in turn implies that the distribution of

the primary moments must have been more favourable in the more lightly
reinforced panels and this tended to compensate for the reduced strength.
3.2.5 Serviceabillty
Because of

compressive membrane action,

crack widths,
slabs.

restrained slabs have smaller

deflections and steel stresses than similar unrestrained

Holowka <81 ), Cairns <82) and others have measured steel strains of

the order of a tenth of those predicted by conventional flexural theory


whilst Kirkpatrick <4-9) observed a similar effect on crack widths.

There is

also wide agreement that compressive membrane action delays the formation
of

the

first

crack,

presumably

because

concrete's stress-strain

departs from linearity before cracks become visible.

curve

However, the effect

of restraint on acceptable service load is not as great as on strength.


Because of

this,

nearly all

the

researchers

- 4-7 -

who have considered

the

implications of using membrane action in design have acknowledged that


serviceability

criteria

would

become

critical.

Despite

this,

the

theoretical studies have concentrated almost exclusively on the prediction


of strength.

The design rules which will be considered in 3.2.8 do depend

on the enhancement of serviceability due to membrane action but, in this


respect, they have no quantitative theoretical basis at all.
Hewitt's approach, which was considered in 3.2.4, is one of the few to have
been applied to behaviour at service loads, specifically to the prediction
of deflection.

However it is very unsatisfactory for this purpose.

It is

an axi-symmetrical model whilst, although the failure modes of deck slabs


are approximately axi-symmetrical, the behaviour at service loads is not.
Also the model considers a compressed volume of concrete which is almost
entirely arbitrary except in its area at the critical section.

In view of

these and other faults, some of which were considered in 3.2.4, it appears
that any resemblance between the deflections predicted by this approach
and

those

However,

which

occur

because

in

the

practice

analysis

is

little

more

linear

assumes

than

coincidental.

load-displacement

relationship, whilst the observed behaviour is often highly non-linear, it


does not

under-estimate deflections under service loads as much, or as

consistently, as at failure.
Although

compressive

membrane

action

tends

to

improve

the

ultimate

strength of restrained slabs more than their service load behaviour, there
are situations in which it may be useful at service loads but not at
failure.

Yield-line theory assumes that the full plastic bending moment is

developed

across

wide

width

of

slab.

This

means

that

helpful

compressive force across this critical section can only be developed by


restraint

which is external to the slab, or at

material well away from the loaded area.


appropriate

to service

load

However, elastic theory is more

behaviour and

bending moment under concentrated loads.

least which comes from

this predicts high peaks of


Thus a beneficial compressive

force across these critical areas could be developed by adjacent areas of


less heavily stressed slab.
----~---stresses

could

be

reduced

This means that maximum crack widths and


by

compressive

unrestrained slabs, such as slab bridges.

membrane

action

even

in

This possibility does not appear

to have been considered before, presumably because of the concentration on


strength and the historical development of compressive membrane theory
from yield-line theory.
- 48 -

3.2.6 Restraint

Compressive membrane enhancement depends on the availability of adequate


restraint strength and stiffness.

Thus the prediction of this restraint is

important.

One reason why membrane action has been so little used in

design

the

is

feeling

that

the

restraint

available

to

slabs

is

unpredictable and perhaps unreliable.


Park is one of the few researchers to give restraint the at tent ion it
deserves.

His work considered building-type slab and beam systems and he

tested many nine-panel specimens <83).

When only the centre panel was

loaded,

before

peak

load

was

achieved

provided the restraint) cracked.


the

concrete

restraint.

in

just

the

outer

panels

<which

This shows that the tensile strength of

the surrounding structure contributed greatly to the

Park assumed that this tensile strength should be considered

unreliable for design purposes, as is usual.

Thus, when Hopkins and

Park (50) designed a nine-panel floor system allowing for membrane action,
they provided extra reinforcement in the beams to resist the restraint
forces.

They showed that this steel was heavier than that which they had

saved by considering membrane action in the design of the slab so they


suggested that design using membrane action was uneconomic.

This arises

because building slabs are designed for all bays fully loaded so the same
load case is critical for all bays.

Bridge decks, in contrast, are designed

for moving loads and hence a different load case is critical for each part
of the slab.

This means that the critical area is always surrounded by

areas for which a different load case is critical.

Thus there is always

under-stressed steel available to provide the restraint and no extra steel


is needed.

This means that the scope for economy from using membrane

action in design is much greater in bridges, and other structures which


are designed for moving loads, than it is in buildings.

This is why recent

research into membrane action, including this study, has concentrated on


bridges.
Park analysed his specimens using his strip approach, which was described
in 3.2.1c.
resist

the

He consistently recommended that steel should be provided to


full

restraint -force- but-he-was-less--consistenc -inhls- - --

assessment of the contribution of concrete to restraint stiffness.

In

reference 83 he used only the steel in assessing axial stiffness, but


ignored

lateral

bowing of

the outer slab panels.

Theoretically this

approach should be conservative where there are wide lightly reinforced


- 4-9 -

outer

panels

panels.

but

unsafe

where

there

are

narrower

This is confirmed by the test results.

heavily

reinforced

In reference 50 Hopkins

and Park used gross concrete properties for assessing restraint stiffness
but compensated for this by arbitrarily increasing the axial flexibility of
the loaded panel by a factor of 4.

This shows that the prediction of

restraint flexibility is highly approximate so it is fortunate that, as was


seen

in

3.2.1c,

the

strength of

slabs

is

not

sensitive

to the exact

stiffness of the restraints.


Apart from Park's study, very little work has been done on the prediction
of restraint.

The approach adopted in the Canadian study was to measure

the restraint available; not by direct measurement of restraint stiffness


or strength, but by observing the behaviour of the slab under a load and
back-calculating the "restraint factor" needed in their theory to predict
the observed behaviour.
A disadvantage of this approach is that it is only possible to measure the
restraint available at the time of the test.
prediction means

that

it

is not

The lack of an analytical

possible to predict

restraint which might occur in the future.

any reduction in

In view of the importance,

according to Park's work, of the tensile strength of concrete in providing


the restraint this is significant; cracking due to loads previously applied
in other positions, or to shrinkage,
Canadians

were

aware

of

this

so

could reduce
they

the restraint.

conducted

tests

where

loads <84) or pre-loading to failure <78) had occurred in adjacent

The
cyclic
bays.

This seems to have had little effect.


Hew it t obtained restraint factors for laboratory specimens and models by
back-calculating from the observed failure loads.

However, in the field

tests on full size bridges <8D, it was not practical to test to failure so
the restraint factors were estimated from the deflections at lower loads.
In view of the doubts expressed in 3.2.5 about the validity of Hewitt's
method

for

predicting

deflections,

this

approach

is

less

satisfactory.

However, the results were similar to those obtained from models although
the variation was much greater.
Kirkpatrick assumed rigid restraint which is obviously an unconservative
assumption.

However, since his approach is essentially empirical and was

calibrated with

tests on real structures which had less than perfect

restraint, this is unimportant from a practical viewpoint except that, as


- 50 -

with the Canadian work, there is no way of allowing for possible future
reductions in restraint.
3.2.7 Global Behaviour
Compressive membrane action in bridge decks is normally considered as a
mechanism for resisting local wheel loads spanning between webs.

However,

as was noted in 2.4.3, bridge decks are also subjected to global flange
forces

and

moments.

Since slab

behaviour

is not

linear-elastic

<and

compressive membrane action depends on this non-linearity> the principle of


superimposition

does

not

apply.

Similarly,

because

the

behaviour

of

restrained slabs under concentrated loads is not ductile, it is not safe to


assume

that

global

stressed areas.

forces

will

re-distribute away

from

locally over-

Thus separate studies of global and local effects cannot

prove that behaviour will be satisfactory under combined effects so the


interaction of the effects has to be considered.
A global flange force which is compressive has the effect of prestressing
the slab.

Thus,

unless the stress is so high that concrete crushing

becomes a problem <which is unusual), it improves the behaviour and can


safely be ignored; as it has been by all the previous research.
Tensile flange forces might be expected to have a detrimental effect on
slab behaviour.
simulated

the

Because of this the Ontario study included tests<78> which


support

region

of

continuous

bridge.

The

resultant

tensile flange force had remarkably little effect on the behaviour which
was still entirely satisfactory.

It can also be shown that tensile flange

forces are unlikely to be serious for another reason: the critical design
load case for global flange tension does not impose any local wheel loads
in the critical area.

Thus, when local wheel loads are imposed, any loss

of longitudinal compressive membrane action due to flange forces is more


than

compensated

for

by

reinforcement

provided

to

resist

the

non-

coexistent worst global moment.


Global transverse moments present a more difficult problem.

It was noted

in 2.4 that, in some types of deck, these moments can be even greater than
the local moments predicted by elastic theory.
these

large

moments

in

combination

with

local

It

is conceivable that
effects

could

premature failures in the very lightly reinforced slabs proposed.

cause

Previous

experimental studies, although comprehensive in other respects, have not


- 51 -

investigated this possibility.


bridges
moments.

with

Many of the tests were on steel composite

cross-frames which greatly reduce

the global

transverse

Most of the tests were performed using single wheels and some,

because of propping off the laboratory floor or reacting against the


adjacent beams, did not even model the global transverse moments which a
single wheel could cause.

Those studies which have considered whole

vehicles used loads which were much less severe than HB.
It would be possible to virtually eliminate global transverse moments by

providing intermediate diaphragms or cross-frames.

However, for reasons

discussed in 2.2.1, this solution is unlikely to be economic in concrete


bridges, except in the rare cases when beam and slab bridges are built
entirely in-situ on falsework.

It is more practical in bridges with steel

girders where cross frames are, in any case, often required to provide
restraint to the compressive flange in construction.

Both KirkpatrickC13)

and the Ontario code<11) require such frames to be provided between steel
girders although they say that this is primarily to provide restraint.

Figure 3.11:

HB Whul
load

Compressive membrane action to resist global moments (J3)

Although Kirkpatrick, like all the other researchers, failed to model full
global effects in his tests, his background in British practice meant that
he was more aware of the problem.

He suggested that compressive membrane

action improves the ability of slabs to resist global, as well as local,


transverse moments as shown in Figure 3.11.

This may be true but there

are no tests to prove it and there are several reasons for believing that
the effect is less pronounced.

One of these is clear from Figure 3.11;

resisting global moments requires the slab to effectively span at least


twice as far as resisting local effects.

This doubles the span to depth

ratio which reduces the effectiveness of compressive membrane action and,


as was shown in 3.2.1c, makes it more sensitive to restraint flexibility.
Another reason is that, unlike local moments, global moments act over a

- 52 -

substantial portion of the span length.


restraining

structure

to

structure

This means that the ratio of

requiring

restraint

is

far

less

favourable.
There is also another effect which is likely to reduce the contribution of
compressive membrane action to resisting global transverse moments.

It

was noted in 2.4.3 that the connection between the top flanges of adjacent
beams tends to even out the compressive stresses.
most heavily stressed beam

<the one which most

This means that the


requires support

from

global transverse moments) effectively has a top flange which is wider


than the beam spacing.
moment

equilibrium

required

as

transverse

In order to keep a wide compression flange in

about

the

vertical

shown

in

Figure

3 . 12.

tension

at

mid-span

and

axis,
These

transverse
put

compression

the
at

stresses

are

flange

into

support:

the

top
the

opposite of what is required to develop compressive membrane action.

Figure 3.12:
Although
action

Transverse stresses in a wide compression flange

it seems likely that the contribution of compressive membrane

to resisting global transverse moments

necessarily mean that


unserviceable.

is small,

the slabs themselves will become unsafe or even

Once a slab cracks, its stiffness reduces and the global

transverse moment starts to redistribute away.


deterioration

this does not

in

However, this leads to a

the distribution properties of

increase in the moment in the critical beam.

the deck and hence an

This could be a

problem as

both Kirkpatrick and the Ontario Code use analysis based on


section

properties

Paradoxically,

the

to

obtain

Ontario

Code

the

design

introduced
- 53 -

moments
this

in

analysis

uncracked

the
<which

beams.
is

departure from conventional North American practice> at the same time as


introducing the empirical slab design method.
Theoretically,
design

the problem of worst

value when

conventional

the slab cracks

methods.

However,

beam moment

increasing above the

also arises in decks designed

the

cracked

and

uncracked

by

elastic

stiffnesses differ by a factor of around three compared with around ten in.
very lightly reinforced slabs designed allowing for membrane action.
the effect is much smaller.

Thus

Also, conventional design methods provide a

safe solution according to plastic theory.

Thus, if a beam did start to

fail, redistribution would bring the transverse moments back into play.
There is no guarantee that slabs designed to the Ontario rules, or even
the Northern Irish rules, will be able to act in this way.
3.2.8 Empirical Design Rules
Both the Ontario and the Northern Irish study noted that the available
"theory" for restrained slabs predicted only their strength, which is not a
critical design criterion.
design method.

Thus there was no theoretical basis for a

However, they considered that there was no need for one

either: the observed load-carrying capacity of deck slabs was so high that
simple,

and

probably

very

conservative,

empirical

design

rules

would

suffice.
Batchelor et al<78) noted that tests suggested that unreinforced slabs
would have adequate strength so they initially recommended 0.21 isotropic
reinforcement
AASHT0<45>.

in each

face;

the

minimum reinforcement

recommended

by

This was later amended to 0.3S for reasons which are unclear.

Curiously, the percentage is based on the effective depth: there is no


logical reason why less steel should be required if it is further from the
face.

However, this is a fault which is shared with the minimum steel

rules in many other codes, including BS 5400 and CP 110 but not BS 8110.
The

Ontario

Code

requires

extra

steel

to

be

provided

in

some

circumstances.

The reinforcement is doubled in the end regions of highly

skewed decks.

Also, but only in decks with box girders, reinforcement

designed by normal means to resist global transverse moments is added to


the nominal steel.

This rule is rather odd since these moments can be

just as great, and just as important, in other types of deck.

- 5.4 -

Also, for

reasons noted in 3.2. 7, it still does not prove that behaviour will be
satisfactory under combined effects.
Kirkpatrick gave specific recommendations for only one slab thickness;
160mm.

This was to provide T12s at 150mm which is approximately 0.6%.

The reason for specifying more steel than the Ontario Code was that
Kirkpatrick realised that the reinforcement required to resist calculated
global transverse moments alone could exceed
covered by his rules.

0.5~

in some slabs which were

It is not clear why he specified the same steel in

the longitudinal direction and his rules appear to be unduly conservative


in this respect: the author has designed a deck slab to conventional rules
which had less longitudinal steel.
Both sets of rules require reinforcement for any deck cantilevers to be
designed by conventional methods, which means they are likely to require
substantially more steel than the rest of the deck.

In his own design

Kirkpatrick<13,49> avoided the resulting awkward detailing by not having

any cantilevers at all.

This was an economic solution for his particular

case because the cantilever formwork and reinforcement would have been
expensive compared with the cost of an extra beam;

This would not apply

to longer span bridges and the need to provide extra reinforcement for the
cantilevers is a significant limitation on the advantage of using the
rules.
The major disadvantage of empirical design rules is that there must be
restrictions on their range of applicability.

These will now be considered.

a. Span and Depth


Both Kirkpatrick and the Ontario Code specify a limiting span to depth
ratio of 15 for the use of their empirical rules.
evidence

that

the

theories

on

which

they

Although there is some


are

based

<particularly

Kirkpatrick's) become unsafe by this span to depth ratio, the observed and
predicted strengths of slabs are so high that the limit is conservative.
However,

it

seems

to have been considered that

this was unimportant

because the limit covered normal practice, at least for beam and slab
decks.

This is not entirely logical; the reason shallower slabs are not

used is that

they are uneconomical, or even impossible, to design to

conventional rules because the reinforcement required increases rapidly


with span to depth ratio.

This does not apply in slabs designed to the


- 55 -

empirical rules,
thickness.

indeed

Thus,

the

required

reinforcement

reduces

with

slab

if one was designing a bridge to these rules from

scratch <that is, without the restrictions imposed by using an existing


range of standard beams>,

it appears that

the optimum solution would

always have wide beam spacings and the maximum allowable slab span to
depth ratio .
.The Ontario Code specifies a minimum slab thickness of 225mm but the
Commentary makes it clear that this is not for structural reasons but
because shallower slabs are not advised for durability reasons; in Ontario,
as in many states in the USA, bare concrete decks are the norm. The
restriction on minimum depth, which is not applied in the assessment of
existing decks, has the unintended advantage of limiting the problem of
global transverse moments since the author's analysis shows that theseare
most significant in shallow slabs on close-spaced beams.
The Ontario Code also specifies a

maximum slab span of 3.7m.

This

requires a slab depth of only 247mm, compared with the absolute minimum
of 225mm,

so the range of slab depths which are likely to be designed to

the rules is very narrow.

There is no advantage in using more than the

minimum slab thickness.


A restriction on span is probably justified because longer spans introduce
effects

which

have

not

yet

been

researched;

significant

deadweight

stresses and a much greater interaction between the effect of several


wheels.

However, even with the Ontario Code's allowance for haunches, 3.7m

is a modest slab span by the standards of modern long-span concrete box


girder bridges.
the

Thus the Ontario rules will not be used for these, indeed

limiting span

to depth

ratio

makes designing

them

to the rules

uneconomic anyway as the extra weight would more than cancel out the
saving in reinforcement.

There is scope for economy in the design of this

type of deck from using membrane action, particularly if this could justify
even longer slab spans or shallower slabs than at

present, but

this

requires further research.


b. Restraint
The two sets of rules are very similar in their requirements to ensure
adequate restraint.
are used.

Both require intermediate cross-frames if steel beams

Both require diaphragms at the supports if concrete beams are


- 56 -

used.

Kirkpatrick also suggests the use of concrete support diaphragms

even with steel beams.

Both require parapet upstands or edge cantilevers

to ensure adequate lateral restraint.


The test results suggest that these requirements are sufficient but give
little indication as to whether they are necessary.
Seal's

deck,

which

did

not

comply

with

the

The outer bays of

requirements

stiffening, did show lesser (but still adequate> strength.

for

edge

All the other

decks tested complied with the requirements, as do most of the decks


currently designed.

Some concrete decks have, however, been built without

diaphragms and this has been advocated by Grans ton <85) because of the
costs of forming diaphragms.
3.3 PRESTRESSED SLABS

One

of

the

earliest,

comprehensive

and

and

in

some

influential,

respects

studies

of

still

the

one

effect

of
of

the

most

compressive

membrane action on the behaviour of bridge deck type structures was


conducted by Guyon <10).

His study is worth reviewing even though he

considered only prestressed slabs whilst, for reasons given in 2.2.2, the
remainder

of

this

thesis

assumes

that

bridge

deck

slabs

will

be

Guyon's slab was cast integral with longitudinal and transverse beams.

It

constructed of ordinary reinforced concrete.

was stressed transversely by concentric wires giving a stress of 1.5N/mm"',


whilst

tendons

2.~N/mm 2

located

in

the

beams

gave

longitudinal

stress

of

These stresses are very low, much lower than the longitudinal

stress applied by global effects to many slabs which are not normally
considered as being "prestressed".
A jack was used to apply a single. central concentrated load to each bay in
turn.

It reacted, via steel girders, against the beams adjacent to the

loaded bay of the slab.


Several

conventional

Pucher's<~O>,

Thus only local moments were applied.


elastic

methods,

including

were used to analyse the slab.

Westergaard's <39 > and

The results were reasonably

consistent both with each other and with the initial behaviour of the slab.
The strain gauge readings started to show some
approximately

the . load

for

which

~he

signs-.~f

calculated stress equalled

-measured flexural tensile strength of the concrete.


- 57 -

non-linearity at
the

However, despite the

use of "a powerful microscope", no cracks were visible until the load was
increased by a further 30 to 40%.
Once formed, the cracks extended very slowly in both width and length;
much more slowly than conventional elastic flexural theory would suggest.
Guyon attributed this to a combination of moment re-distribution away from
the cracked region and redistribution of the prestress force towards the
cracked strips, that is compressive membrane action.
confirmed this explanation.

Strain gauge readings

Initially only the central part of the slab

was subjected to a compressive force and tension in the remainder helped


to restrain it.

As the load increased, the area in compression extended

until the whole of the loaded bay was in compression.


Guyon considered that the behaviour was acceptable from a serviceability
viewpoint up to a load of over 211! times that at which the calculated
stress equalled the measured tensile stress, or 10 times the load given by
Freyssinet's no-tension rule.

Removal of the load at this stage caused the

cracks to close up, but this is the one aspect of the behaviour of such a
lightly stressed slab which could be significantly different from that of a
reinforced slab.
With further increases in load the existing cracks grew wider and new
radial cracks developed.

The load was then carried by "a system of

concrete struts", that is pure compressive membrane action.

A brittle

punching failure occurred at a load of some 25 times the 'no tension" load
or twice the load given by Johansen's yield-line theory.
In add it ion to this qualitative description, Guyon developed some simple

analyses.

He

acknowledged

that

these

were

based

on

"debatable

assumptions" and in many respects they have been superseded by more


rigorous analyses such as those given in 3.2 and Chapter 5.
are still useful as

descriptions of

behaviour.

His

However, they

analysis

of

the

behaviour of strips of slab at relatively low loads is largely confirmed by


the form of analysis considered in Chapter 7.

Although it is difficult to

use in any quantitative way, it is significant because none of the more


recent theoretical studies explore the behaviour at low loads.
Guyon extended Johansen's theory to allow for compressive membrane force.
Instead of calculating the membrane force required to maintain lateral
displacement compatibility, like most of the analyses considered in 3.2.1,
- 58 -

he estimated the maximum available restraint force.

His analysis over-

estimated the failure load but he attributed this to the fact that he
ignored the effect of the vertical displacements on the lever arm at which
the restraint force acts.

Back-calculation confirms this explanation.

He

acknowledged that his analysis would not be valid for a slab with a very
large area of surrounding restraining concrete and he attempted, largely
unsuccessfully, to analyse such a case.
Guyon also gave an axi-symmetrical analysis of the punching failure based
on the assumption of rigid lateral restraint.

This assumed that the radial

struts were elastic and uncracked, except at the outer edge and at the
edge of the loaded area, which is analogous to the elastic-plastic analyses
considered in 3.2.1 c.

It also assumed that the force in the struts was

constant over their length.

This implies that there are no circumferential

forces but this was neither mentioned nor justified.

Guyon assumed that,

at failure,

the whole depth of the slab adjacent to the load was in

compression.

This seems unlikely as there is no mention of cracks closing

up in the description of behaviour.

Another fault in the analysis is that

the calculated concrete stress on the critical section at failure is some


130N/mm2 and a very large portion of the slab is stressed up to more than
the elastic limit.

This shows that there were circumferential forces and

this axi-symmetrical analysis appears to be the least satisfactory aspect


of the study.
As a result of the study Guyon developed a simple design method which he
acknowledged to be ''much too conservative".

This was to analyse the slab

using elastic theory but taking Poisson's ratio as zero and taking the slab
to be simply supported.

The resultant mid-span moment is then shared

between the support and mid-span sections and resisted by bending of the
prestressed sections.

These are analysed ignoring the tensile strength of

concrete and the lateral redistribution of prestress, but allowing cracking


to extend to the level of the centre-line of the cable.

If the cable is at

mid-depth of the slab, this gives twice the allowable moment given by the
no-tension rule.

In the case of a simply supported slab it also gives

twice

load.

the design

In fixed-ended slabs the difference is much

greater because Guyon's method allows designers to take as much moment as


they like at the support, whereas conventional elastic theory only allows a
reduction in the mid-span moment of some 15".
uniform

fixed-ended

slab,

Guyon's

method

- 59 -

The result is that, for a

requires

only

36"

of

the

prestress

that

the no-tension

rule requires.

In

haunched slabs

the

difference can be greater.


Although less radical than the Ontario approach, this design method is
more useful in longer span slabs as it requires no limit on span to depth
ratio.

The reduction in stressing force is so great that, despite the

reservations in 2.2.2, it has had a significant effect on the relative


economy of transverse stressing and ordinary reinforced concrete;

This

has meant that many deck slabs mainly <but not exclusively> in France have
been designed using Guyon's rules.

These now represent a very significant

number of bridge-years of satisfactory experience.


Slabs designed to Guyon's rules are very lightly stressed.
crack

long

Thus

their

before the concrete's


behaviour

is

not

They would

compressive stress became excessive.

fundamentally

different

from

that

of

reinforced slabs so the experience of their satisfactory behaviour is


significant to this study.

However, there is one respect in which stressed

and reinforced slabs differ in their behaviour.

Once cracked, a rein'forced

slab's stiffness is greatly reduced for all subsequent applications of


tensile stress, however small, but a
significantly affected when

prestressed slab's stiffness is only

the applied

tension exceeds the prestress.

Thus slabs designed to Guyon's rules may have better restraint than those
designed to, for example, the Ontario rules.
3.4 CONCLUSIONS

Previous research shows that bridge deck slabs are far stronger than
conventional design methods imply.

Slabs designed to the Ontario rules,

for example, have very much less steel yet they have behaved well both in
service and in load tests.
Two "theories" have been proposed which claim to "predict" the ultimate
punching shear strength of bridge deck slabs subjected to wheel loads.
These

theories

are

essentially

sometimes significantly in error.

empirical

and

their

predictions

are

There is also some indication that the

assumption that the observed failures were "shear" rather than "flexunil"
failures could be incorrect.

However, the observed strengths of bridge

deck slabs are so high that these faults have no practical significimce;
typically it is a question of whether the factor of safety is 5 or 7.

In

practical terms, the only questions over the strength of slabs which are
- 60 -

restrained and which are subjected to single wheel loads relate to span to
depth ratios above those for which these theories clatm to be valid.
Although it

is clear that

the restraint available in bridge decks is

adequate to develop compressive membrane action, there is no quantitative


explanation for this.

Similarly, there is no quantitative theory to explain

the observed satisfactory service load behaviour of decks designed to the


empirical rules discussed.

Since service load behaviour is critical in

design, this means that there is no theoretical basis for a design method.
Another

aspect

reinforced

of

slabs

the
which

behaviour

of

has

been

not

bridge
proven

decks

with

very

theoretically

is

lightly
their

performance under combined global and local effects.

This appears to be

been

far

more

serious

omission

since

experimentally either.

- 61 -

it

has

not

investigated

CHAPTER
ELASTIC

ANALYSIS

4.1 INTRODOCTION

Chapters

and

showed

that

serviceability

criteria,

strength, are critical in the design of bridge deck slabs.

not

ultimate

Elastic theory

is more appropriate to the analysis of serviceability than plastic theory


but the elastic theory of restrained slabs has not been developed.
The complexity of the behaviour of realistic slabs, particularly under
concentrated loads, is such that it is not practical to obtain rigorous
analytical solutions, either elastic or elastic-plastic.
considered

in 3.2.1

empirical factors.

Thus the solutions

all contained gross approximations,

assumptions or

It is, however, possible to determine elastic solutions

for simple cases by making reasonable assumptions.

These cases are not

realistic but they do indicate the sensitivity of the behaviour to the


relevant

variables.

Also, by comparison with conventional analyses of

similar cases, they give some indication of the significance of membrane


action in practical cases.

In addition they can be used for checking

computer programs which can then analyse more realistic cases.


Since these simple analyses cannot be used directly in design, there is
little point in considering a wide range of cases.

Thus, in this chapter,

only one simple case will be considered; the unreinforced symmetrical slab
strip which was considered in 3.2.1, subjected to a single central point
load.
4.2 ASSUJIPTIONS

The analysis is based on conventional elastic engineer's beam theory.


Plane sections are assumed to remain plane and compressive stress is taken
to be proportional to strain whilst concrete is taken to have no tensile
strength.

Unlike the analyses considered in 3.2.1, the deflection is taken

to be small relative to the slab thickness but

the validity of this

assumption will be checked.


Although these assumptions are just as arbitrary as those used in the
analyses considered in 3.2.1, this analysis is more rigorous in the sense
that the assumed moment, stress and strain fields are made consistent
- 62 -

throughout the structure.

The analyses considered in 3.2 . 1 either used

different material properties at the critical sections from elsewhere, as


in Park's approach, or, like Me Dowell, only checked that the assumed strain
field

and material properties

were consistent with the forces at

the

critical sections.

4.3 STRESS

Since the assumed slab system has no tensile strength, 1t can only resist
vertical forces by virtue of the vertical component of the restraint force.
It is thus convenient to consider the system in terms of the line of

thrust

of the

restraint

force.

This must

be straight except at

the

supports and the point of application of the vertical load.


Because of the assumptions, the slab cracks if the line of thrust goes
outside the middle third of the section.

Where the slab is cracked, the

line of thrust must act at the edge of the middle third of the effective,
uncracked, sect ion.

This leads to the geometry shown in Figure 4.1.

P/2

Cracked concrete

Figure 4.1:

P/2

Restrained slab strip

<elastic theory>
In Appendix Al it is shown, by consideration of displacement compatibility
assuming rigid restraint, that the depth of concrete in compression at the
supports and at mid-span is 0.222h and the maximum stress in the concrete
is 2.64-P1/h2

Thus, with a concrete stress fc:, the load P

4.4 COMPARISON Wim OTIR AHALYSES

The rigid- plastic analysis considered in 3.2.1a gave the strength of the
equivalent strip as

f c 1 h2 / 1

- 63 -

Using the BS 5400

des~n

rectangular stress block this gives;

Using the.BS 5400 elastic stress limit (0.5fcu> the elastic solution gives;
p

at the serviceability limit state.

The ratio of design ultimate to

des~n

service load is a function of the load factors and in BS 5400(23> it is

Y3 x y,L <ultimate limit state>


Y<3 x Y<L <serviceability limit state>

Considering the case of HB load and load combination 1 <which is usually


critical in deck slabs) this is;

1.1

1.3

1.0

1.1

This means that a section on the limit of the allowable elastic service
stress would have a design ultimate load

which is only 62% of its strength, confirming that the serviceability check
is critical even without allowing for redistribution.
Using the simple BS 5400

des~n

method, the reinforcement required in each

face to resist this load would be approximately 0.6%.


s~nificant

This is a very

amount of reinforcement, confirming that membrane action is

worth considering.
4.5 CRACK WID11IS
Unlike most other crack width prediction formulae, the BS 5400 formula can
be applied to unreinforced concrete.

With the maximum allowable service

load derived in 4.4 the calculated crack width for our case
"'

0.00027h

For a 160mm deep slab this is 0.43mm.

If it is assumed that the maximum

allowable crack width had to be complied with on the surface <which is not
strictly required as there is no reinforcement> the limiting value would be
0.25mm.

However, this does not have to be complied with under the full HB

load; only under 25 units of HB.

The result is that crack widths would


- 64 -

not be a limitation in a deck designed for '5 units of HB but they would
be in a deck

des~ed

for a lower load.

,,6 DEFLECTION

In Appendix A2 it is shown that the analysis given in ,,3 and Appendix A1


leads to a central deflection

0.173 PP/Eh3

for the load corresponding to a stress of 0.5feu this is


"'

4.5 X 10- 5 P/h

Now the membrane force acts at a total lever arm


=

(1 -

"'

0.852h

0.222

2/3)h

For small deflection theory to be valid the displacement has to be small


compared with this.
101. with ,3,6;

The error is 1" with an 1/h of 13.8, 5" with 30.9 and

In practical terms, this means that small displacement

theory is valid for the serviceability analysis of local effects.

However,

if membrane action were used for resisting global transverse moments the

effect of displacements could be significant.


4. 7 I014X:I' OF RESTRADI1' FLEXIBD...ITY
In Appendix A3 the effect of in-plane restraint flexibility is added into

the analysis.

The result is shown in Figure ,,2 by plotting the load for a

stress of 20N/mm2 against the restraint stiffness expressed as a multiple


of the axial stiffness of the uncracked slab.

The elastic analysis is very

much more sensitive to restraint stiffness than the analysis considered in


3.2.1.

Thus restraint is an even more important factor than Chapter 3

suggested.

- 65 -

Load

120

CkN / m)

100
80

60
40

20
0 --~----~----r---~----,-----r----,
0. 125

0. 25

Figure 4.2 :
Cl

0. 5

Effect of restraint flexibility

= 1.5m,

= 160mm,

= lm>

4..8 CONCLUSIONS

The simple analysis considered in this chapter shows;


1. Compressive membrane action is significant in the elastic range.
2. It is less significant than in plastic analysis, hence serviceability
criteria

are

likely

to

be

more

critical

in

design

allowing

for

compressive membrane action than in conventional design.


3. Small displacement theory is valid for considering local behaviour at
the serviceability limit state.
4. The behaviour under service loads is more sensitive than ultimate
strength to restraint flexibility.
Despite the extreme simplicity of the case considered, the various test
results considered in Chapter 3 suggest that all these conclusions are
likely to remain valid for more realistic cases.

- 66 -

CHAPTER
NON-LINEAR

FINITE

ELEMENT

ANALYSIS

5.1 INTRODUCTION

The major difficulty with analysis allowing for membrane action, at least
as far as flexure is concerned, is not conceptual; it is the complexity of
the mathematics.

This fact,

which is clear from 3.2.1

and Chapter '

suggests that the subject should be amenable to solution by numerical


analysis and the finite element method is the most convenient way of doing
this.
The analysis of membrane action has to be non-linear; even the simple
analysis considered in Chapter 4 is only linear under proportional loading.
Non-linear finite element analysis, NLFEA, is only practical with powerful
computers so it is a comparatively recent method which was not applied to
concrete until the 1960's <86>.

Despite this the literature is extensive

and, although only a tiny fraction is aimed at the analysis of membrane


action, much of it is relevant.
work in detail.

It is thus not possible to review all the

This chapter aims only to introduce the principles and

problems of the method.

A particular, relatively simple, form will be

considered in more detail in Chapter 7 whilst readers requiring a more


comprehensive coverage of the state of the art

should consult recent

specialist works such as reference 87.


5.2 GENERAL APPROACH
The analytical method adopted should be capable of resolving all
problems identified in Chapter 3.

the

Two of the most important of these, the

prediction of restraint and the analysis of global effects, require the


analysis to consider the whole bridge.

Even with very powerful computers,

this puts a severe restriction on the form of analysis which can be used.
In

this

study,

therefore,

only

the

"smeared

layered approach" will be considered.


modelled;

the

cracks

infinitesimal cracks.

are

smeared

crack,

distributed

steel,

In this, individual cracks are not

out

into

an

infinite

number

of

Similarly, individual bars are not represented; the

steel is distributed evenly across the element width.

The significance of

layering is that it enables the material state to be varied over the


element depth whilst still using a two-dimensional element.

The stresses

are calculated independently for each layer as a function of the strains


- 67 -

which are calculated from the displacements at the "reference plane"; the
level at which the elements are tmplicitly located.

The element forces are

then calculated by integrating the stresses over the element volume.

Thus

the forces are calculated directly from the displacements but the correct
displacements can only be obtained from the loads by an iterative solution
scheme.
Although linear-elastic analyses of slabs, particularly for bridge design,
are often performed using alternative structural idealisations, such as
grillage analogy, non-linear analysts have assumed it necessary to use
plate finite elements.

This is because of their desire to produce rigorous

and accurate analyses.


5.3 ELEMENT TYPE
5.3.1 Slabs

Early finite element analyses of slab systems used classical thin plate
theory which assumes that
normal.

This

appr:oach

lines normal to the reference plane remain


is

being

"gradually

supersededn<87)

by

the

Mindlin <88) form which assumes that lines normal to the reference plane
remain

straight

but not

necessarily

normal.

This

enables

shear

deformations to be included in the analysis so the theory is sometimes


described as "thick plate theory".
in reinforced concrete and
straight

prevents

the

However, shear causes diagonal cracks

the assumption

realistic

that

vertical

modelling of shear

lines

failures.

remain
Indeed,

according to Chana<75), shear failures are sensitive to dowel behaviour at


the crack which tmplies that they cannot be realistically modelled by any
form of smeared crack, distributed steel analysis.

Despite this, Mindlin's

theory does give more realistic predictions than classical theory for the
shear forces at a free edge, as has been illustrated by Cope <89).
1t

appears

that

the main reason

for

adopting

it

However,

is one of analytic

convenience; it requires a lower order of displacement continuity across


the element boundaries <87 ).
The

nodal

forces

in

the

elements,

due

calculated using the virtual work approach.

to

nodal

displacements,

are

To do this it is necessary to.

assume a displacement field for the whole element from the known nodal
displacements.

A wide variety of elements can be developed, according to

the number of nodes, the displacement field assumed and the method of

- 68 -

integrating

the

stresses

and

strains

over

the

element

volume.

discussion of their relative merits is outside the scope of this study.


5.3.2 Beams

The beams in beam and slab decks can be modelled using either simple beam
elements or an assemblage of plate elements.

The latter is far more

expensive but 1t enables inclined web cracking and the transverse bending
stiffness

of

the

beam

to

be

modelled.

Edwards(36>

found

the

two

approaches gave very similar results in a bridge with rectangular beams.


However, the transverse bending stiffness of the flange of, for example, an
M beam is much greater so there may be more advantage in using plate

elements for these.

Buckle and Jackson<90> have developed a form of beam

element which can model transverse bending.

However, because 1t assumes

that plane sections remain plane, 1t cannot model the warping stresses
which contribute to the resistance to torsion.
The beam elements are rigidly attached to the plate elements at the nodes.
Since the mesh size is decided by the requirement to model the local slab
behaviour, 1t is smaller than is required to model the beam behaviour.
Thus the analysis is not sensitive to the type of beam element used.
Buckle and Jackson<90) used a displacement function which will be shown in
Chapter 7 to have serious faults, whilst Edwards <36> used a displacement
function

which

was

not

consistent

with

that

used

for

the

slab.

Calculations suggest that neither of these faults had a significant effect


on the results.
5.4 MATERIAL PROPER'IIES
5.4.1 Steel

The reinforcing and prestressing steel is assumed to be fully bonded to


the concrete and to exhibit uniaxial behaviour; that is, it is stressed only
by strain in the direction of the bars.
Any stress-strain relationship can be defined numerically and incorporated
into a program but it is more usual to use elastic-plastic properties,
sometimes with linear strain hardening.
prestressing,

departs significantly

from

Modern reinforcement, and all


this

advantages in using more realistic properties.

- 69 -

assumption so there are

When unloading is considered, it is usual to use a straight line parallel


to the inif:ial portion of the stress-strain curve which gives a permanent
set equal to the previous departure from linearity.
5.4.2 Concrete
a. Uncracked
The

enhancement

compression,

and

of

concrete's

compressive

strength

due

to

the reduction due to orthogonal tension,

biaxial

is normally

considered but the effect of vertical stress cannot be modelled in the


form of analysis considered here.

Despite this, the enhancement can be

significant; up to approximately 20%.


A variety of stress-strain relationships have been used.

Abdel Rahmen<87>

used a simple elastic-plastic relationship with a straight cut-off at a


limiting

strain whilst

Edwards<36>

used

Popovics'

formula(91)

for

the

uniaxial case in beams and Nilson's<92> approach for the biaxial case.
As with steel, unloading is usually modelled with a straight line.
unloading part of the properties are sometimes
under

monotonically

increasing

loads

The

specified even in analyses

in order

to avoid

deformation theory which was mentioned in 3.2.1 b.

the

fault

of

this is done, the

maximum strains have to be stored for all the sampling stations.


The variability of
analysis.

concrete

is a

major difficulty

in a

deterministic

This variability is particularly significant to failures, such as

punching failures, which are affected by local rather than average concrete
strength.

The effect

stress-strain

curves

is

large compared with

and

this,

relationships are used many

combined

the difference between

with

thousands of times

the

fact

that the

in the course of an

analysis, encourages the use of simple relationships.

It also means that

the predictions are unlikely to be precise.


b. Cracked
Although

smeared

crack

infinitesimal spacings,
centres.

analysis

implies

infinitesimal

cracks

real structures have discrete cracks at

at

finite

The concrete between the cracks is able to resist tension and

this stress contributes to the stiffness of the structure.

This effect,

which is known as "tension stiffening, is very significant, particularly in


lightly reinforced elements and at

low loads.

- 70 -

It

is modelled by an

empirical stress-strain curve which has a descending branch after the


concrete has cracked.
Cracks first form in the direction of the maximum principal tensile strain.
If this direction subsequently changes, a shear stress is developed across

the crack.

The shear stiffness is reduced by the crack and can be

modelled by another empirical factor called a "shear retention factor".


However, even with this reduced shear stiffness, the analysis can imply a
tensile stress

in other directions

which exceeds

the cracking stress.

Cope et al<93l used an alternative approach in which the "crack" direction


rotates to follow the principal strain direction.

This appears to give


It

may

formed

but

better results in cases where. the rotation is significant (94).


appear

illogical

that

cracks

can

rotate

after

they

have

presumably the explanation is that the crack direction in a smeared crack


analysis represents only the average or active crack direction so it can
rotate as new cracks form.
The few analysts who have considered unloading in cracked concrete have
used widely different assumptions (95,96,97> reflecting the lack of data in
this area.
Whatever tension stiffening function is used, the predicted behaviour is

very sensitive to the assumed cracking stress.


even

wider

random

variation

than

In addition to having an

compressive

according to strain rate, strain gradient, curing


number of load repetitions and many other factors.

strength,

this

varies

load duration,

r~gime,

This makes accurate

deterministic predictions of behaviour impossible.


Unlike

the

non-linearity

due

to

reinforcement

yielding

and

concrete

crushing, that due to concrete cracking is significant under service loads.


Thus it is the only non-linearity which is important to the design of
structures for which serviceability criteria are critical.

Also, even in

strength analysis, it is not realistic to consider only the cracking due to


a single monotonically increasing load case since cracking could have been
caused by many different service load cases.

In particular, cracking due

to wheel loads previously applied in other positions could reduce the


restraint

available

considered.
particularly

Thus
for

to develop

membrane

action

under

the

case

the lack of an agreed tension stiffening

unloading,

is a

serious obstacle to

being

function,

the use of the

analysis in design so the subject will be considered further in Chapter 6.


- 71 -

5.5 APPLICATION TO MEMBRAHE ACTION


Several analysts have applied NLFEA to slab systems.

Most have considered

only monotonically increasing loads, which restricts the application of


their analyses, but they do give a useful insight into behaviour.

Several

analysts have claimed good predictions for the behaviour of restrained


slabs but some of these might be considered slightly surprising.

For

example, Jackson <98> obtained good predictions for Roberts' tests <53) but
his

analysis

used

small

displacement

theory

and

simple

calculations

suggest that including the effect of the observed <and predicted> large
displacements would have reduced the

predicted strength by some

20~.

Despite these doubts, non-linear analysis has proved better able to predict
the behaviour of complicated slab systems than other methods.

At service

load levels, the predictions for complicated structures actually appear to


be better than those for simple "fully restrained" laboratory specimens.
The reason for this appears to be a fault in the tests rather than the
analysis;

it

is

difficult

to develop

full

restraint

and

service load

behaviour is very sensitive to restraint as was demonstrated in 4.7.


It is not practical or necessary to consider all these analyses in detail

but it is useful to consider a particularly relevant example; that of Cope


and Edwards<99).

They analysed several of the tests which were considered

in Chapter 3, including those of Kirkpatrick.

In view of the inherent

variability of results which are sensitive to local concrete behaviour,


they

considered

their

predictions

to

be

good.

However,

Kirkpatrick

produced enough results to enable the variability to be estimated and this


appears to be remarkably small and certainly smaller than the discrepancy
between the test results and the non-linear analysis.

Despite this, the

analysis is reasonably good with the worst error in the failure prediction
being some 30% with 15% being more typical .. This may not sound that good
compared with Hewitt's or Kirkpatrick's "analyses" but they are largely
empirical

whilst

the

non-linear

analysis

obtained

the

restraint

and

strength only from the geometry and material properties of the specimen
with no empirical correct ions.
The brittle nature of the "punching shear failures was ale~ correctly
predicted even though the analysis is incapable of modelling shear.

This

appears to confirm the suggestion in 3.2.4 that such failures are primarily __
brittle bending compression failures although the analysis did tend to
over-estimate strength slightly, implying that the high shear stress in the
- 72 -

critical region reduced its compressive strength.

In practical terms, the

good predictions for stresses and deflections at lower loads are more
significant as they suggest that the approach is valid for the critical
serviceability analysis.
Cope and

Edwards'

study suggests that

NLFEA is able

to successfully

predict restraint and analyse the local behaviour of bridge deck slabs
allowing for membrane action.
of

global

considered

and

local

this.

