Sunteți pe pagina 1din 202

Precast Concrete Raft Units

Precast Concrete Raft Units

Edited by

JOHN W. BULL
Department of Civil Engineering
University of Newcastle upon Tyne

Springer Science+Business Media, LLC


© 1991 Springer Science+Business Media New York
Originally published by Blackie and Son Ltd in 1991
Softcover reprint ofthe hardcover Ist edition 1991
First published 1991
AII rights reserved.
No part ofthis publicat ion may be reproduced.
stored in a retrieval system. or transmitted.
in any form or by any means-graphic.
electronic or mechanical. including photocopying.
recording. taping-without the
written permission of the Publishers

8ritish Library Cataloguing in Publication Data

Precast concrete raft units.


1. Construction materials: Precast concrete
1. Bull, J.W. (John William)
624.18341

ISBN 978-1-4613-6135-0

Library of Congress Cataloging-in-Publication Data

Precast concrete raft units / [edited by] 1. W. BulI.


p. cm.
IncIudes bibliographical references (p.
ISBN 978-1-4613-6135-0 ISBN 978-1-4615-2644-5 (eBook)
DOI 10.1007/978-1-4615-2644-5
1. Pavements, Precast concrete. 1. BulI. 1.W.
TE278.8.P74 1990
625.8' 4-dc20 89-25107
CIP

Phototypesetting by Thomson Press (India) Ltd., New Delhi


Preface

The use of precast concrete is a well-established construction technique for


beams, floors, panels, piles, walls and other structural elements. The advan-
tages of precasting include excellent quality control, economical large scale
production, improved construction productivity (especially in adverse
weather conditions) and immediate structure availability. These advantages
have been recognized for precast concrete raft pavement units (raft units) since
their introduction in the 1930s.
In the last ten years there has been a considerable increase in the use of raft
units, especially in their range of applications, their analysis and their design.
However, the description of these developments has been published in
academic journals and conference proceedings which are not readily available
to practising raft unit pavement design engineers. Pavement design engineers
are under increasing pressure to produce raft unit designs that are inexpensive,
long lasting and able to allow reorganization to accommodate changing use
and uncertainty offuture loading requirements. This is the first book devoted
to raft unit pavements, and will become a standard work of reference.
The aim of this book is to produce a range of up-to-date information that
allows a raft unit pavement design engineer to satisfy the client's requirements
for increased structural adequacy at reduced capital cost, whilst at the same
time complying with the operator's requirement of reduced maintenance costs
and increased organizational flexibility. Specifically, this book is aimed at
pavement design engineers and specifiers for roads, seaports, airports, floors
and heavy industrial areas where raft units offer financial and constructional
advantages. This book will also provide a guide to precast concrete
manufacturers who are looking towards new product markets, and to research
workers where it will indicate existing boundaries of raft unit research and
point the way to new research areas.
This book describes the present position regarding the theory, analysis,
design, maintenance, research and testing of raft units and links these
positions to future developments. To this end the book is divided into three
sections which consider, (1) analysis and design, (2) performance and per-
formance improvement and (3) good raft unit pavement practice.
In the first section-analysis and design-Chapter 1 reviews and compares
precast concrete pavements and raft units, their types, reinforcement and uses.
Chapter 2 looks at the analytical procedures for concrete pavements taking
into account subgrade support characteristics; this chapter also looks at
closed form solutions, computerized numerical analysis and dimensional
VI PREFACE

analysis used for data interpretation. The chapter goes on to show that further
research is required into raft design to achieve the same level of completeness
as exists for flexible pavements. In Chapter 3 a design method for two metre
square rafts is presented. The method has been validated in the laboratory and
is being extended to other rectangular shapes. The work shows that both
Westergaard's analysis and the equivalent single wheel load (ESWL) concept
can not be used for raft units. Chapter 4 assesses a number of raft unit design
procedures related to port pavements and subjected to fatigue loading.
In the second section-performance and performance improvement-
Chapter 5 looks at ways of improving raft unit impact resistance by using high
yield steel reinforcement and fibre reinforcement. Chapter 6 considers unrein-
forced raft units, by discussing stress and displacement due to moving loads,
climatic conditions, sub-base support and jointing requirements. Chapter 7
looks at the specific requirements of raft unit pavements used for the rapid
repair of runways following attack.
In the third section, Chapter 8 draws together the threads of the previous
chapters and develops a precis of good engineering practice for the analysis,
design, manufacture and use of raft units. The chapter also suggests areas
of future research and development.
In conclusion, I found this a particularly difficult subject area within
which to edit a book: the sources of information and expertise are widely
scattered and difficult to obtain. I would like to thank the chapter authors
for their considerable and successful efforts, and those other persons with
whom I corresponded but who were unable to contribute to this book-they
gave me considerable assistance.

lW.B.
Contributors

Dr H. AI-Khalid Department of Civil Engineering, University of Liverpool,


UK.

Dr J.W. Bull Department of Civil Engineering, University of Newcastle


upon Tyne, Newcastle upon Tyne, UK

Professor M. Fujii Department of Civil Engineering, Kyoto University,


Yoshidahonmachi, Sakyou, Kyoto 606, Japan

Professor A.M. Ioannides Department of Civil Engineering, University of


Illinois, Urbana, Illinois, USA

Lt. Col. L.J. Kennedy Ministry of Defence, Christchurch, Dorset, UK

Professor A. Miyamoto Department of Civil Engineering, Kobe University,


Rokkodai, Nada, Kobe, 657, Japan

Professor M. Poblete Department of Civil Engineering, University of Chile,


Santiago, Chile

Dr R.S. Rollings GeoServices Inc, Boynton Beach, Florida, USA. Formerly


with Pavement Systems Division, US Army Engineers Waterways
Experimental Station, Vicksburg, Mississippi, USA
Contents

A review of precast concrete pavements and rafts


R.S. ROLLINGS

1.1 Introduction I
1.2 New precast pavement construction I
1.2.1 Airfield construction 2
1.2.2 Road construction 2
1.2.3 Precast concrete slabs 4
1.2.4 Slab reinforcement 5
1.2.5 Slab prestressing 5
1.2.6 Fibre reinforcement 6
1.2.7 Precast operations 6
1.3 Pavement repairs with slabs 7
1.4 Concrete pavement design 9
1.4.1 Semi-infinite slabs 9
1.4.2 Finite-sized slabs 10
1.4.3 Concrete fatigue relationships 12
1.4.4 Subgrade loading 14
1.4.5 Slab handling stresses 14
1.5 Joints 15
1.6 Construction 16
1.7 Summary 17

2 Analytical procedures for concrete pavements 18


A.M. IOANNIDES
2.1 Introduction 18
2.2 Concrete pavement subgrade support characterization 18
2.2.1 The dense liquid foundation 18
2.2.2 The elastic solid foundation 19
2.2.3 Can the dense liquid and the elastic solid be compared? 19
2.3 Closed-form solutions for concrete pavement responses 20
2.3.1 Slab resting on a dense liquid foundation 20
2.3.2 Slab resting on an elastic solid foundation 23
2.3.3 Limitations of the closed-form solutions 25
2.4 Computerized numerical analysis methods for concrete slabs-on-grade 25
2.4.1 Computer programs for concrete pavements 27
2.5 Principles of dimensional analysis applied to concrete pavement data
interpretation 28
2.5.1 Dimensional analysis and data interpretation 29
2.5.2 Dimensional analysis applied to pavement systems 29
2.5.3 Implications of dimensional analysis: some examples 30
2.6 Conclusions 35
X CONTENTS

3 The design of precast concrete raft-type pavements 37


J.W. BULL
3.1 Introduction 37
3.1.1 Raft production and use 37
3.1.2 Raft design and laying 38
3.2 Previous methods of analysis 39
3.2.1 Pre-computer methods of analysis 39
3.3 Repair criteria 40
3.4 Loading 42
3.4.1 Standard axle loadings 42
3.4.2 Non-standard axle loadings 43
3.5 Computer modelling 44
3.6 Raft pavement design method 45
3.7 Design example 47
3.8 Conclusions 50

4 The behaviour of precast concrete raft pavements under fatigue


loading with special reference to their use in port areas 51
H. AL-KHALID
List of symbols 51
4.1 Introduction 51
4.2 Port loading 53
4.2.1 Dynamic loading 53
4.2.2 Static loading 62
4.3 Fatigue behaviour of concrete 64
4.3.1 Plain concrete 65
4.3.2 Reinforced concrete 71
4.3.3 Fatigue models 73
4.3.4 Port pavement model 76
4.4 Structural analysis 77
4.4.1 Analysis models 78
4.4.2 Precast concrete pavement models 82
4.5 Finite element analysis 83
4.5.1 Previous work 83
4.5.2 Current analysis 85
4.6 Conclusions 98