In theory it can also model the interaction

moments

He

but

analysed a

only

Edwards(36>

hypothetical

appears

to

have

bridge with rectangular

reinforced concrete beams and with deck slab reinforcement designed to the
empirical rules considered in 3.2.8.
reinforcement

would

His analysis suggested that this

be over-stressed under combined global and

moments, confirming the doubts expressed in 3.2.7.

local

However, because of the

lack of test data, there is no proof that the analysis was realistic in
this respect.

The form of deck he considered was also unrepresentative of

modern practice since ordinary reinforced beams are rarely used and the
moment redistribution behaviour would be very different with prestressed
beams.
5.6 USE IN DFSIGN
Although NLFE has proved capable of predicting the behaviour of reinforced
concrete slab structures,

it has rarely

(if ever> been used in their

design; either directly or for validating simpler design methods.


and Kotsovos<lOO> have said

'~he

Bedard

main reason for this appears to be a lack

of agreement concerning the numerical description of material behaviour".


This reason is supported by 5.4.2b but, in the case of the slabs considered
here, it is not a sufficient reason; the analysis would still produce more
economical designs

than conventional methods if the most conservative

conceivable material properties were

used.

Thus there must

be more

fundamental reasons and these will now be considered.


a. Cost and Complexity
A non-linear analysis of a given structure with a given element mesh is at
least an order of magnitude more expensive in computer time than the
equivalent linear analysis.

Also, because the principle of SUPerimposition

does not apply, every load combination has to be analysed separately.


Similarly, global and local effects cannot be superimposed so the whole
structure has to be analysed with a fine enough mesh, at least in the
- 73 -

critical areas, to model local behaviour.


computer

time

is

several

orders

The result is that the cost in

of

magnitude

analytical methods considered in Chapter 2.

higher

than

for

the

More seriously, the analysis

is also much more expensive in engineer's time.

This discourages its use,

particularly under design fee competition.


A related disadvantage, which is perhaps more serious, is the conceptual
difficulty; NLFEA is difficult for the ordinary designer to fully understand
or control.

This

makes

it

potentially dangerous

as

(at

least at

the

present state of the art> NLFEA is neither fully automatic nor foolproof.
b. Load History Dependence
For

reasons

which

were

discussed

in

2.3.4

and

5.4,

the

behaviour

of

concrete structures, and hence the realistic analysis of such structures,


is

load

history

dependent.

Since

it

is

impossible

to

predict

and

impractical to analyse the load history of a bridge over its entire design
life, this could be a serious problem.
c. Incompatibility with Codes
Existing

codes

methods

in

of

practice

mind.

If

were

the

written

critical

with

design

conventional
criteria

were

analytical
clear-cut

fundamental requirements, such as ultimate strength, this would not be a


major problem.

However, Chapter 2 showed that the fundamental critical

design criterion for bridges is the very ill-defined one that they should
remain "serviceable" for their design life.

It is very unclear what this

means in non-linear analysis terms, except that it appears to confirm that


a whole life analysis is required.

5.7 CONCLUSIONS
Non-linear

finite

element

analysis

is a

powerful analytical

tool which

sheds some light on the fundamental behaviour of slab systems and which
can give reasonably good predictions for their behaviour.

The reported

analyses support the suggestion in 3.2 .4 that "punching shear" failures may
be primarily

flexural.

One also appears

global behaviour expressed in 3.2.7.

to confirm the doubts about

There are, however, major difficulties

in using the analysis in design.

- 74 -

CHAPTER
TENSION

STIFFENING

6.1 INTRODUCTION

In Chapter 5 it was noted that

the lack of an agreed expression for

tension stiffening is a serious obstacle to the use of NLFEA.

Tension

stiffening is particularly significant in lightly reinforced elements and at


low loads; that is, at service load, rather than at failure.

Since bridge

deck slabs designed using membrane action are very lightly reinforced, and
since

serviceability

criteria

are

critical

in

stiffening is particularly important to these.


structures,

tension stiffening

may still be

their

design,

tension

Also, unlike in most other


important at

higher

loads

because of its contribution to the restraint.


Cope et al <37)

Stress
1.0
Cr ac ki ng Stress

Gilbert and Warner

( 101 )

0. 5

15

10

Stra in/Crac king St rai n


Figure 6.1 :

Tension stiffening functions

The tension stiffening functions used in non-linear analyses are purely


empirical.

Many such relationships have been used,

illustrated

in

fundamental

Figure 6.1.

analytical

method,

stiffening

function

appears

analytical

methods

depend

properties.

However

The dependence of
to

NLFEA,
be a

ultimately

major
on

apparently rigorous

totally

in not

empirical

weakness.

empirically

tension stiffening differs from,

tensile strength of reinforcement


property.

on

an

two of which are

being a

tension

Admittedly,
derived

material

for example,

the

fundamental material

It is a property of the composite material, reinforced concrete,

or even of the structure, not of the concrete or reinforcement.


therefore

all

decided

to

investigate

this

subject

at

slightly

It was
more

fundamental level than is justified by the relatively simple analytical

- 75 -

method subsequently adopted in Chapter 7.

The tests reported in this

Chapter were also used to calibrate the analysis used in Chapters 7 and 9
as well as to investigate the effect of scale in the half scale models
considered in Chapter 8.
6.2 TIIEORY
6.2.1 Mechanisms
The tension stiffening functions used in non-linear analysis represent
stress which is transmitted to the concrete between cracks by two, or
perhaps

three,

mechanisms.

The

reinforcement and concrete.

first

of

these

is

the

bond between

This enables some of the force, which is

carried across the cracks by the reinforcement, to transfer to the concrete


between the cracks.

The second mechanism, which only applies to sections

in flexure, is the shear connection between the compression zone and the
teeth of concrete between the cracks.

The third mechanism which affects


Mart he <102 >

tension stiffening is the ductility of concrete in tension.

has shown that even when the strain exceeds that at which- the peak stress
is developed and

cracks

significant tension.

have

started

to

form,

concrete

can transmit

However the effect is often ignored, which may be

justified as the stress is only significant over a narrow range of smeared


strains.
Consideration

of

these

mechanisms

might

suggest

that

particular

empirical expression for tension stiffening stress would only be valid in a


narrow range of circumstances.

The bond contribution, for example, might

be expected to be sensitive to the bond characteristics, size, quantity and


orientation

<relative to the cracks) of the reinforcement.

In practice,

however, many non-linear analysts have obtained satisfactory results using


the same function in a wide range of circumstances.
this apparent

An explanation for

paradox can be obtained by considering the way cracks

develop in a region of constant moment or constant direct stress.


Prior to the formation of the first crack, the bulk of the load is taken by
the concrete <Figure 6.2a).

As the stress approaches the effective tensile

strength of the concrete, f.:t a crack forms at the weakest point.

Here

most of the stress is transferred to the steel but beyond a distance, S0


the stress is unaffected <Figure 6.2b).
cause another crack to form.

A further increase in load will

This cannot occur within S0

crack because the stress is too low.

of the first

Finally, when all the cracks have

- 76 -.

formed <Figure 6.2c>, no two adjacent cracks will be more than 250

apart

because otherwise there would be a section between them subjected to a


stress in excess of
spacing becomes 1.350

Beeby (34) has shown that the average crack


.

Stres~

Cracki ng
------------------------..-- Stres~

----------------------a ) Before First Crac k

n\V
So

----------

b> After First Crack


--cracking Stress

c> After Last Crack

Flgure 6.2:
This description

Stresses 1.n concrete as cracks develop

implies that

anything which improves the transfer of

stress to the concrete on either side of the cracks reduces the final
crack spacing.

It thus reduces the wavelength of the stress distribution

shown in Figure 6.2c, but has no effect on either the amplitude or the
average value, which is the stress used in smeared crack analysis.
The above description can be used to obtain an estimate for the tension
stiffening when all
necessary
cracks.

the cracks have first

to assume

Vetter <103),

shape

for

formed.

To do this

the stress distribution

in an analysis

it is

between

intended for a different

the

purpose,

assumed that the concrete stress increased linearly either side of the
crack.

He also assumed that the rate of increase of stress either side of

the crack was unaffected by the formation of further cracks.


deduced that the average stress was approximately 0.5fc t

From this he
However, this

was based on the incorrect assumption that the average crack spacing was
2.050

Using Beeby's crack spacing of 1.350

becomes 0.3350

the average concrete stress

It has often been assumed that a further increase in strain reduces the

tension stiffening stress but it is not clear why it should.


where

bond

inconsistent

is

the

with

dominant

the

usual

mechanism,
design

the

assumption

assumption

- 77 -

that

For example,

appears

to

be

bond strength

is

independent of strain up to at least the yield strain of the reinforcement.


It

may be argued,

therefore,

that

the smeared stress

should remain constant after cracking.


code

of

practice

formulae<104>

and

in the concrete

This assumption is used in some


is

supported

by

some

researchers

including Hartl <105>.


6.2.2 Steel Stress
An unfortunate consequence of smearing the concrete strain is that the
steel strain is also smeared.
occurs

at

the

cracks,

is

This means that the peak steel strain, which


not

modelled

reinforcement yields is over-estimated.

so

the

load

at

which

the

This has not previously been a

serious problem because the tension stiffening functions used have meant
that the effect became insignificant well before the reinforc'ement became
non-linear.
as

If, however, a constant tension stiffening stress were used,

suggested

in

6.2.1,

the

problem

would

become

more

serious.

Cervenka <106 > avoided this by calculating the steel strain independently,
ignoring tension stiffening.

This approach introduces the reverse error;

that is, it is assumed that all the steel is subjected to the strain which
only really occurs at the crack position and thus the non-linearity is
over-estimated.

It appears that it would be more correct to use some form

of averaging process between the strains <or stresses> calculated with, and
without, allowing for tension stiffening.

However, this would be even more

inconvenient than Cervenka's approach.


Because of these problems, an analysis using one of the tension stiffening
functions shown in Figure 6.1 could give better results than an analysis
using a constant tension stiffening stress, even if the constant stress is
more representative of the real behaviour of the concrete.

This, and the

tendency of researchers to concentrate on behaviour at high loads, could


explain the preference for the type of tension stiffening function shown
in Figure 6.1.
6.2.3 Mesh Dependence
The formation of a crack affects the stress over a distance which is
related to the final crack spacing, but in a finite element analysis it
affects the stress over a distance which is related to the element size.
Because of this it has been suggested that the tension stiffening function
should be varied with the element size so that the energy released by a
crack is independent of the mesh.

However, this was not done in the

- 78 -

analysis considered here and it was found that the results were entirely
independent

of mesh size.

This was because only regions of ,:onstant

moment or constant direct

tension were considered so all the elements

cracked simultaneously, unless a variable tensile strength was used.

Thus,

when a fine element mesh was used, the resulting under-estimate of the
energy released at the position of the real cracks was compensated for by
the over-estimate of the energy released elsewhere.

Consideration of this

behaviour shows that this would not occur in a region of varying moment.
Ideally, therefore,

the tension stiffening stress should be varied both

with element size and with the stress state in the adjacent elements.
This would be very difficult to do in a general solution procedure so it
is fortunate that experience shows that, unless the mesh size is small
compared with the crack spacing, a constant function can be used.
of

the

variability

additional

accuracy

significance.

of

tensile

obtained

strength

from

and

finer

tension

mesh

In view

stiffening,

would

have

no

the
real

The problem does, however, prevent the use of smeared crack

analysis in the study of behaviour which is very local compared with crack
spacing.
6.2.4 Cyclic Loading
The contribution of tension stiffening tends
loading.
zero.

to reduce under repeated

Indeed the crack width clauses in BS 54-00 assume it reduces to

Cope and Rao<107) modelled the reduction by reducing the length of

the tail of their tension stiffening function, leaving the value of the
tensile

strength

unchanged.

This

approach

cannot

be

constant tension stiffening stress suggested in 6.2.1.

used
It

with

the

implies that

cyclic loads reduce the tension stiffening stress but do not cause any new
cracks.

However,

it

is known that concrete is susceptible to fatigue

failures in tension, indeed the conventional design methods for concrete


pavement (108> are based on quite well established fatigue relationships.
Thus an alternative way of modelling the effect of cyclic loads on tension
stiffening would be to reduce the tensile strength used in the tension
stiffening expression but to leave the form of the expression unchanged.
6.2.5 Unloading
The

bulk

of

research

into

both

tension

stiffening

concentrated on monotonically increasing loading.

and

NLFEA

has

This means that

the

tension stiffening functions assume that the tensile strain currently being
- 79 -

In a

experienced is the greatest the concrete has ever experienced.

complex non-linear structure, this assumption may not be valid even when
the structure itself is experiencing a monotonically increasing load.
importantly,

for reasons discussed in 5.4.2b,

it

is not

More

reasonable to

consider only monotonically increasing loads in the type of structure


considered in this study.
Once concrete has cracked, it never re-acquires its tensile strength.

Thus

the tensile properties of cracked concrete are not reversible; a separate


unloading curve is needed.

It seems reasonable that once the crack has

fully closed the compressive stiffness of the concrete will be largely


unaffected by the crack.

It remains only to decide the stress required to

close a crack and the amount of strain, if any, which becomes permanent.
Although, in reality, the unloading curve may have a complex shape the
other errors in the analysis and variability in the behaviour mean that
the use of such a

function

relationship will be used.

cannot be justified.

A simple bi-linear

Unfortunately, at present the values to be used

in this relationship can only be obtained empirically.


A variety of expressions have been used.
Bazant <97), have used a straight
cracks

close

completely

at

Some researchers, such as

line to the origin implying that the

zero

stress.

At

the

other

extreme,

Crisfield<95) used a straight line parallel to the initial, linear, part of


the stress-strain curve, implying that the cracks do not close at all.
This

seems

Cope<96>

extremely

used

the

corresponding
permanent.
data.

to

unlikely,

more
that

particularly

reasonable
at

which

if

assumption

the

concrete

the

cracks

that

only

first

are
the

cracks

wide.
strain

becomes

The wide range of these expressions indicates the lack of

However many structures, because of relatively heavy reinforcement,

or only monotonic loading, are insensitive to the assumptions.


6.3 AHALYSIS OF PREVIOUS TESTS
6.3.1 Direct Tension Tests

Some analysts have derived their tension stiffening functions by obtaining


the

best

structures.

fit

to
This

the

load-displacement

approach

is

not

very

response

of

satisfactory

quite

complex

because

tension

stiffening is only one of many factors which affect the response.

Thus

tension stiffening functions are liable to become "fiddle factors" which .


compensate

for

a wide variety of errors in the analysis.


- 80 -

A better

approach,

which has

been

used

by Cope et

al <37 ),

is

to

test simple

statically determinate beams with long constant-moment regions.


these specimens, however,
displacement

it

relationships

Even with

is possible to obtain very similar load-

with

different

tension

stiffening

functions.

Only tests which subject the whole specimen to the same smeared strain,
that is direct

tension tests, give unambiguous results.

Unfortunately,

because there are theoretical reasons for believing that tension stiffening
could be different in flexure and direct tension, direct tension tests
cannot be used as the sole basis for deriving tension stiffening functions.
However they do give some useful information.
Williams <109) has tested a series of fifteen large slabs in direct tension.
The response was approximately linear until the first crack appeared.

This

occurred at a stress of 0.5 to 0. 7 times the tensile strength of the


concrete as measured by the split cylinder test.

This difference between

the effective tensile strength and the split cylinder strength is partly
due to the random variation of the tensile strength of concrete; cylinders
are constrained to fail on a pre-defined plane whilst a slab is free to
crack at its weakest section.

Statistical analysis suggested, however, that

this alone could not explain the difference.

The remainder was presumably

due to restrained differential strains which had a greater effect on the


slabs than on the cylinders.

A restrained strain equal to only- some 5% of

the total likely shrinkage is sufficient to explain the difference so it


could be due to differential shrinkage across the section.
After the first

crack appeared, the extension increased rapidly with a

relatively slow increase in load.

However, tension stiffening remained

significant even at a load such that the steel behaviour was non-linear.
All the tension stiffening functions previously used by non-linear analysts
under-estimate this effect; indeed most <including both of those shown in
Figure 6.1> have no effect at all on the results of an analysis performed
under load control, as can be seen from Figure 6.3.
The specimens were re-analysed using a constant tension stiffening stress
as suggested in 6.2.1.

This analysis under-estimated the load in the slabs

at extensions just above that required to cause the first crack.

This

could suggest that the stress between existing cracks was higher at low
strains but it seems more likely that it was because of the variation of
the tensile strength of the concrete; that is, because not all the cracks
- 81-

developed at the same load.


split

into

several

equal

To investigate this, the computer model was


elements

distribution of tensile strengths.

and

these

were

given

normal

The number of elements used was ten,

which was approximately the final number of major cracks.

This approach

implicitly assumes that the strengths of the ten elements are independent
variables, whi ch is not strictly correct, but it does give a good indication
of the effect of concrete variability.
Force

Only
----Analysis <displacement control)
---Analysis <load control)
Strain
Figure 6.3: Effect of tension stiffening on
analysis in direct tension

Force
400
Steel Area
Limit of Linearity

<Nimm2

-------

300

-~----

--------------)(

Test
Analysis

-Steel Only

0. 2

0.4

0.6

0.8

1.2

1.0

1.4

Strain x 108
Figure 6. 4:

Analysis of Williams' Specimen 1

<1% steel>
Analysis with a coefficient of variation equal to that obtained for the
split cylinder tests (10%) gave results such as those shown in Figure 6.4.
To

obtain

this

excellent

relationship,

however,

the

average

tensile

strength used in the analyses had to be adjusted for each specimen.

In

the more normal situation, where this cannot be done, the best estimate
for

the

effective

average

tensile strength of

approximately 0.8 times the split cylinder strength.


- 82 -

the

concrete would

be

The actual range used

was 0.7 to 0.9.


effective

The effect of small errors in the value used for the

tensile

strength

of

concrete

is

large

compared

with

the

difference between the analysis and test results shown in Figure 6.4 so
there

is

no

practical advantage

in

making

further

refinement

to

the

tension stiffening function.


10~

Although a coefficient of variation of

gave good predictions for the

load displacement response, a higher variation and a skewed distribution of


strengths were needed to make the analysis model the actual development
of the cracks.
tension

Analyses which did this exaggerated the rate of decay of

stiffening with

increasing strain.

In

the

tests,

developed with no discernible effect on tension stiffening.

new cracks

This seems to

suggest that the stress in the concrete between cracks, even cracks which
are within 250

of each other, increases with strain.

This gives further

confirmation of the suggestion in 6.2.1 that the tension stiffening effect


does not reduce with strain.

It is the development of new cracks which

causes the apparent decay.


Hartl<105) has also concluded that the tension stiffening stress in direct
tension remains constant once the concrete has cracked.
suggested a tension stiffening stress of 0.4 fct

He said his tests

However, because re did

not consider concrete variability in his analysis, this conclusion is closer


to the Author's than it may at first appear.

In effect, Hartl concluded

that the average tension stiffening stress is 40% of the initial cracking
stress.

The Author has concluded that

stress is approximately 30% of the

the average tension stiffening

average cracking stress.

For the

analysis considered here this comes to over 35% of the initial cracking
stress.

In view of the other variables in the analysis this is remarkably

close.
6.3.2 Flexural Tests

Clark and Spiers <11 0) have tested a series of beams and slabs with long
constant-moment

regions.

As

it

was

these tests which were used

to

develop Cope's tension stiffening function<37>, it is not surprising that


his function gives a good fit to the results.
r~analyse

However, it was decided to

some of the specimens using the constant tension stiffening

function suggested in .6.3.1.

The only concession made to the difference

between direct tension and bending was to increase the effective tensile
strength from 0.8 to 1.0 times the cylinder strength.
- 83 -

Predictions using

the two tension stiffening functions are compared with the results of one
of Clark's tests in Figure 6 .5 .
Cope's function appears to give a better fit to the results but it is not
possible to tell conclusively from Clerk's r esults which function is more
realistic.
0

Mome nt 40
(kNm)
30

20

<!

Test

---Cope' s Function
10
-

Constant Function

0
0

Curvature x 10 6
Figure 6.5:

<mm - )

Analysis of Clark's Beam 4-

6.4. TESTS

6.4.1 Design of Specimens

Because the tests would serve to calibrate the analysis and to investigate
the effect of scale for the model bridge tests which will be considered in
Chapter 8 , it was desirable to make
possible

to a

strip of

the test specimens as similar as

the proposed slab.

perform identical tests at full and half size.

It

was also desirable to

In order to facilitate the

cyclic loading of the specimens they were designed so that the full size
specimens would fit into a Mayes testing machine.

This left very little

choice in the design of the specimen which is illustrated in Figure 6.6.


The specimens,

like the deck slabs considered in this thesis, were very

lightly reinforced which made tension stiffening more significant to their


behaviour.

They were provided with 0.49% reinforcement.

This compares

with Clark's most lightly reinforced specimen which had 0.4-4-%.


is

conventional,

this

percentage

is

calculated

from

A. l bd.

However, as
A better

indication of the significance of tension stiffening is given by the ratio


of uncracked to cracked stiffness.

This is a function of bh 3 Ec /A.E.d2

According to this relationship, the specimens were some 40% more lightly
reinforced than any of Clark's.

Some slabs used in practice are, however,


- 84 -

more

lightly

reinforced still,

particularly in

the secondary direction.

Consideration was given to testing more lightly reinforced specimens.

It

appeared, however, that they would not crack until normal service loads
were exceeded so little use could be made of the results.
A single specimen was

tested

typical of the reinforcement

with
used

higher

reinforcement

in current

area, more

practice, to see if the

expression derived from the lightly reinforced specimen was applicable to


these.
800
Load

~c

"'
0

_.

TA

. I

_.._

"'
0
Support

35Cover

1500
f700

390

A-A

.I

-r_j

7T12-0'2-250EF

Figure 6.6:

2T12-0 1 -200EF

Detail of test specimens

<dimensions are for full size specimen)

6.4.2 Loading Rig

All the specimens were tested in the Mayes machine simply supported under
two point loading as illustrated in Figure
under test in Figure 6.7.

6.6.

A specimen is illustrated

Consideration was given to applying known in- 85 -

plane

restraint

applicable to
results

to

the specimens

membrane action.

would

then

be

to

make

However,

extremely

the

results

more

analysis suggested

sensitive

to

the

directly
that

the

of

the

stiffness

restraint and it was not possible to control this well enough to obtain
useful results.
All the tests were performed under load control.
information

could

have

displacement

control.

been

obtained

from

Theoretically, more

tests

performed

under

However, the test rig was not stiff enough to

achieve true displacement control.

Figure 6. 7:

Half size specimen under test

6.'-.3 Materials

a. Reinforcement
The reinforcement used was GKN ''Tor Bar" obtained from normal commercial
sources in the required sizes; 6mm and 12mm for the main tests and 16mm
for the more heavily reinforced specimen.
A stress-strain curve was obtained for samples of the bar using the same
Mayes machine which was used for the tests.
b. Concrete
The

mix

used

for

the

full

scale

specimens

was

intended

to

be

representative of normal practice for bridge deck slabs and to give a


strength close to the nominal design strength.
be mutually exclusive;

These objectives proved to

mix which complied with the minimum cement

content normally specified and which had a reasonable workability always


- 86 -

The exact mix

produced significantly more than the nominal strength.

proportions were varied between specimens as an attempt was made to get


closer to the desired results, but a typical mix is detailed in Table 6. 1.

Quantity (per nominal m3

Material

Half Size

Full Size

10-20mm Thame s Valley Gravel

780kg

5-lOmm Thames Valley Gravel

390kg

995kg

Sand <Thames Valley, zone 2 grading)

610kg

726kg

Ordinary Portland Cement

325kg

365kg

~1901

Water
Table 6.1:

~2101

Typical mixes

The mix used for the half scale tests was intended to be as close as
practical to a half scale model of the mix used for the full scale tests.
A lOmm maximum size aggregate was used and the proportion of fines was
increased

to get

close

to

the scaled grading curve.

greater surface area of aggregate in


content was needed.

the

finer

mix,

Because of the
a

higher cement

The water cement ratio was also increased so that

the strength, particularly the tensile strength, would be no higher than


for the full scale specimens.

A typical mix is detailed in Table 6.1 .

As

with the full size mix, modifications were made over the course of the
test sequence.

However, in order to ensure that this did not affect the

relationship between the full and half size mixes, the same modifications
were made to both mixes.
The small size of the specimens meant that it would have been practical to
mix

the

concrete

in

the

laboratory

using

dried

aggregate.

However,

because it was intended to use the tests to develop a mix design for the
model bridge tests for which this would not be practical, it was decided
to use the same 0.25 cubic metre pan mixer and batching plant which would
be used for the model bridge tests.
Cube tests and split cylinder tests for all the mixes were performed using
150mm cubes and 150mm diameter cylinders.

Some 150mm diameter cylinders

were also tested for Young's modulus in the Mayes machine.

- 87 -

All the cubes

and cylinders were cured with the test specimens, under plastic for seven
days and then in the laboratory.
The split cylinder tests suggested that the tensile strength of the full
and half size mixes were similar.
size mixes at test age

The mean tensile strength of the full

<approximately 28 days> was 3.30N/mm"' compared

with 3.14 for the half sized mixes.

Theoretically, it is more correct to

compare results for the full size mix tested with 150mm diameter cylinders
with results for the half size mix tested with 75mm cylinders.

However,

since no 75mm cylinder moulds were available, this was not done.

Instead,

some IOOmm cylinders from the full size mix and some 50mm cylinders from
the half size mix were tested.
3.94

and

3.49N/mm

The mean results from these tests were

respectively.

Thus,

changing

from

150

to

100mm

specimens for the full size mix gave a 19% higher strength whilst changing
from

150 to 50mm with the half size mix gave only an

Interpolating

between

the

results

for

the

150

and

11% increase.

50mm

cylinders

suggested that the strength of the half size mix measured using 75mm
cylinders would be 3.36N/mm"'; 2% higher than the measured strength of the
full size

The real significance of

mix.

these results is that

they

indicate that both the scale effect and the difference between the two
mixes were small compared with the random variation in the results.
The compressive strength of the full and half size mixes were also similar
to each other but the latter did tend to be slightly lower.

Typical

figures (actually those for the first pair of specimens and for the mix
detailed in Table 6.1) were 54.7N/mm"' for the full size mix and 47.0 for
the half size, both measured with 150mm cubes.
increased the latter to 48.4N/mm

Using half size cubes

6.4..4 Loading

The first pair of specimens, one full size and one half size, were loaded
to a load corresponding to the maximum service moment which BS 5400 would
allow.
the

Next, they were subjected to many cycles of a lower load <55% of

first

BS 5400,

load)
that

corresponding

is 25

to

the

units of HB in a

maximum

HA

equivalent

bridge designed

for

load

in

45 units.

However, the number of cycles <over 100,000> and the intensity of the load
were deliberately excessive.

It was hoped to use the test

to justify

using a much smaller number of cycles in the model bridge tests.

- 88 -

After

the cyclic loads had been completed the specimens were loaded to full
service load, unloaded, then loaded to failure.
The second pair of specimens were treated in the same way except that
they were first loaded to only the reduced, 55%, load.
The third full size specimen was treated in the same way as the second,
except that it was tested upside down to see if this altered the results.
When the third half size specimen was tested, an unloading expression had
been developed which gave reasonable results.

It was realised, however,

that these results were not affected by the stress which was assumed to
exist

in cracked concrete which was subsequently compressed.

It was

desirable, therefore, to test a specimen under reversed moment but the


apparatus did not enable this to be done.

The solution adopted was to

load the speciuien in the same way as the first but to turn it over after
10,000 cycles and start the test again.

It was considered that only the

half. size specimens could reliably be tested in this way because the dead
weight stress involved in turning over the full size specimen would be too
great.
The more heavily reinforced specimen was treated in the same way as the
first

specimen,

the

loads - being

increased

to

allow

for

the

extra

reinforcement.
6.4-.5 Processing of Results

Three columns of "demec" points were 'fixed to one side of the constant
moment regions.

The strain was averaged over the three demec readings in

a row and then a linear regression over the rows was performed to give an
average curvature and extension.

The curvature was also estimated from

deflection readings taken from rows of dial gauges.

These curvatures

differed, typically by 10% but sometimes by as much as 30%.

The curvature

estimated from a row of dials along the edge of the slab on the side to
which

the demec studs were attached

<that

is the three dial gauges

nearest the camera in Figure 6. 7 >, was only marginally closer than that
estimated from a row at the longitudinal centre-line or on the far side of
the- slab.

It was therefore conCluded that

the discrepancy was due to

variation in curvature over the length of the constant moment region,


rather - than_ over the width of the slab.

Since the analysis assumes a

constant curvature and the regression- gives a true average curvature,


89 -

whereas the curvature estimated from the deflection readings is weighted


towards the curvature at the mid-span of the slab, it was decided to use
the demec readings in preference to the dials.

The discrepancy does,

however, give an indication of the relatively low accuracy which can be


expected in the analysis of tension stiffening.
The regression analysis calculated the deviation of the strain readings
from the straight line.

The root mean square deviations varied between

specimens from less than 1% of the maximum strain to over 20%.


former

figure

The

indicated that, on average over the gauge length, plane

sections had remained plane

<as the theory assumes> whilst

the latter

indicated that they had not.

The difference is due to the random nature

of the cracking and the fact that the constant moment region was only
long enough to accommodate some three main cracks.

The best fit was

obtained in specimens for which both ends of the gauge length happened to
be mid-way between cracks, giving the theoretically desirable exact integer
number of cracks.

The worst fit occurred in a specimen in which a sloping

crack crossed the end of the gauge length.

This problem could be reduced

by using a longer constant moment region.

However, because of the effect

of variability of concrete tensile strength, this would give a misleading


impression of the shape of the tension stiffening function.
The main reinforcement in most of the specimens was also provided with
electrical resistance strain gauges.

This provided some useful information

but

that

the short

gauge length meant

the results were not directly

applicable to smeared crack analysis and the gauges were not fitted to the
final specimen.

6.4.6 Results and Analysis

a. First Loading
All the tests were analysed using a non-linear program.
specimens

were

essentially

beams,

in

that

they

were

Because the

subjected

to

constant moment over their width and they were not wide enough to be
forced

to

bend

cylindrically

(rather

than

anti-elastically>

convenient to use beam elements for the analysis.

it

was

The program used will

be described in Chapter 7.
The results of three of the tests are shown in Figures 6.8 to 6.10 along
with the results of analyses using the tension stiffening function which
~

90 -

was

eventually

adopted and

which

is

illustrated

in Figure 6.12.

To

facilitate direct comparison between the figures, the results of the half
size tests are expressed as equivalent results at full size.
The

experimental

predictions

results

obtained

were

using

initially
variety

compared

of

tension

with

the

analytical

stiffening

functions.

Because of the low steel area, and because the predictions for both the
curvature and for the axial extension at mid-depth were considered, this
gave a better indication of the shape of the tension stiffening function
than previous tests.

However, it was still not possible to .obtain totally

unambiguous results.

Both Cope's function and that proposed in 6.2.1 gave

reasonably good predictions.

A close study of the results suggested,

however, that immediately after cracking the true ,tension stiffening stress
was higher than suggested by either function.
up

to and above yield) it

explanation

for

this

appeared

behaviour,

and

At high strains <apparently

to be around 0.1
its

apparent

to 0.2f et

difference

from

An
the

behaviour in direct tension, is proposed.


When the peak concrete stress is reached in direct tension, a crack forms
and the load reduces.

A significant increase in extension is needed to get

back to the load which caused the crack and no new cracks can form until
this has happened.

If the specimen was held at an extension just above

that at which the peak concrete stress was developed, there would be a
significant tensile stress in the concrete even at the sections where the
cracks were forming.
a

test

unless

it

However, this stress has no effect on the results of


is

performed

under

true displacement

control,

which

requires a very stiff testing rig.


In a section in flexure, in contrast, there is always a region near the top
of a crack where there is a significant tensile stress due to the ductility
of concrete.

This gives the observed higher tension stiffening stress at

lower strains.
of

cracks

and,

It also increases the stress in the concrete on either side

as

we

saw

in

6.2.1,

this

reduces

the

crack

spacing.

Paradoxically this means that when the strain subsequently increases, and
the stress at the cracks reduces to zero, the tension stiffening stress is
lower than it would have been without the ductility of concrete in tension.
This explains why, at high strains, the tension stiffening stress is .lower
in flexure than in direct tension.

- 91 -

Moment 10
<kNm )
8

-Analysis

0
2

><

Test (first cycle)

Test (after cycling)

10

12

14

Curvature x 10
Fi!fure 6.8:

16

18

<mm-

Results of first f ull size test

Momen t 10
<kllm)
8
0
0

6
lr

--Ana~ysi s

xTest <first c ycle)

Test <after cycling)

><

0
0

10

12

14

Curvature x 10 6
Figure 6.9:

Theoretically,

this

effect

16

18

<mm-

1 )

Results of first half size test

should

be

more

pronounced

where

the

crack

s pacing is controlled by the depth of concrete in tension, rather than by


the reinforcement .

According to Beeby's theory <34), this means it will be

more pronounced where the reinforcement is widely spaced.

Thus it appears

that the tension stiffening stress at high strains should reduce as the
bar spacing increases.

This has been observed by Clark and Cranston (111 ),

further confirming the theory.


- 92 -

The

ductility

of

c onc rete

in

tension

also

explains

why

the

drop

in

stiffness visible in Figures 6 .8 to 6 . 10 when the concrete first cracks is


far less abrupt than analysis using any normal tension stiffening function
suggests.

This might have been partly explained by the variabi lity of

concrete which means that not all the cracks formed at once.

However,

analysis using an approach similar to that cons idered in 6.3. 1 s howed tha t
this explanation was not suffi cient.

Moment 8
<kNm>

6
4

-20

- 15

10

-1 0

-'

15
<mm-

Curvature x 10 6

-2

Analysi s
(first cyc le >

)(

Tes t

Test <after c yc ling >

..',

Test (after inve rted

-6

..

-8

c yc ling )

Fig ure 6.10:

Resul ts of third half size test

The f irst vis ible crack did not appear in the half s cale specimens until
the strain was

significantly greater than in the full scale specimens;

approximately 300 microstrain compared with 200.

However, there was not a

corresponding differen ce in the effective tensile strength.


and

half s ize

s pecimens

exhibited

significant

Both the full

non-linearity

before

the

cracks became visible but t his was more pronounced with the half scale
specimens.

It was considered that this might have arisen solely because

the cracks in the half scale specimens were half as wide and so did not
become vis ible

until their s cale size was greater.

To eliminate this

pos sibility, the later half scale specimens were inspected thoroughly for
- 93 -

cracks using a magnifying glass and a crack microscope, whilst the full
scale ones were inspected only with the naked eye.

This did not alter the

conclusion.
The results suggested that the half scale tests would give a reasonably
good indication of the load-displacement relationship, and hence of the
stresses, in a full scale specimen.
tension
However,

stiffening
the

most

function

could

obvious

fault

They also suggested that the same


be

used

of

at

the

full

analysis,

and

at

its

half

size.

tendency

to

exaggerate the abruptness of the loss of stiffness as the concrete cracks,


is greater with the half scale model.

The results also suggest that the

use of a half size model to predict the behaviour of a full size bridge is
liable to over-estimate the load at which cracking first appears.
The non-linearity observed before the cracks could be seen suggested that
both moment re-distribution and compressive membrane action could start to
act before cracks become visible.

Thus compressive membrane action should

delay the format ion of the first visible crack.

This has been observed by

both Guyon <10> and

it

Guyon's

specimen,

Kirkpatrick (49).
being

small

However,
scale

model,

now seems likely that


exaggerated

the

effect.

Similarly Kirkpatrick's slab, being only 160mm thick, would have shown a
more pronounced effect than a thicker slab.
Another implication of this non-linearity before cracking is that, in an
analysis which ignores the effect, reinforcement could significantly affect
the apparent

tensile strength.

This is confirmed by the fact

that

the

best fit to the results for the lightly reinforced specimens was obtained
using an effective tensile strength of approximately 0.8 times the split
cylinder strength whilst, for the more heavily reinforced specimen and for
Clark's tests, the full split cylinder strength gave better results.
These effects could be modelled by including some non-linearity before
cracking in the analysis.

It was found, however, that if sufficient non-

linearity was included to model these effects, the non-linearity in the


moment-curvature response prior to cracking was greatly exaggerated.
explanation for

this is that

The

the non-linearity was due to local micro-

cracking which had little effect on the strain averaged over a long gauge
length.

- 94 -

The best

interpretation of the results seemed

to be that

the tension

stiffening function should have the form shown in Figure 6.11.

This ie

essentially a constant tension stiffening function with an expression for


the tensile stress transmitted across the crack added on.

Theoretically,

the latter stress is a function of the width of the crack.

This implies

that, when it is expressed as a function of strain, it should be affected


by crack spacing.

However, the variability of the results and the narrow

range of crack spacing in the specimens made it impossible to detect this


trend.
f.~

Stress

Strain

Figure 6.11:

Because

the stress

Ideal tension stiffening function

reduces significantly on cracking,

and

because

the

subsequent stress is taken to be a function of the initial cracking stress,


the predicted response is very sensitive to the assumed tensile strength.
Small changes in the tensile strength have a much greater effect on the
results

than quite

large changes in

stiffening

function.

considered

in 6.2.2, encourages the continued use of tension stiffening

fun et ions
s tiffening

of

This,

the assumed shape of the tension

the type

stress

reinforcement

in

could

combined with

shown

in Figure 6.1.

concrete
delay

the

the

problem of steel stress

Even with

between

the

yielding

of

neutral
the

these,
axis

tension
and

reinforcement

the
but,

fortunately, the effect is not significant in realistic cases.


For the present study, therefore, it was decided to use the function shown
in Figure 6.12.

This is essentially a compromise between the constant

function s uggested in 6.2.1 and the type of function shown in Figure 6.1
and it was found to give marginally better results than either.

In most

cases, the results could be improved further by stopping the reduction in


stress at approximately 0.1

f e t rather than at zero but, because of the

risk of artificially delaying reinforcement yielding, this was not done.

- 95 -

Stress

FTENSHct/Er
<FTENS normall y = 20)

"'

Stra in

Unloading

/
/

Per man en t s e t = 0 .

Figure 6. 12:

5 *f c t/E~

Tens i on stiffening f un c ti on adopted

In any particular tes t, the results can be improved by minor adjustments


to the tension stiffening function but the behaviour is too variable to
justify this.

An indication of the lack of repeatability of the tests is

given by comparing the maximum curvature on first loading in Figures 6.9


and 6.10.

The curvature of two nominally identical specimens subjected to

identical

loads

differed

by

47%,

even

though

their

measured

material

properties and dimensions were almost identical.


The

tension

stiffening

expression

derived

for

the

lightly

reinforced

specimens appeared to work equally well in the one more heavily r e inforced
specimen.

However, because the behaviour of this was much less sensitive

to tension stiffening,

and

because only one specimen was

tested,

this

result wa s not conclusive.

b. Cyclic Loading
The effect of even 160,000 cycles to 55% of the peak load experienced was
considered to be too small, compared with the other variables and errors
in the analysis, to be worth including.

As expected, the effect of the

same load cycles on specimens which had not previously been subje cted to a
higher load was much greater.

The static load used in this test was not


- 96 -

sufficient

to

produce a

fully

developed

developed during the cyclic tests.


fatigue theory.

crack pattern and new cracks

This is as would be expected from

Analysis using a reduced tensile strength to allow for

fatigue appeared to give good results but tests to a wide range of load
cycle intensities would be needed to check this properly.

It is also not

entirely clear that the degradation was due to fatigue; it could have been
largely due to creep since the tests were conducted under sinusoidal load
variations which gave a mean moment some 55% of the maximum.

However,

since no long term static tests were performed, it was not possible to
separate the effects of fatigue and creep.
A single cycle to a high stress had a much greater effect than many cycles
to a lower stress.

In terms of bridge deck design and analysis this

suggests that it is reasonable to consider only the worst load cases, the
HB load cases, in the stress history analysis and to ignore cyclic loads
completely.

This is fortunate as it means the stress history of a given

point in the structure can, for practical purposes, be recorded by a single


number; the maximum historic strain.
c. Unloading and Re-loading
There was a difference between the unloading and the re-loading path, as
can be seen from Figures 6.8 to 6.10.
this

was

assumed
path,

small

compared

with

However, it was decided that since

either

the

effect

fct or the difference between the first

it

was reasonable to ignore

it.

of

small

changes

in

loading and unloading

Thus the ability to store the

relevant stress history as a single number was preserved.

It should be

noted, however, that this approach may not be valid in a dynamic analysis
because the difference in the paths, the hysteresis loop, represents energy
absorbed by the structure and contributes to the damping.
None
Cope's

of

the

for

unloading

example

expressions

<which

was

the

previously
best

of

used gave good


them)

results.

under-estimated

the

curvature which remained when the load was removed; typically by a factor
of three.

Since these functions were based on data which was either very

inadequate or derived from structures which, because of relatively heavy


reinforcement, were not sensitive to the expression used, it was decided to
ignore them completely.

After trying various relationships that shown in

Figure 6.12 was adopted, the slope of the unloading path being 3.5 times
the slope of the tension stiffening function, that is ci 2 equals 3.5a,
- 97 -

in

Figure 6.12.

This gave reasonably good results such as those shown in

Figures 6.8 and 6.9.


A significant

deformation

implies

when a

that,

remained after

load

is applied

the

load

which

is

was removed.

small relative

This
to the

maximum previously applied load, the strains and deflections relative to


the

initial

<unstressed) condition can

conventional
concrete.

elastic

analysis

which

be greater than predicted by a


ignores

the

tensile

strength

of

It was found that the deformations <both observed in the tests

and predicted by the non-linear analysis) became equal to those predicted


ignoring concrete in tension at a load which would correspond to 25 units
of HB if the section was fully stressed under 45 units of HB.

Since crack

widths in BS 5400 are checked under a load of 25 units of HB, this implies
that

BS

5400

is

calculations in

justified

in

ignoring

bridges designed

for

45

tension

stiffening

in

crack

units of HB load even though

significant tension stiffening was observed under full load after over a
hundred thousand cycles of normal service load had been applied.

However,

in structures designed for lower HB loads, the assumption is conservative.


Unfortunately, the one specimen which was inverted during the tests was
the only
assumed

one
to

0.5f ct/Ec.

which

become

was

sensitive to the amount

permanent

after

unloading,

of strain which was

which

was

taken

to

be

The results for this specimen are shown in Figure 6.10 in which

moments due to loads applied before the specimen was inverted are shown
as positive.

The biggest discrepancy in the unloading and re-loading part

of the plot

is that

the analysis

failed to predict the earlier initial

cracking load under negative moments, that

is in the inverted position.

This earlier cracking appears to have been the result of the cracks formed
by the previously applied positive moments acting as crack inducers since
the new cracks all joined the previous cracks.
this

earlier

cracking

in

the

one

specimen

There was no evidence of


which

was

tested

inverted

throughout.

d. Failure
On completion of the tests, all the specimens were loaded to failure and
they all failed

in

flexure.

The only unusual feature of the failure

behaviour was that the low steel area combined with the low d/h ratio
meant they did not reach peak load until the top steel yielded in tension.
This was predicted by the analysis.
- 98 -

The tension stiffening function used implied that unloading and re-loading
to the same load would have no effect on the deflections.
did have some effect, as can be seen from

Figure 6.8.

In practice it

However, when the

loading was further increased the tension stiffening appeared to recover


and the discrepancy was considered acceptable.

6.5 CONCLUSIONS
Because of its sensitivity to a highly variable quantity, the effective
tensile

strength

of

concrete,

tension

stiffening

cannot

be

predicted

accurately.

However, the functions developed in this chapter appear to be

significant

improvements over

those used

in

the

past,

particularly for

unloading.
The studies of similar full and half size strips indicated that a half
scale model will give a good indication of all aspects of behaviour except
the load to produce the first visible crack.
function can be used as at full size.

- 99 -

The same tension stiffening

CHAPTER
A

SIMPLER

NON-LINEAR

ANALYSIS

7.1 INTRODOCTION

Section 5.6 may give the impression that analysis for design is far more
difficult

than

specimens.

analysis

for

the

behaviour

of

laboratory

However, designers have a major advantage; they have no need

accurate predictions,

for

predicting

they

need

only

safe predictions.