5 Performance improvement of precast, reinforced and prestressed


concrete raft units (beam and slab) under impulsive loading 101
M. FUJII and A. MIYAMOTO
5.1 Introduction 101
5.2 Load characteristics 101
5.2.1 Types of impulsive loads 101
5.2.2 Load modelling 103
5.3 Analytical studies 103
5.3.1 Analytical methods 103
5.3.2 Non-linear model 104
5.3.3 Finite element model for beams 106
5.3.4 Finite element model for slabs and handrails 108
5.3.5 Dynamic solution of equilibrium equation 111
5.4 Experimental studies 113
5.4.1 Test program for beams 113
5.4.2 Test program for slabs and handrails 114
5.4.3 Verification of analytical results for beams and slabs 118
CONTENTS XI

5.5 Concepts of performance improvement for impulsive loading 120


5.5.1 Concepts of performance improvement for beams 120
5.5.2 Concepts of performance improvement for slabs and handrails 123
5.6 Conclusions 130

6 The behaviour of plain undowelled raft-type concrete pavement 132


M. POBLETE

6.1 Introduction 132


6.2 Laying of a plain cement concrete pavement 134
6.3 ElTects of temperature change and reversible moisture movement 136
6.4 Structural response 141
6.4.1 Deflection 142
6.4.2 Joint efficiency 144
6.5 ElTects of pumping 146

7 Rapid pavement repair using precast concrete rafts· 150


L.J. KENNEDY

7.1 Introduction 150


7.2 Preliminary considerations 150
7.3 Surfacing types 152
7.3.1 Mal systems 152
7.3.2 Flush repair systems 152
7.4 Key factors in repair design 153
7.4.1 Loading 153
7.4.2 Durability 154
7.4.3 Specific military needs 155
7.4.4 Ride quality 155
7.4.5 Installation speed 156
7.5 Repair design 156
7.5.1 Bulk fill 158
7.5.2 Base layer 158
7.6 Raft design 160
7.6.1 Raft size 160
7.6.2 Raft edge detail 161
7.6.3 Raft reinforcement 163
7.7 Repair boundary 163
7.8 Special equipment 164
7.8.1 Subsurface screed beam 164
7.8.2 Raft-lifing beam 165
7.8.3 Concrete cutting saws 166
7.9 Practical experience 167
7.9.1 Early tests 167
7.9.2 Recent developments 167
7.9.3 Heavy aircraft tests 167
7.9.4 Current experience 169
7.10 Road repairs 170
7.11 Future developments 171

8 A review of the analysis, design, manufacture and use of precast


concrete raft pavement units 172
lW. BULL
8.1 Introduction 172
8.2 Manufacture of raft units 173
xii CONTENTS

8.3 Materials used in raft manufacture 173


8.4 The analysis of raft unit pavements 174
8.4.1 Concrete stress and subgrade stress 174
8.4.2 Analytical and numerical analysis 175
8.4.3 Loading 175
8.5 On-site laying raft units 176
8.5.1 Subbase 176
8.5.2 Joints 176
8.6 Conclusions 177

References 179

Index 190
1 A review of precast concrete
pavements and rafts
R.S. ROLLINGS

1.1 Introduction
Precast concrete technology is well-established in many fields of construction.
A variety of engineered products such as beams, columns, bridge members,
telephone poles and railroad ties are precast in concrete at a plant and erected
on-site later. This potentially provides a better product at the plant than would
be obtained by in situ casting and may ease and speed construction in the field.
Concrete pavements typically consist of many slabs of identical dimensions
and are therefore viable candidates for precasting. For example, a 7.5 m wide
road could consist of approximately 534 slabs per km with each slab being
3.75 m long per side.
Manufacturing and curing concrete at a plant provides a more consistent
and stronger product than can normally be obtained at a construction site.
Special reinforcing such as conventional steel bars, prestressing cables, or steel
fibres are more easily incorporated in a factory setting than at the construction
site. Also, adverse environmental conditions do not impede precast construc-
tions as they do conventional cast-in-place concrete. Once fully cured precast
units are in place, they can be opened to traffic immediately since no further
delay is needed for curing or strength gain. Consequently, precast concrete
pavements offer potential economic advantages, have the ability to be placed
under adverse conditions and can speed the opening of newly constructed
pavements.
The following sections will review some past applications of precast
concrete for pavement construction and repair, and will discuss some problem
areas that require future development.

1.2 New precast pavement construction


Clay brick surfaced roads probably represent the earliest precast manu-
factured road pavements. High-quality concrete block pavements were
developed in Europe in the 1950s, partially as a replacement for scarce clay
bricks after World War 11, and have spread to many areas ofthe world. These
approximately brick-sized units are manufactured under tight controls and
commonly achieve compressive strengths in excess of 55 MN. This illustrates
2 PRECAST CONCRETE RAFT UNITS

the potential quality control that is achievable in a factory setting or a precast


yard. Because of their relatively small size, these paving blocks are unable to
carry load through bending as is done by conventional concrete pavement, but
instead they distribute surface loads through base and subbase layers to the
subgrade. They are essentially a concrete-surfaced flexible pavement and are
outside the scope of this book. More information on block pavements may be
found in references [IJ-[7].