Realistic

predictions are desirable because they lead to more economical designs but
errors which would be considered excessive to researchers are acceptable
to designers, provided they act in the safe direction.

Research on non-

linear analysis has concentrated on obtaining accurate predictions for the


load-displacement response of structures monotonically loaded to failure.
From the point of view of the design of the type of structure considered
here,

this

is unfortunate; neither deflection nor ultimate strength are

critical design criteria, loads do not increase monotonically and "accuracyu


is

neither

obtainable

Batchelor<78)

and

nor

necessary.

Kirkpatrick<l3>,

Indeed,

slabs

with

according

the

minimum

to

both

practical

reinforcement have over three times the required ultimate strength.

If

this is true, an analysis which under-estimates strength by a factor of


three is not merely adequate for predicting strength; it is as good as one
which is accurate to 0.001%.
The analytical methods considered in Chapter 5 contrast sharply with those
currently used in design and considered in Chapter 2.
sophisticated

and

expensive

potentially

but

able

to

The former are


give

realistic

predict ions based on realistic behaviour models even if, at the present
state of the art, they are not totally reliable.

The latter are cheap and

simple but based on unrealistic models of behaviour.

Their predictions are

not as realistic as those of NLFEA but they are more reliable; they are
always safe.

In the extreme case of restrained slabs, the two forms of

analysis may differ by factors of 5 or even 10 on strength.


therefore, some intermediate form of analysis
understand

and

considered

in

more
Chapter

compatible
5

but

Chapter 2) would be useful.

with

more

<safer, cheaper, easier to

codes

realistic

Clearly,

of

practice

than

those

than

those

considered

in

There is vast scope for making conservative

simplifying assumptions compared with the analysis considered in Chapter 5,


whilst

still

maintaining

greater

realism
- 100-

than

the

forms

of

analys~_s

considered in Chapter 2.

This Chapter aims to develop such a form of

analysis which could be used in design and assessment.


The

Chapter

also

aims

to

develop

program

which

can

be

used

assessing simpler design methods, such as those considered in 3.2.8.

for
The

same program will be used as an analytical tool for investigating the


behaviour of bridge decks, including the models which will be considered. in
Chapter 8.

However, because of the fundamental difference between the

safe estimate analysis needed for design and the best estimate analysis

needed in research to facilitate direct comparisons with test results some


details, including the material models, will differ.

7.2 GENERAL APPROACH


The analysis is essentially a simplification of the approach considered in
Chapter 5; that is, it is a non-linear analysis using the smeared crack,
distributed steel approach.

However, in order to simplify it as much as

possible and to make it more similar to the grillage analyses with which
most bridge engineers are familiar, simple line elements are used to model
both the beams and the slab.

This greatly reduces the size of the program

enabling it to run on a desk top computer; it is perhaps the first time


this form of analysis has been performed on such a machine.
A disadvantage of this "grillage" type of analysis is that it treats the
stresses in the two directions as independent so it has to use uniaxial
material properties.

It thus cannot model the increase in stiffness due to

the Poisson's ratio effect in concrete subjected to biaxial compression, nor


can

it

model

reduction

in

the

enhancement

tensile

strength,

of
due

concrete's
to

biaxial

compressive
stress.

strength,
These

or

faults,

however, generally act in the safe direct ion and are considered acceptable
in design, indeed they are shared with all the analytical methods normally
used in

design.

Another fault is that this form of analysis can only check stresses in the
element direction so the maximum principal stress is not modelled if its
direction does not coincide with an element direction.
faults,

Unlike the other

this one is not acceptable because it could lead to significant

over-estimates of the load to cause cracking or failure.

To avoid this,

torsionless elements are used forcing the principal moment directions to


align with the elements.

This is also sometimes done in conventional


- 101-

linear grillage analyses because it enables the computed moments to be


used directly to design the reinforcement without the need to transform
them

to

the

reinforcement

direction.

Because

the

program

uses

this

principle, the element direction has to be the same as the reinforcement


direction.

This, and the desirability of using an approximately orthogonal

mesh so that the concrete stresses in the two directions are independent,
leads to what is probably the major practical limitation on the use of the
program; it is difficult to use it to model highly skewed bridges.
The problem of the principal moment direction does not arise in the downstand beams so these can be given torsional stiffnesses and elastic values
are

used

for

this.

However,

because

the program assumes

that

plane

sections remain plane and normal to the reference plane, warping stresses
and the effect of transverse bending in the flanges cannot be represented.
In the type of beams considered in this study, the predicted torques were
not excessive and the increase in stiffness due to the transverse bending
stiffness of the bottom flange exceeded any reduction due to cracking.
Thus the errors resulting from using elastic torsional properties were
conservative as well as small.

However,

this would not apply in all

structures and the program has been altered to enable a limiting value for
the torsional strength of beams to be specified <112 ).

This feature was

used in the analysis of the second of the models which will be considered
in Chapters 8 and 9 and the limiting moment was reached in the diaphragms
although not the main beams.
The particular program used was developed from one written by Edwards<36>,
although the modifications are so extensive that analyses have little more
than some basic principles in common.

7.3 DISPLACEMENT FUNCTION


In a beam element which is loaded only at the ends, the axial force is
constant whilst the bending moment varies linearly over the length.

The

displacement function used by Edwards matched thiE by using constant axial


strain and a linear v11riation in curvature over element length.
linear-elastic beam element,

force

In a

is proportional to axial strain and

moment is proportional to curvature so this shape function is ideal.

In a

non-linear element 1t is not quite as good because the stiffness varies


over the length which tends to result, for example, in the analysis underestimating the moment variation over element length.
- 102-

For use in non-

linear analysis, however, there is a more fundamental fault in the shape


function which does not appear to have been considered by other analysts,
such as Buckle and Jackson<90), who have used this type of element in nonlinear analysis.
The displacements are defined at the reference "plane" or <more correctly
for a line element) the reference "line" which, in this study, is at middepth of the slab.
assumed

to

be

This means that the strain at the reference plane is

constant

whilst

at

any

other

variation in axial strain along the element.


to

be

proportional

level

there

is a

linear

This variation is constrained

to the vertical distance

from

the

reference plane.

Since the level of the reference plane is largely arbitrary, this is not
satisfactory; even with a perfectly uniform and linear-elastic element, the
correct displacement field can only be reproduced if the reference plane is
at the neutral axis.
In a typical cracked slab element the actual neutral axis, the level at
which there is no axial strain, is well above <that is, on the compressive
side of) mid-depth.

In the real structure the variation in axial strain

along the element is proportional to the distance from the neutral axis
but, in the computer model, it is proportional to the distance from middepth.

Thus, if the variation in curvature over length is correct, the

variation

in

strain

along

the element

material below the reference plane.

is

under-estimated

for

all

the

As a result, unless the element mesh

is so fine, that the variation in curvature over length is insignificant,


the analysis can

fail

steel

can

and

hence

to predict

reinforcement yielding in the tension

over-estimate strength.

There

is

also a

region

between the actual neutral axis and the reference plane where the real
strain becomes more tensile in the direction of increasing curvature but
that in the computer model becomes more compressive.
the top of the cracks are in this region

This means that, if

<which is often the case> the

computer model will indicate that the extent of cracking will reduce over
element length in the direction of increasing curvature, which is clearly
incorrect.
One solution

to this problem would be to keep the same displacement

function but to put the reference plane at the neutral axis.

This is not

practical because the neutral axis moves as the concrete cracks.

The

effect can, however, be obtained by introducing a linear variation in axial


- 103 -:

strain, as well as in curvature, over element length.


strain

due

to

reproduced at

linear

variation

all depths

in

curvature

in the element,

can

The variation in
then

be

correctly

rather than at only one.

In

addition, a linear variation in neutral axis depth over element length can
be modelled.
Introducing

the

linear variation in axial strain gives

the displacement

function illustrated in Figure 7 . 1.

Since the stresses are calculated at

only

of

two

sections

in

the

length

the

elements,

it

simply

means

increasing the axial strain at one section and reducing it at the other by
the same amount.

However, because this linear variation does not alter

the element length, it cannot be defined from the displacements of the two
end nodes.

The modification effectively amounts to introducing a

third

node at mid-length with only one degree of freedom; the axial displacement

o"'

in Figure 7.1.

-8,

Node 2

Node 1

jo, 1 1
L-----------------~

1.

Due to

o,

10:2/ 1
~--------------~-

2 . Due to 0 2

4-0 c /]

~40c/ ]
3. Due to Oc
Axial Strain
<at reference level>

Axial Displacement

Figure 7.1:

Displacement fun ction

- 104-

Introdt1cing

this

complicated

the

node

into

analysis

the

and

globsl

stiffness .matrix

increased

the

computer

would

storage

have
space

required.

This was avoided by considering the internal equilibrium of the

elements.

The fault in the original displacement function meant that the

axial forces calculated for the two sampling sections were not necessarily
equal.

This leads to a criterion for the correct value of liei the value

which equalises the forces.


comparing the forces at
calculation.

However, this value can only be obtained by

the two sections which requires an iterative

Performing this iterative calculation for every element each

time the forces in the structure are calculated would have greatly slowed
down the analysis.

To avoid this, the number of iterations for lie is

limited to two, but a vector of lie for all the elements is stored and used
as the first estimate the next time the element forces are calculated;
that is in the next iteration of the whole structure.

The modification has

effectively increased the number of degrees of freedom in the analysis by


some

30%

without

proportional

increase

in

the

required

computer

capacity.
The modified version of the program was tested by analysing the simple
case considered in Chapter 4 and the results are shown in Figure 7.2.
the Figure the percentage error in predicting the restraint

In

force or

displacement, whichever is greatest, is shown for analyses using different


numbers of elements.

For comparison, the same case was also analysed

using the previous version of the program.

Because this beam has no

tensile strength, and hence the formation of a crack does not release any
energy, the problem of mesh dependence which was considered in 6.2.3 does
not arise.

Thus, as the mesh is refined, both programs converge on the

"exact" analytical solution which was derived in Chapter 4.

However, the

modified form of the program converges very much more quickly and 3
elements with this give better results than 6 with the original program.
In most of the structures considered in this study, the improvement is
more fundamental because, with the old program, the mesh size required to
reduce the discretisation errors to acceptable levels is too fine by the
criteria considered in 6.2.3.
Having adopted the principle of defining extra degrees of freedom by
considering internal equilibrium of elements,
extend it to develop higher order elements.

it would be possible to

For example, one could use a

quadratic variation in both axial displacement and curvature.


- 105-

This would

give two extra degrees of freedom which could be defined by calculating


both the axial force and the bending moment at a third section at midlength and checking that
sections.

they were consistent with those at the other

This would undoubtedly enable a beam to be modelled with a

coarser element mesh.

However, Figure 7.2 shows that with the existing

program a remarkably coarse mesh gives satisfactory results.


only 3

elements

in a

half span

in which

there

is

Even with

complete moment

reversal, the worst error is around 3% which is small compared with the
variability

of

behaviour observed

in Chapter

6.

Also,

to

model slab

behaviour with a grillage (even a linear grillage ) a finer mesh would be


r equired.

Thus

the extra complication of higher order elements is not

justified.

Vcrs t Err or

<% )

30
---Orig1nal Program
- - - - Modified Program

20

\
10

'-...
1

32

16

Number of Elements
<in ha l f model)
Figure 7.2:

Effect of change t o di splacement function

7.4. ELEMENT INITIAL STIFFNESS CALCULATION


The program cal culates the initial stiffness matrix elastically, as in a
conventional linear grillage, using the gross- concrete section properties.
Althoug h the displacements are defined from the reference plane at middepth of the slab, Edwards' program cal culated the initial stiffnesses of
the down-stand beams about their own neutral axes.

It then treated them

as though they were calculated about the reference plane.


lead

to

errors

in

the

final

results

because

the

This did not

non-linear

force

calculation correctly calculated the forces from the displacements allowing


for the eccentricities.

However, because the initial stiffness matrix did

not model composite action between the beam and slab, and thus did not
- 10 6 -

represent the true behaviour even in the elastic range, it did result in
very slow convergence.

With the simple solution scheme used by Edwards,

analyses of structures in which the beams were large compared with the
slab would not converge at all.

The solution to this was to use a rigid

body transformation and this was done by defining the element stiffness
about the reference plane by using the stiffness matrix for an off-set
beam which is g iven in Ta ble 7.1.

x,

w,

Reference Level
= - -- -

8,
X

= EA

Axial Stiffness

Flexural Stiffnes s = El
Length

M,

R,

F,

6EI I P

- EAXI l

=1

4EI 1 1

8,

+
EA X2 11

w,

R2

F..,

2EI/l
- 6EI I J2

EAXI l

- 12EI IJ2

EAP 11
12EI/ P.

x,

M2

EA /1

6EI I P'

EAX I l

- EA i l

4EI 11

e."

<sy mmetri cal )

+
EAX2 1 l

w2

- 6EI IF

12EI/ F

X :.'

-EAX I l

EA/1

Table 7.1:

Stiffness matrix of an off-set beam element

<For simplicity an element in a plane frame is illustrated)


- 107 -

7.5

~PLANE

FORCES

Although Edwards described his program as a "grillage" this is not strictly


The non-linear analysis leads to axial forces in the elements

correct.

In order to distribu te these forces

which a true grillage cannot model.

correctly, it is necessar y to consider horizontal displacements and the inplane shear in the slab.

Thus Edwards' program considered five degrees of

freedom per node instead of three as in a true grillage.


The

in-plane

transverse

shear

in

displacement

the

elements

of

the

was

nodes;

calculated
the

from

horizontal

the

relative

displacement

perpendicular to the element direction of the node at one end of the


element relative to the node at the other end.

This implied that all of

this transverse displacement was resisted by shear even though it may


have actually been largely due to rotation of the whole element about the
vertical axis with no shear deformation, that is as shown in Figure 7 .3b
rather than 7.3a.

It also meant that

the complimentary shear and the

resulting axial forces, such as the transverse forces illustrated in Figure


3.12, were not modelled.

This led to errors in the treatment of in-plane

forces which were serious, not so much because they were large <although
they could be), as because they tended to act in the unsafe direction.

In

the finite element programs considered in Chapter 5, this fault is avoided


because there are enough nodes in an element to define its horizontal
shear

deformation

from

the

horizontal

displacements

of

the

nodes.

However, with only two nodes per element, the shear deformation can only
be defined if the rotation of the nodes is also known.

Thus, it was

necessary to introduce this sixth degree of freedom, rotation about the


vertical axis, into the program.
Trans verse displacement

b. Due t o rotation

a . Due to shear deformation

c. Due to uniform bending


Figure 7.3:

d. Due to non- uniform bending

Transverse displacements of a line element

- 108-

If

the

slab

was

modelled

with

normal

line

elements,

as

used

in

conventional space frame analysis, the transverse shear force would have
caused

transverse

bending

in

the

individual

elements

as

shown

in

Figure 7 .3d; it would have introduced Vierendeel frame type displacements.


These displacements do not arise in the real slab because the "elements"
cannot bend independently; they act compositely.

Thus the displacements

due to this local transverse bending of the elements had to be suppressed


in

the

computer

model.

To

achieve

this,

it

is

assumed

that

if

the

elements are subjected to a transverse displacement without rotation of


the nodes <that is as shown in Figure 7.3a and d> the only deformation is
due to shear flexibility and the deformation is as shown in Figure 7 .3a.
The shear force is calculated from this shear deformation, as in Edwards'
program, but the moments required to keep the element in equilibrium about
the vertical axis <an equal and opposite moment at each end) are applied.
In Edwards' program, these moments were not applied to the structure.

In

effect they were resisted by a totally artificial restraint to rotation of


the nodes about the vertical axis.
In order to preserve the basic simplicity of the elements, the individual
elements are assumed not

to provide any resistance to uniform bending

about the vertical axis; they do not resist the form of deformation shown
in Figure 7.3c.

This means that

the stress state can be taken to be

constant across the element width and avoids the need to perform a stress
integration over width as well as over depth and length.

The relatively

small moments required to maintain equilibrium with the in-plane shear are
the only moments about the vertical axis within the elements.

The bending

stiffness of the structure about the vertical axis is, however, modelled by
the differential axial forces in the elements.

The approach is to split

the transverse deformation of the elements into two components; a uniform


bending about

the vertical axis as shown in Figure 7.3c, which is not

resisted,

and

shear

deformation

as

shown

in

Figure

7.3a

which

is

resisted by the transverse shear stiffness of the concrete in the slab.


The mathematics of the assumed deformation state are given in Appendix B.
In

practice,

nominal

bending

stiffness

was

added

because

otherwise

rotation about the vertical axis is completely unrestrained in some models.


This treatment of in-plane forces is inherently approximate.

It might also

be argued that including in-plane shear is inconsistent with the reasons


given in 7.2

for

using torsionless elements since it


- 109-

implies that

the

maximum principal tensile stress may not align with the element direction.
However, the program is intended for modelling structures whose behaviour
is primarily flexural so it is appropriate to use a lower order of analysis
for the in- plane forces .
Load 60

<kN >
- - - - - 6 D. 0. F.

20

--6 D.O.F <2X s he a r modu l us )

10

15
Defle ction

Figure 7. 4- :

Figure

7.4

(actually
modified

shows
the

the

program

original program.

with

two

( DID)

Eff ect of in-plane shear

results

structure which

20

of

the

will

analysis

of

be considered

different

shear

a
in

moduli

simple

structure

7.10 .3) using

and

also

using

the
the

The effect of doubling the shear modulus is small, which

implies that errors in the treatment of in-plane shear have little effect
and justifies the use of approximate analysis for in-plane shear.

Even the

apparently fundamental fault in Edwards ' program has only a small effect
on this particular structure, although the artificial restraint to rotation
about

the vertical axis

modulus.

is equivalent

to more than doubling the shear

However, it is possible that the effect could be greater in some

other structures s o it was considered prudent to use the modified program


for all subsequent analyses to ensure that the results would be safe.

For

the same reason, and unlike in Edwards' program, a reduced shear modulus
is used for cracked concrete.
Because the elements are fixed together at slab level, and because inplane forces in the slab are represented, the program is able to model
both shear-leg and the effect of the shear connect ion between the beams
which was discussed in 2.4.3 and 3.2.7.
checks

moment

equilibrium about

Unlike Edwards' program, because it

the vertical axis,

it

also models

resulting transverse stresses which were illustrated in Figure 3.12.


failure

of

the

original

program

to

model

effect further justifies the modification.


- 110-

this

potentially

the
The

significant

7.6 LARGE DISPLACEMENTS

When a slab deflects relative to the restraining beams, the le1er arm at
which the restraint force acts is reduced.
significant

compared with

the

Once the deflection becomes

thickness of the slab,

reduces the slab's load carrying capacity.

this significantly

Curiously, most of the NLFEA

studies mentioned in Chapter 5 did not consider this effect whereas all
the <otherwise far less sophisticated) analyses considered in 3.2 did.
this study,

was originally decided

it

to follow

In

the NLFEA studies and

ignore the effect and, because the analysis is conservative in other ways,
the predictions still tended to err on the safe side.
reasons,

it was eventually decided to modify the program to make some

allowance
always

However, for three

for

leads

large displacements.

Firstly,

ignoring the effect nearly

to

in

unsafe

errors

which

act

the

direction

so

it

is

undesirable in a design situation even if the errors are relatively small.


Secondly, some of the tests on model bridges which were considered in
3.2.3,

notably Seal's <77>,

reached

such

large deflections

before

failing

<around h/2) that an analysis of these which assumes the deflection to be


small relative to slab thickness is clearly invalid.

Thirdly, for reasons

discussed in 3.2.8a, it would be desirable to be able to use membrane


act ion in the design of slabs with longer span to depth ratios than the
empirical design rules allow.

However, financial and time restrictions on

this project prevented an experimental study of such slabs.


design

was

to

be

justified

purely

by

analysis,

it

was

Thus, if their
particularly

important to ensure that the analysis was safe and, since longer span to
depth ratios increase the significance of deflections, this meant allowing
for the effect of deflect ions in the analysis.
Because

of

the

use

of

line

elements,

include the deflection in the analysis.

it

was

comparatively

simple

to

It was done within the elements by

adding the vertical component of the axial force to the shear force.

As

is illustrated for a simple case in Appendix Cl, this has the effect of
modelling

the

moment

in

the

elements

<that

is,

about

the

deflected

reference level) due to the axial force acting at the undeflected reference
level: it models what in a column would be called the "buckling", "added". or
"P6" moment.

The vertical component of the axial force is calculated only

from the difference in the vertical displacements of the two nodes.

The

effect of curvature over the length of the element is not included but
this is only significant if an excessively coarse element mesh is used.
- 111 -

To

maintain consistency, the vertical component of the in-plane shear was also
added and this is calculated from the rotation about the longitudinal axis
averaged for the two ends of the element.

As an additional allowance for

finite displacements, the axial strain in the elements is also corrected


for the effect of the slope as detailed in Appendix C2.

That is, the axial

strain used to calculate the forces in the element allows for the increase
in length of the element due to its slope.
These effects are modelled only in the non-linear force calculation, not in
the stiffness matrix.

This must reduce the convergence rate but was

considered acceptable.
7. 7 MATERIAL MODELS
7.7. 1 Steel

A tri-linear stress-strain relationship is used for steel as indicated in


Figure 7 .5.

In analyses for research purposes, the factors are chosen to

give the best approximation to the actual stress-strain curve of the steel.
The pre-strain is included primarily to enable prestressing to be modelled
but,

in simple slabs,

shrinkage.

a negative pre-strain can be used to represent

When modelling prestress, the pre-strain has to be reduced to

allow for losses because the program does not consider long-term effects.

Stress

f..,1t

Evlt

0. 2%

Strain
<including Pre-Strain)

Figure 7.5:

Steel properties

The same properties are used in compression as in tension, except for a


limit on strain hardening in compression.
function only of the present strain.

The stress is taken to be a

It would be simple to adjust the

- 112-

program to allow for the permanent set in steel which has been stressed
beyond its elastic limit but this would require the maximum strains to be
stored for all the steel layers in all the elements.
structures analysed
under

the

final

had steel stressed above

failure

load

case

when

Since none of the

its elastic

the

strain

limit, except

was

increasing

monotonically, the facility to model permanent set was not implemented.


In analyses for serviceability design, in order to avoid the problem of
stress history dependence as much as possible, the steel is taken to be
linear-elastic.

To

justify

this

assumption,

service stress has

limited to the elastic limit and this becomes a design criterion.

to

be

Thus no

advantage can be taken of re-distribution due to reinforcement yielding


under service loads.
is

considered

However, this is not a disadvantage as such yielding

undesirable

anyway.

The approach

has

the advantage of

making the analysis more compatible with current codes of practice.


In analysis for design at the ultimate limit state, the tri-linear stressstrain

relationship

can

be

used

to

represent

properties or the code specified properties.

either

the

actual

steel

It is normally assumed that

only reinforcement yielding due to the load case being analysed needs to
be considered.

This is justified if one assumes that only one load case

above design service level is applied.

However, in a bridge deck slab, this

is not very logical since the design vehicle cannot get to the critical
position without

first

being applied in other positions which are only

marginally less severe.


strength analysis) has

Fortunately,

this problem

(like all aspects of

little practical significance since serviceability

criteria are critical.

7.7.2 Concrete in Compression


The

stress-strain

relationship

illustrated in Figure 7.6.

used

for

concrete

in

compression

is

Various curves have been proposed which are

more realistic, but when one allows for the variability of concrete the
improvements

are

not

significant

and

Abdul-Rahmen<87)

used

the

even

simpler elastic-plastic relationship.


It

is

assumed

that

when

the

concrete

is unloaded,

parallel to the initial part of the loading diagram.

it

follows

line

Thus it takes on a

permanent deformation which is equal to the departure from linearity on


loading.
- 113-

Stress

0.7fc
0.5fcu

0+---_.------------~-------------r-----------

0.0025

0.0045
Strain

Figure 7. 6:

Properties used for concrete in compression

The strain at which the stress is taken to start to reduce is lower than
in many other analyses.

This was a reflection of the results obtained

f r om the cylinder tests and may have been due to the relatively fast speed
at which these tests were performed.

The rate of reduction of stress

after

intended

the

peak

has

been

passed

was

to

represent

the

true

behaviour of the concrete, which can only be observed with a very stiff
testing machine, rather than its apparent behaviour.

However, because of

the nature of most of the structures considered, analyses performed with


the more usual form of curve, with a longer plateau followed by a more
abrupt cut-off, gave very similar results.
The same basic approach to serviceability design is used as for steel; the
material is taken to be linear-elastic and a stress limit is imposed to
ensure that this is reasonably true.

However, although this limit is also

given in codes of practice, it is far


limit .

As

noted

in

2.3.4c

and

less satisfactory than the steel

demonstrated

in

reference

35 ,

even

structures designed to BS 5400 can be stressed well above the limits, yet
their behaviour is satisfactory.

This presents a problem.

If the stress

limit is imposed on structures designed using non-linear analysis it is


unduly conservative; if it is not imposed it will be more difficult to get
the approach accepted and it is also difficult to decide what the design
criteria should be.

A possible compromise is to impose a limit but t o

make it less conservative.

This can be done without departing from the

principle of linearity if the increase is justified by the biaxial stress


state in the critical area.
enhancement

due

to

There is a precedent for explicitly considering

multiaxial

stress
- 114-

states

in

BS

4975 <113).

An

alternative approach is to make a totally arbitrary increase in allowable


stress and then to justify it by comparison with test results.
For analysis for design ultimate strength the code specified stress block
could

be

used

but

tri-linear

conventional design methods,

the

approximation

was

employed.

In

characteristic material properties are

used in the analysis of the structure and the design strengths, with the
partial safety factors applied, are used only for
critical sections.

the analysis of the

In non-linear analyses, the analyses of the structure

and of the critical sections are not separated so this approach, although
recommended by BS 8110, is not appropriate.
safety

factors

were applied

to all

viewpoint, this is not justified.

In this study, therefore, the

the material.

From a

statistical

However, it is conservative (except for

some cases where restraint stresses are dominant> and, since serviceability
criteria are critical, this is acceptable.

A disadvantage of this approach

is that in most codes, including BS 5400, the design ultimate stress in


concrete is less than the limit of linearity used at serviceability.

Thus,

unlike in analyses for research, completely separate analyses have to be


performed for serviceability and for ultimate strength.
7.7.3 Concrete in Tension
The properties used for research analyses are illustrated in Figure 6.12
and were discussed in the last chapter.
In choosing properties for analyses for design the major problem is that

the

desirable

characteristics

of

the

properties,

that

they should

be

reasonably representative of real behaviour and that they should not be


strain history dependent, are mutually exclusive.

The simplest solution to

this problem is to abandon realism in favour of avoiding strain history


dependence and ignore the tensile strength of concrete completely.

As was

noted in 2.4.2, this approach has the major practical advantage of being
compatible with current codes.
disadvantage

is

that

it

is

It is also normally conservative, indeed a

liable

to

lead

to

an

unduly

prediction of the distribution of moments between the beams.

pessimistic
However, it

is not possible to prove that the approach is always conservative.


peculiarity of membrane action is that

the restraint force,

the effect

which leads to the enhanced behaviour, is a direct result of cracking.


Thus tensile strength, by reducing the extent of cracking, can reduce the
restraint force and hence the degree of enhancement.
- 115-

To investigate this,

a simple slab was analysed using concrete tensile strengths of zero and
3N/mm~.

The

latter

analysis

used

the

material

properties

Figure 6.12 and the results are shown in Figure 7.7.


used and the load quoted is that on the half model.

given

in

A half model was


For comparison, the

results of a conventional analysis ignoring the restraint as well as the


tensile strength are also shown.
The stresses in the slab which was analysed using a tensile strength of
3N/mm2

are calculated in two ways.

The lower lines give the stresses

directly from the computer program; that is the smeared stresses.


are

always

less

than

those

calculated

ignoring

the

tensile

although, once the concrete has cracked, the margin is small.

These

strength

The reason

for the discontinuous plot is that in the numerical analysis the cracking
advances, both in depth and along the slab, in discrete steps.

In the real

structure, the cracks can grow more smoothly in depth but the cracked zone
can only advance along the slab in discrete steps as individual cracks
form.

In order to make the analysis as realistic as possible, the element

length was matched to the estimated crack spacing giving five elements in
a half model.

It

was found that an analysis using a

finer mesh

(20

elements in place of 5> predicted very similar behaviour, the deflections


being within 2%.

However, the smeared steel stress was up to 50% higher ..

A study of the results revealed that there were two reasons for this.

The

first was that, on first cracking, the fine mesh predicted unrealistically
localised

cracking

and

hence

an

unrealistically

small

restraint

force.

However, because of the tension stiffening function used, the extension on


initial cracking was very limited even in the coarse model.

The effect of

mesh size was therefore far less pronounced than in an earlier analysis
performed with concrete tensile strength but without tension stiffening.
The second reason for the effect of mesh size is that the analysis gives
the

stress

section.

only

at

the

last

integration station,

not

at

the

critical

In the coarse mesh, the last integration station is 40mm from

the critical section and thus is subjected to a 4% lower moment.

This

might normally be expected to make only a 4% difference to the stress.


However, the concrete properties used in the analysis make the momentsteel stress relationship non-linear whilst the 4% difference in moment is
not accompanied by a difference in the enhancing axial force.

The peak

stress predicted by the fine mesh is very localised and the effect is far

- 116-

less pronounced in the analysis of real slabs because of the finite width
o f the applied loads.

Stress 300
<Nimm2

200

100

I
f~t

10

15

=3

20

25

<smeared stress)

30

Load <kN/m width)


a . Steel Stress

=0

Stress 20

<unrestrained)~c~

=3

<stress at crack)

=0

(N/ID1!f2)

15

10

=3

<smeared stress)

10

15

20

25

30

Load <kN/m width)


b. Concrete Stress

Figure 7. 7: Effect of concrete tensile strength


<Central line load, simply supported but with rigid in- plane restraint
1

= 2000,

= 160,

= 119,

T12-250 reinforcement)

It was noted in 6.2.2 that smeared crack analysis under-estimates the peak

stress in the reinforcement.

This is clearly undesirable if stress is to


- 117-

be used as a design criterion so it appears to be more appropriate to


calculate the peak stress at the crack at the critical position using only
the forces, not the stresses, given by the computer analysis.

This also

has the advanfage of eliminating one of the effects of mesh size.


stresses calculated in this way are given by the upper solid lines.

The
In

accordance with normal practice, the critical sect ion was analysed ignoring
both the tensile strength of the concrete and the effect of the top steel.
Separate

calculations

relatively small,

confirmed

provided that

that

the

effect

of

these

would

be

the post-peak part of the stress-strain

relationship used for the analysis of the structure was not included.
Until the analysis predicts cracking, the stress calculated in this way has
no real physical meaning and is not

plotted.

When the analysis first

predicts cracking, the extent of the cracking is very limited.

Thus the

restraint force is small and the calculated stress at the critical section
is similar to that given by the conventional analysis and substantially
greater

than

completely.

is

predicted

ignoring

the

tensile

strength

of

concrete

As the load increases, the extent of cracking <and hence the

restraint force) increases disproportionately.

Because of this, the steel

stress calculated for the critical section does not increase substantially
until concrete non-linearity comes into effect and the plot is discontinued
because the elastic sect ion analysis used is invalid.
The difference between the various calculation methods is much less for
the

concrete

stress

which,

using

critical for the restrained slabs.

BS

5400

serviceability

criteria,

is

The restrained analyses also converge

to give similar failure loads of around 70kN compared with 25kN for the
unrestrained

analysis.

Nevertheless,

the

difference

in

the

allowable

service loads implied using t"he stress at crack approach, 21kN, and the
smeared crack approach, 30kN, is disturbingly large and it appears prudent
to

use

the

former

approach.

It

should

be

noted,

however,

that

substantial part of the difference is due to the aforementioned effect of


the difference between the stress at the critical section and at the last
integration station.

This has two important

implications.

Firstly the

effect will be less pronounced under patch loads <as opposed to point or
line loads) so the difference between the two approaches will normally be
less than implied by this study.

Secondly, the peak concrete stress is too

localised for normal material models, based on the behaviour of specimens

- 118-

in a uniform state of stress, to be valid.

Because of this the use of the

stress at the critical section is conservative.


The analysis of these slabs appears to confirm that ignoring the tensile
strength

of

concrete

strength

of

concrete

may

not

does

be

have

conservative.
a

major

However,

beneficial

the

effect

tensile
on

real

structures which does not arise in the rigidly restrained slabs considered
here;

it improves the restraint.

Thus the tensile strength of concrete is

far less likely to have a detrimental effect on the stresses in realistic


bridge deck slabs.

7.8 SfRESS INTEGRATION

The

element

forces

element volume.

are obtained

by

integrating

the stresses over

the

The stress is taken to be constant over element width and

the integration over length is performed using two integration stations at


the Gauss points, 21% of element length from each end.

The forces at the

nodes are then obtained as a function of the forces at the integration


stations using the shape functions.
In analyses of this type, it is usual to perform the stress integration
over

depth

numerically

sometimes as
smooth

few as

curve

with

high

order

integration

five sampling stations.

between

the

stations.

As

the

function

This effectively
stress

and

fits

functions

used,

particularly for concrete in tension, are highly discontinuous it appears


that this could lead to significant errors and Ganaba and May<114) have
confirmed this.

In many of the sections considered in this study, with

their very light reinforcement, a five point integration scheme gave only
one station in uncracked concrete.

This, combined with the fact that the

tension stiffening function used was more discontinuous than that favoured
by

Ganaba

and

May,

suggested

that

the

integration

errors

would

be

particularly significant.
Two solutions to this problem were used.

For analyses which did not

consider stress history, an exact analytical integration was developed.

In

addition to eliminating integration errors, this was significantly faster


than

numerical

integration.

However,

neither

this

solution

nor

that

suggested by Ganaba and May <splitting the integration at the r.oot of the
crack),

could

be

used

for

stress history analyses.

It

was

therefore

decided to increase the number of integration stations from five to eight,


- 119-

to improve accuracy, and to change from high order Newtonian integration


to trapezoidal integration to reduce the effect

of the discontinuities.

The stresses in the down-stand beams were integrated separately using the
same integration scheme.
Comparison with the exact analytical version of the program showed that
these changes made the integration errors in the analyses of structures
insignificant compared with the other errors.

However, it appeared that

this was partly because the errors were essentially random and so tended
to cancel out;

the error in the forces calculated for a single element

could still be significant.

This tendency of the errors to cancel out

explains why, despite the large errors observed by Ganaba and May in the
forces calculated for individual elements, other analysts [such as Abdel
Rahmen<87)l have found their results to be insensitive to the number of
integration stations used.
In the analyses of the constant moment regions considered in Chapter 6
there was no scope for the integration errors to cancel out so they could
be

more

significant.

Because

of

this,

and

because

"accuracy"

was

considered more important for a fundamental study of tension stiffening, a


special version of
trapezoidal

rule

the program was developed which employed 32

integration.

This was

used

for

all

point

the analyses

in

Chapter 6, except those which did not consider stress history and so could
be performed with analytical integration.
eliminated integration errors completely.

For practical purposes, this


Indeed, since they were spaced

at only a quarter of the maximum aggregate size, the integration points


were unrealistically close.

However, because the maximum historic strains

are stored for all the integration stations, this version of the program
required more storage space as well as more computer time and it was not
used for the analysis of more complex structures.

7. 9 SOUJI'ION

~HEME

In non-linear analysis,

the

forces can be calculated directly from the

displacements but the displacements can only be obtained from an iterative


solution

scheme.

Incremental

iterative

schemes

are normally

used to

enable the behaviour of the structure under increasing loads to be studied


and also because the behaviour is sometimes "path dependent" so analyses
using very large increments could give incorrect solutions.
- 120-

A detailed study of solution schemes is beyond the scope of this thesis.


However, some problems were experienced which are peculiar either to the
type of structure considered or to the form of analysis used.

These will

now be considered along with a brief review of the scheme adopted.

7.9.1 Control
Solution schemes using displacement<97) or arc-length(ll5) control are now
favoured

by

analysts

but

this

has

arisen

primarily

because

of

the

It is difficult to

concentration on ultimate and post-ultimate behaviour.

achieve convergence with analyses using load control as failure approaches


and

impossible

through".

to

model

softening,

post-ultimate

behaviour

or

"snap-

However, with the type of structures considered in this study,

ultimate strength is a secondary consideration and neither "post-ultimate


behaviour"
because

nor

the

true

"displacement

failures

are

local and

energy is stored in the beams.


under

perfect

control"

displacement

suddenly and completely.

have .much

brittle whilst

physical

most

of

meaning

the strain

Thus, even if the bridges had been tested

control,

the

slabs

would

still

have

failed

Also, the temporary reduction in load which can

occur under monotonically increasing displacements as cracking occurs has


no

practical

control.

significance since

real

structures are

loaded

under

load

There is thus little practical advantage in departing from using

load control, at least in a pragmatic study such as this.

As structures

are designed for specified loads, and neither strength nor displacement are
critical

design

criteria

for

the

type

of

structures

considered

here,

analysis under load control is far more convenient for use in design.

7.9.2 Initial Stiffness Method


As

serviceability

criteria

are

critical

there

is

no

need

to

take

an

analysis for design up to failure, only to design ultimate load which is


just 30% above design service load in BS 5400.

Since this is normally

well below the actual collapse load, the demands on the solution scheme
are

comparatively

modest

so

relatively

simple

scheme

can

be

used.

Edwards used the simplest possible scheme; the initial stiffness method
with no accelerators.
degree of

freedom

In this approach, which is illustrated for a single

system

matrix is used throughout.

in

Figure

7.8,

the

initial elastic stiffness

The displacements are calculated from the

loads using the inverted initial stiffness matrix.

The forces are then

calculated

non-linear

from

these

displacements,
- 121 -

using

the

material

properties, and compared with the applied loads.

The difference

(which

represents the forces released by cra cking, crushing and yielding> is then
used to calculate a new set of displacements which are added to the first
set.

A new set of forces is then calculated for these displacements and

the whole process is repeated until the forces match the applied loads.
results

The

are

then

printed

out

and

the

next

increment

of

load

is

applied.
The approach has the advantage of being numerically stable and reliable as
well as simple.

However, if the actual tangent stiffness matrix of the

structure is substantially different from the initial stiffness matrix, the


convergence rate is very slow.

This normally occurs as failure approaches.

However, in the case of some of the slabs considered in this study, the
very low steel areas meant that cracking changed the stiffness so much
that

the

convergence

rate

service load was reached.

became

excessively slow

even

before design

In a typical analysis of a 25 node model, over a

hundred iterations per increment were needed.

This, and the desire to use

the analysis as a research tool <which meant that failure behaviour had t o
be considered ) me an t t hat the convergence rate had to be improved.
!..oad

:n.:r eroem:s )

---------- Analysis
-

- Str uc ture

0 ~----------------------------~--------------

Displacement
Figure 7.8:

Initial stiffness method

7.9.3 Accelerat ors


The first modification to improve convergence was to use a simple approach
suggested by Cope et al<37>.

In this, the displacements due to the last

load increment are used as the first estimate for the displacements due to
the

present

load increment

as

illustrated in Figure 7 .9.


- 122-

Because the

stiffness

of

the

structures

considered

tended

to

degrade

reasonably

progressively, this meant that the first estimate was much closer than it
would have been if calculated from the initial stiffness matrix.

Thus the

number of iterations required to achieve convergence was much reduced.


This approach is particularly effective in analyses using very low or zero
tensile

strength

for

concrete

because

the

displacements

due

to

each

increment are then equal until reinforcement yielding or concrete crushing


occurs.
Although

this

modification

greatly

reduced

the

number

of

iterations

required, it was still excessive so a number of acceleration schemes were


considered.

Some were found to be very effective on some structures but

they had erratic results and prevented analyses of other structures from
converging

altogether.

procedure instead.
stiffness
optimised.

matrix

Eventually, it was decided to use a "line search"

In this, the displacement vector calculated from the


is multiplied

by

scalar

factor

and

this

factor

is

In co-ordinate geometry terminology, the stiffness matrix is

used only to obtain the search direction in "n" dimensional space and the
line search attempts to find a scalar multiplier for this vector such that
the component of the error energy in that direction is zero.
Load
~ ncr<?ments>

- - - - - Ana l ysis

- Struc ture

0 ~------------------------------------Displac ement
Figure 7. 9:

Modi f ied initial stiffness method

Since the value of this scalar can only be obtained iteratively, which
involves

calculating

all

the

element

forces

calculation would require excessive computation.

for

each

iteration,

exact

However, by using a very

slack optimisation criterion, the number of iterations or "searches" can be


- 123-

reduced.
better

In this study, the criterion used was that attempts to obtain a


scalar

factor

were

made

only

if

the

sum

of

the

error

forces

multiplied by the ... r respective iterational displacements, that is the error


energy

in

the

search

direction,

was

in

excess

of

60%

of

similar

summation performed using the error forces from the previous iteration.
The

line

required.

search

procedure

greatly

reduced

the

number

of

iterations

However, because of the computer time used in the line searches,

the effect on the time to achieve convergence was less dramatic although
still very significant.

Perhaps more importantly, the procedure means that

when the structure has failed the analytical deflections become very large.
With a pure initial stiffness scheme, failure was sometimes indicated only
by

failure

of

distinguish

the

failure

analysis
of

the

to

converge

structure

which

from

made

it

difficult

numerical problems

with

to
the

program.
Line

searches

are

used

in

most

recent

NLFEA

programs,

sometimes

combination with other more sophisticated acceleration schemes.

in

However,

for the analysis of cracking, they do have a theoretical fault which does
not appear to have been fully resolved.

When a crack first occurs, the

true displacement is greater than predicted by the stiffness matrix so a


line search factor substantially greater than one is applied to all the
displacements.

This can cause cracking in elements which were previously

uncracked and in perfect equilibrium.

The cracking leads to error forces

which are eventually reduced by the iterative solution scheme.

However,

this could be done by increasing the deformations until the force is taken
up

by

the

uncracked

reinforcement,
state.

The

rather

than

fundamental

by

returning

problem

is

the element

that

there

can

to

its

be

two

different deformation states in a section which give the same forces; one
cracked and one uncracked.

The initial stiffness method always under-

estimates displacements and so always arrives at the uncracked equilibrium


state first.

However, once a line search is included in the analysis, it is

theoretically possible

for

the analysis to predict

cracking in concrete

which has never been stressed up to its tensile strength.

In practice it

was found that this did not occur in the analysis of highly redundant slab
systems; analyses with the line search converged on the same solution as
those without.

However, it did arise in the analysis of direct tension

tests using variable tensile strength.


was not used in the analyses for 6.3.1.
- 124-

For this reason, the line search

Line searches, and other accelerators, can give displacements

<and hence

strains) within an iter at ion which exceed the final equilibrium values.

If

these strains were used in the stress history analysis, false results could
be obtained.

For example, concrete could be taken to have cracked, and

thus to have lost most of its tensile strength, as a result of strains


which only occurred in iterations which had over-shot the true solution.
To avoid this, the maximum strains are updated only after convergence has
been achieved.

7.9.4 Stiffness Recalculation


With these improvements, the convergence rate was acceptable for small
problems and for analyses for design.

However, it was still too slow to

use the program to analyse large computer models up to failure.

It was

therefore decided to depart from using the initial stiffness method and a
numerical

recalculation

program.

Ideally, this should calculate the exact tangent stiffness for

the

current

represents

of

deformation

the

the

stiffness

state

structure's

so

response

that
to

matrix

the

was

added

stiffness

small

changes

of

into

the

matrix

truly

load.

Some

analyses <98) have been performed using a "Newton-Raphson" approach,

in

which the stiffness matrix is recalculated for every load increment or


even every

iteration.

This

approach gives

much reduced number of

iterations but the computer time required to recalculate and invert the
stiffness matrix more than uses up that saved by reducing the number of
iterations.

In the analyses of cracking, the true current stiffness matrix

can also contain negative diagonal terms which would lead to numerical
instability.
For

these

reasons,

in

the

present

study

the

tangent

stiffness

was

calculated only infrequently and approximately and the concrete was always
given a significant positive stiffness; usually not less than 3% of the
full elastic value.

It appears that most studies have attempted to obtain

a closer estimate and used a lower tangent stiffness for cracked concrete.
This

is

possible

in

an

analysis

under

monotonically

increasing

loads.