1.2.1 Airfield construction


Unreinforced precast concrete hexagons were used for the first concrete
airfields in the Soviet Union in 1931-32. These hexagons were 1.25 m long and
100-140 mm thick. Larger hexagons 1.5 m long and 140-220 mm thick were
later introduced for heavier aircraft. Problems with rocking and spalling of
these early hexagonal units led to their general replacement with conventional
cast-in-place reinforced concrete [8].
The first airfield use of prestressed concrete pavement occurred at Orly in
Paris and used precast slabs that were triangular with side lengths of 1 m and a
thickness of 160mm (9, 10). Later a 61 m square section of airport pavement
was constructed at Finningley, UK, of 9.1 m by 2.7 m, 150 mm thick precast pre-
stressed slabs [9]. Also a taxiway at Melsbroek was constructed of 1.25 m by
12 m, 75 mm thick precast prestressed slabs [11]. These slabs were pretensioned
and cast as parallelograms. After placement the slabs were post-tensioned with
transverse cables to obtain interaction between slabs and to increase their
structural capacity. Six experimental precast prestressed slabs 2.3 m by 10 m
and 200mm thick were constructed for DC-8 aircraft in Japan [12]. In the
Soviet Union precast, prestressed slabs have been acceptable for airfields
subject to twin tandem gears of 55000 kg and single gear loads of 30000 kg
[8]. Their use is particularly beneficial when dealing with non-uniform
swelling or settlement, construction during freezing temperatures, awkward
project geometry that limits the use of conventional paving equipment,
requirements for rapid construction, and strengthening existing pavements.
Precast, prestressed slabs have been used a number of times in the past for
airport construction. The heavy aircraft loads require structurally strong
pavement where prestressed concrete can be used very effectively.

1.2.2 Road construction


The Soviet Union has also used precast concrete slabs for road construction.
Over 180 miles of temporary roads such as forest roads were constructed and
gave good performance during a 10-year evaluation [13]. Soviet technical
literature favourably reports other precast road construction around
Moscow, under heavy industrial traffic in the Donbass, on the Kiev-Odessa
Highway, and elsewhere [14-18]. Between 80000 and 100000m of precast
hexagonal precast slabs were placed in the Moscow area during the period
A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 3
1968-74, and their use was reported to be increasing [19]. These slabs were
hexagonal with side lengths of 1.15 m and 180 mm thick.
In 1968 the South Dakota Department of Highways and the US Federal
Highway Administration built a 7.3 m by 274 m section of highway with
precast prestressed concrete [20]. The precast slabs were 1.8 m by 7.3 m in plan
and 114 mm thick. They were reinforced longitudinally with 9.5 mm cables
pretensioned to provide 2760 kPa prestress and reinforced transversely with
conventional No.3 steel bars (9.5 mm diameter). A 38-89 mm thick asphalt
overlay was placed over the concrete to provide the required surface slope and
to improve smoothness.
The Ohio River Division Laboratory of the US Army Corps of Engineers
designed and tested a precast sectional mat for military missile carriers [53].
This mat was to be capable of being assembled by military troops, should
support traffic from missile launchers with wheel loads up to 11 365 kg, and
must withstand the 400 kN thrust of the missile firings. Figure 1.1 shows the
OJ m wide, 5.5 m long ribbed beams developed to meet this requirement. Low-
weight sintered shale aggregate was used to keep the beam weight down to
250 kg so that it could be handled by a crew of 8-10 men. The concrete
obtained 28-day compressive strengths in excess of 40 MN and was preten-
sioned longitudinally to 8275 kPa of prestress with the 6.4 mm cables. Both
rods and cables were used successfully to transversely post-tension the
individual beams together into monolithic mats of the desired size. Testing of
these units found that wheel loads were generally distributed over three beams.
The mats withstood the missile blast tests but developed some spalling at
beam edges under traffic. Three failures occurred in the thin plank sections of
the beams under the heaviest 10900 kg wheel loads tested, but the mat was

BEAMS SYMMETRIC ABOUT /


CENTER UNE TOTAL /
lENGTH 5.4 m ~
r-----------------.,. /

28 mm HOlES FOR POSTTEN·


SENSIONING ROO OR C/J8lE

Figure 1.1 Mat made of precast, prestressed concrete (dimensions are in metres unless otherwise
marked) [53].
4 PRECAST CONCRETE RAFT UNITS

adequate for all other test loads. Some of the transverse post-tensioning cables
and rods lost up to 17% of their prestress under traffic loading. The weight of
the beams and lengthy assembly times precluded the use of this concept for a
portable missile pad, but the concept was thought to have potential for
temporary roads and storage areas. It was not pursued further.

1.2.3 Precast concrete slabs


Precast concrete slabs, commonly called rafts, are commercially produced in
many areas, particularly in Europe, and are used for industrial pavements such
as container storage areas and warehouse floors. They are particularly
appropriate for container terminals where large loads must be supported but
ground conditions are often poor, resulting in large settlements over time.
Concrete rafts provide the strength to support the large loads, but they are also
a flexible structure that can tolerate the settlements. They can be lifted while
the settlement is repaired and can then be relaid. These rafts are typically 2 m
square and 120-200 mm thick. They are generally reinforced with 0.3-0.5%
steel distributed in two layers at the top and bottom. The edges of the slab are
commonly chamfered or lined with steel to prevent spalling. A study found
that these rafts generally performed well but were usually economical only
when large settlements were a problem [21].
Table 1.1 summarizes the characteristics of several of the slabs used in the
past to build precast pavements. The length to width ratios vary considerably.
A square slab of sufficient dimension to allow two way bending to develop is
structurally more efficient than a long, narrow slab that bends predominately
in one direction like a beam. Several of the projects in Table 1.1, such as the
South Dakota highway and the sectional missile mat, used transverse post-

Table 1.1 Selected examples of precast pavements

Source Location Size (m x m x mm) Reinforcing

Glushkov and USSR Hexagon 1.25-1.5 m Plain


Rayev Bogol [8] long, 100-220mm thick
USSR 3.2 x 6.0 x 140 Prestressed
USSR 2.0 x 6.0 x 140 Prestressed
Mednikov et al. [19] Moscow Hexagon 1.15 m long, Plain and
180mm thick reinforced
Mellinger [53] USA 0.3 x 5.5 x 140 Prestressed*
Larson and Haug [20] USA 1.8x7.4x114 Prestressed
Patterson [21] Hamburg 2.0 x 2.5 x 140 Reinforced
2.5 x 2.5 x 140
Vandepitte [II] Melsbroek 1.25 x 12.0 x 75 Prestressed*
Hanna et al. [9] Finningley 2.8 x 9.2 x 150 Prestressed*
Stott (1955) London 0.9 x 0.9 x 165 Prestressed*
Harris [10] Orly 1.0 x 1.0 x 160 Prestressed*
Sato et al. [12] Japan 2.3 x 10.0 x 200 Prestressed*

*Transversely post-tensioned in the field.


A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 5
tensioning to tie individual elements together and obtain two-way action from
the pavements. Units with length to width ratios greater than 2 develop
predominately one-way bending and would have less structural capacity
without this transver:;e post-tensioning in the field. Smaller units, such as the
earlier Soviet slabs, are lighter and easier to handle than large units but are less
effective for carrying loads. These small units require a stronger base to help
support the load than would a larger slab that distributes the load over a larger
area.

1.2.4 Slab reinforcement


Most of the slabs in Table 1.1 are reinforced, and in general precast concrete
slabs will require some form of reinforcing to withstand handling stresses if not
traffic stresses. Several smaller slabs have been used without reinforcing, but
their small size reduces the bending stresses that develop during handling.
Conventional steel reinforcing does not prevent cracking from traffic in
concrete pavements, but it does change the pattern of cracking and retards
deterioration of the crack. Because of this improved post-cracking behaviour,
the US Corps of Engineers' rigid pavement design method allows some
reduction in required pavement thickness for reinforced concrete, but this
reduction in thickness usually does not pay for the cost of the steel [22]. Most
other pavement design methods allow no thickness reduction for reinforce-
ment. Consequently, a designer would view the reinforcement used in the slabs
in Table 1.1 as protection against handling stresses and against deterioration
after unexpected cracking but would not reduce the required pavement
thickness for the reinforcing. This places precast concrete slabs at some
economic disadvantage compared to conventional cast-in-place concrete.