However, when unloading is considered, it leads to complications since the


material

models

used

give

different

tangent

whether the strain is increasing or decreasing.

stiffnesses

according

to

Thus the exact tangent

stiffness matrix can only be calculated if the direction of change, as well


as the value, of the strain is known for all the sampling stations.

- 125-

It

proved

much

simpler

to

use

only an

approximate

calculation giving

stiffness matrix which could be used for both loading and unloading.
Having

adopted

periodic

recalculation

of

the

stiffness

matrix,

necessary to adopt a criterion to decide when to do this.


approach

is

to recalculate at

the

beginning of an

it

is

The usual

increment

if some

"current stiffness parameter" is substantially different from that implied


by the stiffness matrix currently in use.
it

However, in the present study,

was found that this approach did not work very well.

cracking occurred

in

particular

increment

the

always recalculated for the next increment.

If extensive

stiffness

matrix

was

However, if little further

cracking occurred in that increment, the use of the displacements due to


the last increment as a first estimate for the displacements due to the
. current increment meant that the analysis would converge quickly whatever
stiffness matrix was used; provided the previous increment had converged.
Thus recalculating the stiffness merely wasted time.

If, however, the

previous increment had not converged, it would have been better to make it
converge by.' recalculating the stiffness matrix earlier.

Thus it was found

more satisfactory to recalculate during the increment.

It was decided to

do this at

iteration eight

if the convergence rate was slow and the

remaining errors were significant.


The_ choice of iteration eight was a compromise between early recalculation,
which could

mean

unnecessary

recalculation,

and

delaying

recalculation

until much computer time had been used up in iterations using the old
stiffness matrix.

However, late recalculation has the advantage that the

deformation state, and hence the calculated tangent stiffness, is closer to


that in the final equilibrium state so the final convergence tends to be
faster.
The

stiffness

matrix

recalculation

improved

the

although not by as much as the line search.

rate

of

convergence

However, the greater effect

of the line search may not indicate that it is a superior method; rather it
appeared to be due to the line search having been incorporated first.
recalculations

had

much

greater effect

analyses performed without the line search.


effect of
structures,

recalculating
being

the

generally

cracking in the beams.

stiffness
greatest

the

This implies that

convergence

rate of

It was also apparent that the

matrix

where

- 126-

on

The

the

varied

greatly

softening

was

between
~ue

to

the details of the pptimum

solution scheme, such as when to recalculate the stiffness matrix, are


different for different structures.

It was thus clear that the solution

scheme adopted was not the optimum for all the structures considered and
there was certainly scope for improvement.

However, the convergence rate

achieved was considered acceptable for the project.

7.9.5 Convergence Criteria


In an iterative solution scheme, it is necessary to adopt a criterion to
decide when the solution is sufficiently accurate to stop the iterations,
without

knowing

the exact

forces,

iterational displacements

energy> can be used.


local

convergence.

solution.
or

Criteria based on out-of-balance


the

product

of

the

two

<that

is

It is also possible to consider either overall or

Analysts

tend

to

favour

overall

energy

criteria,

primarily because finite element analysis is an energy based approximation


method and there is a useful norm with which to compare the error energy;
the work done by the loads on the structure.

However, in the present

study two difficulties were experienced with energy criteria.

Firstly, as

Cope and Cope<94> have noted, the in-plane forces tend to be the last to
converge and,
flexural

since the

stiffness,

significant

in-plane

the

in-plane stiffness is large compared with the


energy

error

associated

forces

can

converged according to energy criteria.

with

these

remain

in

is

small.

analyses

which

Thus
have

Although these forces might be

considered unimportant, since eliminating them usually has little effect on


the displacements, they can represent a significant force in the critical
elements.

Thus, if local stresses are to be used as design criteria, it is

important to limit the error forces.


The second problem encountered is a peculiarity of the type of structure
considered.

The failures were local and brittle.

The energy associated

with a failure, an individual wheel load multiplied by the displacement of


the slab relative to the beams, thus represented only a small fraction,
typically

1~,

of the total work done by the loads.

The combined effect of

these problems was that there could be significant local force errors in
an analysis when the error energy was less than

0.0001~

of the work done

by the loads.
Another disadvantage of both energy and displacement criteria is that they
depend on the iterational displacements which (unlike the out-of- balance
- 127-

forces) are a

function of the solution scheme used as well as of the

displacement.
desirable

Thus,

to

use

iterational

if

the

tighter

displacements

initial

stiffness

energy

convergence

systematically

method

is

used,

criterion

under-estimate

it

is

because

the

the

true

displacement errors.
The major difficulty with force criteria is defining a norm with which to
compare

the

out-of-balance

forces.

The

standard

of

comparison

for

moments and axial force has to be different otherwise the criteria become
dimension dependent.

The out-of-balance moments could be compared with

the maximum element moment.

However, in the type of structure considered

here, this would lead to either unduly slack criteria for the slabs or
unduly severe criteria for

the beams.

Comparing

in-plane forces

with

maximum element forces is even less satisfactory because the axial force
in a slab element is obtained from the difference between similar tensile

and compressive forces.


cracking

is

not

In the early stages of a slab analysis, when the

extensive,

the

in-plane

forces

are

very

small

so

criterion based on a percentage of these forces would be unduly severe.


Conversely, in an analysis of a beam and slab deck, the axial forces in the
down-stand beams are too large to use as a standard of comparison.
Consideration of these problems led to the decision to use both an overall
energy

and

local

force

convergence

criterion.

stopped only when both criteria were satisfied.


based

on

comparison

structure whilst
avoid

dimensional

with

the

total

work

The

iterations

The energy criterion was

done

by

the

loads

on

the force tolerances were specified by the user.


problems, separate

force

were

and

moment

crite~ia

the
To
were

specified.
Despite

using

very

tight

energy criterion,

typically

0.01~,

and the

slackest force criterion considered reasonable, the in-plane force criterion


was nearly always the last to be satisfied.

7.10 CALIBRATION
Although

the program

was not

intended

to be highly accurate,

it

was

considered desirable to check it by comparison with test results and other


analyses to ensure that the results were reasonable.

In addition to the

studies mentioned in 7.3, 7.7, and also Chapter 6, as well as a. check


against

linear

grillage

to

ensure that
- 128-

the

program

was 'at

least

numerically correct,

number

of structures which

had been

tested

by

others were analysed to investigate the behaviour.

7.10.1 Duddeck's Slabs

Duddeck<116) tested a series of three square corner supported slabs under


single

central

point

loads.

These

have

been analysed

by

both

Abdel

Rahmen<87) and Cope and Cope(94) using non-linear plate finite element
programs.

Thus they enabled the program to be compared both with test

results and with more sophisticated analyses.


Two of the slabs were analysed using a four by four node quarter model.
The results for the first slab, which had 0. 7% isotropic reinforcement, are
shown in Figure 7.10 and, for comparison, Abdel Rahmen's results are also
shown.

Duddeck gave little material data so Abdel Rahmen used Mueller's

estimate <117) for the tensile strength.

In order to make the analysis

directly compara ble the author used the same figure, but it is improbably
low for the quoted compressive strength which probably explains why Abdel
Rahmen's analysis under-estimates stiffness.
study

under-estimates

stiffness

still

At

more.

low loads, the present


This

might

be

expected

because of the torsionless elements, particularly as the principal moment


direction

in

the

critical

area

is

at

45

to

the

elements,

the worst

possible direction.
Both analyses give good predictions for the failure load.

The present

study is marginally conservative but this is not significant compared with


material variability.

--

:..oad <kN ) 60

50
40

30

- - - -Test

20

10

-----Abdel Rahmen's Analys is

- Author's Analysis

10

20

30

40

50

Deflection (mm)
Figure 7. 10:

Analysis of Duddeck's slab 1

- 129-

Duddeck's second and third slabs had the same total quantity of steel as
the first but in orthotropic arrangements.

The results for the third slab,

which had 1 0% steel in one direction and 04-% in the other, are shown in
Figure 7.11.

Abdel Rahmen's results for this are far less satisfactory;

they significantly over-estimate the strength.

In contrast, the present

study gives better results for this slab than for the first.

Both these

changes are due to the fact that, as failure approached, the direction of
maximum princ ipal moments rotated towards the direction of the heavier
reinforcement.

Since the use of torsionless elements in the present study

means that the principal moments in the analysis act in this direction
from

the

outset,

the

rotation

Conversely, Abdel Rahmen's

improves

the

realism

of

the

analysis.

analysis correctly predicted that the principal

moments in the uncracked slab, and hence the initial cracks, would be at
45 to the reinforcement.
was then fixed.
rotated,

However, it assumed that the crack direction

As failure approached, and the principal moment direction

the shear retention

factor

in the analysis gave a significant

shear stress across these cracks which implies a significant tension in the
reinforcement
approximately

direction.

In

perpendicular

to

yielded, they became very wide.

fact,
the

new

cracks

formed

secondary steel and,

which

as

were

this steel

Thus the real concrete was incapable of

resisting tension in this direction so the slab was weaker than Abdel
Rahmen's analysis suggested.

Cope and Cope<94) have shown that this fault

can be avoided, either by using a shear retention factor which becomes


very

low at high smeared tensile strains or by using a rotating axis

material model.

However, Abdel Rahmen failed to identify the cause of the

error and it illustrates the danger of using this form of analysis in the
design o f even simple s labs without some calibration against tests.
45

Load <kN > 40


30

- -...;Test

20

---Author's Analysis

10

--- -- Abdel Rahmen's Analysi s


0

10

20

30

40

Deflect ion (mm)


Figure 7.11 :

Analysis of Duddeck's slab 3

- 130-

Abdel Rahmen noted that both the tests and his analysis gave failure loads
for all three slabs which were higher than predicted by yield-line theory.
He attributed this to the contribution of the tensile strength of concrete.
However, even with the tensile strength set to zero, the author's analysis
gave

failure loads which were higher than yield-line predictions.

The

reason for this is that, as in the slab strips considered in Chapter 6, the
depth of concrete in compression was substantially less than the depth to
Thus the strength was enhanced by the tensile force in the

the top steel.


top steel.

Although, with the top steel removed, the non-linear analysis gave almost
identical failure

loads to yield-line theory, it did not give the same

moment distribution.

At peak load, it predicted a moment in the element

under the load which was substantially above the yield line value; the
extra strength coming from a net compressive force on the element.
force was resisted by tension
moments.

in outer elements which resisted lesser

Thus the analysis suggested that compressive membrane action

affected the behaviour of even these unrestrained slabs.


be

This

confirmed

by

other

test

results.

For

example,

This appears to
Regan

and

Rezai-

Jorab1<118J measured strains in the reinforcement of a one-way spanning


slab

subjected

transverse

to

single

reinforcement

transverse curvature.

concentrated

indicated

that

load.
there

The
was

strains
very

in

the

significant

Thus the longitudinal curvature must have varied

significantly

over

the

slab width;

reinforcement

did

not

vary

yet

the

significantly

strain

over

in

slab

the

longitudinal

width.

The

only

possible explanation for this appears to be that the neutral axis depth
varied across the slab width because of the compressive membrane force in
the centre of the slab and the tension at the edge.
7.10.2 Taylor and Hayes' Slabs
Taylor and Ha yes (55 J tested a series of square slabs under single central
concentrated

loads.

These

enable

the

program

to

be

assessed,

by

comparison with test results, for both restrained and unrestrained slabs.
a. Unrestrained Slabs
Taylor and Hayes tested a series of slabs with two different reinforcement
percentages

<0.9~

and 1.8%) under patch loads of three different sizes.

These were all analysed using both a four by four and a five by five node
quarter

model.

A typical

load

displacement
- 131 -

relationship

is

shown

in

Figure 7 . 12.
predicted

will be seen that

It

well.

The

two

the displacements at

analyses

gave

very

similar

low loads are

deflections

but,

because it modelled the stress concentration under the load, the finer
mesh

always gave

slightly

lower

failure

load;

the

difference being

greater with the smaller load patches and the heavier r einforcement.
Load <lrN J 100
80
60

;""
/

40

Test

/
/

----

20

Ana lysis

0
6

10

Deflect ion (mro )


Figure 7. 12:

Analysis of Taylor and Hayes' slab 254

The predicted failure loads were generally lower than the actual failure
loads,

which

discrepancy

is
was

Nevertheless,

desirable
up

to

in
30%,

an

analysis

which

f or

might

design.

be

However,

considered

the

excessive.

the analysis still gave failure loads which were typically

30% higher than are implied by the elastic analyses currently used in
bridge design.
use

of

the

Thus, even though arguably excessively conservative, the

analysis

in

bridge design

would still l e ad t o significant

economies compared with current practice.


Although

the

ratio of predicted to actual failure

cons i s tent,

with

s ystematic

fault

coefficient

of

variation

of

load was reasonably


approximately

7%,

could be observed in the res ults ; the analysis under-

estimated the effect of load patch size.

It gave errors in the unsafe

direction in only one case; the heavily reinforced slab with the very small
load pat ch.

This was a rather extreme example with a load patch only

50 mm acr oss , compared with the slab thickness of 75 mm.

This, combined

with the heavy reinforcement, resulted in very high stresses round and
under the loaded area at

failure.

The calculated shear stress on the

critical section at the fa ce of the load reached 6N/mm 2


s tress under the load was 31N/mm2
analysis

which

ignores

these

and the vertical

It is thus not surprising that an

stresses

over-estimates

the

strength.

However, an alternative explanation is that since even the finer mesh gave
- 132-

elements which were over twice as wide as the load patch, the analysis had
failed to model the stress concentration round the load.

To test this,

the slab was re-analysed using a nine by nine node quarter model.
gave a significantly lower failure load, below the actual value.

This

For other

slabs, with larger loaded areas, it gave only a very slight reduction in
failure load.

It

thus appears that in order to correctly model failure

load without a separate shear check there is an additional criterion for


the size of the elements in the critical area; they should not be much
bigger than the loaded area.

This criterion would be difficult to comply

with in the analysis of complicated structures.

However, further tests

showed that a rather coarser mesh can safely be used if the concentrated
load

is applied at a single node, rather than being distributed in an

attempt to model the patch size as it was in the analysis of Taylor and
Hayes' slabs.
In contrast to Duddeck's slabs,

the failure loads for Taylor and Hayes'

unrestrained slabs were lower than predicted by yield-line theory.

They

said that this was because the slabs failed in punching shear, rather than
flexure.

However,

the

analysis

suggested

that

essentially brittle bending compression failures.


grade concrete <typically 30N/mm2
higher

steel

percentage,

the

slabs

were

than Duddeck's.

However,

Petcu

Stanculescu's<79>

ductility

theory.

failures

were

Because of the lower

compared with 43N/mm 2

reinforced
and

the

effectively

as well as the

),

far

more

heavily

they were still only just outside


requirement

for

using

yield-line

The analysis predicted, apparently correctly, that the behaviour

would be less ductile


critical

section,

under

compressive force.

than Petcu and Stanculescu assumed because the


the

load

patch,

would

be

subjected

to

net

This effect does not appear to have been considered

previously, apparently because this type of failure has been attributed to


shear.
loss

The analysis slightly over-estimated the detrimental effect of this


of

ductility

on

strength.

Other

reasons

why

the

analysis

was

conservative for these slabs, and more so than for Duddeck's, include the
under-estimate of concrete crushing strength due to ignoring the multiaxial

stress

state

<which

has

greater

effect

in

more

heavily

reinforced slab), the torsionless elements' failure to model the diagonal


hogging moments in the corners <which do not arise in a corner supported
slab) and the use of a finer element mesh relative to load patch size in
analysing Taylor and Hayes' slabs.

Despite all these faults, the analysis


- 133-

was entirely satisfactory from a design point of view and its use would
lead to significant economies compared to current practice.

b. Restrained Slabs
In addit i on t o the unrestrained slabs, Taylor and Hayes tested restrained
s labs .

They tested a series to match the unrestrained set plus a set of

othe rwis e s imilar unre inforced slabs .

These were also analysed and t ypical

results are s hown i n Figure 7 .1 3.


:..oad ( kN ) 140

120
100

80

I/~

60
40

--

Test

- - - - Analys is <full restraint )


- - - --Analysis ( in plane r estraint onl y )

20
0

10

Deflection (mm)
Fig ure 7.13:

In

the tests,

the

Analysis of Taylor and Hayes ' s lab 2R4

restraint greatly

increased the

init ial analys i s predic ted a much smaller effec t .

failure

loads.

The

This appe ars to be due

t o differences between the real and assumed restraint conditions.

Taylor

and Hayes used a steel frame to provide the restraint and the slab was
ins erted j ust prior to the test, the gaps being packed out with mortar.
This

was

apparently

intended

to

give

full

in- plane

restraint

with

negligible rotational restraint so these restraint conditions were used in


the analysis.

In fact, there clearly was significant rotational restraint;

hence the greater than predicted increase in strength.

However, when full

rotational restraint was used in the analysis, it over- estimated strength.


It appears that the steel frame gave partial restraint to both in-plane

and rotational movement.

This could not be predicted satisfactorily by the

analys is.

- 134-

7.10.3 Batchelor Md Tissington's Speciaens

The analysis of Taylor and Hayes' slabs confirmed, as had been found in
designing the specimens for Chapter 6, that it is difficult to produce
known restraint conditions artificially.
not

It therefore appeared that it was

possible to check the analysis of compressive membrane action for

simple

laboratory

specimens

before

complicated bridge structures.

going

However,

on

to

use

it

to

analyse

Batchelor and Tissington have

tested a series of simple bridge models with only two beams each and
these provided a useful intermediate case.

They also had the advantage of

having been analysed by Cope and Edwards(99) so they enabled the analysis
to be compared with a plate type finite element program.
Ba t chelor and Tissington's largest specimen is illustrated in Figure 7.14
and the load-deflection response under a single central load is illustrated
in Figure 7.15.

It will be seen that Cope and Edwards' analysis gives good

results whilst the author's, a s expected and intended, is conservative.