1.2.5 Slab prestressing


Prestressing offers a more efficient use of reinforcing steel and was used in
many of the slabs listed in Table 1.1. When a prestressed pavement cracks, the
crack forms a plastic hige and redistributes the load by increasing the negative
moments in the slab. After the load is removed, the prestress closes the crack,
and the slab remains functional. If a slab is prestressed to a minimum of
700 kPa in both the transverse and longitudinal directions, the slab will exhibit
this improved post-cracking behaviour and will have two-and-a-half to three
times the structural capacity of a plain concrete slab [22]. Several of the slabs
in Table 1.1 applied longitudinal prestress to the slab during manufacture and
then used transverse post-tensioning to assemble the slabs in the field.
Prestressing offers a method ofobtaining a large increase in structural capacity
for a given cross-section of concrete. The compressive stresses applied during
prestressing counter some of the tensile stresses caused by loading, and when
cracking does develop, the prestressed concrete remains structurally sound for
more loading. Prestressing is a common procedure in precasting yards, and
6 PRECAST CONCRETE RAFT UNITS

post-tensioning is a normal field operation in many areas. Consequently,


precast prestressed concrete slabs would appear to have considerable
potential.

1.2.6 Fibre reinforcement


Steel fibre-reinforced concrete offers another potential material for precast
concrete paving units although it has not been tried in the slabs listed in
Table 1.1 [23]. Conventional paving quality concrete may achieve flexural
strengths of 4100-5200 kPa, while steel fibre-reinforced concrete can achieve
flexural strengths of 6200-7600 kPa in the field. Also, once a crack forms the
steel fibre continues to bridge across the crack and the concrete has additional
load-carrying capacity. Extensive research has found that failure for these
pavements is the opening of the crack that allows corrosion ofthe fibre and not
the formation ofthe crack as in conventional unreinforced concrete pavement
[24]. The higher strength of steel fibre-reinforced concrete, together with the
improved performance after the initial crack forms, allows the design of these
pavements to be much thinner than conventional concrete pavements [25].

1.2.7 Precast operations


Precasting has potential for new pavement construction. A pavement consists
of many identical slabs with the same dimensions, thereby making precasting
feasible. The concrete strength achieved in a precasting operation should
exceed that achieved with in-situ construction and therefore the precasting
operation should make better and more economical use of the concrete
materials. Small paving jobs such as parking areas around commercial stores
or plants often encounter problems obtaining consistent paving quality
concrete from ready mix producers who are simultaneously supplying a
variety of different jobs besides the paving work. Under these conditions the
consistent high-quality concrete that is available from precasting is a major
advantage. However, transportation of the precast products becomes a major
consideration, so probably only an urban area would have a sufficient volume
of these smaller paving projects to make precasting cost effective. The ability to
continue construction during adverse environmental conditions such as
freezing temperatures and to open pavements to traffic immediately after
construction without lengthy curing periods are additional valuable benefits
of precasting. Similarly, the ability of individual precast slabs to be removed to
allow corrective repairs to the underlying materials and then to be replaced
can be a decisive advantage for projects such as port pavements that must be
built on soft, settlement-prone soils. Past projects have not taken full
advantage of special reinforcing methods such as prestressing or steel fibre
reinforcing that develop the maximum structural capacity for the volume of
concrete used, and more could be developed in this area.
Precast pavement construction has generally been more costly in past
A REVIEW OF PRECAST CONCRETE PA VEMENTS AND RAFTS 7
Table 1.2 Relative costs of precast highway construction at Brookings, South Dakota

Construction task % of total cost*


Slab manufacture 73.0
Transporting and unloading slabs 6.0
Fine grading subbase and placing sand layer 3.7
Placing sl1!bs 3.3
Grouting slabs 9.0
Asphalt concrete overlay 5.0

*Based on costs reported by Larson and Haug [20].

projects than has conventional construction and this has been a limitation in
its use. For instance, the cost per square metre of pavement for the Brookings,
South Dakota, highway project, even after allowing for possible savings
through improvements in the procedures, was 1.9-2.7 times that of other
highway projects in the surrounding area [20].
The Brookings project was an experimental effort and, consequently, a
premium price was paid for the work. Experience would determine the most
cost-effective construction procedures, and costly experimentation plus on-
the-job learning would be eliminated. Consequently, costs could be substanti-
ally reduced. Table 1.2 shows the Brookings project costs as reported in [20].
The most costly items are the precast, prestressed concrete slabs. If mass
produced, the cost of individual slabs should be reduced. Also, the high
structural capacity of the prestressed members and high strength of the
concrete possible in a precasting operation must be evaluated when compar-
ing the cost of precast slabs to that of conventionally constructed pavements.
Transportation and unloading the slabs is a significant part of the pavement
cost in Table 1.2. This implies that precast construction will be most cost
effective for large projects where a precast yard can be set up on site or in urban
areas where there will be sufficient smaller projects requiring identical slabs to
keep transportation costs to a minimum. Improved design and construction
procedures could also reduce the cost of placing the slabs and making the joint
connections. If sufficiently smooth pavements can be built, the cost of the
asphalt overlay can be eliminated.

1.3 Pavement repairs with slabs


Precast units offer major advantages for pavement repairs. Traffic is only
interrupted during the repair itself, and as soon as the repair is completed the
pavement can be reopened to traffic. A notable example of such a repair was in
1981 at San Diego's Lindbergh Field [26]. During this effort, 116 damaged
slabs were replaced, at night, with precast slabs, while the field was kept open
to aircraft traffic during the day. The damaged slabs were removed and the
subgrade excavated 150 mm. The excavated subgrade was filled with lean
8 PRECAST CONCRETE RAFT UNITS

concrete, and a precast slab of the same dimension as the removed slab was
lowered onto the lean concrete. The slab was seated with a 9100 kg roller.
Patented load transfer devices were used to interconnect the slabs. When all
repairs were completed, the runway was strengthened with a 200mm asphalt
overlay.
The ability to reopen a pavement to traffic immediately upon completing
the repair is a major advantage of precast repair slabs. The cost of traffic delays
generally far outweighs the cost of the repairs themselves, so precast repairs
can be relatively costly on a per-square-metre basis compared to conventional
concrete but still be very economical over all. Similarly, it can be cost effective
to cast only a few precast slabs of a given dimension under these conditions
rather than have a large market for identical slabs, as would be needed for
economical construction of new precast pavement facilities.
The advantages of precast repairs has led several agencies in the US to
experiment with their use. Some of these agencies include the Michigan
Highway Department [27,28], Virginia Highway Department [27,29], New
York Thruway Authority [30], Florida Department of Transportation [31J,
California Department of Transportation [32], Texas State Department of
Highways [33] and the Department of Defense [34,35]. Table 1.3 presents
additional detailed information on some of these repair slabs. Reported costs
varied from $29.90/m 2 to $89.70/m 2 . More important in most cases than the
cost of the repair is the fact that the road was closed to traffic for as little
as 1 h 25 min to overnight during the repair. Most repair slabs were reinforced,
and the New York Thruway slabs were prestressed. The repairs described in
[29] were partial depth repairs to repair localized surface damage. A Klarcrete
cutting machine removed 64 mm of concrete, and precast slabs 50 mm thick
and 0,31 m by 0,31 m or 0.62 m by 0.92 m in plan were placed on a bed of epoxy.
The repair ofcontinuously reinforced concrete pavements with precast units
poses a particularly difficult problem as steel continuity must be maintained
from the original pavement to and into the repair. High stresses develop due to
the restraint necessary for steel continuity and a stress reliefjoint made with a
weakened plane in the precast slab will be needed for slabs longer than 2.1 m

Table 1.3 Selected examples of precast repair slabs

Source Location Size (m x m x mm) Cost $/m 2

Trans. Res. Board Michigan 1.8-3.7 x 3.7 x 203 49.04 to


[28] or 229 62.20
Grimsley and Morris Florida 3.7 x 6.1 41.87
[31]
Better Roads [32] California 3.7 or 5.3 x 3.5 x 203 89.71
Creech (1975) Virginia 0.3-0.6 x 0.3-0.9 x 50 269.11
Overcracker [30] New York 4.0 x 9.1 x 230 Not available
Meyer et al. [36] Texas 1.8 x 1.8 and 3.7 x 3.7 Not available
A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 9
[33]. This repair technique was successfully demonstrated in Texas by
repairing punchouts in a continuously reinforced pavement with slabs 1.8 m
by 1.8 m and 3.7 m by 3.7 m in plan in one afternoon [36]. This repair concept
requires removal of damaged concrete, placing the precast slab on a levelling
material, connecting steel in the precast slab to the steel in the continuously
reinforced pavement and then filling the space around the steel connections
with a rapid setting material. This rapid setting material controls the time
when the repair may be opened to traffic, but rapid setting polymer concretes
were effective for this task in the Texas repairs.