0 )J"I. osotropoc
r eonforcement

IDI~Dilf~.,..
I

L060

~~~-J'\ As=920mml
fy : )10N/ mm 1

u
Fi gure 7. 14:

Ba t chelor and Tissington's specimen

The analysis also gave good

predictions

for

the cracking response. It

predicted cracking due to hogging moments along the edge of the slab,
where these moments are resisted only by torsion in the beams.

It also

predic ted, as observed in the tests, that just before peak load was reached
the main beams would crack right through under the restraint forces.
Unlike Taylor and Hayes' slabs Batchelor and Tissington's, with their large
span to depth ratio, reached deflections which were significant compared
with

their

thickness

before

failing .

Thus

the

displacements

had

significant effect on the lever arm at which the restraint force acted.
However, in order to make the analysis directly comparable with Cope and
Edwards', the correction for this, which was described in 7.6, was not used
in the analysis shown in Figure 7 .15.

The analysis was, however, repeated

- 135 -

with the correct ion.

This increased the deflection at a load of 20kN by

only 1% and at 30kN by 6%.


25%.

However, it r educed the failure load by some

This implies that both the author's and Cope and Edwards' analyses

under-estimated the basic static strength of the slab.

This may have been

due partly to under-estimating the material strengths since Batchelor and


Tissington gave little data for this; for example, they gave no indication
as to whether the reinforcement strain hardened.
Lo ad <kN )

80

60

4- 0

- - - - Test
Author's Analysis

20

- - - - Cope & Edwards ' Ana l ysis

10

15

20

Deflec t ion (nun )


Figure 7. 15:

Analysis of Bat chelor and Tissington 's Specimen

7 .10.4. Kirkpatrick's Model


Kirkpatrick's bridge model, which was considered in Chapter 3, enabled the
program's prediction of local bridge deck slab behaviour to be compared
both with test results and with Cope and Edwards' analysis.

A major problem with the analysis of this type of structure is the size of
the computer model required.

In order to model local effects safely the

element mesh has to be fine.

In order to obtain the correct restraint the

whole of the bridge should be represented, even in an analysis for local


effects.
which

To reduce the computer time required,

restricted

it

to

the

analysis

of

a half model was used

symmetrical

load

cases.

The

computer model was also banded so that only the half of the loaded slab
span was modelled with a fine enough mesh to represent local behaviour.
Despite this, a model with 288 nodes was required.

The banding, combined

with the variety of reinforcement areas and different slab spans used by
Kirkpatrick, meant that the model required 53 element types.
The result of the analysis of bay C2, which had 0.5% reinforcement, is
shown

in

Figure 7. 16.

The analysis gave


- 136-

conservative

predictions for

deflection and strength.

It was slightly more conservative than Cope and

Edwards', although this was partly due to including the effect of large
displacements.

The analysis of bay A2, which had 1. 7% reinforcement, also

gave conservative deflection predictions but over-estimated failure load by


some 10%.

Although a 10% error would normally be considered acceptable,

as it is small compared with the variability of concrete behaviour, in


combination
program

with

the 20% under-estimate for C2

over-estimated

the

effect

of

it

indicated that

increasing

the

the

reinforcement

However, a study of the results revealed that much of the

percentage.

increase was due to the contribution of the top steel, not only in tension
over the beams but also in compression under the load.

The latter is

unusual; in a thin bridge deck slab the cover required usually means that
the steel on the compression face is too near the neutral axis to make a
sig nificant contribution.

Load CkN ) 100

--

80

60
40

- -- - -:rest
Author's Analysis

20

- -- - Cope & Edwards' Analysis

Deflec tion <mm >


Figure 7. 16:

The

steel

in

Analysis of Kirkpatrick's panel C2

Kirkpatrick's

model

was

given

only

6mm

cover

which

is

equivalent to 18mm at full size; approximately half the cover required by


BS 54-00.

Kirkpatrick was careful to maintain and check the bottom cover

but attributed less significance to the top cover.

Thus the true top

cover is uncertain and the effect of the top steel is very sensitive to
its position.

The steel would need to drop only a few millimetres to

explain the discrepancy.

Another effect of having significant compression

steel is that, unlike the other analyses, the analysis of this bay was
- 137-

sensitive

to

the assumption

compressive strains.

made

for

the stress

continued

to

reinforcement.

increase

as

compressive

However, in this bay the


force

relatively low strain.


effect

of

analyses.

transferred

to

the

Thus the unsafe prediction could have been avoided by

using a concrete model which gave an abrupt

the

high

Other analyses failed soon after the. compressive

stress in the critical region started to reduce.


load

in concrete at

reduction in stress at a

This may explain why the tendency to exaggerate

reinforcement

was

less

pronounced

The greater ductility given

by the

in Cope and Edwards'


relationship shown

in

Figure 7.6 may be more representative of the behaviour of concrete loaded


uniaxially under displacement control.

However, the critical concrete in

these slabs was subjected to biaxial compression and also to shear which
reduced its ductility.
slab,

to

impose

compressive stress.

Thus it may be prudent, in the analysis of such a


limit

on

the

strain

at

which

concrete

can

carry

A limit of approximately 0.0045, which is still higher

than used in most analyses, would eliminate the unsafe predict ion for this
bay but have little effect on any of the other analyses.
that

However, given

this bay had more effective compression reinforcement and as much

tensile reinforcement as any practical bridge deck. slab, the analysis can
be considered safe for practical slabs despite over-estimating the effect
of steel.

It also appears, once again, that a flexural analys-is has proved

capable of predicting a "punching shear" failure.

This suggests that the

failures were primarily brittle bending compression failures, although the


shear force did precipitate the final collapse.
Although

the reasonably good predict ions for the failure

load of these

slabs are reassuring, they have little practical significance.

Kirkpatrick

acknowledged that design should be controlled by serviceability criteria.


Applying conventional 85 5400 stress criteria to the element forces from
the

analysis

of

approximately 22kN.

bay

C2

suggested

an

allowable

service

load

of

This compares with a design service 45 unit HB wheel

which, at this scale, is 13.75kN.

This is interesting as the reinforcement

in this bay was similar to Kirkpatrick's eventual recommendation.


analysis has given further support to Kirkpatrick's proposals.

Thus the

However, it

remains to consider the influence of global transverse moments.


From Kirkpatrick's observations of the behaviour of his model, and of his
full scale bridge,

it

would appear

that,

in the absence of any global

effects, the behaviour of this bay would certainly be satisfactory under a


- 138-.

service

load

significantly above

22kN.

analysis was unduly conservative.


it

should

be

noted

that

Thus

it

may appear

that

the

However, to put this into perspective,

conventional

analysis

of

this

bay,

using

Westergaard and BS 5400, gives an allowable service load of only 9kN and
implies a failure load of less than 14kN.

7.11 cor:L.USIONS

The

form

gives

satisfactory

predictions for the behaviour of realistic slab structures.

Provided an

element

of

mesh

concentrations

analysis

is

used

around

considered

which
the

is

applied

in

this

fine

chapter

enough

loads,

it

to

model

appears

local
to

stress

give

'~

safe

predictions for the failure loads even of slabs which fail in "punching
shear".
In some cases the analysis under-estimated strengths by up to

30~.

The

allowable service loads calculated from the analysis also appear to be


conservative.

However, despite this, the use of the program in design and

assessment would still lead to significant economies compared with current


practice.

- 139-

CHAPTER
MODEL

BRIDGE

TESTS

8.1 INTRODUCTION
The analytical methods considered in Chapters 5 and 7 are potentially very
useful, but they have not yet reached the point where they can justify
radical changes in design practice without some calibration against tests.
It was therefore necessary to perform some tests.

These were designed to

investigate the key areas identified in Chapter 3 as requiring further


research;

service

load

behaviour,

restraint

and

the

effect

of

global

moments.
8.2 DESIGN OF MODEl.S
8.2.1 Scheme
Although small scale models have proved successful for

predicting the

strength of slabs <51>, the cracking behaviour of concrete does not scale
well.

Thus,

in order

to obtain

reliable

predictions of

service

behaviour, it is desirable to use the biggest practical scale.


full size models would be used.

load

Ideally,

However, financial constraints on this

project, combined with the need to model a whole bridge and a whole HB
load, made this impractical.

It was therefore decided to use half scale

models of relatively small M beam type bridges.


these were

Analysis suggested that

the type of structures in which global transverse moments

would be most significant.


The first model, which is detailed in Figure 8.1 and illustrated in Figure
8.2, was designed to be a worst case for restraint so it had four beams
<the minimum practical number for a bridge of this type>, no parapet upstands and no diaphragms.

The last point is particularly significant since

3.2.8 noted that previous researchers have said that diaphragms are needed
to

provide

the restraint,

yet

no tests have

without diaphragms to confirm this.

been performed on decks

Also, analysis using the program

described in Chapter 7 suggested not only'that diaphragms were not needed


to provide the restraint but also that, because of the effect illustrated
in Figure 3.12 and discussed in 3.2.7, the slab near the ends of the bridge
would be subjected to transverse compression when the full HB load was
applied near mid-span.
- 140-

I
250 I

6000

250

ELEVATION

500

1000

10 00

1000

500

SECTION
Figure 8.1:

Details of first deck

Figure 8.2:

First deck under test

To give a worst case for local effects, the maximum practical beam spacing

was used with the standard slab thickness,

160mm at

full size.

The

spacing of the beams was li.mited by their shear strength under the design
~5

unit HB vehi.cle and the slab's span to depth ratio, although greater

than normal for this type of deck, was qui.te modest at 12.5.

However,

analysis suggested that larger, wider spaced beams would be a less severe
test because of the smaller global transverse moments.

As these moments

were a key area requiring investigation, it was considered better to use a


deck which was a severe case for these.

- 141 -

c
L

ELEVATION

50

, .... 225

1000

-(-- --.....

~-

..........

8000

400

1000

1000

SECTION

Figure 8 .3:

De tails of second d eck

1000

1!.

250

1 150
I

400

The second model, which is detailed in Figur e 8 .3 and illustrated in Figure


8 .4, was designed to be more typical of current practice so it had parapet
up- stands, support diaphragms and an extra beam.

The same beam spacing

was used as for the first deck and the overall width approximated, at full
size,

to

that

shoulders.

of

two

lane

bridge

for

footways

nor

hard

It thus represented the narrowest bridge which is likely to be

designed for 45 units of HB load.


case

with neither

restraint,

analysis

Although a narrow bridge is the worst

suggested

that

would have been greater in a wider deck.

global

transverse

moments

However, it was considered that

the behaviour of a wider deck could safely be predicted with the aid of
the results of tests on the deck and the program described in Chapter 7 .
It was not possible to test a wider deck in the laboratory.

Figure 8.4:

Second deck under test

After the two models had been tested, a single beam with the appropriate
width of slab was tested on its own to help calibrate the analysis.

8.2.2 Beaas
Since the slab behaviour was the main concern of the project,
modelling of the beams was not required.

perfect

However, pre- tensioned beams

were used so that the global behaviour was reasonably similar to that of
the prototype bridge.
half scale models of

Standard inverted T beams were used as approximate


M beams.

These had the s ame advantage in the

research project which they have in practice; the multiple use of formwork
makes them much cheaper than specials.

- 143-

In Figure 8.5 the section of the inverted T beam is compared with a true
half scale M beam.

The thicker web of the inverted T beam was considered

an advantage as it was desirable to avoid shear failures.

However, the

lack of rebates for the slab formwork was a disadvantage, not only because
they were needed to support the formwork, but also because their absence
improved the support to the slab.

Thus non-standard rebates were provided

as shown in Figure 8.5.

25

~L
0

n
'

.. ''

103

"'

100

!9

0
Ll"l

40

53

0
0
0
Ll"l

,------

0
0

0
~

......,

C)
C)

Ll"l
N

C)

<X)

'\.

'
248

Figure 8.5:

When

243
T2

M4

<full size>

<half size)

Comparison of full size T2 and half size M4 beams

the design of the beams was fixed,

which was very early in the

project, it was considered desirable to avoid global failures so the beams


for the first deck were provided with approximately 25% more prestress
than the conventional BS 5400 based design method required.

The beams

for the second deck were provided with the same prestress which, because
of the improvement in distribution properties due to the diaphragm, meant
they had approximately 35% more steel than BS 5400 would have required.
Because of the interest in the interaction of global and local effects.
under service loads, it was desirable to provide a realistic beam size near
- 144-

the minimum which could be used, within the code, for this type of deck.
Thus

the beam size was not

increased

to match

over~provision

the

of

prestress so the beams were stressed at transfer to a higher stress than


would normally be allowed.
The

shear

reinforcement

was

designed

to

the

normal

BS

5400

rules.

Because Hughes(1!9) has found that these are conservative for this type of
beam,

the

shear

reinforcement

was

not

increased

to

match

the

over-

provision of prestress.
The beams

for

the second deck were provided with standard transverse

holes to accommodate reinforcement for the diaphragms.

In order to get

the diaphragms down to the correct scale size, the holes had to be nearer
to the end than is recommended by Green<120>, so extra links were provided
to control the expected cracking.
6.2.3 Diaphragms
The diaphragms

for

the second deck were designed to the conventional

BS 5400 rules.

However, a considerable variety of approaches are used for

calculating the torsional inertia used in the analysis to obtain the design
moments.

This significantly affected the design.

It was decided to follow

the recommendations of Clark and West <121> and use half the Saint Venant
value for the gross-concrete sect ion.
6.2.4 Slab Re:lnforcement
Because

global

investigation,
reinforcement

behaviour
it

was

and

restraint

considered

were

that

using

percentages was undesirable.

major

areas , requiring

bays

with

different

The more heavily reinforced

bays would have provided extra restraint and distribution which would have
given an optimistic impression of the behaviour of the lightly reinforced
bays.

This meant the choice of steel area was important.

The original idea was to provide the first deck with 6mm high tensUe
steel bars
secondary

at
steel

lOOmm
in

centres
both

<that

faces.

is T6-100) main
This

compares

steel and T6-125


with

Kirkpatrick's

recommendations <13) which are equivalent to T6-75 at this scale.

However,

later analysis suggested that even .T6-100 was slightly excessive and it
was decided to reduce the ma:ln steel to T6-125 as well.
deck

was

20%

less

heavily

reinforced
- 145-

than the- 'strips

This meant the.


considered

in

Chapter 7.

However, since doubling the reinforcement appeared to have had

little effect on the tension stiffening, this was unimportant.


that the secondary steel could also have been reduced.
reinforcement
125mm

was

was

the

considered

largest

undesirable

spacing

which

for

It appeared

However, smaller

practical reasons

complied

with

the

code

whilst
maximum

<300mm at full size) and which kept the reinforcement spacing in phase
with the beam spacing.
The reinforcement is detailed in Figure 8.6.

The reason for providing an

8mm longitudinal bar over each beam was to provide a proper anchorage for
the 8mm links projecting from the beams.

Since these links stopped short

of the edge of the top flange of the beam <as is usual because they have
to fit inside the formwork when the beams are cast) it was considered that
they would not greatly affect the slab's flexural behaviour and they were
ignored in its analysis.
i,.TB-01-IOOOT

l7 T6 -03- 115'

l7T60U08-11S'ff

fA
63T6 -03 -I 25'

-- -- -

1-1-

- -. k- 1-

1'-

- - - - - - - - - --

63T6-0I-/15' Er
2Th-(JL-IOOF f

- "" - j;- - -

- - - - - - - - -

--

m.
-

2TI0-05 -25EF

Quarter Plan

Ol C? Ol

le

og

07 Ol

oe

02 08 Ol

?" 1

08

Ol

02

07 01 09
I i

A-A
01

~!

01

01

r
\

01

01.

01.

; ; ; ~ ; ~ J.tm.
01.

01.

Ol

Figure 8.6:

Detail of reinforcement in slab of .first deck

- 146-

O!i 05

O!i 05

The real bridge would have to support wheel loads right up to the end of
the deck, where there were no diaphragms and where membrane enhancement
would be reduced.

It was therefore assumed that extra reinforcement would

be required.

It was also considered desirable to loop the secondary steel

round

bars

these

to

provide

the

correct

detailing

for

free

edge.

Because of the very thin slab the resulting detail, which is shown in
Figures 8 .6 and 8 .7, was slightly awkward.

Figure 8 . 7:

Reinforcement in corner of first deck

It was clear that the diaphragms in the second deck would make it stronger
than the first.

Since the first slab had behaved well, it was decided to

reduce the steel area in the second.

The opportunity was taken to look

into the possibility of using only one layer of steel each way.

This has

advantages for durability, since it greatly increases the cover, and it also
halves the steel fixing cost.

It had been rejected by Beal<122) but it

seemed probable that the thinner slab in the type of deck considered here
would make it more viable.
The main steel was increased to T8- 125 giving nearly 90% of the total
steel area in this direction used in the first deck.

However, since both

the first test and the analysis suggested that the secondary steel would
be very lightly stressed, this was not increased and just one layer of T6125 was used.
One effect of diaphragms is to apply a support moment to the most heavily
stressed beam.

This relatively small moment is

frequently

ignored in

design and the steel in the deck slab would normally be ample to resist
it.

However, the longitudinal reinforcement in this deck was so light that


- 147-

it was necessary to provide some designed reinforcement to resist this


moment.
It was also decided to use the second deck to conduct a small test, which
has

been

described

briefly

elsewhere<l23),

reinforcement corrosion on deck slabs.

on

the

effect

of

local

To simulate the effect of severe

local corrosion, eight adjacent main slab bars were cut right through at
mid-span of the slab.
slab was cast.

This was done using using bolt croppers before the

The position of the cuts was chosen so that the "damage"

would have the minimum effect on the behaviour under the load case used
for the initial failure test.

However, it would be possible to conduct a

service load test with one wheel of the HB vehicle immediately over the
cut bars.

It was also hoped to perform a

failure test using a single

wheel over

the area if it was in reasonably good condition after the

failure test.

8.2.5 Bearings
The beams were supported on normal commercial laminated bearings which
were PSC"370132"<124>; the smallest size of this type made.

These had a

greater movement capacity than a single span bridge of this type would
require.

As a result, they were less stiff than a true half scale model of

bearings for a single span bridge.

Their behaviour was close to that of

the bearings which would be required for a two-span bridge.


The stiffnesses of the bearings were checked in the Mayes machine, first
under concentric loading, which gave results very similar to the specified
stiffness, then under eccentric loading to measure the flexural stiffness.

8.3 MATERIALS
8.3.1 Concrete
The mixes used for the deck slabs and for the other in situ concrete were
similar to those used in the half scale beam strips considered in Chapter
6 and the nominal mixes are detailed in Table 8.1.

The mix for the first

deck used a realistic cement content but even with a high water content,
giving a very wet-looking mix with a slump of some lOOmm, it gave a 28
day cube strength of 44N/mm 2
model.

obtained from 150mm cubes stored with the

The cement content was reduced for the second mix givihg a .28 day

strength of 33N/mm2 with a lower water content and a more typical slump
of around 40mm.

The change in properties between the two mixes was much


- 148-

greater than the change in the nominal mix proportions would s uggest .
However, this was a consequence of the long delay between the tests <which
meant that both the cement and the aggregate came from different batches)
and

the

use of a

normal

commercial type

batching

plant

without

the

advantage of using dried aggregate as in smaller scale tests.

Quantity <per nominal

Material

First Deck

113 )

Second Deck

5-10mm Thames Valley Gravel

875kg

905kg

Sand

900kg

930kg

Ordinary Portland Cement

300kg

275kg
et1601

=1651

Water

Table 8.1:

Mixes for in situ concrete

Similar control specimens were used as for the beam strips considered in
Chapter 6 .

In addition, two sets of six 150mm cubes were tested, one

cured in a tank in accordance with BS 1881 <125> and the other cured with
the specimens.

The second deck used 12 batches and a pair of cubes, one

for each of these sets, was taken from every other batch.
results are shown in Tables 8.2 to 8.4-.

The test

The BS cured 150mm cubes gave

higher crushing stresses than the dry cured cubes of either 70 or 150mm
size showing that curing had a greater effect than size.
Because the beams were 500mm deep, compared with only 80mm for the slab,
and

because

the

precise

reproduction

of

their

behaviour

was

less

important, it was considered acceptable to use 20mm aggregate for these so


a normal commercial mix design was used.

This is detailed in Table 8.5,

and the results for the control specimens are given in Tables 8.6 and 8. 7.
Although the beams were cast in four separate pours, there was so little
difference between the test results for the different pours they have all
been considered together in the tables.

Although it was

nominally a

50N/mm2 mix, the actual strengths were much higher with a characteristic
strength of over 65N/mm2

This arose in part from the specified minimum

cement content and in part from the mix being designed to achieve transfer
strength, 40N/mm 2 , in the minimum time.
- 149-

Age

Test

(days>
28

Size

Number

(N/mn,2

<mm>

Veriation

Mean

Cube

150

'3

43.8

Cube

150

53.2

150

46.0

70

48.2

Indirect

1500

3.66

tension

1000

3.42

500

3.66

1500

27000

150

49.8

70

45.3

150

60.3

Indirect

1500

3.66

tension

1000

3 .20

500

3 .76

1500

3.92

1500

29000

(~)

(wet cured)
61

Cube

<start of
tests)

Elastic
modulus
90

Cube

<end of

5 .0

tests)
Cube

1.6

<wet cu.r ed)

Indirect
tension
<wet cured>
Elastic
modulus

Table 8.2:

Test results for slab concrete from first deck

- 150 -

Age

Test

<days)
28

Number

Cube

Variation

Me1jn

(N/m[l} 2

<mm)

(%)

150

33.4

70

33.0

150

35.1

4.5

Indirect

1500

2.52

14.8

tension

500

3.42

Elastic

1500

28500

150

36.2

70

33.5

Indirect

1500

3.01

tension

1000

3.20

Elastic

1500

24400

<start of
tes ts)

Size

5 .2

f---

Cube
<wet cured)

modulus
43

Cube

<end of
test>

modulus

Table 8.3:

Test results for slab concrete from second deck

- 151 -

Element

Test

Size

Number

(N/mm2

<mm)
Diaphragm

Mean

Cube

100

34.2

end of deck

Elastic

1500

23100

in figures>

modulus

Diaphragm

Cube

100

37.2

Cube

100

33.3

Elastic

1500

23500

<right-hand

<left-hand end)
Parapet

modulus

Table 8.4:

Test results for other in situ concrete from second deck

<all tested at end of test; approximately 40 days old)

Material

Quantity
<per nominal m3)

10- 20mm Crushed Limestone

819kg

5-lOmm Crushed limestone

352kg

Fines; Crushed Limestone

629kg

Rapid Hardening Portland Cement


P2 additive

400kg

Water

1. 121

::!1701

Table 8.5:

Mix for precast concrete

- 152-

Age

Test

Size

Number

<Nimm2

<mm>

(days>

Variation

Mean
)

(%)

2-5
<transfer)

Cube

100

43.6

6.0

28

Cube
<wet cured)

100

12

71.6

2.6

180+

Cube

150

70.1

Indirect

1500

4.08

1500

35500

Cube

150

71.5

Indirect

1500

3.92

1500

37900

<start of
tests)

tension
Elastic
modulus
210+

2.2

<end of
tests>

tension
Elastic
modulus

Table 8.6:

Test results for precast concrete from first deck

- 153 -

Age

Test

Size

<days)

Number

(N/mm2

<mm>

2- 4

Mean

Variation
)

<r.)

Cube

100

43.6

6.0

Cube

100

12

71.6

2.6

150

72.4

2. 4

1500

3.87

1500

39500

<transfer)

28

(wet cured)

320+

Cube

<end of
tests)

Indirect
tension

Elastic
modulus

Table 8. 7:

Test results for precast concrete from second deck

8.3.2 Reinforcement
The reinforcement for the deck slabs was GKN Tor- Bar in 6, 8 and lOmm
sizes.

The first of these sizes is no longer available commercially and

sufficient steel was in stock for the main steel of the first deck only.
For the remaining 6mm steel, all secondary steel, hard drawn wire was
used.

In order to give reasonably similar bond characteristics to the

normal steel, this was specially indented.


Stress-strain

curves

for

all

the steel were obtained using

the Mayes

testing machine and these are shown in Figure 8.8 in which each line
represents the average of three test results.

The hard drawn wire had a

significantly higher yield stress than the equivalent Tor-Bar.

Although it

appeared that this would have little effect on the behaviour of the slab
because the secondary steel was lowly stressed, the different properties
were modelled in the computer analyses.

- 154-

Stress

600

(JI/JDJ!i2)

500
400
- - - - 6mm Indented bArd drawn wire

300

<ultimate stress
200

- - - - 8mm Tor-Bar

<622)

6mm Tor-Bar

<594)

= 6511/~)

100
0

Figure 8 .8 :

0.6

0.4

0.2

Stress-strain relationship for reinforcement

8.3.3 Prestressing
The prestressing was provided by 12.7mm Bridon Dyform strand stressed up
to 70% of characteristic strength at transfer.

It was intended to take a

stress- strain curve for this using exactly the same procedure as for the
reinforcement.

However, two problems were experienced with this.

Firstly

the steel was too hard for the points on the clip-on 50mm gauge length
strain gauge.

It was

Plastic Padding as in

therefore necessary to attach demec points with


the concrete tests.

machine caused premature failures.

Secondly,

the jaws on the

This problem could have been solved by

using normal commercial wedge anchors but these would not fit in the jaws
of the testing machine.
properties
stress

at

for
1~

the

It was therefore decided to obtain the material

computer analysis

elongation

manufacturer's certificates.

and

the

from

the 0.2% proof stress,

ultimate

strength

given

on

the
the

The results obtained in the laboratory were

used only for the elastic modulus, E,..

In the event, this was the only

important property since the structures failed before the steel was fully
stressed.

8.4 CONSTROCTION

The

beams

were cast

commercial manner.

by Costain Concrete, South Wales,

in

the normal

Demec points were attached to the centre section of


- 155-

the beams and were read before and after stressing, as well as at the
start of the tests, to enable prestress losses to be estimated.

The beams

were transported to the laboratory, stored outside until required, and


placed on

the bearings

in

accordance with

the bearing manufacturer's

instructions <124).
The slabs were cast on plywood formwork supported off the beams.

Thus

the stresses due to the normal unpropped construction were reproduced but,
because of the lack of deadweight compensation, they were under-estimated
by a factor of two compared with a full size bridge.
In the case of the second deck, the diaphragms were poured first, then the
slab and finally the parapet up-stands.

In both cases, the deck slab was

cast in one pour and this required some 12 batches of concrete, slightly
more than would be used in a real deck where the batches would normally
be 6m3 truck mixer loads.

The concrete was placed by skip and Figure 8. 9

shows the first deck under construction.

Figure 8.9:

First deck under construction

It was considered very important

to give the slab an even and correct

thickness since analysis suggested that the local strength would be very
sensitive to this.

Two spare beams of each type were cast and those used

were selected for equal camber.

In the event the cambers of the beams,

although greater than normal due to the high prestress, were unusually
equal and this precaution was unnecessary.

The formwork was adjusted to

give as flat a soffit as possible and the concrete was finished using a
- 156-

screeding rail which spanned the full width of the deck.

After completion

of the tests, the thickness of the slab was checked by drilling a number
of holes and measuring through.
measured

on

the

second

deck

The mean thickness of the

was

79.65mm

and

although

17 depths

there

was

significant variation, from 76 to 84mm <which was a considerably greater


variation

than

was

observed

in

the

first

deck),

this

appeared

to

be

entirely random with the mean depths for four bays being 80.0, 80.5, 79.8
and 78.3mm.

It was therefore decided to base the analysis on the nominal

dimensions.
The

top of

the

concrete

was

covered

in

plastic

for

seven days

then

uncovered whilst the soffit formwork was struck after a minimum of four
days.

Real

bridge decks of

this

type are normally

constructed using

permanent formwork but access to the soffit was required to enable the
cracking to be observed and the surface strain gauges to be attached.

For

the same reason, neither water-proofing nor surfacing were provided.

This

made the tests conservative and Cairns<82> has found that surfacing alone
reduces the live-load steel stress by some 30%.
No

attempt

was

made

to

match

experienced in a real bridge.

the

curing

conditions

which

would

be

The lack of permanent formwork or water-

proofing, the small scale, the unusually wet concrete <particularly in the
first deck) and the dry laboratory air all had the effect of increasing the
shrinkage of the slab whilst the beams were rather older than usual when
they were placed.

Thus the shrinkage of the slab, and the differential

shrinkage between the slab and the beams, was significantly greater than
in a real bridge.

Because of this, if

<as has been suggested) shrinkage

has an adverse effect on the development of membrane action, the test


results would be conservative.

8.5 LOADING
8.5.1 Loads Applied
Since the slab behaviour was of prime concern, and since analysis indicated
that HA load would have a relieving effect on the slab whilst dead weight
would have an insignificant effect, only the HB load was applied with no
HA load or dead weight compensation.

These loads would have increased the

moments in the beams, thus the degree of over-strength in the beams was
slightly greater than that due to the over-provision of prestress.

- 157-

The loading sequence was designed to first apply the design service HB
load in a critical position, then to simulate the full load history due to
the service life of a real bridge before returning the load to its original
position.

The service load would then be re-applied, enabling the effect

of cracking and loss of restraint due to other load cases to be assessed.


The load on the HB rig would then be increased until failure occurred.
Whilst the design static service load which should be applied to the deck
was well established,
simulate

the

and defined

service

life

of

in BS 5400,

bridge

Part 10 <126) defines fatigue loads.

the loading required to

was

less

clear.

BS

5400

However, these are intended for use

with defined fatigue relationships for steelwork details whilst the primary
concern in this project was the cracking behaviour of the concrete.

This

is much more sensitive to small numbers of large load cycles, as has been
found in Chapter 6.

Thus if the fatigue loads had been used, they would

have been used well outside the range for which they were intended or
calibrated.

When relatively small numbers of cycles are considered, the

design fatigue loads can be locally more severe than the design ultimate
load.
small

This does not matter in normal fatigue assessment, since these


numbers

calculations.

of

cycles

have

little effect

on

the

cumulative

damage

However, it is clearly illogical to require a structure to

resist a thousand cycles of a load in excess of design ultimate.

Since

bridge deck slabs are likely to be most sensitive to the few loads of near
design service level which are applied in their life, it was decided to
base the cyclic loads on BS 5400: Part 2 loads.
Unlike the long span HA loading, the HB loading and the short span HA
loading, which are relevant to these decks, have no statistical base<127>.
It

was therefore necessary to make some gross assumptions in order to

decide how many cycles, and of what magnitude, to apply.


assumed that

It was initially

the design service loads should have the same chance of

occurrence as their

long span HA equivalents;

occurring once in 120 years< 12 7 ).


loading should be applied.

that

is a 5% chance of

This implied that only one cycle of this

However, it was decided to apply a more severe

sequence to ensure that the tests would be conservative.


Another difficulty with simulating the load history of a bridge was that
real bridges are subjected to rolling loads whilst the loading rig was only
able to apply pulsating loads at di?crete positions.
- 158-

In order to ensure

that this would be at least as severe as applying the intended load at all
positions along the length of the deck, the test load . was increased.

load of 1.2 times design service load was therefore applied to all the
positions.

The original intention was to apply two cycles of this load

followed by 10000 cycles of a reduced load, simulating 25 units of HB


<again with a 20% excess) then 100 cycles of design service HB.

The

significance of the 25 unit HB load is that it was used to represent HA in


the then current loading standards <23,24-).

Finally, a cycle of 1.2X design

service load . would be applied, enabling the effect of the cyclic loads to
be assessed by comparing the behaviour then with that under first loading.
In the event, the 10000 cycles had very little effect so, after the first
position, the number applied was reduced to 5000.
critical

parts

of

the

slab

were

subjected

to

However, because the

wheel

loads

under

two

different load positions, these were subjected to at least 10000 cycles of


wheel loads.
The maximum load applied in the service load tests was approximately equal
to the design ultimate load.
<which

This, combined with the nature of HB load

is particularly severe

for

this type of structure and probably

unrealistic), the lack of surfacing and the large number of load cycles
applied, meant that the load history to which the bridges were subjected
was

excessively

severe

and

made

the

tests

conservative,

as

intended.

However, Perdikaris and Beim's work<128>, which was published after these
tests were completed, suggests that rolling loads are more severe than
fixed pulsating loads.

They suggested that one passage of a rolling load

could have the same effect on the fatigue life of a slab as 34- to 1800
cycles of a fixed load.

As they considered the number of cycles to 60% of

static strength to cause failure, whilst the tests considered here are
investigating the effect of cycles of service load level, their conclusions
may not be applicable here.
be

related

to

the

crack

Also, the difference they observed appeared to


patterns;

pulsating

loads

gave

local

patterns whilst rolling loads gave extensive grid-iron patterns.


a consequence of the use of large single wheel loads.

radial

This was

Under the HB

service loads used in the author's tests, the cracking extended over a
greater length of the bridge but was purely longitudinal.
no

reason

to

anticipate

that

rolling

fundamentally different crack pattern.

loads

would

There is thus

have

led

to

However, even if <as Perdikaris and

Beim suggested for this type of reinforcement> one pass of a rolling load
was equivalent to 34 cycles of a static load, the use of 20% over-load
- 159-

meant that the load history used was still conservative.


of

1.2X design service

load had more effect

than

One application

100 applications of

design service load .

8.5.2 Loading Rig


The

16

wheels of the HB load were loaded by

four one-hundred tonne

hydraulic jacks acting through spreader beam systems which are illustrated
in Figure 8 . 10.

The four jacks were interconnected and connected to a

hydraulic pump system which enabled cyclic loads to be applied.

This

s ystem was rated at only 40% of the hydraulic pressure for which the
jacks were designed.
applied, equivalent

Although this enabled a load of 400kN per jack to be


to 2. 7 times the design ultimate load, calculations

suggested that a slightly higher load would be required to fail the deck.
Separate hand pumps were therefore provided for the final failure test.

Figure 8.10:

Spreader beam assembly

- 160-

The jacks reacted against

two large steel universal beams which were

supported by double channel stanchions bolted down to the strong floor of


the laboratory.

The loading frames can be seen in Figure 8.2.

Because the

anticipated loads were close to the calculated capacity of the floor, it


was necessary to position the bridge to minimise the moments in the floor
under the load case which would be used for the failure test.

It was also

desirable to spread the load on each leg of the frame evenly between four
floor bolts.

Since the standard spacing of the HB bogies did not match up

with the bolt centres, it was only possible to achieve this for the legs of
one of

the

bogies.

The

load

from

each of

the other

two

legs was

therefore spread unevenly amongst six bolts.


The HB bogies could easily be moved sideways to any required position by
moving the spreader beams and jacks.

However, to move them longitudinally

it was necessary to move the whole loading frame.

It could have been

moved to any position but this would have required a re-arrangement of


the anchorage system.

In practice it proved adequate to move the bogies

only by multiples of the bolt spacing.


8.6 INSTRUMENTATION

The loads were measured using four 800kN load cells located below the
jacks.

Separate figures were recorded for the four cells but no facilities

to adjust the relative loads were incorporated in the system.


A

50 mm travel linear voltage displacement transducer was provided under

the centre
provided

of each

over

beam.

each

In addition,

bearing

and

under

10mm travel

transducers

some

positions.

wheel

were
The

transducers under the wheels were supported off the top flanges of the
beams and thus measured only the slab displacement relative to the beams.
Vibrating wire strain gauges were used both on the surface and in the
concrete at selected positions.

These have the advantage of remaining

stable over long periods, which was important as it was intended to record
the

total

positions.

strains

due

to

the

application

of

several

different

load

However, their strain capacity was not sufficient to use them

to measure smeared strains in cracked concrete.

Thus "portal" gauges

developed by Cook<129) were used in positions where cracking was expected.


Because it was considered undesirable .to estimate curva.tures or extensions
in concrete sections

from

top and bottom gauges with different gauge


- 161-

lengths and other characteristics, portal gauges were also used in some
positions where cracking was not expected.
A disadvantage of surface strain gauges is that

because their thermal

inertia is much less than that of the specimen, they are very sensitive to
temperature
coefficient

changes;
of

unlike

demecs,

expansion since

the

portals

have

significant

they are made of aluminium.

However,

although the laboratory was not air conditioned, it proved possible to keep
the temperature constant to within some 2c for the tests for which the
strain data was used.

In order to avoid the problem of sunlight warming

the gauges directly, all the blinds on the South side of the laboratory
were closed for the duration of the tests.
Because portal gauges have not previously been used for long-term tests,
it was decided to monitor their long-term performance using readings off
demec points mounted as close as possible to each portal.

The original

idea was to use the portals only to record the change of strains during a
test and to add these on to long term changes recorded by the demecs.

In

practice, the changes of reading in the portals were close to those in the
demecs so this extra complication proved unnecessary.
The reinforcement under one wheel in the first test and two in the second
was also strain gauged, using electrical resistance gauges.

Unfortunately,

some of these gauges were damaged during the construction of the deck and
few of the results were usable.
Two gauge lengths were used for the portal gauges: 200mm for the beams,
which is the largest size made, and lOOmm for the slab.

The latter length

was a compromise between the requirement for a short gauge length, to


monitor local peaks in the bending moment distribution, and a long gauge
length

to

make

the

results

comparable

with

smeared

crack

analysis.

However, the latter objective was not achieved since the crack spacings
were greater than lOOmm.
as

indicating only

gauges

on

the

Thus it is more realistic to consider the gauges

the movement

reinforcement

of

individual cracks.

represented

only

the

Similarly,
strain

at

the
their

particular location and were not directly comparable with smeared. crack
analysis.
All the electronic instrumentation, a total of 74 channels,

wa~

connected

to a ''Compulog" data logging system which converted the results to digital


- 162-

strain, displacement and force readings before storing it on disc and tape
for later processing.

Some key strain and deflection readings, as well as

the load cell readings, were printed out whilst the tests were in progress.
8. 7 TESTS ON FIRST DECK

8.7.1 Global Service Load Tests


a. First Load Position
The loading

frame

was

first

positioned to apply the HB load in the

position indicated in Figure 8.11.

The design service load was then

applied in ten approximately equal increments.


examined for cracks after each increment.
critical areas

of

the slab with an

The structure was carefully

However, despite studying the

illuminated

magnifying glass,

cracks were seen until the full load had been applied.

no

Under the previous

increment the strain measured by the portal gauge immediately under the
wheel nearest the centre of the deck was 575 microstrain; some three times
the strain at which cracking normally first becomes visible.

This was

partly a consequence of the thin slab and high strain gradient.

However,

this also applied to the half scale specimens considered in Chapter 6


which cracked at lower strains.

Another explanation is that under the

concentrated load the scope for stress redistribution, both by moment


redistribution and by membrane action, was so great that the concrete in
the critical area was effectively being stressed under strain control, even
though the structure was loaded under load control.

Thus the cracks did

not become visible until the concrete stress had dropped significantly
below the normal cracking stress.
The crack widths were measured using a crack microscope.

Under full

service load the maximum width, which occurred under wheel 10 in Figure
8.11, was 0.05mm;

equivalent to O.lmm at full size.

This would certainly

be acceptable in practice and may appear to be very small considering that


conventional design methods implied that the slab should have failed by
this

stage.

However,

other

studies,

notably

Kirkpatrick et

al's <49>,

suggested that the slab should have been uncracked under this wheel load.
The fact that the outer bay of the slab <where global transverse moments
were

less

significant>

was

indeed

uncracked,

suggested

that

global

transverse moments were the reason for this difference from Kirkpatrick's
result.

- 163-

After

this

test,

the

deck

was

unloaded

and

the

cracks closed

completely that they were invisible even using the microscope.


Figures 8.12 and 8.13

indicate that

up

so

However,

the local strains and deflectionss

under the critical wheels did not fully recover.

The strain reading under

wheel 10 was marginally greater than under wheel 9 and the reason for
plotting the deflection under wheel 9 <rather than 10) in Figure 8.13 was
that the displacement transducer under wheel 10 failed .

.900
1056

J50

900

2250

1056

1056

1056

Beam A
BeamS
Beam(
Beam 0

I
0

HB wheel positions for first and last service load test


and f or failure test.

1(

Other HB wheel positions.

Single wheel test positions.

Figure 8. 11:

Load positions for first deck

- 164-

Load
150
<lri/Jaclr)

Cycle 2

125

/
Cycle 1

/
/

100

~After

/
75

Loading in

Other Positions

/
/
/

50

Loading

U1

- -

- Unloading

25

/
/
0
0

500

1000

1500

2000

2500
Strain x 106

Figure B. 12:

Transvers e sofflt s train under wheel 10


<Servi c e l o ad tests)

Load
150
(kJI/Jac k>

Cyc le 2

125
Af ter Loading in
Cycle 1
Other Positions
lOO

/
/
75
Loading
O'l

en

50

Unloading

25

0 ~~~~---,----------~--------~~--------~---------,

0.0

0. 2

0. 4

0. 6

0. 8

1.0

Deflection <mm relative to beams>

Figure 8 . 1 3 :

De fle c t i o n unde r

<Serv ic e l o ad tests )

wheel

The deck was then loaded to 1.2 times design service load.

As the load

increased above design service level, the cracks under wheels 9 and 10
grew longer and at a load of 140kN per jack they joined up.
maximum load

<150kN/jack> the crack width under both these wheels was

approximately 0.13mm whilst


wide.
been

Under the

mid-way between them the crack was 0.08mm

Since the maximum crack width under the design service load had
less

than 0.08mm,

this

<and

the similar relationship between the

strain readings> suggested that applying the increased load in this one
position was at

least equivalent, as far as this area of the deck was

concerned, to rolling the service load 0.9m along the deck.

If similar

relative widths occurred in subsequent tests <which they did) this meant
that applying 1.2 times service load in just the three positions along the
deck illustrated in Figure 8. 11 would be equivalent to rolling the service
load along its full length.
After the cyclic loads described in 8.5.1 had been applied, a load of 1.2
times design service load was again applied.

The change in the strains

and displacements, compared with the load application before the cyclic
tests had been performed, was so small that this application could not be
plotted on

Figure 8.13

without

making

it

illegible

and,

for

the same

reason, only the peak part of this load cycle is shown in Figure 8.12.

The

strain measured at the start of the cycle was marginally smaller than that
measured at the end of the second cycle to 1.2. times service load.

Thus

the 10000 cycles to "HA" service load plus 20% and 100 cycles to full HB
service load had had a small effect on the behaviour compared with just
two

cycles

to

1.2

times

HB

service

load;

the

structure

had

actually

recovered some of its strain whilst the cyclic loads were being applied.
Under full load, the cracks were not significantly wider than under the
first

load application and no cracks were visible on the top surface of

the slab; the only visible cracks were the four longitudinal soffit cracks,
one

under each

pair of wheels.

On

unloading,

the cracks were again

invisible even with the microscope.

b. Other Load Positions


On completion of the tests in the first
moved sideways by 500mm.

position, the loading rig was

The same load sequence was applied in this and

subsequent positions, except for the reduction from 10000 to 5000 cycles
of the HA equivalent load.

As can be seen fr.om Figure 8.11 some of the


- 167-

wheel positions in this load case were the same as in the first case. Thus
cracks were visible under these wheels at a much earlier stage than in the
previous test.

Cracks were also visible in the newly loaded bay one load

stage earlier than they had been in the first test.

However, apart from

this the behaviour was very similar.


On completion of these tests, the whole loading frame was moved along the
deck by 1056mm to apply the HB load in the third position illustrated in
Figure 8.11.
similar.

The same load sequence was applied and the behaviour was

Because the instrumentation had been positioned to suit the first

load position,

the behaviour could not be monitored so closely.

Crack

widths were measured, however, and they were marginally greater than under
the

first

load

position;

the maximum width

being 0.15mm against 0.13.

Since the first load position was a worse case for global effects, and
both were identical for local effects, this suggested that

the loss of

restraint and distribution due to the cracking caused by the previous load
cases was affecting the behaviour.
The

same

load

sequence

was

illustrated in Figure 8.11.

then

applied

in

the

remaining

positions

By the completion of these tests, all three

bays of the deck slab had cracked along almost the full length of the
bridge.
seen.

However, these three cracks were the only cracks which had been
They were visible with a

magnifying glass

when

the deck was

unloaded, with a maximum width of 0.05mm and a more typical width of


0.02mm.

c. Return to First Load Position


For the final service load test, the loading frame was returned to its
original position and the load was re-applied.

As will be seen

from

Figures 8.12 and 8.13, the deformations were greater than under the first
applications but still not excessive.

The maximum measured crack width

was 0.2mm which, as in all the tests, was slightly <25%) less than would
be assumed from the strain gauge reading, indicating that the concrete on
either

side

of

the

cracks

was

still

under

significant

tension.

The

maximum crack width was equivalent to 0.4-mm at full size, compared with an
allowable width of 0.25mm in BS 54-00: Part 4-.

However, that document only

requires crack widths to be checked .under a much lower load; 25 units of


HB compared with
subjected.

the

1.2

times 4-5

unit

load to which the model was

Under the load used . for crack width calculation, the measured
- 168-

crack width was 0, 12mm; the scale equivalent of 0.24mm compared with the
allow ab le width of 0.25mm.

Although it is unreasonable to expect this

level of precision in crack width predictions, and a model is likely to


under-estimate crack widths, the many conservative features of the tests
which have been mentioned earlier mean that it is reasonable to conclude
from this that the crack widths would be acceptable in a full size bridge.
Thus,

by

this

criterion,

the

service

load

behaviour

of

the

deck

was

satisfactory although it clearly did not have the enormous margin of overcapacity which previous research implied it should have.
will

be

considered

in

the

next

chapter,

and

also

Analysis, which

observation of

the

behaviour suggested that this difference was due to the global transverse
moments resulting from the use of full HB load in this study, compared
with only single wheels in other studies.
conclusively

that

it

was

not

due

to

However, it remained to prove


the

absence

of

the

diaphragms

recommended by others.
It

was noted in Chapter 2 that crack widths are an unsatisfactory, and

perhaps unnecessary, design criterion.

However, by any other fundamental

design

deformations>

criterion

satisfactory.

<such

as

permanent

behaviour

was

Similarly, the stresses estimated from the strain readings

were well within the BS 5400 criteria.

Thus the behaviour of this very

lightly reinforced deck slab was clearly satisfactory.


the

the

over-provision

of

prestress

and

the

However, because of

conservative

nature

of

conventional design rules for prestressed concrete, the lack of cracks in


the beams did not prove that the distribution properties of the deck were
either satisfactory or similar to those which had been assumed in the
design

of

investigated

the
by

beams.
detailed

This

aspect

comparison

of

the

with

behaviour

analyses

could

only

be

thus

will

be

and

considered in Chapter 9.

8.7.2 Global Failure Test

After the service load tests had been completed, the design ultimate HB
load was re-applied and

the HB

load

was

then

increased

in

steps of

approximately 25kN per jack, that is 17t of design ultimate load.

The

displacements of the beams are shown in Figure 8.14 whilst that of the
slab under wheel 9 is shown in Figure 8.15.

The loading to failure was

not continuous and the points where the load was removed and re-applied
are indicated by breaks in the plots in the figures.

- 169-

Beam ~

(k~~~=ck)

:::

Beam~

Beam A

300
250
200
150
100
50
0 4-----,-----~----~------~----~----0
10
20
30
40
50
60

Deflection (mm)

Figure 8. 14:
Load

Beam deflect1ons

4 00

' kN/Jac k )
350

:wo
250
200

150

lOO
50
0
0

Deflection <rn1n )

Figure 8.15:

Deflection under wheel 9

As the load increased the longitudinal cracks under the slab grew wider
but , at a

load of 245k.N per jack

cracks were visible.

(1.67

times design ultimate), no new

At this stage the largest strain recorded by the

portal gauges across the crack in the centre bay of the slab was 4600
microstrain, 80% higher than under design ultimate load,

indicating

(as

will be seen from Figure 8. 15> that the slab was beginning to depart from
- 170-

the near linear behaviour it had exhibited since the completion of the
service load tests.
this crack was

Even mid-way between the bogies, the strain across

1900 microstrain but, in contrast, the highest

tensile

strain recorded by the portals on the top of the slab over the edges of
the webs of the beams was 300 microstrain adjacent to wheel 10.

The

gauge over the web of Beam B on the outside, that is adjacent to wheel 2,
was reading 136 microstrain compression indicating that in this region the
sagging moment due to transverse global effects was greater than the
local moment.

Longitudinal portal gauges positioned under wheels 9 and 10

were showing very small strains but


microstrain tension,

that under wheel 12 showed 600

the difference presumably being due to the lower

global compression in this area.


A number of transverse strain gauges and demec points had been positioned
in the slab near the expected points of transverse contraflexure in an
attempt to estimate the membrane forces.

Due to the small and erratic

readings, the proximity of cracks and the transverse strains resulting from
the

Poisson's

ratio

extremely difficult

effect

of

to interpret.

the

global
However,

flange

forces,

these

were

there did appear to be a

transverse compression adjacent to the wheels.


After

this load stage,

the

load was removed and there was over .80%

recovery on all the significant readings.


increased further.

The load was then re-applied and

At 250kN per jack, a shear crack appeared in the right

hand end of Beam B <as shown in the figures) and a flexural crack was also
just visible in the soffit of the same beam under wheel 6.
and flexural cracks formed in the same regions at 275kN.

Further shear

At this stage, a

shear crack also appeared in the right hand end of Beam C and in the left
end of Beam B.

There was also a very fine horizontal crack running along

the outside of the web to Beam A adjacent to wheels 1 and 2, due to the
beam's action in restraining the hogging moment in the slab.

A second

longitudinal crack had appeared in the soffit of the slab but still no
cracks were visible on its top surface.
At 300kN per jack, twice design ultimate load, the first shear crack which
had appeared in Beam B extended right through the bottom flange.

What

looked like shear cracks also appeared in the left end of Beam C between
the support and wheel 9.

However, cracks on the opposite side of the web

sloped the opposite way, indicating that the cracks were largely due to
- 171-

torsion although there were only the flexible bearings and thin deck slab
available to resist, or apply, the torque.
Longitudinal cracks in the top of the slab also became visible at this
stage.

These then extended over most of the length of the deck on either

side of Beam B and on one side of Beam C.

A crack also appeared over the

inside edge of the web to Beam A, adjacent to wheels 1 and 2, but this
only extended a short distance either side of the bogie.
At a

load of 2.39 times design

through

the

flange

of

Beam

discontinuity of approximately

ultimate

(350k.N/jack),

became

1mm across

the shear crack

very

wide

with

vertical

it.

This was due

to bond

failure with the strands which could be seen to have drawn in by some
2mm.
than

Since the strands used were larger in size and smaller in number
in

true half scale

relative to a true model.

model,

this

may

have occurred prematurely

The shear cracks are illustrated in Figure 8 .16.

Figure 8.16:

Shear cracks in Beam B

By this stage, flexural cracks extended over 2.2m of the length of Beam B
and had also developed in Beam C.

There were now five longitudinal cracks

in the soffit of the outer bay of the slab under wheel 4.


out

beyond

These fanned

the wheel towards the beams and the end of the deck as

illustrated in Figure 8 . 17.

This <and less pronounced fanning at the far

end of the same bay beyond wheel 1) was the only sign in the slab of
transverse cracks or of the characteristic radial crack pattern observed by
other researchers in single wheel tests.

- 172 -

Figure 8.1 7:

Figure 8.18:

Soffit cracks under wheel 4

First deck under 400kN per jack

By this stage, Beam A had developed a substantial rotation which is visible


in Figure 8.18.

This was largely due to

the differential deflection

between

the still uncracked Beam A and the much more heavily loaded

Beam B.

The local transverse hogging moment due to the wheels may have

also contributed, but the crack pattern showed clearly that the slab was
subjected to a net transverse sagging moment right out to Beam A adjacent
to wheels 3 and 4.

The slab in this region was thus contributing to

restraining the rotation; the global transverse moment was dominating over
the local moment.

The rotation of the beam about its longitudinal axis,

combined with the resulting transverse movement, was well in excess of the
intended capacity of

the bearing.

At

- 173-

load of 400kN per jack the

resulting transverse force

became too much for the bearings under the

other beams and the whole deck suddenly moved sideways by some 20mm.
The resulting unintended eccentricity led to elastic torsional buckling in
the main girders of the four load spreading rigs.

The load was therefore

removed and the structure recovered remarkably welli


the

local deflections

and

90~

of

the

approximately

beam deflections.

80~

However,

of
the

maximum strain recorded by the gauge under wheel 10 reduced only from
11500 to 4.400 microstrain.
On re-loading, the deck settled down to its new position but the buckling
re-occurred so the load was removed again and the ball bearing under one
end

of

the

offending

girders

was

replaced

with

rocker

bearing.

Calculations showed that because of the low torsional stiffness of the


girder <which had been the cause of the problem) the resulting unevenness
of the load distribution between the four wheels would not be significant.
The modified rig was therefore used for all subsequent tests, including
the service tests on the second deck.
By this stage the limiting pressure of the electric pump had been reached
so

the

movement

loading
of

was

resumed

the deck had

using

moved

hand

pumps.

Because

the bearings off

the

sideways

their seatings,

the

pronounced step in the plots in Figure 8.14 may have little significance;
the displacement transducers over the bearings had come off their points
so this could not be checked.

However, there clearly was a deterioration

in the distribution properties of the deck as well as in the stiffness of


Beam B.

The large differential deflections across the deck are clearly

visible in Figure 8.19.

Figure 8.19:

View across deck as failure approached

- 174-

When the load level was only marginally higher than before, a number of
"bangs" were heard from around the right hand end of Beams A and B; one
so loud an observer assumed it to be due to rupture of the links either in
Beam B Cdue to shear) or Beam A (due to separation from the slab).
However,

when the concrete was

proved not to be the case.

later broken out

in this region,

this

It may have been due to further bond failure

in the end of Beam B as the draw- in was now approaching lOmm.


By this stage difficulty was being experienced in holding the deck up to
load, indicating that failure was imminent .

The appearance of the right

hand end of Beam B, combined with the loud bangs, suggested to some
observers that this would take the form of a shear failure in this beam.
However, a line of crushing concrete was just discernible between wheels 3
and 4-.

At a

load of approximately 4-14-kN per jack

ultimate> wheel 4- punched through the deck.

(2.83 times design

The resulting release in the

load on the beams caused the beam deflections to reduce; hence the wheel,
although

loaded

only

by a

jack and

thus

under

displacement

control,

punched right through the slab as shown in F'igure 8 .20, rupturing the
steel as it went.

,.

Figure 8.20:

Failure; Wheel 4 punched through deck

The hand pumped hydraulic systems used for the two bogies were separate.
The reduction in beam deflection caused by the failure therefore led to an
increase in the load on the other bogie.

Several minutes later, when a

reading was taken, the load on this was some 25% higher than that which
- 175 -

had caused failure under the first bogie.

Because of the high creep in

concrete structures approaching failure, this indicates that the failure


load would have been substantially higher.
After the structure had been unloaded, the concrete remaining under the
wheel which had punched through was removed.

This revealed the classic

conical form of a punching shear failure as can be seen in Figure 8 .21.


However, the suggestion in earlier chapters that such a failure can be
precipitated by concrete crushing had been reinforced by the fact that t he
line of crushing concrete
noticed

before

the

failure

(which

is

visible

occurred.

It

in Figure 8 .20) had


also

appeared

that

been
global

transverse moments had significantly reduced the local strength of the


slab.

The rotation of the edge beam <which was due to the differential

beam deflections at

mid-span) had clearly led to a transverse sagging

moment in the slab over the web near the wheel which failedi a region
where the local moment would have been sagging.
transverse moment

Near mid- span, where the

in the slab was causing rather than restraining the

rotation of the beam, there was a transverse hogging moment over the
beam; hence the higher local strength.

Figure 8.21:

Failure cone viewed from below

Further confirmation of the significant reduction in strength due to the


global moments <and hence, by implication, of the flexural nature of the
- 176-

failures)
methods

was

given

considered

estimated

when
in

the

slab

3.2.3.

the _strength

by

strength

Kirkpatrick
nearly

80%

was

et

whilst

estimated

al's

using

approach(13)

Hewitt

and

the

over-

Batchelor's

approach <72>, using a restraint factor of 0.6, was only marginally better.
The

restraint

factor

back-calculated

from

the

failure

load

was

approximately 0.2, compared with the "conservative" figure of 0.5 used in


the Ontario Code<1ll for assessing existing decks.

Thus, although the test

results did not suggest that decks designed to the empirical rules would
be unsafe,
could be.

they did

imply that

the Ontario assessment recommendations

However, since all previous research into compressive membrane

action in bridge decks had suggested that support diaphragms are needed
<or at least desirable> to provide the restraint, a plausible alternative
explanation for the reduced strength was that the restraint in this deck
was inadequate.

It was decided to perform local tests to investigate this.

8.7.3 Local Failure Tests


Two single wheel tests were performed using the test rig illustrated in
Figure 8.22.

The position of these, which is shown in Figure 8.11, was

chosen for convenience in testing and also to avoid areas of the slab
which had been significantly damaged in the previous tests.

However, the

slab around the wheel tests had been cracked by the previous tests whilst
the adjacent bay, which could be important to the restraint, had apparently
been loaded very close to failure.

Thus the test situation was extremely

unfavourable compared with the normal situation in a real bridge deck.

It

was considered that the behaviour at low loads was so greatly affected by
. this

that

it

had

no

real

significance.

Thus

serviceability

was

not

considered in such detail as in the global tests and the slab was loaded
monotonically to. failure.
The same instruments were used as . for the global tests but some were
repositioned

and

all

were

re-zeroed..

Thus

strain

and

displacement

readings were taken relative to the start of the test, rather than relative
to the initial <uncracked) state as in the global tests.
were experienced with

the logger during the first

Some difficulties

test whilst in the

second test the displacement transducer under the wheel stopped working.
Since the behaviour in the two tests was very similar, only the behaviour
of the second test will be described in detail but the load-deflection
response for the first test is illustrated in Figure 8.23.
- 177-

9
Fig ure 8.22:

Single wheel test rig

Load <k!D 2 00

150

100

50

0
0

8
Deflection

Figure 8.23:

10
( :mm)

Result o f single wheel test

A longitudinal crack was visible under the wheel before the test started
and by a l oad of 43kN <which corresponds to 1.15 times the design ultimate
wheel load ) this crack was 0.08mm wide.

This crack was thus significantly

narrower than under the same load per wheel in the global test, despite
the pre-cracking.

At the s ame stage, the crack on the top of the slab

over the web was 0 .05mm wide, whilst the equivalent crack did not appear
in the global test until nearly twice the load per wheel had been applied.
However, this difference was probably largely due to damage sustained in
- 178-

previous tests; the crack was visible before the load was applied.

transverse crack was just visible in the soffit under a load of 67kN.

In

the global tests, such a crack had not appeared until the wheel load was
some 30% higher.

This difference could not be explained by pre-cracking;

it was due to the global flange force prestressing the slab in the local

tests.

Figure 8.24:

Crack pattern under single wheel

At a load of 113kN, by which point failure had occurred in the global


tests, the maximum crack width was 0.3mm.

By 145kN the cracks in the

soffit had taken on the characteristic radial form which is illustrated in


Figure 8.24.

At a load of 202kN the crack in the top of the slab along

the web of beam 3 had joined the similar crack due to the previously
performed

single

wheel

test.

However

there

was

no

sign

of

this

interaction reducing the strength; the wheel finally failed under a load of
approximately

226kN

compared

with

204kN

in

the

previous

test.

comparison, the failure load in the global test was 103.5kN per wheel,

- 179-

For

The wheel punched a neat hole through the top of the deck, coming to rest
only

some

10mm

below

the

top

surface

of

the

slab.

However,

this

difference from the failure mode under full HB was not indicative of any
fundamental difference in the local behaviour.
different

post-failure

behaviour

caused

It was purely due to the

by

the

much

smaller

displacement and force which was released by the local failure .

global
Removal

of the loos e concrete from under the wheel revealed the classic conical
failure s urface illus trated in Figur e 8 .25 which is very similar to that
observed in the global tests.

The behaviour was also very similar to the

global tes ts in that although there was plenty of warning of failure, in


the sens e that the structure was clearly unserviceable when subjected to
only half its final f ailure load, the final collapse was very sudden.
with

Even

the advantage of the strain readings, and of having observed an

iden t i cal test only hours earlier, it was not easy to tell when failure was
imminent.

Figure 8.25:

Failure cone viewed from below

The difference between the two results may not be significant as it is not
unusual to obtain

much greater strength differences between nominally

identic al concrete

spec imens.

However,

it

may

have

been due

to

the

inferior restraint available to the first test which was nearer the end of
the deck.

The failure loads were, however, close to the 185kN predicted by

Kirkpatrick et al's approach; the ratios being 1.22 and 1.10 compared with
an

average

of

1.19

for

Kirkpatrick's

own

tests<13 ).

Similarly,

the

res traint factor back-calculated using Hewitt and Batchelor's approach, at


approximately 0 .65, was close t o t heir findings.
- 180-

This confirms that both

approaches give good predictions for the failure load under single wheel s
and suggests that the restraint needed to develop this local strength is
not dependent on diaphragms.

Thus the lower local strength observed when

all 16 wheels of the HB l oad had been applied must have been the result
of global effects.

8.8 TESTS ON SECOND DECK


8 .8 . 1 Global Service Load Tests
a . First Load Position
The second deck was s ub jected to a very similar load history to the first.
However, only 5000 cycles of the reduced load were used and, because of
the greater width of the deck, t he load was applied in a maximum of four
different posit i ons across the width instead of two.

The loading positions

are illustrated in Figure 8 .26.

1-

r
I

)(

--

)I')(

r
)(

~-

I
,_
f--

-~

~
-

s
~

)(

- -=- ~
..

)0(

Jt

I"

I"'

111(

-~

J(

l' I

-- - --

-:it

~- - --*"- .

10

14

1(-

.If)(

..
-

,.

- --

-7

le@

--,

)(

- - -....... -

--

)(

4
.
-8~

0
12

-=-o

e
t-----i

-_;_-?t---..:..-

X::

BeamB

-.- Beam(

Beam 0
Beam E

HB whe e l pos i t i ons for firs t and last s ervi ce l oad test
and f or f ailur e t est.

Beam A

l
J
-=-~ -0- l

1(1(

~ .

11

-xr~-

'&.

)('

)(

""3

-- - -_-:Jt
- - r-(

lt:=

}(

13
.q

XI(

"

-~

Ot he r HB wheel positions.
Wheel posi t i ons fo r local tes t s.
Line o f cut s i n re inforcement .

Fig ure 8. 26:

Load p ositions for second deck

- 181 -

Like the first, this deck was initially subjected to design service load
then to two cycles of a 20% higher load.

A soffit crack appeared at

approximately the same stage as before and under full design service load
it

had

width

of

0.1 mm:

equivalent

stage

in

the

approximately

previous

test.

twice

the

However,

concrete are very variable, particularly at

width

as

at

the

the crack widths

in

loads just above that which

causes cracking, so this may have had little significance or may have been
due to the lower tensile strength of the concrete.
The difference in crack width was less pronounced at the higher load of .
the second loading cycle.

A more important difference from the behaviour

of the first deck was that a top crack appeared at the same time as the
soffit crack, whereas in the first it had required a load some 150% higher.
The measured local deflect ions, crack widths and transverse strains were
greatest adjacent to wheel 14.

These strains are shown in Figure 8.27,

whilst the deflection of wheel 14 relative to the beams is shown in Figure


8.28.

Initially the soffit strain exceeded the top strain by some 50~ but

this percentage reduced once the behaviour departed from linearity and the
top strain overtook the soffit strain when the cracks became visible.
remained greater
greatest

<even

throughout
in

the subsequent

absolute

terms>

when

tests,
the

It

the difference being

structure

was

unloaded.

However, these high strains were confined to the region over the inside of
the web to Beam D which was the location of the only top crack.

Figure

8.27 shows that the strain over the edge of the adjacent Beam C was very
much lower.

Indeed, nearly all the tensile strain in that region could. be

explained by the Poisson's ratio effect of the longitudinal compressive


strain due to the global flange force.
The difference
particularly
analysis

between

significant

would

treat

the strains at
when
the

either end of the slab span is

it

is

realised

that

slab

as

symmetrical

and

conventional
so

would

local

predict

identical strains at either end, whilst an apparently more sophisticated


local analysis <treating the slab as continuous over simple supports) would
predict a greater hogging moment over Beam C than over Beam D.

- 182-

150
Load
<kN/ Jack>

Cycle 3

Cycle 2

;I
;I
;I

125

100

I
/

Af ter Loading in

Ot her Positions

/
/

/I

75

/I

():)

.p.
I

50

~//

25

~/

'l

/
/

//

Loading

//

Unloading

Deflection <mm relative to beams>

Figure 8 . 28:

Derlection under wheel 14

<Ser v i ce l oa d tests )

The explanation .for the greater strain over Beam D was that there the
global transverse moment was hogging, adding to the local moment, whereas
over Beam C the global moment was sagging and therefore acting against
the local effect.

Further evidence to suggest that the top crack over

Beam D was largely due to global effects was given by the length of the
crack.

As soon as it appeared, it extended over most of the length of the

deck.

The

width

even

mid-way

between

the

bogies,

where

Pucher's

charts (40) indicated that the local moment should have been sagging, was
some two thirds of the maximum width adjacent to wheel 14.

In contrast,

as in the first deck, separate soffit cracks appeared initially under each
wheel.

The soffit cracks formed by the two wheels of a bogie, such as

wheels 13 and U, joined together as the load increased but the cracks
formed by the two bogies did not join until the load had been applied in
other positions along the length of the deck.

Thus it appeared that the

soffit cracks were primarily due to the local effect whilst the top crack
was largely due to the global effect.
This

also explains

soffit.
strain

Initially,
was

behaviour
smoothed
concrete.

as

greater
departed

out

as

the difference
predicted

than
from

force

the

in strain behaviour of the

by elastic

maximum

linearity

top

the

redistributed

theory,
strain.

stress

peak

top and

the maximum soffit


However,
in

the

once

the

soffit

was

to the surrounding under-stressed

There was less scope for redistribution of the top stresses

because global moments are relatively uniform over the length of the deck;
hence

the

rather

greater

increase

in

strain

on

cracking.

Uncracked

concrete surrounding the local cracks and trying to push them closed would
also lead to better recovery of the soffit strain on unloading.

However

the much inferior recovery of the top strains <the top strain after each
cycle of the first load position .was over double the soffit strain) was
undoubtedly exaggerated by the reinforcement detailing since the single
layer of main steel was located some 10mm below mid-depth.
The reason why the top cracks had not appeared in the first deck until a
much higher load was applied was that the lack of diaphragms meant that
the beams were free to rotate.

Thus local hogging moments were relieved

by rotation and differential displacements of the beams led to lesser


global transverse moments.

The diaphragms in the second deck contributed

to its superior distribution properties and the maximum beam displacement

- 185-

was some 25% lower than in the first

deck even though a static load

distribution would predict identical deflections.


As with the first deck, the application of 5000 cycles of a reduced load
had remarkably little effect on the behaviour.

b. Other Load Positions


After completion of the tests in the first position, the HB loading rig was
moved a metre sideways in the direction towards the top of Figure 8.26 to
apply load in the second position.
that in the first position.

The behaviour was generally similar to

The one new top crack appeared over the edge

of the centre beam when the design service load was applied, whilst the
new soffit crack appeared in the bay at the top of Figure 8.26 under the
full load; that is 1.2 times design service load.
The original intention had been to apply the HB load in a total of three
different positions across the width of the deck.

This loading sequence

was designed to induce all the soffit cracking which was likely to occur
in service.

However, it was now apparent that top cracks due to global

effects could be equally significant.

The intended load sequence would

have failed to induce top cracks over the inside of the web of Beam B.
Since such cracks could .be significant to the behaviour of the adjacent
bay of slab

<that between Beams B and C> and since that bay would be

loaded in the final test, it was considered that this was a fault of the
sequence.

An intermediate load position was therefore used.

The loading

rig was moved back 1.5m towards the bottom of Figure 8.26 to apply the
same load position as in the first test but opposite hand.

Apart from the

effect of the pre-cracking, which led to a softer initial response, the


behaviour was very similar and the maximum crack widths were similar.

The

maximum strains and deflections were also similar although few direct
comparisons could be made as few gauges were in equivalent positions.

The

global deflections were very similar; within 5%.


After completion of the tests
moved a

in this position,

the loading frame was

further metre sideways to the position nearest

shown in Figure 8.26.

the bot tom as

This position was particularly significant as it

included a wheel directly over the region of the cut reinforcement.


Under the maximum load which had previously been applied, there were no
cracks visible in the region of the cut reinforcement.
- 186-

This was slightly

surprising as the load case was identical to one which had already been
tested, apart from being opposite hand, and cracks had appeared in that
test at the equivalent position and load stage.

It was also unfortunate

as it had been intended to use the test to investigate the behaviour,


under

cyclic

reinforcement.

loads,

of

cracked

bridge

deck

slab

with

damaged

It was therefore decided to increase the load slightly to

crack the slab; the strain of 500 microstrain measured by the portal gauge
under the wheel indicated that

cracking was imminent.

However a

10%

increase in load failed to produce visible cracks despite giving a strain


of 610 microstrain.
the

reinforcement

Curiously, the strain under the adjacent wheel, where


was

intact,

was

986

microstrain which

indicated,

comparison with other cases, that cracks would have been visible.

by

These

were not noticed although the area was not inspected as thoroughly.
Clearly, cutting the main steel had not advanced the formation of cracks:
indeed it appeared to have delayed it although, in view of the variability
of concrete behaviour, this was probably not significant.
A further increase in load would have applied a significantly higher global
load

than

had

been

intended

and

may

significantly alter the behaviour under

have

caused

the later

enough

damage

load cases.

It

to
was

considered that the intended load sequence was over-severe and a further
increase would have made it too unrealistic.

It was therefore decided to

unload the structure and disconnect three of the four jacks.


a

higher

load

to

be

applied

on

four

of

the

This enabled

wheels without

causing

significant damage in regions where it was likely to affect the behaviour


under the later load cases.
At a load of 163kN, a marginally lower local load than that which had
previously failed to crack the region, a 0.05mm crack was visible in the
soffit under the cut reinforcement.

The load was then increased to 178kN,

equivalent to 1.22 times the design ultimate HB wheel load, which increased
the crack width to 0.1mm.

It also induced a crack, approximately 0.05mm

wide and 1.5m long, in the top of the slab over Beam D.
The bridge was then unloaded, the other three jacks re-connected and the
cyclic loads applied as in the previous positions.

Despite the initial

over-loading, the behaviour was entirely satisfactory and appeared similar


to that when the load case had been applied opp'osite-hand . over .intact
- 187-

re in f orcemen t.

There was thus no evidence that cutting the steel had

adversely affected the behaviour.


For the nelCt load position the loading frame was moved 1056mm along the
deck and back to its original transverse position.
and

subsequent

positions

was

very

similar

to

positions so it will not be described in detail.

The behaviour in this


that

in

the

previous

Because the load cases

were less severe on the beams, the top cracks were not greatly elCtended.
It was therefore decided that it was not necessary to apply the load in as
many transverse positions to simulate all the damage which could occur in
practice and only two were applied.

It appeared that applying 1.2 times

service load in just one longitudinal position was equivalent, as far as


top

cracks

were

concerned,

longitudinal positions.

to

applying

service

load

in

all

possible

However, it was still necessary to apply the load

in all the positions shown in Figure 8.26 in order to ensure that the
soffit cracking would be correctly simulated.
For the final loading position, the bogies were positioned to give wheel
loads 250mm off-centre to the slab span.

A new crack formed under some

wheels but not until the load elCceeded design service load.
appeared

that

loading

only at

mid-span of

the slabs,

as

It therefore
in all

the

previous tests, had given a reasonable representation of the extent of


cracking which would have occurred if the service load had been applied in
all positions.
c. Return to First Position
The loading frame was returned to its original position and the full load
re-applied.

As will be seen from Figure 8.28, the local deflection under

wheel 14 <the greatest local deflection recorded> was substantially greater


than under the first loading.

The maximum local deflection was also 42%

greater than the equivalent deflection at the same stage in the tests on
the first deck, whilst the deflection on unloading was over four times
greater.

It will also be seen, by comparing Figures 8.28. and 8_.13, that

loading in other positions had had a greater effect on this deck than it
had on the first.
One reason

for

the greater effect

on

this deck of loading in other

positions can be inferred from Figure 8.27:. it had opened .a crack over the
web of Beam C.

This crack tended to close as load was applied, indicating


- 188-

that the global sagging moment in this region under this load case was
greater

than

remained

the

local

s~nificant

hogging

moment.

thus

continued

and

However,
to

the

tensile

contribute

to

strain

the

local

deflection throughout the load cycle.


The maximum crack width recorded was for

the crack over the edge of

Beam D and was equivalent to approximately 0. 7mm at full size, whilst the
maximum soffit cra ck width was equivalent to approximately 0.4mm.

The

pat tern of top cracks after the completion of the service load tests is
illustrated in Figure 8.29.

Under the reduced load used for checking crack

widths in BS 5400, the soffit crack width was equivalen t to 0.2mm compared
with

the allowable width of 0.25mm.

The top crack was equivalent to

approximately 0.4mm but the top cover was over twice the "C""'"' " required
by BS 5400.

Thus, since BS 5400 allo ws the crack widths to be calculated

"on a hypothetical surface at a distance

Cn o rro

from the outermost bar", the

crack width was, at worst, very close to being acceptable.

Nevertheless

the condition of the slab was not as obviously satisfactory as that of the
first deck had been.
the

permanent

occurred in

The permanent deformations were greater and although

soffit

the

cracks were

first

deck at

equivalent, even on unloading .

no

wider

top cracks,

all, were up

which had

to 0.3mm wide

not

full scale

It was thus somewhat debatable whether the

condition of the slab could be considered acceptable.

f-

r
I

--

- --

.---

- --

Beam A

-----=-- --

- - -

"---

- --

---

,.-

-l J

----- - l
___, J
--=
- -:. 1--

"------,-

BeamS
Beam C
Beam 0

J BeamE
- - -.;_-_

,_

Figure 8.29:

Top cracks on completion of service tests

The fact that top cracks had occurred in t he second deck but not in the
first

was

undoubtedly

due

to

the greater

resulting from the presence of diaphragms:


- 189-

transverse

hogging moments

the load which was required to

induce top cracks in the first deck was so much greater <over two and a
half times as great) that none of the other differences between the decks
could have been more than minor contributory factors.
for

the

poorer recovery of

However, the reason

the second deck on unloading,

greater maximum crack widths, was less clear.

and for its

It could have been due to

the nature of global as opposed to local moments, but the difference in


the main reinforcement and in the tensile strength of the concrete may
also have been important.
essentially

parallel

to

However, since the significant cracks all ran


the

beams,

it

appeared

that

the

substantial

reduction in the secondary steel could not have been a major factor.
Unfortunately, it was not possible to test a third deck so the relative
significance

of

the

differences

investigated analytically.

between

the

decks

could

only

be

Analytical investigation was also essential to

assess the distribution properties and to see how they compared with those
which would normally have been assumed in design.

Nevertheless, it was

clear that the distribution properties were superior to those of the first
deck since, despite the lower grade concrete in the slab, the deflect ion of
the heaviest loaded beam had been consistently some 25% lower.

8.8.2 Global Failure Test


On completion of the service load tests, the load on the model was reapplied and then increased.

The strains adjacent to wheel 14 are shown in

Figure 8.30 whilst the beam deflections are shown in Figure 8.31.

Both are

plotted relative to the original zeros, which explains why they do not pass
through the origin, and the break in the plot indicates a point where the
load was removed before being re-applied.
The initial strain response shown in Figure 8.30 is approximately linear.
However, once the load exceeded 150kN per jack, the highest load which had
previously been applied,

significant departure

from

observed as the cracks extended in depth and width.

linearity can be
At this point, the

tensile strain over Beam C, which had previously been reducing slightly,
began to increase slightly.

. This would appear to indicate that the local

moment in this region was increasing as moment redistributed away from


the more heavily cracked regions.
The load-deflect ion response of Beams B and C is dist in et ly non-linear
from a load of approximately 150kN per jack.
- 190-

However, this was clearly

due

to

deterioration

softening

of

deflections

these
does

of

the

beams,

not

approximately 250kN.

dist r ibution

since

become

the

properties,

plot

noticeably

of

the

rather

sum

non-linear

of

until

t han
the

to

beam

load

of

As the load increased, the proportion which the outer

two beams carried clearly reduced and, above a load of 275kN per j ack, the
deflection of Beam E began to reduce in absolute, as well as in relative
terms.
to

the

This deterioration of the distribution properties was largely due


extensive

diaphragm,

cracking

wh i ch began

contributed .
Load
500
<kN /J ack )
450

400

in

the

to appear

slab

but

fr om a

torsional

cracks

in

the

load of 225kN per jack, also

Soffit at Kid- Span

over Beam C

------

350
300
250
Top over Beam D
2 00

150
100

50
0

10

12

16

14

Strain x 10 3

Figure 8 .30:

Transverse strains adjacent to wheel 14

Although, despite the lower grade concrete in the slab, the maximum global
deflec tions

were s maller than at equivalent stages in the test s on the

first deck, a comparison of Figures 8 .31 and 8.14 shows that they began to
depart

from linearity at an earlier stage.

This was apparently largely

because the earlier formation of top cracks meant that the distribution
properties

began

to

deteriorate

at

an

earlier

stage.

However,

an

additional but closely- related reason was that the distribution properties
of

the

first

deck had been so poor.

Thus

the heaviest

loaded beams

carried s uch a high percentage of the load which a static distribution


would predict that even a total loss of distribution properties would have
had little effect on the deflec tions.
- 191 -

Beam

Beam C

Load
450
<kN/Jack)
400
350
300
250
200
150

Sum of all Beams

100
50
0

10

20

30

40

50

60

Deflection (mm)
Figure 8 .31:

Beam deflections

As the load increased, new cracks formed.

The distinct kink in the plot of

the soffit strain in Figure 8.30 at a load of around 260kN per jack is due
to the formation of another longitudinal crack close to, but outside, the
gauge length.

At a load of 300kN, there were four longitudinal cracks in

each of the loaded bays of the deck.

They extended over much of the

length of the deck; three of the initially separate cracks caused by the
two bogies in each bay having joined together.

These cracks fanned out in

a radial pattern at either end of the deck, between the bogies and the
diaphragm, but there were no transverse cracks in the soffit and only one
longitudinal crack in each bay deviated

from the longitudinal direction

between the bogies to meet the webs.


Whilst the pattern of soffit cracks was generally similar to that in the
first deck at equivalent load stages, the pattern of cracks in the top of
the slab, which is illustrated in Figure 8.32, was very different .

The

longitudinal crack over the web to Beam A formed at a load of 225kN per
jack whilst the diagonal cracks at either end of this appeared at between
250 and 400kN per jack.

Since there was no load at all applied directly

to this bay of the slab, it is quite clear that these cracks were the
result of global effects.
in other bays.

Global effects also dominated the crack pattern

Thus the cracking in the bay between Beams C and D, which


- 192 -

had very different deflections, was far more extensive than that between
Beams

and

whose

deflectionss

were

more

similar.

The

crack

perpendicular to Beams B and C at the right hand end of the Figure was
clearly due largely to global hogging in these beams as torsion in the
diaphragm <which was restrained by the lightly loaded Beams A, D and E>
attempted to resist the rotation of Beams B and C.

Similarly, th e extreme

asymmetry <about the deck's longitudinal axis) of the crack pattern in the
bay between Beams D and C indicated t hat 'here too, global effects were
dominant.

Beam A

- -0:

BeamS

- --

CE

_- 0:::::

Beam C

Beam 0

~~~~~~~~~~~~~
Figure 8.32:

BeamE

Top cracks immediately prior to failure

<note: cracks which were formed by other load positions and


which were closing under this load case are not shown)
At a load of 300kN per jack, twice the design ultimate load, the first
crack was observed in a beam; a flexural crack in the soffit of the centre
beam under wheel 10 .

A shear crack appeared in the right hand end of the

same beam (as shown in the Figures ) at a load of 325kN.

A shear crack

appeared in Beam B at a load of 350kN and a flexural crack followed at


approximately 375kN.

By this stage, the capacity of the hydraulic system

had

The

been

reached.

bridge

was

therefore

unloaded

and

the

jacks

connected to a new electric pump which had not been available at the time
of the tests on the first deck.

The load was then re-applied.

By this stage, the cracking in Beams B and C had caused a significant


reduction

in

their

stiffness

as

can

be seen

from

Figure 8.31.

The

resulting increase in the differential deflections of the beams led to


further cracking in the top of the slab and the strain represent ed by the
- 193-

crack over the web to Beam D had reached the limit of the capacity of the
portal gauge.
Figure 8.32

By a

load of 400kN per jack, all the cracks shown in

had formed except

the

longitudinal one over Beam E.

The

widest of the cracks, as throughout the test, was that over Beam D.

By a

load of 425kN this crack was some 3mm wide and the crack over Beam A was
1.5mm wide.
By a load of 460kN per jack, 10% higher than the failure load of the first
deck, there were flexural cracks over much of the length of the centre
beam joining the shear cracking at the right hand end.

The flexural cracks

crossed the soffit of the beam at right angles to its longitudinal axis.
In contrast, those in the adjacent Beam D crossed at an angle of up to 45
degrees and tended to form first on the outside, that is away from. the
loaded bay.

Since the very wide longitudinal crack over one side only of

the web to Beam D suggested that it was subjected to a very substantial


torque, this was not surprising.

However, the crack pattern implied that

this torque was almost entirely resisted by transverse bending and shear
in the bottom flange and not by torsion as such; unlike in the first deck,
the shear cracks on opposite sides of the web sloped in the same direction
indicating relatively low torsional stresses in the web.

The asymmetrical

loading of the beam had, by this stage, caused a longitudinal crack in the
web on the outside of the beam.
Up to this stage, despite the very extensive cracking, there had been no
difficulty in loading the deck or in holding it up to load.

However, as

the load increased further this became increasingly difficult, indicating


that failure was imminent.

At a load of approximately 475kN per jack, a

line of crushing concrete could be clearly seen on the soffit of the slab
extending

along

the edge of Beam D for some one and a half metres

adjacent

to wheels

13 and

14.

reaching

the

of

moment

limit

its

This section of the slab was clearly


capacity

and

it

is

presumably

the

resulting redistribution of local moments which caused the increase in the


strain over Beam C which can be seen in Figure 8.30.
At a load of approximately 490kN per jack, 3.35 times design ultimate load
and some

18% higher than the failure load of the first

deck,

failure

occurred in the form of wheel 14 punching through the deck.. The. resulting
sudden

reduction

in

the

global

load

on

the

deck

reduced .the global

deflections and hence increased the load on the other three jacks.
- 194-

The

local behaviour was so brittle and the failure so sudden that this caused
wheels 5, 8 and 16 <one under each jack) to punch through as well, despite
the fact that the four jacks were inter-connectedi they punched through in
such quick succession there was not time for the hydraulic pressures to
equalise.

The deck after failure is illustrated in Figure 8.33.


I

Figure 8 .33:

Second deck after failure

After the failure, a line of crushed concrete could be seen on the top of
the slab between wheels 9 and 10 as well as between wheels 13 and 14-.
There was also more localised crushing adjacent to the other two wheels
which punched through.

However, despite a thorough inspection at the final

load stage before the failure,

this had not been observed until after

failure although there had been some sign of very local crushing by wheel
U.

It appeared that the soffit crushing which had been observed before

failure had been the root cause of the collapse yet, despite this, the
failure once again looked like a classic "punching shear" failure.

This

further confirmed that such failures could be caused by flexural effects.


The fact that the critical section of slab had so clearly reached the limit
of its moment capacity when the strains and crack widths at the equivalent
position at the other side of the local slab span were so modest, as can
be

seen

from

Figure

8.30,

again

confirmed

the

importance

of

global

transverse moments.
Another

interesting

feature

of

the results

was

that, despite a

lower

concrete strength and less reinforcement, this deck slab had failed at an
18Z higher load than the first.
through

the

slab

at

As in the first deck, wheels had punched

substantially
- 195 -

lower

load

than

predicted

by

Kirkpatrick et al's approach but the margin was much smaller; approximately
a

factor

of

1.22

against

l. 78.

It

appeared

that

this

was

due

to

differences in the global behaviour but again an alternative explanation


was the superior restraint to in-plane forces in the second deck with its
diaphragms.

It was decided to perform some local tests to investigate

this.
8.8.3 Local Failure Tests
Despite the extensive damage caused by the global tests, it was considered
that the two outer bays of the slab were in sufficiently good condition
for

local failure

tests

to

be

useful.

perform a total of four local tests.

This gave

the opportunity to

It was decided to perform two tests

identical to those which had been performed on the first deck.

These

would enable the failure loads to be compared with that in the global
tests and in the tests on the first deck.
controls for
single

the other two

wheel

test

over

In addition, they would act as

tests which would be performed;

the

region

with

the

cut

investigate its effect, and secondly a two wheel test.

firstly a

reinforcement,

to

The latter was

considered important as it was not clear how much of the reduction in


local strength which had occurred in the global tests could be attributed
to interaction of the local effects of adjacent wheels.
Since the slab was cracked much more extensively in the bay between Beams
A and B than on the other side of the deck, it was decided to perform both
control tests in this bay.

This meant that any reduction in the failure

load per wheel of the other tests could be clearly identified as due to
the effect being investigated, rather than due to the effect of previous
damage.

The positions used for the tests are illustrated in .Figure 8.26.

The first test performed was the control single wheel test at position A
in Figure 8.26.

This was followed by the test over the cut reinforcement

at position C in the Figure and the load-deflection response of both these


tests is illustrated in Figure 8.34.

The behaviour in the two tests was

very much alike and also very similar, apart from the lower failure load,
to that in the equivalent tests on the first deck.

The fact that most of

the main steel had been cut right through beneath the wheel at B appeared
to have had very little effect, indeed the initial response was softer iri
the test with intact reinforcement although this was probably due to the
greater damage sustained by this region of slab in the global tests.
- 196-

As

in the tests on the f1rst deck, the failure loads were very similar to that
predicted by Kirkpatrick et al's approach.
184-kN

Load \. kN ) 200

150
100

.....:::

50

----------- A<intact

reinforcement )

C (c ut reinforc emen t >

.1.

0
0

10

Deflect ion <mm >


Figure 8.34:

Singl e wheel

t es t s A and C

176kN

200
Load
<kNt wh eel)
150

...-

100

...-

U6kN

B <single wheel)

- D <two Wheels)

50

10

Deflection (mm)
Figure 8 .35:

The

next

test

performed was

Local tests B and D

the

control test

expected, this behaved in a very similar fashion.

at

B and,

as might

be

For the final test, the

two wheel test, a new loading arrangement was required.

Since neither the

time nor the money was available to fabricate this, one of the four load
spreader rigs used for the global tests was employed, the main beam being
re- positioned so that over 95% of the load was applied to two of the
wheels .

The load-deflection response for the two tests is illustrated in

Figure 8.35 and it will be seen that the presence of the second wheel had
a significant effect on both the deflection and the failure load although
the failure once again took the form of one wheel punching through the
- 197 -

deck.

The reduction in strength was approximately 17'1. <or 19'1. using the

average of the two control tests) which is slightly less than experienced
by Kirkpatrick(13) in bays with equivalent span and depth.

However, he

suggested

due

that

the

reduction

particular support conditions.


a

minor

contributory

he

observed

may

have

been

to

his

It now appears that this was probably only


If,

factor.

as

his

approach

appears

to

imply,

punching failures are caused directly by excessive shear stresses in the


region immediately round the wheel, it is hard to explain why the presence
of a second wheel should affect strength.

However if, as suggested here,

the failures are primarily flexural one would expect that any load which
increased the bending moment would reduce the strength.
Although

the

presence of a

second wheel

reduced

the

failure

load

per

wheel, it remained substantially (23%) higher than in the global tests.

It

thus appeared that the lower failure load per wheel in the global tests
was indeed partly due to global effects although the interaction of the
local effects of the two wheels was also significant.
As in the first deck, the single wheel tests were remarkably consistent
with

Kirkpatrick's,

the average ratio of

failure

load

being 1.20 compared with 1.19 in his own tests(13).


single wheel tests,
ratio was

that

1.183 with a

remarkably good

result,

to his

Considering all the

is those performed on both decks,


coefficient

of variation of 0.04.33.

even

lftrgely

for

predict ion

empirical

the average
This is a

formula developed

from tests on structures which were similar to those considered here.


small variation means that
observed

when

more

wheels

variation can be eliminated.

The

the possibility of the reduction in strength


were

applied

being

purely

due

to

random

However, paradoxically, even the fact that an

approach based on shear stresses gave such good predictions could be used
as an argument for saying that the failures are primarily flexural: shear
failure loads are inherently more variable than flexural failure loads and
it

would be extremely unusual to obtain such a small variation in the

shear strength of even apparently identical specimens.

8.9 TESTS ON SINGLE

BEAM

Although this is primarily a study of deck slab behaviour, it is clear from


the previous section that global behaviour is important to this.
behaviour of
important,

the

before

Thus the

beams has a significant effect and it was considered


analysing

the

complicated
- 198-

decks,

to

ensure

that

the

analysis was capable of modelling the relatively simple beam behaviour


correctly.

It was therefore decided to test a beam on its own.

The precast beam which was tested was similar to those used in the first
deck, and had been cast in the same batch.

It was provided with a one

metre wide in situ top flange cast in the same way as the models using a
similar mix which gave a cube strength at time of loading of 45. 7N/mm 2

The reinforcement provided in this flange was like that used in the second
deck.

It was considered that the flexural behaviour, which was of prime

concern,

would

be

very similar

for

the two

types of

beams so their

features were chosen for convenience.


Although it may appear obvious that the single beam should have a flange
width to match the beam spacing in the decks, this is less obvious when
the in-plane shear stiffness of the deck slab is considered:
in 3.2.7

that

the heaviest

loaded beam in a

it was noted

beam and slab deck can

effectively have a flange which is wider than the beam spacing.

This

means that concrete crushing failures may be less likely in bridge deck
tests than in single beam tests but no allowance was made for this effect.
The beam was positioned on bearings in the same way as in the deck tests.
The loads were applied with the same loading frame and jacks, although the
loading rig was modified to bring the wheels closer to the longitudinal
centre-line of the beam to avoid over-loading the slab.

To make the

results directly comparable, the longitudinal position of the loads was


kept the same as in the global failure tests.

Since neither the bearings

nor the loading rig provided significant restraint to rotation about the
longitudinal axis, a steel beam was placed across the top of the beam and
held down to the floor.

However, it proved acceptable to allow this system

to go slack in the test.


Two 100mm travel displacement transducers were provided to measure the
deflection.

These were mounted over the top of the beam to avoid damage

in the event of the sudden failure which was anticipated.

A number of

demec points were provided but no electronic strain gauges were used.

The

beam under test is illustrated in Figure 8.36.


Since the previous tests indicated that the beam behaviour was, for all
practical purposes, perfectly linear elastic until well above design service
load, there was little point in applying complicated load histories.
- 199-

The

beam was therefore loaded to 150kN


monotonically to failure.
load in Figure 8.37.

p~r

jack then unloaded, then loaded

The mid-span deflection is plotted against the

The response was linear elastic up to a load of

approximately 200kN per jack so the first load cycle cannot be seen in the
Figure.

Figure 8 .36:

Single beam under test

Load
500
<kN/Jack)
4.00

300

200

lOO

20

40

60

80

lOO

120

Deflection <mm >


Figure 8.37:

Load-deflection response of single beam

The behaviour prior to failure was remarkably similar in many respects to


that of Beam B of the first deck.

The crack patterns were very similar,

- 200-

including the shear cracks which are illustrated in Pigure 8.38.

Even the

loads at which the cracks appeared, expressed in kN per jack, were similar.
However,

the deflections were slightly higher at each load stage which

suggests that the load distribution in the first deck was slightly better
than that predicted by simple statics.

Figure 8.38:

Shear cracks in single beam

Despite the extensive shear cracks and the draw-in of the tendons which
had occurred by a deflection of 60mm <the global beam deflection at which
failure

occurred in the first

deck) the load increased by another 24%

before a sudden explosive failure occurred at a deflection of 110mm.

The

only warning of this was the increasing creep and the failure was so total
that
meant

the beam fell some 500mm even though the hydraulic system used
that

the load was

removed as soon as failure occurred.

After

failure, as illustrated in Pigure 8 .39, there was no concrete left at all in


the critical section.

Figure 8.39:

Beam after failure

It was clear that the failure had been caused by concrete crushing in the

top flangei what had been observed was the classic brittle bending failure
- 201-

of an over-reinforced <or in this case over-prestressed) concrete section.


However, it is easy to imagine that a two-dimensional version of such a
failure,

that

is in a slab rather than in a beam, would look like a

punching shear failure.


crushing concrete

Indeed, in a sense it is a shear failure because

fails

on

inclined planes.

Thus

there is no clear

distinction between the two types of failure.


8.10 DISCUSSION AND CONCLUSIONS
8.10.1 Service Load Tests

The condition of the first deck was satisfactory at the completion of the
tests.

Its behaviour during the tests was also apparently satisfactory

although it remains to check the distribution properties by comparison


with analysis.
The behaviour of the second deck was less satisfactory because relatively
large cracks opened in the top of the slab on loading and failed to close
on unloading.

Although the failure of these cracks to close may have been

due to the reinforcement detailing, the fact that they occurred only in the
second

deck

was

clearly

result

of

the

higher

transverse

moments

resulting from the presence of the diaphragms.


It appears, although it remains to check this by comparison with analysis,
that

compressive

membrane

action

did

not

greatly

contribute

resistance of either deck to global transverse moments.

to

the

However, as

predicted by previous researchers, compressive membrane action clearly did


contribute
suggestions,
diaphragms.

to

the
this

resistance

to

contribution

did

local
not

moments.
depend

Contrary
on

the

to

their

presence

of

The result was that the cracks in the top of the second deck,

the feature which led to its behaviour being considered less satisfactory,
were the direct result not only of an effect which has been largely
ignored by previous research <global transverse moments>, but also of a
feature which had

been positively recommended

<diaphragms>.

Although

these diaphragms improved the distribution properties, they appear to have


had a detrimental effect on the serviceability of the slab.
8.10.2 Failure Tests

The most significant aspect of the failure tests was not so much the
individual

failure

modes

or

loads

as

the

relationship

between

them.

Failure occurred in the first test when a single wheel punched through the
-202-

deck under a wheel load which represented only half the local strength of
the slab: not only as predicted by previous research but also as measured
by the subsequent single wheel tests.

Despite its diaphragms the second

deck, with its weaker concrete and smaller steel area, was weaker in the
single

wheel

tests

yet

it

was

able

to

resist

higher

global

load.

Clearly, the failures under full HB load were greatly influenced by the
global behaviour but this does not mean that they did not occur until beam
failure was imminent.

The maximum beam deflection in the global tests was

little more than half that at which failure occurred in the beam tested
alone.

That

beam had reached only 80% of

its failure

load when

its

deflection matched that at which the decks failed.


All this fits the hypothesis that the slab failures were primarily brittle
bending compression failures.

In the global tests, the global transverse

moments induced by the beam's differential displacements or <in the case


of the first deck) rotations used up some of the bending strength of the
slab.
though

Because, by virtue of the membrane forces,


locally

rein forced,

heavily

reinforced

even

though

their local behaviour was brittle.

the slabs behaved as


actually

very

lightly

Thus redistribution was

very limited and the safe theorem of plastic design did not apply.
slabs

failed

under

combined

global

and

local

The

transverse moments

even

though, at the failure load, the global transverse moments were not needed
to maintain equilibrium.
The single wheel tests confirmed the work of previous researchers.
they

suggested

that

the

enhancement

to

local

strength

Indeed

caused

by

compressive membrane action is remarkably tolerant of features which might


be expected to reduce restraint.

All the tests were performed 1n outer

slab bays which had already been extensively cracked for their full length
by previous tests; half were performed after other failure tests in the
same bay.

Two of

remaining tests

the tests were

1n

a deck without diaphragms;

the

(all those on the second deck) were performed close to

points where wheels had punched through the adjacent bay of the slab.
Despite all this, the behaviour had been entirely satisfactory.

The only

thing which significantly reduced the strength relative to that predicted


by previous research was the presence of an adjacent wheel under load.
Even with this, the failure load for the very lightly reinforced slab was
equivalent to four times design ultimate load.

-203-

The global tests confirmed that the contribution of compressive membrane


action to resisting global transverse moments is, at best, substantially
less than the contribution to local behaviour.

Because the local behaviour

was so enhanced by membrane action, the crack pattern prior to failure


tended

to

effects.

be

greatly

influenced,

and

in

places

dominated,

by

global

As in the service load tests, global- transverse moments <which

have been largely ignored by previous research) had a major influence on


behaviour.

However, despite the large reduction in local strength caused

by this, the failure loads were still very high; a minimum of 2.83 times
design ultimate load.

It might, therefore, be thought that the reduction

had no practical significance.

This may not be the case.

The failures occurred when the combined local and global moments became
too great for the slab.
local

moments

are

This has important implications because, whilst

direct

effect

of

the

load

on

the

slab,

transverse moments are only indirectly an effect of this.


direct effect of the differential deflections of the beams.
which

increases

these

differential

strength of the slab.

deflections

could

The implication is that

global

They are a
Thus anything

reduce

the

local

if the beams had been

weaker, or less stiff, the slab would have failed in the same way but at a
lower load.
Perhaps the most important conclusion from the tests is that, as predicted
in 3.2.7, global and local behaviour are not independent.

Most previous

research on bridge deck behaviour has implicitly <or sometimes explicitly)


assumed that they are.

The result is that most of the previous research

on membrane action in bridge deck slabs is only strictly applicable under


single

wheel

loads.

This

does not

necessarily

mean

that

recommendations resulting from that research are unsafe.


only

because

of

the

large

reserve

of

strength

of

the design

Indeed, even if

prestressed

beams

designed to current rules, it seems likely that bridges designed using the
empirical rules discussed in 3.2.8 will have more than adequate strength.
Nevertheless it does mean that caution is required.
bridge

assessed

to

the

Ontario

assessment

It appears that a

rules(ll)

as

having

just

adequate global and local strength could actually have a much lower safety
factor

than

intended.

As

for

service

investigation is required.

-204-

load

behaviour,

analytical

CHAPTER
ANALYSIS

OF

MODEL

9
BRIDGE

TESTS

9.1 INTRODUCTION
The tests described in the last chapter provide useful empirical evidence
for the contribution of compressive membrane action to the behaviour of
bridge

deck

slabs.

They

could

help

with

the

development

of,

and

justification for, empirical design rules such as those considered in 3.2.8.


However, in order to appreciate the significance of the behaviour, it is
necessary to compare it with analyses.
conventional

analyses,

to

quantify

the

Firstly it will be compared with


potential

savings

from

using

membrane action in the design of deck slab reinforcement and to see if


the distribution properties predicted by conventional methods for global
analysis, based on uncracked slab properties, were realised.
will be compared with the form of analysis considered in

Secondly it

Chapter 7, both

to see if that analysis would have provided a suitable design method for
the models and to obtain some understanding of the behaviour.

9.2 CONVENTIONAL ANALYSIS


9.2.1 Analysis for Design of Deck Slabs
The deck slabs of both the bridges were checked using BS 54.00 and the
analytical methods which would normally be used with it and which were
considered

in

2.4.

A linear grillage model

was

used

for

the global

analysis and Westergaard's formula(39) was used for the local analysis.

a. First Deck
The

allowable

load

on

the

approximately 14. units of HB.

first

deck

slab

using

this

approach

was

For the intended design load of 4.5 units of

HB the reinforcement required was Tt0-87.5, the odd spacing giving the
minimum steel area and being equivalent to 175mm at full size.
nearly four

times

the steel area actually provided.

This is

The failure

load

implied for the reinforcement provided (setting all Ym values to 1.0) was
14..3kN per wheel compared with the actual failure load of approximately
103.5kN per wheel in the global tests.

The implied failure load under a

single wheel was 21 kN compared with the actual failure load of over 200kN.
It is thus clear that the conventional analytical approach under-estimates
local strength, apparently by a factor of up to ten.
- 205-

It was nofed in

Chapter 2, however, that the use of linear analysis at the ultimate limit
state is merely a convenient way of avoidirig the requirement to check
stresses

under

service

loads.

It

may

therefore

be

considered

more

realistic to compare the failure loads with the predictions of yield-line


theory.

This approach, which is normally considered to give an upper-bound

solution, under-estimated the failure load in the single wheel tests by a


factor of approximately two.

The failure load of the slab in the global

tests was reasonably close to that predicted by yield-line theory although


the failure mechanism was so different that this can be little more than
coincidence.
Even

if

yield-line

reinforcement

analysis

used

in

design

to

BS

5-iOO,

the

provided would be little reduced since the stress limits

would become critical<35).


reinforcement
approach.

were

as

It

the basis of

is thus reasonable to use the Tl0-87.5


comparison with

the conventional design

Since it has been noted in previous chapters that the critical

criteria are serviceability criteria under full global load,

it

is most

realistic to compare the observed behaviour with the conventional design


approach on
considered

this basis.

As

just satisfactory,

the serviceability of the first

deck was

the best comparison with the conventional

design approach is to say that it over-estimated the steel required by a


factor of nearly four.
The steel area provided approximated very closely to that required by the
conventional approach to resist the global transverse moments alone.

b. Second Deck
The allowable load calculated for .the second deck was approximately 16
units of HB, which is slightly higher than for the first deck.

However,

the calculation was based on the strength of the slab in sagging.

It is

common, when designing this type of slab using Westergaard's approach, to


analyse only sagging and then to provide.the same steel in the top.

Since

the single layer of steel provided was lOmm below mid-depth, this approach
was not valid for this deck and analysis of hogging would have given a
lower allowable load.
It was not possible to design a single layer of steel to resist 45 units

of HB using normal design methods because the calculated bending moments


exceeded the concrete capacity.

Using two layers of steel, the requirement


-206-

was !ust marginally higher than for the first deck, due to slightly greater
global transverse moments.

However, this reinforcement was designed using

a nominal vnlue for concrete cube strength of 40N/mm 2

which is the normal

value used in deck slabs and was slightly conservative for the concrete
used in the first deck.

The concrete in the second deck was weaker and

using

strength

the

actual

cube

in

the

design

calculations

made

it

impossible to design the reinforcement for 45 units of HB.


The steel area provided in this deck also approximated very closely to
that required by the conventional British design approach to resist global
transverse

sagging moments.

However,

due

to

the reinforcement

being

located below mid-depth, the moment capacity in hogging was only some 50%
of the maximum transverse global moment given by the grillage.
As in the first deck, the failure loads in the single wheel tests were
approximately double the values predicted by yield-line theory.

9.2.2 Analysis for Design of Beams


Since

the beams were provided with more prestress

satisfactory

behaviour

in

the

service

tests

proves

than normal,
very

little.

their
The

distribution properties of the deck can only be investigated by comparing


predicted and measured displacements or strains.
a. First Deck
In Figure 9.!, the maximum mid-span beam deflection in the test on the
first deck is compared with the prediction of the linear grillage analysis
which used 9 nodes per beam.
load distribution is also shown.

For comparison, the prediction of a static


In order to eliminate the effect of error

in predicting the stiffness of the beams, as opposed to error in predicting


the distribution properties, the displacement is expressed as
the average displacement of the four beams.
whilst

the

behaviour

of

the

beams

is

are shown for the first

factor of

This approach is only valid

linear-elastic

throughout the range plotted in the Figure.

a_

but

this applied

The distribution properties

time that each load level was applied and the

breaks in the plot indicate points where, as described in Chapter 8, the


loading was not continuous.
Figure 9.1 appears to show that the grillage prediction using the grossconcrete properties for the slab is very good.
-207-

Initially, the distribution

was slightly better than the analysis predicts but this might be expected
because the analysis ignores the shear connection between the top flanges
of

the

beams.

As

the

load

increased,

the

distribution

deteriorated

slightly due to the reduction in transverse stiffness caused by concrete's


non-linearity and cracking in tension.

By 120kN per jack, the design

service load, the distribution was worse than the analysis suggests but
the deflection of the heaviest loaded beam was only 3.7% higher than the
prediction and, even under design ultimate load, the discrepancy was less
than 5%.

The global transverse moments predicted by the grillage implied

concrete stresses in excess of

5N/ mm ~

in the slab concrete over the beams

but this concrete was apparently uncracked.


Static Load Distribution
of Beam B 2.0

t:;

Mean
(~:;

= mid-span

~:;

beam 1.8

deflection)

Test

from start of

(t:;

this loading)

1.6

--- - - - -~---- - ----

Linear Grillage

-~--

1.4

Test (/:; from start of tests)


1.2
1.0
25

50

75

100

125

150

175

200

Load (ki/Jack )
Figure 9. 1.

Beam deflections of first deck

<conventional analysis )
The final part of the plot in the Figure, that for loads above 150kN/jack,
relates t o the load application at the end of the service load tests when
the load was returned to its original position.

By this stage, there was a

significant permanent deflection and, as will be seen from the Figure, it


makes a difference to the apparent distribution properties whether the
total deflections <that is, the deflections relative to the original zero)
or

the deflections only from

considered.

of this

load application are

The latter probably gives a better indication of the live- load

stresses so
continues

the start

to

it

appears

deteriorate

that,

as

might

as

the

load

be expected,
increases.

the distribution
Despite

this,

distribution remains very much better than the static load distribution.
- 208-

the

Deflections

are

frequently

used

as

properties of beam and slab decks.

an

indicator

of

the

distribution

However, it is the tensile stresses in

the soffits of the beams which are normally the critical criteria in design
so these are a better indicator.

In Figure 9.2, the predicted strains on

the top and the bot tom of each beam at mid-span are compared with the
measured strains for a particular load level.

Because the beam behaviour

was linear- elastic at this stage, it is reasonable to assume that stress


was proportional to strain.
As in Figure 9.1, the results are expressed as a factor of the average
value f or the four beams.

This makes the grillage predictions identical

for the top and soff it.


Strain at Beam
Mean Strain at Beams

Test <top of slab )

Test <soffit >


Linear Grillage

1. 5

1.0

+
0

0.5

+
0.0
Beam A
Figure 9.2:

Beam C

Beam B

Beam D

Beam strains 1n firs t deck

(end of service load tests: load

= 150kN/jack)

Because the distribution properties of the deck are poor, the analysis
predicts
beams.

very significant

differences

between

the stresses in adjacent

This is reasonable for the soffit stresses because there is no

direct connection between the bottom flanges of the beams.

However, the

analysis implies that the stress distribution across the slab is literally
as shown in the Figure; with large discontinuities between adjacent pieces
of concrete.

This is impossible and shear in the slab evens out the

longitudinal stresses in the slab as can be seen in the Figure.


this does not even out the soffit stresses.
- 209 -

However,

Indeed, because it reduces the

differential dieflections of the beams and hence reduces the contribution


of global transverse moments to distribution, it can cause a deterioration
of the distribution properties as expressed by soffit stresses.

The result

is that whilst Figure 9.1 suggests that the distribution at this load stage
is only 5% worse than the grillage prediction, Figure 9.2 shows that it is
16t worse.

This may suggest that the distribution properties of this deck, with its
very light reinforcement, were unsatisfactory and that the fears expressed
in 3.2.7 are confirmed.

However, the increase in the percentage load

carried by the heaviest loaded beam was only approximately

5~

between the

initial linear condition and the load stage considered in Figure 9.2 which
is the design ultimate load.

This implies that, even if the deck slab had

been so heavily reinforced

that

it remained effectively linear-elastic

after cracking, the grillage based on gross-concrete slab properties would


have

under-estimated

the maximum soffit

stress by approximately

10~.

Thus most of the discrepancy was due to a normally accepted error in


analysis for design, not the reduced steel area.
b. Second Deck
In Figure 9.3, the greatest mid-span beam deflection in the tests on the
second deck is compared with the prediction of a linear grillage.

The

prediction of the static distribution is not shown but it corresponds to a


value of 2.5 in the Figure.

As in Figure 9.1, the maximum beam deflection

is expressed as a factor of the average deflection of all the beams.


However, because of the parapet up-stands in this deck, the four beams
were not identical.

The approach is not, therefore, quite such a reliable

guide to distribution because errors in predicting the difference between


the stiffness of the edge and inner beams would show up in the Figure.
The longitudinal strains in the slab over the beams are illustrated in
Figure

9.~

Figure 9.5.
first

deck,

and

the

soffit

strains

in

the

beams

are

illustrated

in

The reason for using two figures, rather than one as for the
is that

because of the different edge beams the grillage

predictions in the two figures are different.

The reason for choosing a

different load stage to illustrate is related to difficulties experienced in


both tests with the strain gauges.

- 210-

A of Beam C 2 . 0

Mean
(A=

Test

from start of

(A

this loading>

mid-span beam 1.8

--

deflection)
1.6

Linear Grillage

1.4

Test

(A

from start of tests)

1. 2

1. 0

75

50

25

100

125

150

175

200

225

Load <kN/ Jack)


Figure 9.3:

Beam deflections of second deck

<conventional analysis)
Figure

9.3

appears

to

suggest

that

the

distribution

significantly better than the grillage prediction.

was

initially

Assuming that this

difference was due to the shear connection between the top flanges the
extent of the difference is surprising: theoretically it should be less
than

for

the

first

deck.

However,

a detailed study of

the results

suggested that part of the discrepancy may have been because the analysis
exaggerated the effect of the parapet up-stands and hence exaggerated the
stiffness of the edge beams.

This was despite the analysis using the

measured

beam

and

than

the

values

E"'

substantially

more

for

the

different

strength difference would suggest.

parapet
nominal

concretes

which

(or

the

even

were

actual>

The reason for the small effect of the

parapets was probably that they were cracked due to plastic settlement .
The first break in the plot in Figure 9.3 corresponds to the end of the
first

load cycle to service load.

At the end of this cycle the more

heavily loaded beams did not return to their original positions.

The outer

beams

permanent

had

small

negative

deflections

suggesting

that

the

deflection was mainly due to some of the transverse curvature becoming


permanent.

This implies that there were locked-in stresses in the beams.

Since the next part of the plot relates to a load application immediately
after the first, these stresses would not have been relieved greatly by
creep.

Thus the solid line, relating to the distribution calculated from

the total deflections,

probably gives the best


- 211 -

indication of the load

distribution in the next section of the plot.

In contrast, the permanent

de flections at the s t art of the load application represented by the fina l


part of the plot were relatively uniform across the deck, implying that
they were

largely due

dotted line gives a


stage.

to permanent beam curvature and

thus that

the

better indication of the load distribution at this

It will thus be seen that the distribution properties deteriorated

progressively throughout the test and that the deterioration was greater
than

for

the

first

deck.

The greater deterioration

might

have

been

expected both from the extent of apparently global cracking in the slab
and from the fact that the steel provided was inadequate to resist the
global transverse moments predicted by the analysis.
Stra in at Beam

2.0

Test

Mean Strain at Beams

Linear Grillage
1.5

1.0

0.5

0.0

Beam A

Figure 9.4:

Be a m B

Beam C

Beam E

Beam D

Slab strains over beams in second deck

(f irst loading;

load = 120kN/jack)

Figure 9.4 shows that, as for the first deck and for the same reason, the
stresses
predicts.

in

the

slab

were

more

evenly

distributed

than

the grillage

Figure 9.5 shows that the maximum soffit stress was higher than

the analysis predicts.

The margin is so small it would not normally be

considered significant.

However, Figure 9.3 shows that there was further

deterioration

in

the

Figure 9.5 relates.


that

the

worst

distribution properties after

the stage to

which

By the completion of the service load tests it appears

soffit

stress

was

approximately

10%

higher

than

the

grillage prediction although still some 30% lower than the static load
distribution suggests.
- 212-

2. 0 .

Stra in at Beam

Test
Linear Grillage

Mean Strain at Beams


-t-

1. 5

+
1.0

...

-
+

0.5

0.0

Beam A
Figure 9.5:

Beam B

Beam C

Beam D

Beam E

Beam soffit strains in second deck

(first loading: load = 120kN/jack)

9.3

NON-L~

ANALYSIS

9 .3.1 Single Beam

The s ingle

beam

test

served

to

check

that

the

analysis modelled the

behaviour of the beams correctly, and thus to ensure that any errors in
the predictions for the behaviour of the model bridge decks were not due
to failure to model the behaviour of the beams.

Because of this it is

convenient to consider these tests first.


The predicted and observed load-displacement relationships are shown in
Figure 9 .6 .

They are as close together as can reasonably be expected;

larger discrepancies are frequently observed between the behaviour of two


nominally identical concrete specimens even when they both come from the
same batch of concrete.

- 213-

Load
500
<kN/Jack)

400
~

300
200

- - - - Analysis
-Test

100

60

40

20

80

100

120

Deflect ion
Figure 9.6:

<mm>

Analysis of single beam tesc

9.3.2 First Deck

a. Global Tests; coarse mesh analysis


The deck was initially analysed using a 15 by 12 node model.

This gave

four

considered

transverse

reasonable.

elements

between

each

beam,

which

was

However, these elements were 727mm wide so it was assumed

the model would be too coarse to model the local behaviour correctly.

The

reason for using such a coarse mesh was that this was the largest model
which would fit in the 386 desk- top computer used.
The computer model was loaded monotonically to failure under the load case
used in the tests and the predicted central deflections of the beams are
shown along with the test results in Figure 9. 7.

Considering the many

approximations in the analysis, the complexity of the behaviour and the


usual variability of the behaviour of concrete structures, the predictions
are

reasonably

good.

Because

the

analysis

was

performed

under

load-

control using relatively large increments, it did not pin-point t he failure


load precisely.

However, the best estimate of the failure load whi ch could

be obtained from the analysis was 400kN per jack compared with the actual
value of 414.

The analysis also predicted correctly several features of

the behaviour which might otherwise have been considered surprising.

It

predicted that the first cracking would occur in the soffit of the slab
under wheel 10 (as shown in Figure 8.11) but that at later stages of the
loading the slab would be much more highly stressed in the outer bay, that
is under wheels

to 4.

It

also predicted correctly that, as

failure

approached, the slab adjacent to wheel 4 would be subjected to transverse


- 214-

sagging moments right out to Beam A.


at

failure

and

the

mode

of

Finally, it predicted both the load

failure

surprisingly well;

in

the

final

increment plotted in the Figure, concrete was crushing most extensively


under wheel 4. and it was
converge

in t his regi on t hat the anal ys is

failed to

gave l arge deflect i ons in the next increment.

an~

Beam D
400
Load
<kN/ Jack >

Beam A

Beam C

Beam B

350
300
250
2 00
150

Analysis

100

Test

50
0

10

20

30

40

50

60

Deflection <mm>
Figure 9. 7:

Beam deflections o f first deck

(from non-linear analysis using a coarse element mesh )


With such a

coarse element

mesh,

the correct

prediction of such an

apparently local failure might be considered surprising.


be three explanations for it.

There appear to

Firstly, although the final collapse took the

form of local punching due to the local shear stress, it was apparent that
the concrete was crushing along a line which extended from wheel 3 to
slightly

beyond wheel 4.

This crushing,

which apparently caused the

failure and which was predicted by the analysis, thus extended over the
width of two elements of the computer model enabling the resulting failure
to be predicted.

Secondly, as was found in Chapter 7, the analysis tends

to be conservative in its prediction for punching failure loads and this


cancelled out the failure to model the peak of the stress concentration
under the wheel; thus a finer mesh would have led to an under-estimate of
failure load.
model,

the

Thirdly, in an attempt to cancel out the fault of the coarse


loads

were

applied

as

points

distribution over the patch area.

- 215-

with

no

allowance

for

the

This

ability

of

the

analysis

to

predict

failure

load

and

mode

is

reassuring; in particular, the fact that it still gives low predictions even
when a coarse mesh is used means that

is safe for

it

use in design.

However, the model tests suggest that the limiting service load for the
bridge was equivalent

to around 120 to 150kN per jack.

In BS 5400, a

design service HB load of !50kN corresponds to a design ultimate HB load


of 204kN*.
should

be

requiring

Even allowing for the fact that the material safety factors
applied
an

debatable),

in

extra
it

the

analysis

15% strength

is

clear

that

for

for

the

ultimate

brittle

serviceability

failure

is

limit

state,

mode

critical.

and

(which

is

Thus

the

there was local soffit cracking under wheel 10 at

the

predictions for the lower load stages are more important.


In the analysis,

first increment, 50kN per jack, and there was limited top cracking in the
slab by the second increment, IOOkN per jack.

In the tests, such cracking

was not observed until loads of 110 and 300kN per jack respectively.
main

reason

for

this

very

large

discrepancy

appears

to

be

that

The
the

analysis assumed the concrete to be linear-elastic until cracking, whereas


in fact

there clearly was a significant departure from linearity before

cracking.

The fault was exaggerated by the use of a low tensile strength

for the concrete in the analysis, 0.67 times the split cylinder strength.
This was chosen because it gave the best results in Chapter 6 for lightly
reinforced specimens.

However, because of the scope for redistribution to

the steel, the use of the full split cylinder strength gave better results
in heavily reinforced specimens.
lightly

reinforced

there

was

Although the slab of this bridge was

great

scope

for

redistribution.

It

thus

appears that in this respect, as in their failure mode and load, lightly
reinforced

restrained

slabs

behave

like heavily

reinforced

unrestrained

slabs.

*Footnote
The ratio of design ultimate to design service load implied by this is
higher than that used in Chapter 8.
This arises from the Author's
interpretation of the factor y~ 3 in BS 5400.
y," is a partial safety
factor for errors in analysis which, in BS 5400: Part 4, is applied to the
loads. Since the form of analysis considered here is not elastic, a factor
of 1.15 is used at the ultimate limit state as specified by the code.
However, when considering the load to be applied to models, the author has
assumed that, since no analysis is involved, y; 3 can be 1.0. This might be
considered debatable since y, 3 also covers errors in dimensions. However
applying y,,. to the loads used in Chapter 8 would not alter any of the
conclusions.
- 216-

The

behaviour

linearity of

in

the

tests

concrete

in

was

clearly greatly

tension

tension after cracking.

and

Although

by

affected

its ability

to

by

the

non-

transmit

some

the analytical prediction of cracking

under wheel 10 by a load of 50kN per jack was greatly premature, the
measured strain in this region at this stage was 100 microstrain which,
with

E~

the measured

value,

implied a

stress of approximately 3N/mm"'

Assuming concrete to be linear elastic in tension until it cracks, this


would

certainly

imply

that

cracking

was at

least

imminent.

In

fact,

although non-linearity is visible in Figure 8.12 from approximately this


stage and becomes very pronounced by a load of 75kN per jack, cracking was
not visible until a load of 1l0kN.
Although the analysis exaggerated the amount of cracking, there was not a
corresponding exaggeration of the stresses in the slab.

Using the analysis

in the same way as a conventional linear analysis, that is calculating the


stresses from the element

forces given by the program using a cracked

elastic section analysis ignoring the tensile strength of the concrete, the
allowable service load from the BS 54.00 criteria was approximately 1l0kN;
the critical criterion being the steel stress.

Although the actual steel

stress in the model was unknown and probably substantially lower than the
34.5N/mm2 implied by this, it was concluded in Chapter 8 that the behaviour
was just acceptable for the load history applied which had been intended
to simulate the life of a bridge with a design service load of 120kN.
Thus the analysis was conservative although it still allowed nearly three
times the load on the deck that a conventional analysis would allow.
For reasons discussed
material

models

concrete's

to

ductility

in Chapter 6,

make
in

development of cracking.

it

the

analysis

tension

and

is not

possible to adjust

reproduce

hence

to

the

full

predict

the

effect

correctly

of
the

Indeed, it is debatable whether this is desirable

since the effect is probably size dependent and thus an analysis which did
this for the model would be incorrect for a full size bridge.
analysis appeared
development

of

to be as good as possible.

cracking

did

have

an

However,

undesirable

effect;

Thus the

the premature
it

made

the

analysis exaggerate the rate of decay of the distribution properties, a


trend which can be observed from Figure 9. 7.
Before

cracking,

substantially

the

better

predictions
than

those of

for
a

- 217-

the

strain

conventional

in

the

beams

linear grillage;

were
not

because of non-linear it ies but

because of modelling

shear connection of the top flanges of the beams.

the effect

of the

However, by a load of

150kN per jack, the error in the predicted soffit strain was as great as
that of the linear grillage although, unlike for the linear analysis, the
error was in
properties
analysis.

the safe direction.

was

due

to

the

The premature decay of distribution

premature

development

of

cracking

This was not entirely due to the material model used.

partly due to the failure of the analysis


finite width of the beam webs.

in

the

It was

to model the effect of the

The predicted hogging moment in the slab

at the position corresponding to the face of the web in the model was
little more than half

that over the centre-line of the beams.

Thus a

length of transverse element which was, in fact, uncracked and effectively


very deep was modelled as being shallow and cracked.

b. Global Tests; fine mesh analysis


The model bridge was re-analysed using a finer mesh to give a better
indication of the behaviour.

Because the previous analysis had shown that

the most highly stressed regions of the slab were not confined to small
areas

and

that

the

most

highly stressed region

moved as

the

loading

progressed, it was undesirable to restrict the fine mesh to local critical


areas,

as

Chapter 7.

had

been

done

in

the

analysis

of

Kirkpatrick's

tests

in

Because of this the computer model was too large to run on

the desk top computer and it was transferred to a Vax 111750 machine.
This

greatly

increased

the

space

available

but

the

machine

was

significantly slower than the 386 and this imposed a practical limit on
the size of model which could be analysed.
Six transverse elements were used between each beam and those adjacent to
the beam were made shorter so that their ends coincided with the face of
the web.

They were given a full width lOOmm deep web to represent the

presence of the top flange of the beam.

32 nodes were used along the

span of the bridge giving 258mm wide elements and a total of 672 nodes.
The full split cylinder value was used for the effective tensile strength
of the concrete and, unlike in the coarse mesh analysis, the finite size of
the load patches was represented.
The

load history of

the service

tests

was

simulated

removing the test load from the six different positions.


- 218-

by applying and
However, in order

to keep the computer time used within reasonable limits, only the first
and last loadings were analysed in detail.

The loads were applied to and

removed from the other positions in single increments.


The predicted displacement of wheel 9 relative to the beams is shown in
Figure 9.8.

On first loading, the analysis still under-estimated the load

to cause visible cracking although not by as large a margin.

It also

over- estimated the loss of stiffness caused by this cracking.

This is

again due to its failure to model the ductility of concrete in tension.


This

is

far

more

significant

factor

in

this

highly

indeterminate

structure than in the simple strips considered in Chapter 6.

In a simple

beam, despite the ductility of concrete in tension, as soon as cracks form


they extend well above the soffit.
despite

the

relatively

brittle

In the analysis of this bridge deck,


concrete

model

used,

the

scope

for

redistribution meant that it was common for cracks to form at only one of
the eight integration stations through the depth of the slab.

This meant

that the area of concrete which is strained out of the linear range but
still resisting tension is larger and further from the neutral axis.

It is

thus far more significant t o the behaviour.

250

Load
<kN/ Jac k )

200

---Analysis

150

-Test

100
50
loading in other positions

0
0

0.5

1.0

1.5

2.0

Deflection <mm relative to beams)

Fig ure 9.8:

Deflection under wheel 9

(from non-linear analysis using a fine element mesh)


The over-estimate of the deflection in the final loading is also probably
due to the failure to model the effect of the ductility of concrete in

- 219-

tension.

However, considering the many unknowns in the analysis and the

complexity of the behaviour, the prediction is remarkably good.


Unfortunately, the analysis could not be taken up to failure.
300kN

per

jack,

when

the

effect

of

cracking

in

the

By a load of
beams

became

significant, the convergence rate became excessively slow and it was not
possible to obtain a sufficiently accurate solution using a reasonable
amount of computer time.

However, the analysis appeared to confirm that

the allowable service load on the deck was approximately 120kN per jack
and the design ultimate load corresponding to this is only 172 .5kN per
jack.

Thus the analysis had shown that the deck had at least 50% more

ultimate strength than could be used in design.


The beam deflections predicted by the analysis are shown in Figure 9.9.
Because of the use of a higher tensile strength for the slab concrete, and
because of the modelling of the finite width of the beams, this analysis
predicted better distribution properties than the coarse analysis.
predicted

beam

deflections are as close to

the

test

The

results as can

reasonably be expected.
Beam D Beam A Beam C
Load
<kN /J ack )

25 0

Beam B

200

150

Analysis

100

--Test
50

0
5

10

15

Deflection
Figure 9.9:

20
(mm)

Beam deflections of first deck

(from non- linear analysis using a fine element mesh)


Since the computer models had given good predictions for the behaviour of
the deck, it seemed reasonable to use them to obtain some insight into the
mechanism

by which

this behaviour was achieved.


- 220-

In Figure 9.10 the

transverse restraint force across the longitudinal centre-line of the deck


predicted by the computer model is shown for a particular load stage.

As

might be expected from previous research, there is a compressive force in


the region of the wheels.

However,

the compressive force

in the end

regions of the deck (which is apparently due to the effect illustrated in


Figure 3.12) is much greater.
of the deck is in tension.

Thus, with these forces to resist, the rest

The behaviour is different from that described

or implied by other researchers: with significant compressive force in the


end regions their assumption that diaphragms are needed to resist tension
cannot be correct.

The tension required to resist the compression in the

critical areas comes


critical areas.

from

material which

is

relatively

close

to

those

The analysis also suggests that t he restraint force is

more localised than previously supposed in the transverse direction.

The

plot in Figure 9.10 is based on the average of the forces in the elements
on

either

side

of

the

centre-line.

predicted for a section at

the

The

transverse

restraint

face of the web is markedly different,

barely going into tension at all adjacent to the wheel positions.


appears to confirm

force

the suggestion in 3.2.5

This

that membrane action under

service loads is not dependent on external restraint and could still be


significant in unrestrained slabs.
It is also clear from Figure 9.10 that the resistance to global transverse

moments is not enhanced by compressive membrane action since much of the


relevant area of the deck is in tension.

.Wheel Posmons

Resrraint 200
Force (kNAN
1
100

-100

Figure 9.10:

Predicted force across centre-line of first deck

(from fine analysis; load

= 250kN/jack)

The coarse mesh analysis suggested that the area of slab in compression
around the wheels extended in both directions as failur e approached.
- 221-

The

restraint

force also increased disproportionately as the load increased.

However, the general form of Figure 9.10 remained unchanged and the end
regions of the deck were subjected to an increasing compressive force.
As the predicted distribution of restraint forces was so different from
that

implied

or

described

by

previous researchers,

was considered

it

highly desirable to check it by attempting to measure the restraint forces


in the model.
concrete,

Because of the many variables in the behaviour of cracked


was

it

considered

best

to do

this

for

section where the

concrete would be uncracked.

Accordingly, a series of transverse demec

points

top

were

attached

to

the

and

bottom of

the slab at

matching

positions close to the assumed point of transverse contraflexure in the


slab.

Unfortunately,

results.

Firstly,

sufficient

to

be

there

were

although

the

highly

many

in

interpreting

the

indicated

in

Figure

are

forces

significant

reinforced cracked section,

difficulties

to

the

they represent

concrete section, typically 1N/mm 2

behaviour

of

9.10
a

lightly

low stress on the gross-

which makes them difficult to measur.e.

This was made worse by the significant longitudinal compression in the


deck due to global moments.

It was necessary to correct for the Poisson's

ratio effect of this and, although longitudinal demec points were provided
to enable the longitudinal strains to be measured, the corrections were
inevitably inaccurate if only because of the uncertainty in the Poisson's
ratio used.

Since the correction was often significantly greater than the

measured transverse strain, errors in the correction had a large effect on


the estimated transverse forces.
A second difficulty was caused by the effect of cracking.

Although there

were no visible cracks within the gauge lengths, some of the measured
tensile strains were in excess of 100 microstrain which would normally be
taken to imply that there would be some non-linearity in the behaviour.
More seriously, cracks outside the gauge length but close enough to affect
stresses within
considered

in

it

<that

6.2.1)

could

concrete transferring it
making

the

is cracks within S0

restraint

release

some of

to the steel.

force

estimated

of

the

the gauge
tensile

length as

stress

in

the

This would have the effect of


from

compressive than the actual restraint force.

the

strain

readings

It appears that

more

this must

have been a significant effect since integration of the restraint forces


estimated

from

all

the

readings

appeared

to

imply

that

significant net transverse compression across the bridge.


-222-

there

was

Since the beams

were

restrained

impossible.

only

by

the

flexible

elastomeric

bearings,

this

was

However, it is perhaps significant that if the bridge had been

provided with

diaphragms

one

might

reasonably have supposed

that

the

"compression" in the slab was resisted by tension in the diaphragms.


The one case where the demec readings did give a reasonable indication of
the transverse force w11s for the ends of the slab when the

bridg~

was

loaded in the first position.

Since this was a free edge, the longitudinal

stress was clearly zero so no Poisson's ratio correction was required.


Similarly, since the nearest visible crack was over a metre away it seemed
reasonable to suppose that the strain readings could not have been much
affected

by

cracks.

Another

advantage

was

that

it

was

possible

to

position demec points at mid-depth of the slab as well as on the top and
bottom surface.

Unfortunately, where this was done, the mid-depth gauge

gave a strain which was significantly different from the mean of the top
and bot tom gauges.
behaviour
plane.

which

This was presumably due to non-linearities in the

invalidated

the

assumption

that

plane sections

remain

The maximum measured tensile strain, over 150 microstrain, also

implied that concrete non-linearity was possible.

These difficulties meant

that, even for the. ends of ttie slab; the restraint


_estimated to within plus or minus some 50%.

force could only be

Nevertheless the results were

significant; all four_ demec sets showed a compressive strain which was

approximately as predicted by the analysis.

Given that previous research

implied that this region should be in -tension, this alone appeared to be


_sufficient to show that Figure 9.10 was closer to reality than were the
implications of previous research ..
The restraint
behaviour

but

enhancement

forces
not,

predicted by the
on

their

own,

analysis. are significant

sufficient

t_o

explain

the

to the
enormous

relative to the predictions of conventional design methods.

An equally significant mechanism is moment redistribution.


factor here is the orthotropic nature of the cracked slab.

An important
It has already

been noted that, with such light reinforcement, the cracked stiffness is
only some 10% of the uncracked stiffness.

rt is clear from the figures in

Chapter 6 that the tangent stiffness of the cracked section is lower still.
Under

global _load,

the deck slab was subjected

to a

very significant

longitudinal compressive stress which delayed the_ formation of transverse


cracks,
value.

hence the longitudinal stiffness remained at

its full uncracked

The result was that the distribution of the transverse moments,


-223-

both in the real slab and in the analysis, was very much more uniform than
implied by a conventional analysis.

This redistribution of the moments is

due to cracking hence, unlike redistribution due to reinforcement yielding,


it

starts to take effect before there is any material damage which is

unacceptable under service loads.

c. Local Test
The

distribution

of

restraint

forces

indicated

in

Figure

9.10

significantly different from that implied by previous research.


the

load

case

considered

was

also

significantly

different

is

However,
from

that

investigated by previous researchers in that 16 wheel loads were applied


It is possible that this was the reason for

instead of only one or two.


the difference.

To investigate this, it was decided to re-analyse the deck

for a single wheel load.

The wheel was positioned in approximately the

position of single wheel A in the tests but the analysis was not directly
comparable with the test.

In the analysis the load was applied to the

undamaged bridge whereas in the tests it was not applied until after the
bridge had been loaded to failure under full global load.
Because

of

the

stiffer

initial

uncracked
response

slab

than

the

was

analysis

observed

predicted
in

the

significantly

tests.

As

failure

approached, the crack pat tern in the test began to be dominated by the
single wheel and consequently the difference between the test and analysis
reduced.

The analysis converged under a load of 220kN but indicated that

failure due to local concrete crushing round the wheel would occur before
230kN.

This is remarkably good agreement with the failure load in the

test which was approximately 226kN.

However, although very fine for the

analysis of a whole bridge deck, the element mesh used was still slightly
too coarse for a local analysis as is indicated by the large difference
between the restraint forces in adjacent elements in Figure 9.11.

It is

likely that a finer mesh would have given a slightly lower failure load.
It is also possible that

higher

if

failure.

the failure load in the test would have been

the deck had not

been damaged by

the

previous

loading

to

However, other tests and analyses suggest that this effect would

have been very small.


In Figure 9.11 the restraint force predicted across the centre of the bay
of slab between Beams D and E is illustrated for two different load levels.
The

first

of

these,

60kN;

is

close

to

- 224-

the

wheel

load

considered

in

Figure 9. 10.

For this relatively low load, the forces are plotted only for

the region around the wheel.


Wheel Position

FEs traint 400


force (k NJrrJ

200

-200
Fig ure 9.11:

Restraint forces predicted under single wheel load

There is a compressive force in the end regions of the deck but this is
very much smaller relative to the wheel load than in Figure 9.10.
confirms

that

the

This

force is due to global effects which are much less

significant with only one wheel

loaded.

The compressive force in the

region of the wheel under 60kN is greater than in Figure 9.10.

This is

partly due to the lack of the global tension force near mid-span.

However,

a comparison of Figures 9.10 and 9.11 reveals another explanation.


compressive

restraint

superimposed on

the

force
tensile

under
force

each

wheel

in

Figure

due to an adjacent

wheel.

9.10

The
is

This is

another reason why behaviour in single wheel tests is an unreliable guide


t o behaviour under multiple wheel loads.
As

the

load

increases,

the

restraint

force

in

Figure

9.11

increases

disproport ionately, particularly for the elements either side of the wheel.
However, even as failure approaches, the area in compression around the
wheel is comparatively localised.

The restraint required to resist this

compression comes from the slab i mmediately on either side of the wheel
and the end regions continue to be subjected to compression.
the global tes ts, the analysis shows
provide the restraint .

Thus, as in

that diaphragms are not needed to

The source of restraint is substantially different

from that implied by previous research.

However, if the cr itical region is

considered to be applying a compressive force to the r est of the sla b, t he


rest of the slab is analogous to a slab with a hole in it across which a
compressive force has been applied.

Figure 9.11 is remarkably consi stent

- 225-

with the r e sult of an elastic analysis of such a case which is shown in


Figure 9.12.

St ress Across Centre -Line

Unit Applied St ress

Figure 9.12 :

Elastic analysis of stresses around a hole

[from finite element analysis by Mehkar-Asl <130 )]


Under 220kN, the analysis was very close to f ailure and the stress in the
concrete immediately below the wheel was j ust starting to reduce from its
peak va lue a s the concrete began to crush.

The maximum restraint force

predicted was only approximately half of that suggested by the rigidplastic strip theory considered in 3.2.1.

This was

entirely due to the

requirem en t for compatibility and to the lack of ductility of the concrete.


The fact

that

the slab failed in the analysis before reaching the full

plastic momen t capa city could not have been due to shear as the analysis
does not model this effect.
As with the global tests, an attempt was made to ascertain whether the
real restra int forces were as predicted by the analysis.
was

already

cracked

difficult to do this.

before

the

tests were st arted,

Because the slab


it

was

even more

All that could be established with any certainty was

that the general form of Figure 9.11 is reasonable.

d. Mod ified Decks


In the last chapter it was suggested that if the beams had been provided
with less pres tress, the deck could still have failed in the same way but
at a l ower l oad.

To investigate this, the bridge was re- analysed using

on ly 60% of the prestress area.

To reduce the computer time required, and

to enable the analysis to be taken up to f ailure, the coarse element mesh


was used.
The beams were not predicted to crack until the load was 175kN per jack
and up to this point the behaviour was not affect ed by the reduction in
- 226-

prestress.
load

Thus the analysis confirmed that the behaviour in the service

tests

would

have

been

identical

without

the

over-provision

of

prestress.
The analysis predicted the same form of failure as before but at a lower
load of approximately 320kN per jack.
failure

was greater,

indicating that

The maximum beam deflection at


the slab could withstand greater

differential beam deflections when subjected to smaller local loads.


The load at which the slab was predicted to fail was below that predicted
by yield-line theory.

This is entirely consistent with the explanations of

the behaviour given earlier in this thesis.

It also suggests that the

supposedly conservative approach to design allowing for membrane action of


using yield-line theory, which was proposed by Tong and Batchelor(51>, is
potentially unsafe.
Analysis using the coarse element mesh was also used to investigate the
effect of varying the quantity of reinforcement in the slab.

Two analyses

were performed, one using the actual quantity of secondary steel with
double the quantity of main steel and another in which both the main and
the secondary steel were reduced to half that which was actually provided.
In both cases, the prestress and also the additional transverse bars in the
end regions of the slab were as provided in the model.

The allowable

service load implied by these two analyses were approximately 190 and
60kN per jack respectively whilst the failure loads were approximately 440
and 375kN.
The service loads were obtained from normal BS 5400 criteria using the
worst stress at a crack calculated ignoring the concrete in tension; the
"stress at crack approach" described in 7.7.3.

The steel area in the

lightly reinforced slab was so low that this approach predicted high steel
stresses as soon as the concrete cracked, that is before the cracking was
extensive enough to develop much membrane action.
service

load

of 60kN predicted

probably too low.

in this

way

Because of this the

is very approximate and

However, this behaviour might be taken to imply that

the steel area was below the desirable absolute minimum.

At 0.18% it was

above the code nominal steel area but the low d/h ratio in a thin deck
slab means that the minimum steel area expressed as a percentage of the
net section should be higher than normal.
-227-

The analysis with increased steel area suggested that doubling the area of
main steel increased the service load by over 50%.

Although well below

the near linear relationship given by normal design methods, this is a


greater effect than implied by previous research.

This is because the

steel contributes to the resistance to global moments and also to the


restraint.

Under single wheel loads, as tested by previous researchers,

the global moments are insignificant and


available

to

provide

the

restraint.

there is relatively more slab


Thus

these

effects

are

less

deck

slab

pronounced.
A

final

analysis

was

performed

using

single

layer

of

reinforcement as provided in the second deck, although still with the extra
bars in the end regions of the slab.
higher allowable service load than

This analysis suggested a slightly


for

the steel actually provided,

in

contrast to the analysis of the second deck which will be described in the
next sect ion.

The

failure

load was, however,

reduced to approximately

375kN per jack.


All the analyses

predicted the same

punching through the deck.

failure mode with the same wheel

However,

the failures were clearly greatly

influenced by global transverse moments.


steel

area

was

to

improve

the

A major effect of increasing the

distribution

properties

of

the

decks,

particularly in the later stages of the analysis as failure approached.


Thus, although the most heavily reinforced slab had the smallest rot at ion
capacity,

and hence failed

when

the differential beam deflections were

relatively small, it failed at the highest load.

At failure, the predicted

deflection of the heaviest loaded beam <Beam B> was similar to that in the
test and analysis of the actual model but the deflections of all the other
beams were significantly greater.

e. Analysis with no Concrete Tensile Strength


The

coarse mesh analysis

was also used

to investigate the effect

reducing the tensile strength of concrete to zero.

of

This analysis gave a

failure load of approximately 375kN which is a reduction of less than 10%


compared with the original analysis.
by a similar percentage.

The implied service load was reduced

These relatively small reductions indicate that

the tensile strength of concrete is not as important to the restraint as


might have been supposed.
that

the

slab

was

This arises because the global moments ineant

cracked

over

much

-228-

of

its

length, reducing

the.

It implies that much of the restraint

contribution of concrete in tension.


actually

comes

from

the

under-stressed

reinforcement

away

from

the

critical areas and confirms that reinforcement is necessary in deck slabs.


It

was

noted

in Chapter 7

that

this

form

of analysis has

practical advantage of not being load history dependent.

the major

In this case, it

did provide conservative answers and would have been a reasonable design
approach.

However,

always be the case.

it was also noted in Chapter 7 that

this may not

Another disadvantage of this form of analysis is that

it gives over-conservative predictions for the distribution properties.


this case it

over~estimated

In

the worst beam moment under service loads by

10%.

9.3.3 Second Deck

Only one computer model was used for the second deck.

This used the same

width of transverse elements as the fine mesh analysis of the first deck,
258mm, but it used only four elements across a slab span.
total of 608 nodes.

This gave a

The use of six elements across a slab span, as in the

fine mesh analysis of the first deck, would have required 864 nodes and a
significantly greater band width,
amount
elements

of

computer
across

time.
slab

which would have needed an excessive

The major disadvantage of using only


span

is

that

it

prevented

the

model

four
from

representing the finite width of the beam webs.


As with the fine mesh analysis of the first deck, a complete load history
analysis was performed.

The predictions for local deflection were not as

good as

deck with

for

consistently

the

first

over-estimated,

the deflection

typically

by

50%.

under wheel 14
This

was

being

undoubtedly

largely due the failure to represent the finite width of the beam web.
The analysis implied a transverse moment at

the face of the beam web

which was only some 50% of that over the centre-line of the beam.
was more significant

than in the

first

This

deck because the critical slab

section was over the beam rather than at mid-span of the slab.
The

analysis

predicted

lower

steel

stresses

than

in

the

first

deck.

However, using conventional BS 5400 design criteria, the allowable service


load would still have been lower at just under 100kN per jack compared
with 120kN for the first deck.

The critical criterion was the concrete

stress in the soffit of the slab over Beam D adjacent to wheel 14.
- 229-

This

contrasts with the hypothetical analysis of the first deck with only one
steel layer which gave a higher service load.

This is due to the greater

hogging moments in the second deck and the fact that the single layer of
steel was below mid- depth.

Another factor was the higher concrete grade

in the first deck.


By

150kN

per

jack

the

concrete

in

the

critical

significantly beyond the limit of linear elastic


computer model.

region

was

stressed

behaviour used in the

This, and the more extensive cracking in this bridge,

meant that the analysis predicted a much greater difference between the
behaviour under the first and last applications of the service load than
for the first deck.

This reflected the real behaviour of the bridge and

confirmed the implication of both the analysis and the test results that
the 150kN per jack applied in the tests was above the desirable service
load for this structure.
Beam E

4-50
Load
<k N/Jack )
4-00

Beam B
/

,...........Beam

.,/"

.,/"

350
300
250
200
150

Analysi s

100 -

Test

50
0
0

10

20

30

40

50

60

Deflect ion (mm)


Figure 9. 13:

Beam deflections of second deck

<from non-linear analysis)


On completion of the analysis of the service load tests,
model

was

loaded

monotonically

to

deflections are shown in Figure 9.13.

failure

and

the

the computer

predicted

beam

As for the first deck, the analysis

correctly predicted the failure mode; it predicted concrete crushing on the


slab soffit over Beam D followed by excessive local deflection under wheel
- 230-

14.

However, the prediction of failure load was not quite as good, the

analysis under-estimating

this by over

10%.

Indeed it would be more

realistic to say that the analysis under-estimated failure load by nearly


20% since
plotted

in

it

did not

Figure

converge properly under the last

9.13

and

at

this

stage

it

load increment

also gave

an

excessive

deflection under wheel 14 of over 20mm relative to the beams.


This greater conservatism of the analysis compared with the coarse mesh
analysis of the first deck might have been attributed to the finer element
mesh or to the greater significance to this deck of the failure to model
the web width.

However, the fact that the analysis predicted significant

increases in the deflections of Beams B and D as well as C in the final


load increment plotted in Figure 9.13 suggests that it was the prediction
of the beam behaviour which was at fault.
load had

been due

to

If the low prediCted failure

the analysis under--estimating slab strength and

consequently under-estimating distribution properties, the analysis should


have under-estimated the deflection of Beam D.
estimated the deflection of that beam.

In fact it slightly over-

It appears that

the reason the

analysis predicted an earlier failure than actually occurred was that it


predicted that the concrete in the slab would start to crush due to the
global flange forces under a lower load than was the case.

In the tests,

there was no obvious sign of this crushing although it seems likely that
it

was

beginning

to

occur

when

the

bridge

failed.

The

happened at a lower deflection than in either the first

reason

this

bridge or the

single beam test was that the slab concrete was significantly weaker and
the analysis appears to have exaggerated the effect of this.

However,

although this global crushing was a major reason for the failure in the
analysis, the analysis still correctly predicted that

the final collapse

would look like a local failure; once again it showed that global and local
behaviours are not independent.
Since the predicted failure load, although 20% below the actual failure
load,

was

nearly

four

times

the

allowable

service

analysis its value had no practical significance.


predictions for

the behaviour at

load given by

the

The reasonably good

lower loads are more important.

The

analysis suggested that the slab as tested was inadequate for the intended
load and this confirms the findings from the tests suggesting that the
analysis would have provided a satisfactory design method.

- 231-

As

with

the

sufficiently

first
good

deck,

the

predict ions

analysis
of

considered

was

to

behaviour

to

suggest

have

that

it

given
was

reasonable to use it to obtain some insight into how that behaviour was
obtained.
In Figure 9 . 14. the predicted transverse force across the centre of the
slab span between Beams C and D is shown for a particular load stage in
the final loading.

In order to make the plot directly comparable with

Figure 9.10 <the equivalent plot for the first deck) the same load level is
used.

It will be seen that

the restraint

force in the region of the

wheels is much greater and much less localised than for the first deck.
The diaphragm at the right hand support, which is relatively close to a
wheel,

is

resisting

significant

tension

as

implied

by

previous

researchers but that at the opposite end of the deck is resisting very
little axial force.

The distribution of forces might be considered less

different from that implied by previous research than was that predicted
for the first deck.

However, the central portion of the deck between the

two bogies of the HB vehicle is still subjected to a significant transverse


tens i on, showing that compressive membrane action is not contributing to
the resistance to global transverse moments.

Wrea Pos;t;cns

Res tram 200


Force lkNiml
100

-100
Figure 9. 14:

Predicted force across deck slab of second deck

<load = 250kN/jack )
At first sight. the obvious reason for the greater compressive membrane
forces in this deck than in the first deck is that the diaphragms provided
better restraint

to these

forces.

However,

this

is not

a satisfactory

explanation; adding in-plane transverse stiffness to a region of the first


deck which which was subjected to significant compression could not have
this effect.

Also, Figure 9 .14 shows that only one of the diaphragms was

resisting

significant

tension.

There

are

two

other

explanations.

Firstly, the weaker concrete and less effective reinforcement in the second
- 232-

deck meant that it was more extensively cracked at this load stage and
consequently there was more membrane action.

Secondly, the diaphragms

reduced

in

the

difference

between

the

moments

adjacent

beams

and

consequently the global effect which led to tension at mid-span of the


first deck was less pronounced.
As with the first deck, an attempt was made to see if the membrane forces
predicted by the analysis were realised in practice.

However, the more

extensive cracking made this even more difficult.

All that could be

determined was that the form of Figure 9. a was reasonable.


Although

the

compressive

membrane

forces

shown

in

Figure

9.14

are

sufficient to cause a very significant enhancement in the behaviour they


are not, on their own, sufficient to explain all the difference between the
actual behaviour of the slab and that predicted by normal design methods.
The re-distribution of moments away from the critical region is equally
significant.
9.4 CONCLUSIONS

Conventional

analyses

of

the

models,

as

expected,

conservative predictions for the slab behaviour.


gross-concrete

slab

properties

are

used

the

give

extremely

Also as expected, if
predicted

distribution

properties are slightly better than were realised in practice.

However,

the discrepancies are relatively small and no greater than other faults of
conventional analysis which are normally considered acceptable.
The non-linear analyses gave reasonably good predictions for behaviour and
appear to give a reasonably good basis for design.
insight into the behaviour.
develop

compressive

They also give a good

They suggest that the restraint required to

membrane

action

comes

relatively close to the areas being restrained.

from

material

which

is

This explains why, as was

clear from the results of the tests on the first deck, membrane action is
not
\

dependent

on

the

presence of diaphragms.

It

also

confirms,

as

suggested in the last chapter, that membrane action does not contribute to
the resistance to global transverse moments.
The

analyses

also

confirm

that

the

failures observed were

primarily

brittle bending compression failures and that they were greatly influenced
by global behaviour.

They suggest that a large part of the difference


- 233-

between the real behavic.ur of bridge deck slabs and that predicted by
conventional elastic analysis

is actually due

rather than to pure membrane act ion.

to moment

redistribution

This redistribution, like membrane

action, is not dependent on reinforcement yielding; it starts to occur as


soon as the behaviour of the concrete becomes non-linear in tension which
is well before the slab becomes unserviceable in any way.

-234-

CHAPTER
USE OF

10

MEMBRANE

ACTION

IN
DESIGN AND

ASSESSMENT

10.1 OO'RODUCTION

Previous chapters have shown that membrane action, and the closely related
mechanism of moment redistribution, have a
the behaviour of bridge deck slabs.

s~nificant

beneficial effect on

They have also shown that the effect

is sufficiently reliable to justify its use in design and assessment.


chapter will consider the use of the effect in

des~n

This

and assessment.

Only the application to concrete bridges will be considered as steelconcrete composite bridges are considered to be outside the scope of this
thesis.
10.2 USE IN DESIGN
10.2.1 M Beall . Type Decks

Under present

des~n

slabs of otherwise

rules, the quantity of main reinforcement in the deck


identical bridges designed

for

identical loads

in

Northern Ireland and in the rest of Britain differ by a factor of over two.
This is clearly unsatisfactory and should be resolved.
It appears that non-linear analysis such as the form of analysis described
in Chapter 7, is needed to give a realistic prediction of the behaviour of
a deck slab under full HB load.

Although it is feasible to use this form

of analysis in design, it is probably not justified for such a routine,


simple and relatively standardised structure as the deck slab of an M beam
type bridge.

That standardisation enables simple prescriptive rules to be

developed.
Although Chapter 8 showed that deck slabs can fail at substantially lower
wheel loads than are predicted by the research on which the Northern Irish
rules are based, the rules are so conservative compared with that research
that they remain adequate.

Indeed they appear to be over-cautious.

The

first of the two decks tested in this study remained serviceable after the
deliberately excessively severe load history had been applied, despite
having only 60t of the steel area recommended by the Northern Irish rules.
- 235-

It also had more than adequate ultimate strength.

It might be argued,

considering the observed significance of global and local interaction, that


the tests were unrealistic because of the over-provision of prestress.
However, since the beams' behaviour remained linear elastic up to some 1.3
times design ultimate load, the behaviour under service loads would have
been virtually identical with substantially less prestress.
strength undoubtedly would have been lower.

The ultimate

However, the analysis in

9.3.2d suggested that even with less than the normal amount of prestress,
the bridge would have been over twice as strong as was required.

It thus

appears that Tl2-250 reinforcement is adequate compared with the Tl2-150


specified by the Northern Irish rules.

Nevertheless, and allowing for the

fact that analysis shows that global transverse moments could be greater
in a wider deck, it is prudent to continue to specify Tl2-150 main steel.
If this reinforcement is provided in M beam deck slabs there is no need to
do any analysis for the design of the slab.
Although this steel area can be justified from the test results alone,
there may be a preference for a design method which is based on some form
of analysis.

Such a method can be obtained by consideration of the tests

described in Chapter 8.

The first deck, whose. behaviour was considered

satisfactory, was provided with just enough transverse reinforcement to


resist the global transverse moments predicted by a conventional grillage
analysis.

The second deck, whose behaviour was less satisfactory, was

provided with substantially less steel.

A possible design approach is thus

to require that the reinforcement be designed for the global transverse


moments only; the opposite of the conventional North American approach.
This would give very light steel areas in some decks so a minimum nominal
area would also have to be specified.

Although one might put a case for

using the steel percentage specified by the Ontario Code, it is considered


prudent to specify a minimum of Tl2-250 which corresponds to the steel
area used in the first test deck and is the lightest steel area which has
been

demonstrated

requirement

to

be

satisfactory

by

tests.

In

practice,

the

to resist global moments means that the main steel would

normally be slightly heavier than this.


It is more difficult to justify the continued specification of the same

quantity of secondary steel.

It appears that the Ontario researchers had

two reasons for specifying isotropic reinforcement.

Firstly, their research

used an axi-symmetrical analysis and so they chose to use isotropic


-236-

reinforcement to get closer to the assumptions of the analysis.

Secondly,

and more significantly, they began by specifying main steel equal to the
minimum nominal steel required by their code so they could hardly have
specified less secondary steel.

When Kirkpatrick et al wisely specified a

larger area of main steel, to allow for global transverse moments, they
rather arbitrarily decided
Both

the

tests

reported

to continue to use isotropic reinforcement.


in

Chapter

and

the

analysis

reported

in

Chapter 9 suggest that the secondary steel in the deck slab of a simply
supported M beam deck is very lightly stressed and contributes little to
the behaviour.

It could be reduced to that provided in the first of the

models considered here, equivalent

to T12-250 at

approach is used for the design of the main steel.

full size, whichever


Even this is probably

over-conservative; there is no evidence from this study that any secondary


reinforcement is required.
The same basic approach to deck slab reinforcement design is valid in
regions of global longitudinal hogging.

This is

clear from previous

research and also because, as was discussed in 3.2.7, the critical load
cases for global longitudinal hogging do not impose any wheel loads in the
region of the slab which is in tension.

It is prudent, although probably

conservative,

longitudinal

to

additional to that

require

the

nominal

slab

steel

to

be

required for global moments and also to require a

proportion of the latter, say 30%, to be placed close to the bottom face of
the slab.

The reason for this restriction is that, although intended only

to resist

local effects, the nominal steel will be stressed by global

effects.

Thus the reserve strength available for local effects could be

very small if only this very small quantity of already highly stressed
steel was provided in the soffit.
The basic limitations imposed by Kirkpatrick et al on the use of the
empirical rules appear to be reasonable; one could debate the limiting span
given but since, with M beams, this is well above the limitation imposed by
web shear strength there is little to be gained by so doing.

The one

restrict ion which is worth reconsidering is the requirement for diaphragms.


The analyses and tests reported in this thesis show that diaphragms are
far less important to the development of compressive membrane action than
has previously been believed.

The empirical rules could be extended to

cover bridges with only nominal diaphragms, or with no diaphragms at all.


-237-

In the latter case the end section of the slab would not receive the full
benefit

of

restraint

and

would

require

extra

reinforcement.

is

It

suggested that a strip of slab extending 0.5m from the end of the deck
should be provided with enough reinforcement to enable it to support a
wheel acting as a single beam.

This is marginally more steel than was

provided in the first test deck considered in Chapter 8.


From a purely theoretical viewpoint, the use of these empirical design
rules in combination with global analysis based on gross-concrete section
properties cannot be justified.

However, analysis shows that the use of

cracked transformed transverse properties is conservative and in 9.2.2 it


was found that the errors resulting from the cracking are no greater than
other normally accepted faults of grillage analysis.
is to

use half

the

concrete properties.

transverse stiffnesses

A reasonable approach

calculated for

the gross-

The economic consequences of the slightly worse

distribution properties resulting from this compared with the conventional


approach are extremely small and significantly less than might be inferred
from the results of the tests considered in this thesis.

This is because

the tests considered HB alone, the worst case for the slab, whilst the
critical load case for the beams is HA plus HB.

Improving the distribution

properties reduces the effect of the HB load in the critical area but it
increases the effect of the associated HA.

Due to a continuing increase in

the HA load which is applied in combination with the HB load, the benefits
of good

distribution

standard

introduced

properties
in

Britain

have
since

reduced
the

with

1950s.

every new

loading

Nevertheless,

the

suggestion that reduced transverse properties should be used does imply


that the beams of bridges designed to the existing Northern Irish rules
could be subjected to slightly greater moments than those for which they
were designed.

In the author's view this is unimportant since the design

criteria currently used for this type of beam (class 1 and 2 criteria in
combination with an extremely severe service load) are unduly conservative.
However, a discussion of this subject is outside the scope of this thesis.
Where half the gross-concrete transverse properties are used in the global
analysis, it appears prudent to continue to require the transverse steel to
be capable of resisting the global transverse moments predicted by a
conventional analysis based on gross-concrete properties.

To avoid the

need to perform two separate analyses, these can be taken conservatively

-238-

to

be

double

the

moments

calculated

using

half

the

gross-concrete

properties.
10.2.2 other Beam and Slab Decks
The slabs of bridges built with U Beams or the proposed new Y Beams are
so similar to those of M beam decks that the same design rules can be
applied.

The only modification required being that, with U beams, the main

steel may have to be increased to enable it to act as part of the torsion


links of the beams.
Other types of beam and slab bridges normally have thicker deck slabs with
wider-spaced beams.

This means that the global transverse moments are

likely to be less significant but it is difficult to prove this.

It is

prudent, therefore, to recommend a check that the main steel in the deck
slab is always sufficient to resist the transverse moments given by the
global analysis.

The suggested minimum steel area to be specified is 0.3"

of the gross-section.

This corresponds to T12-250 <the minimum suggested

for M Beam slabs> for a thickness of 160mm so the rules are consistent.
These suggestions are more conservative than the Ontario rules but this is
justified due to the significance of global transverse moments noted in
this thesis and by the nature of the HB load which is exceptionally severe
for this effect.
The restriction on the use of these rules can be as for the Ontario rules
except for relaxing the requirement for diaphragms as with M Beam decks.
However, where these restrictions are not complied with it does not mean
that membrane action cannot be used in design; merely that the empirical
rules are not applicable.
could still be used.

Analysis such as that described in Chapter 7

Where the span to depth ratio is outside that

required to use the empirical rules the analysis should consider large
displacements.
10.2.3 other Types of Deck
It has been noted in earlier chapters that compressive membrane action is

potentially significant to other types of bridges, apart from beam on slab


structures.

These range from simple slab decks to major concrete box

girder structures.

The detailed consideration of these is considered

beyond the scope of this thesis and, in any case, they are probably not
- 239-

sufficiently standardised to enable prescriptive rules to be developed;


non-linear analysis would be required.

However,

a simple conservative

approach can be developed using normal analytical methods.

This approach

could also be used for beam and slab decks if desired.


The mechanism by which the behaviour of deck slabs is enhanced relative to
the predictions of elastic plate theory is essentially one of stresses
redistributing
Chapter

away

that

from

the

the

critical areas.

restraint

force

It

required

was demonstrated

to

develop

in

compressive

membrane force comes from material which is relatively close to these


critical areas; not

from the diaphragms.

This suggests a very simple

over-conservative way of allowing for the effect.

Design could be based

on a normal elastic slab analysis but ignoring, or rather smoothing out,


the peaks in the moment over a finite width.

If it

was only moment

redistribution which was being considered this width would be related to


the span and to the ductility of the sections.
however, the critical factor is the depth.
elastic analysis

could be

used

with

With arching action,

It is suggested, therefore, that

the

design

based

on

the

moment

averaged over a width equal to the lesser of 6d or half the slab span.
This is undoubtedly extremely conservative; it was demonstrated in 8.8.3
that removing the steel completely over a width of 12h had little effect
on behaviour.

10.3 .ASSESSMENTS

The approaches suggested in 10.2 are equally applicable to the assessment


of

existing

bridges.

However,

purely

empirical

approaches

are

less

suitable for assessment because it is not possible to adjust the structure


to

fit

the

limitations

imposed

for

the

rules.

It

will

therefore be

necessary to resort to non-linear analysis more frequently than in design.


The use of the assessment approach given in the Ontario Highway Bridge
Design Code< 11 >, which relies on the strength predictions of Hewitt and
Batchelor's approach,

is not normally advised.

This is because of its

failure to consider global transverse moments.

However, in assessing a

bridge which has intermediate diaphragms, it is reasonable to assume that


the global transverse moments in the deck slab are insignificant and so
the approach is more reliable.

Even then, if the spacing of the design

wheel loads is less than the slab span, some allowance should be made for
the effect of the second wheel.
- 240-

CHAPTER
CONCLUSIONS

11

RECOMMENDATIONS

AND

11.1 CONCLUSIONS
The first conclusion to be drawn from this study is that bridge deck slabs
are able to support loads by compressive membrane action and, as a result,
that they are able to support very much greater loads than is suggested
by conventional design methods which are based on flexural theory.

Judged

against the background of the research which was reviewed in Chapter 3,


this conclusion is unremarkable.

However, the conclusions to be drawn from

any study depend as much on the the background against which the study is
assessed

as

on

the

study

itself.

Judged

against

the

background

of

conventional design practice, which was reviewed in Chapter 2, the enormous


strengths

of

deck

slabs,

particularly

lightly

reinforced

deck

slabs,

compared with the predictions of conventional flexural theory remains the


most significant

conclusion.

is re-stated here to put some of the

It

other conclusions into perspective; it should be remembered, for example,


that when the deck slab of the first model considered in Chapter 8 failed
at little over half the load which might have been expected from some
previous research,

it

was

resisting some

five

times its ultimate load

according to normal design methods.


The remaining conclusions are:
1.

Compressive
moment

membrane

action

and

the

closely

allied

mechanism

of

redistribution start to enhance the behaviour of deck slabs

relative

to

the

predictions

of

linear

analysis

as

concrete's behaviour becomes non-linear in tension.


thin slabs,

soon

as

the

This, at least in

is well before there are visible cracks.

It

does not

depend on any material behaviour which is unacceptable under service


loads.

Because of this, membrane action significantly increases the

service load, as well as the ultimate load, which a slab can carry.
2.

Compressive membrane act ion is sufficiently reliable to justify its


consideration in design and assessment.

The model tests described in

Chapter 8 were an .exceptionally severe test yet


substantially

better

than

could

be

analysis.

- 241 -

anticipated

the behaviour was


by

purely

flexural

3.

The restraint required to develop compressive membrane action comes


from

the under-stressed reinforcement and concrete surrounding the

critical areas of the slab.

It is not dependent on the presence of

diaphragms.
4.

Compressive

membrane

action

could

even

enhance

behaviour of slabs with no external restraint.

the

service

However,

load

it cannot

increase the failure load of such slabs above that predicted by yieldline theory.
5.

Compressive membrane action does not greatly enhance the resistance


to global transverse moments.

6.

Because of 3 and 5 above, and contrary to the implications of some


earlier

research,

reinforcement

is

needed

in

bridge

deck

slabs.

However, because it is required to resist global transverse moments


and to provide restraint

<rather than to resist local moments), the

behaviour is not sensitive to the exact position of the reinforcement.


Thus the behaviour of bridge deck slabs is remarkably insensitive to
local reinforcement corrosion.
7.

The failure loads of bridge deck slabs subjected to single wheel loads
are reasonably well predicted by the approaches which were considered
in 3.2.3.

The cases where these approaches gave unsafe predictions

were restricted to impractically lightly reinforced slabs with large


span to depth ratios and relatively poor restraint.

The methods do

not, however, give good predictions of other aspects of behaviour; for


example, Hewitt's approach under-estimated the deflection at failure by
a factor of up to 10.
8.

Non-linear analyses of the forms considered in Chapters 5 and 7 are


also capable of predicting these failure loads and are better able to
predict other aspects of behaviour.

The form considered in Chapter 5

is theoretically more rigorous and realistic than that considered in


Chapter 7 but the latter has many practical advantages in a design
situation;

it

is simpler, more compatible with design standards and

also appears to be more consistently safe.


9.

The local failures observed in deck slabs are primarily brittle bending
compression failures.

They can be predicted by analyses which do not

consider shear, the load at which they occur can be reduced by the

- 242-

presence of other moments <such as global transverse moments) and, in


many cases, crushing concrete is visible before failure.
10. Bridges which are subjected to multiple wheel loads, such as HB, can
fail by wheels punching through their slabs at wheel loads which are
substantially

below

the

local

strength

of

their

slabs;

both

as

measured in single wheel tests and as predicted by the approaches


developed_by previous research.
11. The form of failure considered in 10 above can .occur even when the
beams have a reserve of strength and the global transverse moments
are thus not needed to maintain equilibrium.

This is contrary to the

safe theorem of plastic design but the behaviour is too brittle for
this to apply.
12. Non-linear analysis is capable of predicting the behaviour of bridge
decks reasonably well.

In particular, it appears to be the only form

of analysis which is capable of modeliing the interact ion of global


and local effects and of predicting the restraint.

11.2~RECO~ATIONS

11.2.1 Recommendations for Design and Assessment

Less conservative 'design methods for bridge deck slab reinforcement should
be

introduced

which

membrane action.

allow

for

the

beneficial

effects

of

compressive

Possible details of these methods were considered in

Chapter 10 and will not be discussed here.

11.2.1 Recommendations for Further Research

There are many aspects. of the behaviour of the type of slabs considered in
detail in this and previous studies which could be considered to require
further research.

However, such research is not needed to justify the use

of membrane action in design or assessment.

To recommend it would merely

serve to perpetuate the use of conventional design methods which have


been shown to be extremely unrealistic and conservative.
This study has suggested that membrane action could have a significant
beneficial effect on the behaviour of a wide range of bridge deck slabs in
addition to those for which it has so far been investigated in detail.

Any

future studies of membrane action in bridge deck slabs should consider


types. of slab which have not previously been researched.
- 243-

This includes

the thin long-span slabs typical of longer span concrete bridges.

However,

this study has shown that the restraint required to develop compressive
membrane

action

comes

from

critical areas of the slab.

under-stressed

material

of

simply

the

At service load levels, which are critical in

design, it is not dependent on any external restraint.


behaviour

surrounding

supported

and

even

It follows that the

cantilever

slabs

could

significantly enhanced by the effect and this should be investigated.

- 24.4-

be

REFERENCES
1.

Moments and stresses in slabs.

WESTERGAARD, H.M. and SLATER, W.A.

Proceedings, American Concrete Institute.


2.

TURNER,

Advance

C.A.P.

in

Vol. 17. 1921.

reinforced

concrete

pp. 415-538.

construction:

argument for multiple-way reinforcement in floor slabs.

News.
3.

Vol. 61, No. 7.

Feb. 18, 1909.

SOZEN, M.A. and SEISS, C.P.


concrete

floor

Engineering

pp. 178-181.

Investigation of multiple-panel reinforced


Design

slabs.

their

methods

evolution

Proceedings, American Concrete Institute.

comparison.

an

Vol. 60.

.and
1963.

pp. 999-1020.
4.

NICHOLS.

Statical

J.R.

limitations

5.

6.

LORD,

A.R.

Vol. 77.

1914.

A test of a

flat

the steel

slab floor

in a reinforced concrete

National Association of Cement Users

Institute).

Vol. 7.

A proposal

BEEBY, A.W.

slabs.

in

pp. 1670-1736.

building.

1910.

requirement

Transactions, American Society of

rein forced concrete floor slabs.

Civil Engineers.

upon

<American Concrete

pp. 156-179.

for: change in the basis of the design of

Wexham Springs, C & CA, April 1982.

p. 28.

Technical Report

42.547.
7.

STATE CONSTRUCTION COMMITTEE OF THE USSR.


regulations

& R II-21-75.

BS

The

Building standards and

structural use

of concrete.

Moscow, 1976.
8.

BRAESTRUP,

Dome effect

M.W.

Journal of the Structural Division, American Society

plastic analysis.

of Civil Engineers.
9.

OCKLESTON,

A.J.

Vol. 106, No. ST6.

Load tests on a

10.

GUYON,

Y.

June 1980.

Vol. 33, No. 10.

pp. 304-322.
Tests

Prestressed Concrete.

on

continuous

Vol.

slabs.

. Method

of

analysis.

2, Statically Indeterminate Structures.

English translation by C. Von Amerongen.


1960.

pp. 1237-1253.

three-storey reinforced concrete

The Structural Engineer.

building in Johannesburg.
October 1955.

in re1nforced concrete slabs: rigid-

pp. 515-567.

-245-

London, C. R. Books

Ltd.

11.

ONTARIO

MINISTRY

OF

TRANSPORTATION

AND

COMMUNICATIONS.

Downsview, Ontario, Canada,

Highway Bridge Design Code.

Ontario

1st Edition

1979, 2nd Edition 1983.


12.

DEPARTMENT OF THE ENVIRONMENT FOR NORTHERN IRELAND.


beam

bridge decks.

Amendment

Road Servic:es Headquarters.


13.

September 1984.
14.

J.,

KIRKPATRICK,

PRITCHARD,

B.P.

bridges

the

17.

LONG,
of

A.E.,

spaced

GEE, A.F.

and

THOMPSON,

M-beam

bridge

Dec. 1984.

Combatting
way

Transportation

April !986.

Strength evaluation of MVol. 628, No. 3.

The Structural Engineer.

Vol. 628, No. 4.

Engineer.

16.

March 1986.

pp. 60-68.

characteristics

15.

3 to B.D.C, Belfast, Northern Ireland

KIRKPATRICK, J. RANK IN, G. I. B. and LONG, A.E.


beam bridge deck slabs.

Design of M

road

ahead.

National

Load

decks.

distribution

The

salt

corrosion

on

in

Institution

U.K.

pp. 119-159.
Bridge winners and losers.

BEEBY, A.W.

The function of research in the construction industry.

pp. 141-145.

FREYSSINET,

E.

Journal

the

of

EMERSON,

M.

A.

Institution

concrete
of

Civil

principles

and

applications.

Engineers.

Vol.

33,

No.

4.

in

basis of design

pp. 331-380.

Crowthorne,

Prestress
April 1983.

DEPARTMENT
BE5/73.

22.

Berlin, Ernst

differences

Transport

and

bridges:

Road

Research

Laboratory,

Laboratory report 765.

Journal.

21.

Prestressed

Temperature

requirements.

LOW,

1984.

February 1950.

20.

Vol.

The Structural Engineer.

April 1987.

1977.

&

Bridge Maintenance Initiatives.

65A, No. 4.

& Sohn.

19.

concrete

of Highways

testimonial volume for Professor Rehm's 60th birthday.

18.

Structural

pp. 86-88.

Proceedings,

Workshop

A.

OF

THE

design

for

continuous

concrete

members.

Arup

pp. 18-21.
ENVIRONMENT.

Standard Highway Loadings.

DEPARTMENT OF TRANSPORT.

Technical

Memorandum

London, August 1973.

Departmental Standard BD17/83.

Concrete Bridges. Use of BS 5400: Part 4: 1978.

-246-

<Bridges)

Design of

London, 1983.

23.

BRITISH

STANDARDS

BS

INSTITUTION.

54-00:

Part

BRITISH STANDARDS INSTITUTION.

BS 5400: Part 2: 1978.

DEPARTMENT OF TRANSPORT.

BRITISH

STANDARDS

INSTITUTION.

BS

54-00:

practice for the design of concrete bridges.

27.

28.

DEPARTMENT OF TRANSPORT.

Part

4:

1984.

Code

Departmental Standard BD24/84.

Design of

Memorandum

OF

THE

ENVIRONMENT.

Technical

OF

THE

ENVIRONMENT.

Technical

Memorandum

The new concrete code -

use of concrete: Part

is it all it's cracked up to be?

Vol. 64A, No. 8.

BRITISH STANDARDS INSTITUTION.

<Bridges)

London, 1973.

Prestressed concrete for highway structures.

BEAL, A.N.

<Bridges)

London, 1973.

Reinforced concrete for highway structures.

DEPARTMENT

of

London, 1984.

DEPARTMENT

The Structural Engineer.

31.

London, October 1982.

London, 1984-.

B2173.
30.

Loads for

Concrete Bridges. Use of BS 5400: Part 4: 1978.

BE 1/73.
29.

Specification

Departmental Standard BD14/82.

Highway Bridges. Use of BS 5400: Part 2: 1978.

26.

of

London, 1978.

for Loads.

25.

Code

1978.

London, 1978.

practice for the design of concrete bridges.

24.

4-:

August 1986.

CP110: Part 1: 1972.

pp. 199-201.
The structural

London,

1, Design materials and workmanship.

1972.
32.

BRITISH STANDARDS INSTITUTION.


of concrete: Part

1,

BS 8110: Part 1: 1985.

Code of practice

for design

Structural use

and construction.

London, 1985.
33.

BEEBY,

A.W.

Cracking

BEEBY, A.W.

JACKSON, P.A.

Vol. 57 A,

EDWARDS,

in

the

oceans.

K.R.

No. 1.

The

January 1979. pp. 9-17 ..

The stress limits for reinforced concrete in BS 54-00.

The Structural Engineer.

36.

Concrete

The prediction of crack widths in hardened concrete.

Structural Engineer.

35.

corrosion.

Wexham Springs, CIRIA, C & CA, DoE, 1978.

Technical Report 1.
34-.

and

Vol. 65A, No. 7.

Non-linear

analysis

reinforced concrete bridge decks.

July 1987.
of

pp. 259-263.

eccentrically

stiffened

PhD Thesis, University of Liverpool.

1983.
37.

COPE,

R.J.,

structures.

RAO,

P.V.,

and CLARK, L.A.

Non-linear design of. concrete

CSCE-ASCE-ACI-CEB International Symposium.

of Waterloo, Ontario, Canada.

1979.
- 247-

University

38.

CLARK,

LA., Elastic analysis of concrete bridges and

limit state.
39.

WESTERGAARD, H.M.
loads.

40.

pp. 22-24.

Computation of stress in bridge slabs due to wheel


Vol. 2, No. 1.

Public Roads.

PUCHER, A.

October 1977.

The Highway Engineer.

the ultimate

March 1930.

pp. 1-23.
Wien and New York,

Influence surfaces of elastic plates.

Springer Verlag, 1964.


41.

MORICE, P.B. and LITTLE, G.

London, C & CA, July 1956.

abnormal loading.

42.

PENNELLS, E.

Concrete

Publications, 1978.
43.

WEST, R.

Analysis of right bridge decks subject to

Publication 32-002.
London, Viewpoint

bridge designer's manual.

p. 111.

The use of grillage analogy for the analysis of slab and

London, 1973.

pseudo-slab bridge decks.

C & CA Research Report 21,

Publication 4-1.021.
44.

HAMBLY, E.C.

London, Chapman & Hall, 1976.

Bridge deck behaviour.

pp. 83-85.
4-5.

AMERICAN ASSOCIATION OF STATE HIGHWAY AND TRANSPORTATION OFFICIALS.


Standard

specification

for

highway

Washington

bridges.

DC,

13th

Edit ion, 1983.


46.

CONCRETE SOCIETY.
highway bridges.

47.

Design

London, 1986.

DORTON, R.A., HOLOWKA, M. and KING, J.P.C.


design and testing.
1977.

48.

The Conestogo River Bridge

Canadian Journal of Civil Engineering.

Vol. 4.

pp. 18-39.

BUCKLE, I.G., DICKSON, A.R. and PHILIPS, M.H.


reinforced

concrete

KIRKPATRICK,

highway

Vol. 12.

Engineeering.

49.

and cost studies of lightweight concrete

J.,

bridges.

1985.

RANKIN,

Ultimate strength of three


Canadian

Journal

of

Civil

pp. 63-72.

G.I.B.

and

LONG,

A.E.

The

influence

of

compressive membrane action on the serviceability of beam .and s_lab


bridge decks.

The Structural Engineer.

Vol. 64B, No. 1.

March 1986.

pp. 6-12.
50.

HOPKINS, D.C. and PARK, R.


floor

designed

deflection

and

with

Tests on reinforced concrete slab and beam

allowance

ultimate

for

membrane

loads_ of concrete

American Concrete Institute, 1971.


- 248-

slab

action.

Cracking

systems.

Detroit,

Publication SP 30.- pp. 223-250.-

51.

TONG, P.Y. and BATCHELOR, B. de V.


in two-way bridge slabs.

Cracking deflection and ultimate loads of

concrete slab systems.

Publication SP 30.
52.

BEAL, D.B.

Detroit, American Concrete Institute,

Load capacity of concrete bridge decks.

No. ST4.

April 1982.

ROBERTS,

E.H.

54.

Journal of the

American Society of Civil Engineers.

Vol.

108,

pp. 814-832.

Load-carrying

capacity

against longitudinal expansion.


1969.

1971.

pp. 271-286.

Structural Division,

53.

Compressive membrane enhancement

strips

restrained

Vol. 3, No. 9.

September

of

Concrete.

slab

pp. 369-378.

PARK, R.

Ultimate strength of rectangular concrete slabs under short

term uniform loading with edges restrained against lateral movement.


Vol. 28.

Proceedings, Institution of Civil Engineers.

June 1964.

pp.

125-150.
55.

TAYLOR, R. and HAVES, B.

Some tests on the effect of edge restraint

on punching shear in reinforced concrete slabs.


Vol. 17, No. 50.

Research.

56.

BROTCHIE, J.F. and HOLLEY, M.J.


deflection

57.

March 1965.

and

ultimate

Magazine of Concrete

pp. 39-40.

Membrane action in slabs.

loads

of concrete

slab

Detroit,

systems.

345~378.

American Concrete Institute, 1971.

Publication SP 30.

JOHANSEN, K.W.

London, C & CA, 1962. <Translation

Yield-Line Theory.

pp.

Cracking

from Danish first published in 1943 ).


58.

WOOD,

R.H.

Plastic

particular

reference

and

elastic

to

reinforced

Thames and Hudson, 1961 .


59.

MORLEY,

C.T.

Yield

line

60.

PARK,

December 1967.
R.

and GAMBLE,

concretete slabs.

61.

MASSONNET, C.H.

of slabs

concrete

and

floor

plates

with

London,

slabs.

p. 344.

moderately large deflexions.


No. 61.

design

theory

for

reinforced

concrete slabs at

Magazine of Concrete Research.

Vol. 19,

pp. 211-222.
W.L.

Membrane action

in

New York, J. Wiley & Sons, 1980.

slabs.

Reinforced

pp. 562-612

General theory of elastic-plastic membrane plates.

Engineering Plasticity.

Cambridge, Cambridge University Press, 1968.

Editors Heyman and Leckie.

pp. 443-471.

- 249-

62.

MOY,

S.S.J.

and

MAYFIELD,

B.

reinforced concrete slabs.


No. 81.
63.

December 1972.

CHRISTIANSEN, K.P.

Load

deflection

characteristics

Vol. 24,

Magazine of Concrete Research.

pp. 209-218.

The effect of membrane stress on the ultimate

strength of the interior panel in a reinforced concrete slab.


Vol. 41, No. 8.

Structural Engineer.

64.

RANKIN, G.I.B.
in

slabs.

No. ST2.

SKATES,

PhD

A.,

membrane

pp. 261-265

Arching action theory in

March 1956.

pp. 915/1-18.

Punching failure and compressive membrane enhancement

Thesis,

University of Belfast.
66.

The

Journal of the Structural Division, American Society

of Civil Engineers.

65.

August 1963.

Me DOWELL, E.L., Me KEE K.E. and SEVIN, E.


masonry walls.

of

RANKIN,
action

Department

of

Civil

Engineering,

Queens

1982.

G.I.B.

in

and LONG,

the

design

A.E.

of

Utilising the effects of

offshore

cellular

structures.

Proceedings, Marine Concrete '86 International Conference on Concrete


in

1986.
67.

London.

the Marine Environment.

Concrete Society,

pp. 79-90.

WHITNEY, C.S.

Design of reinforced concrete members under flexure or

combined flexure and direct compression.


Concrete Institute.

68.

subjected

MOE, J.

to concentrated

load.

Publication SP 30.

70.

deflection

and

ultimate

Detroit, American Concrete Institute,

Skokie, Portland Cement Association, April

Development Department Bulletin D47.


The strength of two-way slabs with fixed edges subjected

concentrated loads.

Ontario, Canada.
71.

Cracking

Shearing strength of reinforced concrete slabs and footings

YOUNG, D.M.
to

pp. 483-498.

pp. 103-126.

under concentrated loads.

1961.

of the American

Shearing strength and cracking in two-way slabs

loads of concrete slab systems.

69.

Journal

Vol. 33, No. 2. March/April 1937.

AOKI, Y. and SEKI, H.

1971.

September

MSc Thesis, Queens University of Kingston,

1965.

BATCHELOR, B. de V. and TISSINGTON, I.R.

Shear strength of two way

bridge slabs Journal of the Structural Division, American Society of


Civil Engineers.

72.

HEWITT, B.E.
Restrained

Vol. 102, No. ST12.

and BATCHELOR,
Slabs.

Journal

December 1976.

pp. 2315-2331.

de V.

Punching Shear Strength of

of . the

Structural Division, . American

B.

-250-

Vol. 101, No. 6T9.

Society of Civil Engineers.

Septembet

1975.

pp.

1837-1852.
73.

KINNUNEN, S.
shear

7 4.

and NYLANDER, H.

reinforcement.

Transactions

Journal of the American Concrete Institute.

February 1975.

CHANA, P.S.

October 1986.

Bridge

Engineering,

Performance

and

Development

BATCHELOR,

July 1981.

B.

de

Investigation

Bureau,

V.,

of the

steel/concrete

New

B.E.,

State

CSAGOLY,

ultimate strength

PETCU,

V.

and STANCULESCU, G.

P.

and

Critical steel ratios

ultimate loads of concrete slab systems.

Publication SP 30.

NIBLOCK, R.A. and LONG, A.E.

HOLOWKA,

M.

September

in the limit

Cracking deflection and

Detroit, American Concrete

pp. 287-300.

A Design Method for Uniformly Loaded

Laterally Restrained Concrete Slabs.

Proceedings,

6th International

Offshore Mechanics and Arctic Engineering Symposium.

HOLOWKA,

of

pp. 162-179.

design of two-way reinforced concrete slabs.

81.

Department

Bridge Engineering Conference.

Transportation Research Record 664.

1987.

Engineering

of deck slabs of composite

1978.

Institute, 1971.

pp. 189-210.

New York.

York

Advances ..

Research Report 89.

HEWITT,

bridges.

and

Editor Cope, R. J.

Strength of Concrete Bridge Decks.

Transportation.

13.

PhD Thesis,

Membrane enhancement in top slabs of concrete

Concrete

BEAL, A.N.

Vol.

pp. 37-45.

Shear failure in reinforced concrete beams.

BA TCHELOR, B. de V.

Research

80.

of

A two-phase approach to the prediction of the punching

London and New York, Elsevier, 1987.

79.

Institute

LONG, A.E.

bridges.

78.

Royal

Stockholm, 1960.

University College London.

77.

the

No. 158.

72, No. 2.

76.

of

Technology.

strength of slabs.

75.

Punching of concrete slabs without

Houston.

Vol.

pp; 177-184.
M.

and CSAGOLY,

P.

concrete AASHTO girder bridge.

Testing

of a

composite prestressed

Downsview, Ontario, Canada, Ontado

Ministry of Transportation and Communications, 1980.

Research Report

222.
82.

CAIRNS, J.
highway

Measurement
bridge.

Wexham

or service strains in

Springs,

Research Seminar.
- 251 -

C & CA,

the deck slab or a

1986.

pp.

127-133.

83.

PARK,

R.

The

lateral

stiffness

and

strength

required

to ensure

membrane action at the ultimate load of a reinforced concrete slaband-beam floor system.
50.

84.

March 1965.

Magazine of Concrete Research.

fatigue

of

strength

slabs

deck

Engineering

Bridge

bridges.

CRANSTON,

W.B.

Developments

of

An investigation of

composite

Conference.

Transportation Research Record 664.


85.

17, No.

pp. 29-38.

BATCHELOR, B. de V., HEWITT, B.E. and CSAGOLY, P.


the

Vol.

in

steel/concrete

September

1978.

pp. 153-161.
concrete

bridges.

Proceedings,

Institution of Highway and Transportation National Workshop on Bridge

86.

Construction and Maintenance.

April, 1984.

NGO,

Finite element analysis of reinforced

D.

and SCORDELIS,

concrete beams.
No. 3.
87.

March 1967.

ABDEL RAHMEN, H.H.

MINDLIN,

R.D.

pp. 152-163.
Computational models for the non-linear analysis

PhD Thesis, University College, Swansea.


Influence

of

rotary

89.

COPE,

1951.

R.J.

1982.

inertia and shear

motions of isotropic elastic plates.


Vol. 18.

Vol. 64,

Journal of the American Concrete Institute

of slab systems.

88.

A.C.

pp. 40-51.

Journal

on

flexural

of Applied Mechanics.

pp. 31-38.

and

CLARK,

L.A.

slabs,

Concrete

analysis

and

design.

London, Elsevier, 1984.


90.

BUCKLE, I.G. and JACKSON, A.T.


analysis

of

Proceedings,

reinforced

91.

concrete

International

Shanghai, 1982.
POPOVICS, S.

A filament element for the non-linear


beam

and

on

Finite

Conference

slab

structures.

Element

Methods.

pp. 4-15.
A review of stress-strain relation ships for concrete.

Proceedings, American Concrete Institute.

Vol. 67, No. 3.

1970.

pp.

243-248.

92.

NILSON,

A.H.,

TASUJI,

M.E.

relationship for concrete.


No. 109.
93.

1969.

and

SLATE,

F.O.

Vol. 31,

pp. 217-224.

behaviour

for

finite

element

Numerical Methods for Non-Linear Problems.

1980.

stress-strain

Magazine of Concrete Research.

COPE, R.J., RAO, P.U., CLARK, L.A. and NORRIS, P.


concrete

Biaxial

pp. 457-470.
- 252-

Modelling of reinforced

analysis - of

bridge

slabs.

Swansea, Pineridge_ Press.

94.

Design and assessment of one way prestressed

COPE, M. and COPE. R.J.

Vol.

bridge slabs.
Contractor's

2.

report

Analytical investigations and conclusions.


Crowthorne,

74,

Transport

and

Road

Research

Laboratory, 1988.
95.

CRISFIELD,

M.A.

reinforced

Local

concrete

Civil Engineers.
96.

COPE,

R.J.

instabilities

beams and

Part 2.

in

Proceedings,

slabs.

March, 1982.

Non-linear

the non-linear analysis of

analysis

Institution of

pp. 135-145.

of

reinforced

concrete

Computational models of reinforced concrete structures.


Pineridge Press, 1986.
97.

Editors Hinton, E. and Owen, R.

BAZANT, Z.P., PAN, J. and PIJAUDIER-CABOT, G.


concrete

beams

and

Journal

frames.

American Society of Civil Engineers.

slabs.
Swansea,

pp. 3-43

Softening in reinforced

of Structural

Vol. 113, No. 12.

Engineering,
Dec. 1987. pp.

2333-2364
98.

Finite Element Analysis of Compression Membrane Action

JACKSON, A.T.

in Slabs.

University of Auckland.

1979.

School of Engineering

Research Report 207.


99.

COPE.

R.J.

and EDWARDS K.R.

Non-Linear Finite Element Analysis of

Eccentrically Stiffened Bridge Decks.

conference on

finite elements in computational mechanics.

Pergamon Press.
100. BEDARD,

C.

1985.

and

structures.

M.D.

Vol. 111, No. 12.

and

concrete slabs.

of

NLFEA

to

concrete

WARNER,

Dec. 1985.

R.F.

Tension

pp. 2691-2707.

stiffening

in

reinforced

Journal of the Structural Division, American Society

of Civil Engineers.

Vol. 104, No. ST12.

Dec. 1978.

Stress-strain characteristics of concrete in tension.

PhD Thesis, University of Leeds.


103. VETTER, C.P.

Application

Journal of the Structural Division, American Society of

R.I.

102. MARATHE, M.

Oxford,

pp. 431-448

KOTSOVOS,

Civil Engineers.
101. GILBERT,

Proceedings of international

1967.

Stresses in reinforced concrete due to volume changes.

Transactions, American Society of Civil Engineers.

No. 98.

1933.

pp.

1039-1080.
104. BRITISH STANDARDS INSTITUTION.

BS 8110:

19B5.

Structural use of

concrete: Part 2, Code of practice for special circumstances.


1985.
-253-

London,

105. HARTL,

G..

Die arbeitslinie

kurzzeitbelastung
first

eingebetteter st!ihle

<Force-Deformation

diagram

time and short term loading).

No. 8.

1983.

106. CERVENKA,

unter

"embeded

erst-und

bars"

under

Beton-und Stahlbetonbau.

78.

pp. 221-224.
Constitutive

V.

model

for

cracked

Journal of the American Concrete Institute.

reinforced

concrete.

Vol. 82 No.6.

Nov./Dec.

pp. 877-882.

1985.

107. COPE, R.J. and RAO, P.V.

Proceedings, Institution of Civil Engineers, Part 2.

of slab bridges.
Vol. 75 No. 6.

A two-stage procedure for non-linear analysis

December 1983. pp. 671-688.

AASHO interim guide

108. American Association of State Highway Officials.

for the design of rigid pavement structures.

Washington D.C., 1962.

Tests on large reinforced concrete elements subjected

109. WILLIAMS, A.

to direct tension.

Wexham Springs, C & CA, April 1986.

Technical

Report 562.

Tension stiffening in reinforced concrete

110. CLARK, L.A. and SPIERS, D.M.

beams and slabs under short-term load.


1978.

Wexham Springs, C & CA, July

Technical Report 521.

111. CLARK,

L.A.

AND CRANSTON,

W.B.

The

influence

of

bar

spacing

on

tension stiffening in reinforced concrete slabs, Advances in concrete

slab technology.

Oxford, Pergamon Press, 1979, Editors R. K. Dhir and

J.G.L. Munday, pp. 118-128.

The assessment of Mound Bridge for Transport

112. GIFFORD AND PARTNERS.

and

Road

Research

November 1988.
113. BRITISH

Laboratory.

Report

GP

3988.02,

Southampton,

<written by P.A. Jackson>.

STANDARDS

INSTITUTION.

BS

4975:

1973.

Specification

prestressed concrete pressure vessels for nuclear reactors.

for

London,

1973.
114. GANABA, T.H. and MAY, I.M.

An improved numerical integration over the

thickness of plates and shells in finite element analysis.


of Warwick.

1984.

115. CRISFIELD, M.A.


"snap through"

University

Report No. 8.

A fast

incremental/iterative procedure that handles

Computers and Structures.

116. DUDDECK, H. GRIENOW, G. and SCHAPER, G.

No. 13.

1981.

Material and time dependent

nonlinear behaviour of cracked reinforced concrete slabs.


-254-

pp. 55-62.

Nonlinear

behaviour

of

reinforced

concrete

Preliminary Report !ASS Symposium.


1978.
11 7.

Vol.

structures.

DUsseldorf,

Werner-Verlag,

Editors G. Mehlorn, H. RUhle and W. Zerna.

MUELLER, G.

July

pp. 101-113.

Department of Civil Engineering, University of California,

September 1977.

Report No. UC SESM, 77-5.

118. RAEGEN, P.E. and REZAI-JORABI, H.


concentrated

119. HUGHES, G.

loads.

Shear resistance of one way slabs

Structural

March/April 1988.

Concrete Institute.

Longitudinal shear in

Journal

of

the

52.

Crowthornne,

American

pp. 150-157.

composite concrete

bridge

Transport

and

Road

beams.

Contractors

Part 2; Experimental investigation and recommendations.

report

1,

Numerical Problems in Nonlinear Analysis of Reinforced

Concrete.

under

spatial

Laboratory,

Research

1987.
120. GREEN, J.K.
London,
121. CLARK,

Detailing for standard prestressed concrete bridge beams.

c & CA,
L.A.

December 1973.

and

WEST,

Publication 4.6.0 18.

R.

The

torsional

diaphragms in beam and slab bridges.

stiffness

of

support

Wexham Springs, C & CA, August

1975. Technical Report 4.2.510.


122. BEAL,

A.N.

Reinforcement

Engineering

Research

for

and

Bridge

Development

Department of Transportation.
123. JACKSON, P.A.

Concrete

Bureau,

July 1983.

New

Decks.

New

York

York.
State

Research Report 105.

Some aspects of the assessment of concrete bridges.

Highways. Vol. 56 No. 1936.

124.. P.S.C. FREYSSINET,

April 1988.

pp. 32-33

Elastomeric Bearings.

Iver, P.S.C. Freyssinet, 1982

and 1983.
125. BRITISH STANDARDS INSTITUTION.
curing test specimens.

BS 54.00: Part

128. PERDIKARIS, P.C. and BEIM, S.


moving

1980.

Code

of

Background and synopsis of proposed new

loading for highway bridges in the UK.

and

I 0:

London, 1980.

127. DEPARTMENT OF TRANSPORT.

pulsating

Methods for

London, 1983.

126. BRITISH STANDARDS INSTITUTION.


practice for fafigue.

BS 1881; Part Ill; 1983.

load.

London, 1983.

Reinforced concrete bridge decks under


Journal

American Society of Civil Engineers.

pp. 591-607.
- 255-

of

Vol.

the

Structural

114-, No. 3.

Division,

March 1988.

129. COOK, C.F.

An electrical demountable strain transducer.


Society

Conference,

British

Polytechnic.

September 1979.

130. MEHKAR-ASL,

s.

of

Strain

Measurement.

15th Annual

Bristol

Direct measurement of stress in concrete structures.

PhD Thesis, University of Surrey, September 1988.

- 256-

APPENf)ICES
A. RESTRAINED SLAB STRIP TO ELASTIC THEORY

Al Stresses
112
~----------------------------------~p
u
"0

X:z

Figure: Al:

Restrained slab strip under line load

<half section: elastic theory>


Consider the slab strip shown in Figure A.l.

For convenience the origin of

the x axis is located at the intersection of the extended line of thrust


of the restraint force and the projection of the soffit of the slab.
Since the support is fully fixed in rotation, and the section at mid-span
cannot rotate either because this would violate the symmetry of the
system;
=

x,

and

1/2

Now, from the geometry of the line of thrust:

Therefore
Now, from x

=0

to x

P/2

=x

the depth of concrete in compression, de,

=
=
since

hx/x2
3hx/x4

=
A 1

Now the stress on the compression face of the section, f.,.,,


=

2F/d.,

2F
3hx/x4

Substituting for F, this gives:


=

f.,.,

Now, the strain at mid-depth of the slab

Note: For a wide slab, the Young's modulus, E.,, of the concrete should
strictly be replaced by E.,/ <1-v 2

since the slab is forced to bend

cylindrically because the transverse strains, which occur in a narrow


slab due to the Poisson's ratio effect, are prevented.
Substituting for f.,c, d., and F, leads to:
Px 4 2
3E., h2 x
=

Px 4 2
3E.,h 2 x

=x

/3 to x

at mid-depth.

= 2x

h
2<3hx/x..,)

/3 the section is uncracked so its centroid is

Hence the stress there is:

=
so

[1x - 6xx... J

3E.,h 2

From x

p
p- ~;]

E.,

F/h

Now the slab's extension from the centre-line to the support;

x./2

J E.,

-2

-2Px 4
E.,h 2

dx

x,

[f"

f" ]

x...

! - x... dx + 1

6x 2

x,

x./3

x./3

-2Px 4
E h2
"
A

x... [lnx + x...


3
6x

x,
2

x/2

[~]
x./3

~2]

=
=

-2Px42 [ln x4 - ln3 + 3 - x4


~
x,
4 6x,

-2Px 42 [ln<x 4 /x 1 )
3E.,h 2

0. 3486 - <x4/6x,

If the slab is rigidly restrained this must be equal to zero.

and numerical solution of this equation leads to;


=

13.54

so, at the support and at mid-span;


de

= ...L!L
13.54

now;

=
=
=
=

0.222h
1/2

2 [h-(2/3)d.,]
.p

1/2

h[l- (2/3)

Pl/3. 41h

and the maximum concrete stress


=

2F/d.,

2 Pl
3. 41x0. 222h 2

2. 64P l/h 2

A2 Deflection

From x

=0

to x

= x~/3
=

and;

now, the curvature

f cc /Ec

a;;

A 3

0.2221

>l

Hence;

Px4 2

x4

Ec3h'"x 3hx

From x = x4 /3 to x = x4 /2,
curvature

where the section is uncracked,

where the eccentricity, e;

=
=

and

Thus the curvature

Px4
2h

12Px4
2Ech 3

6P
Ech 3

[i - :J

X]

[~4 -

From x = x, to x = x4/3 the slope


X

curvature dx

x,

since it is zero at

thus the slope

x,

!X

dx
Px4"'
9Ech 3 X2

x,
X

Px4"'
9Ech 3

[-~]
x,

slope

Px4"'
9Ech 3

[_!_~
X
X
1

<slope at x

= x4 /3)

+ J curvature dx
x./3

A '

the

.!. [ X,

3J +
X

6P <x... /2 - x) dx

Ech 3

x./3

[ .!..x,

- 2x2].

~] +
x...

x./3
=

=
From x = x, to x

=x

/3, the deflection

= I
=

slope dx

x,

"'J

Px ...
9Ech 3

1 - 1 dx

X,

x,

=
at x

=x

Px ... "' [xx,- lnx]


9E c h"'

x,

/3 this is
=

Px ... "' [x... - lnx ... - 1 + lnx,J


9Ech 3 3x,
3

From x

=x

/3 to x

=x

/2, the deflection relative to that at x

=X

= I

slope dx
x./3

fx . "'
X

Ech 3

X4

+ 3x 4 x - 3x2 dx

9x 1
x./3

Ech 3
at x

=x

/2 this is

A 5

xx ... "' [ 9x,

XX 4

+ 3x4 x2
-2-

x"']

x./3

/3

[X

Ech

1BX 1

x4:a -

2
=

3x4 "'
8--

x,."' + x.. "' + x.. "' - x .. "'


6 + ;;""]
27x,
3

Px,."'
Ec h"'

[ x,. - 5
54x,
108

so the total deflection, w


=

2Px,.'" [ x,. - 1 + 1 ln 3x, + x,. - 5 J


x,.
54x, 108
9
9
Ech 3 27x,

2Px,."'
Ech"'

13. 54

[~
18x,

17 + ..!_ ln 3x,J
108
9
x..

now, with full restraint,


(from Appendix A1)

substituting this into the expression for deflection gives;

now

0.8547 Px,."
Ech 3

1/2

x4[1 -

2~,

x..

so

1. 7046 x..

and

Pl"'
Ech"'

0. 1726 pp
Ech 3

0.8547
l. 7046"'

Note. This expression can also be obtained by an algebraically simpler


method using the virtual work approach.
A3 Effect of Restraint Flexibility on Stress
In Appendix A1 it was shown that the slab's extension from the centre-line
to the support

This is equal to the lateral movement of each support so, if the supports
develop a restraint force, F, of K times the movement whilst stili giving
full rotational restraint, this leads to;
F/K

-2Px,."'[ln<x 4 /x,
3Ech2

>-

0.3486- <x.. l6x,

and, substituting for F using Appendix A1, this leads to;


A 6

>]

Px 4
2Kh

-2Px.. "' [ln(x 4 /x,) - 0.3486 - (x 4 /6x, >]


3Ech 2

3Ech

-4Kx 4 [ln<x.. tx,>- 0.3486 - <x.. l6x, >J

-3E.,h
4x.. [ln <x.,lx,) - 0.3486

Therefore:
Therefore
now
Therefore,

1/2
substituting

<x 4 /6x, )J

x... [ 1 - 2x, /x4 J

for

X4

and

expressing

the restraint

stiffness

relative to the axial stiffness of the uncracked slab strip:


=

Kl

-3[1 - 2x. /x.J

E.,h
Numerical solution of this equation gives a value of x,/x 4 for any given
restraint

stiffness.

By

substituting

this

into

the

expressions

Appendix A1 the restraint forces and the stresses can be obtained.

A 7

in

APPENDIX B. TRAHSVERSE SHEAR DEFORMATION OF LINE ELEMENTS

Undeflected cent re-line

R...,.,

Figure Bl :

Plan of line element

Assume the element illustrated in Figure Bl has only a uniform curvature,


C, and shear deformation, S.

Then:

and:

Rw, + Cl

t., + CP/2 + Sl

Substituting for Cl gives

=
Rearranging gives:

Therefore:

Sl

This deformation is used to calculate the shear force, F, using the elastic
shear stiffness of the slab.

In the program described here, this stiffness

is based on the gross area of uncracked concrete plus one third of the
area of cracked concrete.
To maintain equilibrium, the moments

=
=
are applied to the nodes.

Fl/2

This automatically results in the complimentary

shears being applied to the orthogonal


A 8

element~

The element provides no resistance to the uniform curvature, C.

Thus the

structure provides no resistance to the form of deformation shown in


Figure B2 and this deformation would not affect the results.
could lead to numerical instability.

However, this

To avoid this a nominal resistance to

uniform curvature is added.

Figure 82: Unrestrained deformation


The relevant terms of the stiffness matrix are then as follows;

M...,
R... ,

F,
ASG/2

-ASG/2

ASG/1

ASG/ 2

<symmetrical>

-ASG/1

-ASG/2

ASG/1
Where:
AS

is the the effective transverse shear area of the slab taken as the
width of the slab in the element multiplied by [de: + <h - de )/3J;
that is the uncracked slab area plus one third of the cracked area.

is the shear modulus of the concrete.


A 9

El"

is

the

nominal

transverse

bending

stiffness

element which is set to a very low value,


elastic transverse bending stiffness.
F

is the transverse force

and the other notation is as used previously.

A 10

of

the

slab

in

the

less than 1% of the

APPENDIX C. I..ARGJ: DISPI..ACEMENTS

Cl Example Showing Effect of Vertical Component of Axial Force

~4 P
a. deflected shape

b. bending moment about deflected centre-line

PMl

PMl

c. vertical component of axial force applied to nodes

Figure Cl:
Assume

the

initially

Three element strut

horizontal

strut

subjected only to the axial force P.

illustrated

in

Figure

Cla

is

Clearly, the true bending moment

about the deflected centre-line of the strut is as shown in Figure Cl b.


The axial force in the strut is equal to P and, in an analysis using small
displacement

theory,

this

is

taken

as

acting

along

the

line

of

the

elements.
The vertical component of the axial force in the outer elements is Pt:./ 1,
where 1 is the length of each element.

If this force, which is ignored in

an analysis using s mall displacement theory, is applied to the nodes of a


computer

model

<which

otherwise

uses
A 11

small

displacement

theory)

the

resulting vertical forces are as shown in Figure C1c.

These give the

bending moments which are illustrated in Figure Clb and which are the true
bending moments in the real strut.

They also give shear forces in the

outer

the resultant

two elements

of P6/ 1.

Thus

line of thrust

acts

horizontally because, as in the real strut, the vertical component of the


thrust in the line of the elements is equal and opposite to the shear in
the elements.

Thus adding the vertical component of the axial force has

reproduced the true forces in the elements.

C2 Effect of Slope on Axial Extension

Figure C2: Inclined element

Consider the element shown in Figure C2.

For convenience the left hand

node is assummed to be undeflected.


According to small displacement theory, the axial extension of the element
is equal to the x displacement of the right hand node.
an additional extension due to the slope.

Taking the horizontal length of

the element as 1 and the inclined length as L. we obtain:

=
Therfore

1,.

Neglecting second order terms gives:


1,.

Hence the total axial extenxsion of the element

A 12

However, there is

APPENDIX D. NOTATION

Because of the many references to BS 5400, the notation used has been
made consistent with that document wherever possible.

The main symbols

used are as follows:


reinforcement area
b

width of section

diameter of circular contact area of load


depth to tension steel
depth of concrete in compression
Young's modulus of concrete
Young's modulus of steel
stress

f
fc

f., t

concrete stress on rectangular stress block (0.6f c >


tensile strength of concrete <normally effective value)
cube strength of concrete
cylinder strength of concrete
yield stress of reinforcement

force <normally restraint force)

overall depth of section

restraint stiffness

span <also used as element length>

bending moment

load

restraint factor (used in reference 72)


distance over which a crack affects the stress

vertical deflection

)(

x direction (always horizontal, normally along element>

y direction (horizontal and perpendicular to x)

Y<L

partial safety factor for loads


partial safety factor for errors in analysis
partial safety factor for materials
displacement

strain

reinforcement area (as percentage of concrete area, bd)

A 13

You might also like