1.4 Concrete pavement design

1.4.1 Semi-infinite slabs


The structural design of concrete pavements generally requires the calculation
of the concrete pavement tensile stresses caused by a specific design vehicle.
These stresses are calculated with an analytical model and then related to the
pavement performance through the use of some fatigue relationship. Probably
the most widely used analytical design models for concrete pavements today
are the Westergaard edge-loaded model [39] or the Westergaard centre-
loaded model [38]. These models developed by Westergaard [38, 39]
represent the slab as a thin elastic plate that is characterized by a modulus of
elasticity and a Poisson ratio. The plate is resting on a bed of springs
characterized by a spring constant k called the modulus of subgrade reaction.
Solutions are available for the edge load case where the load is adjacent to the
free edge of a slab infinite in the other directions and for the load located in the
centre of an infinite slab. Considerable rigid pavement design work has also
been done using the layered elastic model [40,41] and a hybrid finite element
model that combines a finite element representation of the slab with a
Westergaard spring foundation [42,43].
Existing design methods using the above analytical models include a
number of assumptions and adjustments to reflect the using agency's
experience, desired level of performance, and necessary simplifications.
Consequently, it is not possible to blindly apply these methods to new
problems such as precast pavements without first checking to see that basic
assumptions used in the method are not violated. Those design methods using
the Westergaard analytical solutions all include the assumption of infinite slab
action. Model studies [44] and finite element analytical studies [45] have
found that this assumption will be met for slabs having dimensions not less
than 3l to 5l where
4 Eh3
1= 12(1 _ J12)k (1.1)
where I is the radius of relative stiffness, E is the modulus of elasticity of the
10 PRECAST CONCRETE RAFT UNITS

slab, h is the thickness of the slab, tt is the Poisson ratio of the slab and k is the
modulus of subgrade reaction.
Therefore, precast slabs can only be designed using any of the Westergaard-
based design methods if the slab dimensions in plan exceed 51. Many of the
precast slab units, such as the 2 m square rafts, will not meet this assumption.
Like the Westergaard-based design methods, the layered elastic model
design methods include inherent assumptions on slab size and joint construc-
tion owing to the way calculated stresses are related to pavement performance.
Therefore, the design methods based on layered elastic theory should not be
used for precast pavement design.

1.4.2 Finite-sized slabs


Finite element methods offer the most effective way of analysing precast
paving slabs that do not meet the infinite slab assumption necessary to use the
Westergaard models. This is illustrated in Figure 1.2 where finite element
calculated stresses and deflections for four slabs ranging from 1 m square to
2 m by 6.2 m are presented. The finite element model used for this analysis was
a hybrid model using four node plate elements to represent the slab and using
springs as in the Westergaard models to represent the material under the slab.
For this analysis the concrete modulus of elasticity in the slab was 41400 MPa,
the Poisson ratio was 0.15, and the modulus of subgrade reaction was
54.3 kPa/mm. The load for this analysis was placed adjacent to the free edge as
shown and was a 13640 kg wheel load (133 kN force) applied at a tyre pressure
of 724 kPa. This is a severe load representative of some heavy fork-lifts,
container handling equipment or some aircraft.
The slabs for this analysis were I m sq uare, 2 m square, 2 m by 4 m and 2 m
by 6.1 m. All slabs were 140 mm thick. The radius of relative stiffness for these
slabs was 0.65 m, so none of the slabs exceeded the 3-5 times the radius of
relative stiffness needed to ensure infinite slab action for Westergaard analysis.
In Figure 1.2 it is clear that the small I m square slab is unable to develop
much bending and develops lower tensile bending stresses but much higher
deflections than do the larger slabs. There is no reduction in stresses and only a
slight reduction in deflection when the 2 m square slab has its length increased.
This illustrates the statement made earlier that a square shape is structurally
the most efficient for precast slabs.
Under conventional concrete pavements that meet the Westergaard
criteria, the subgrade is subject to low stresses, and generally the subgrade
capacity is not checked in most conventional rigid pavement designs. How-
ever, the smaller precast slabs distribute their load through bending over a
smaller area, and now the subgrade under the rigid pavement may be critical
and must be checked. Spring foundations such as those used in the
Westergaard models and in the hybrid finite element model used to develop
Figure 1.2 allow reasonable calculation of stresses in the surface slab, but the
A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 11
DISTANCE ALONG THE V-AXIS. mm

0 250 500 750 1.000


-2.800

0 0
Q.

'";,

{t'5-
y
b
Vl

~x
Vl
w 2.800
eo
Vl
lS50mm
w
-' ~c,
iii
z
w
>-
5.600 I I
SLAB WxL. m
A 2 x 6.1
8.400 B 2 x 4
C 2 x 2
D I x I

11.200

(a)

DISTANCE ALONG THE Y-AXIS. mm

0 250 500 750 1.000


0
I 1 1 1 I I I

AB
E 2.5
E C
Z
0
;= ~
u
w
.zw 5.0
0

-
0
I r I
7.5
(b)
Figure 1.2 Effect of slab size on calculated slresses and deflections; (a) tensile stress, (b) deflection
Rollings and Chou [28].

validity of calculating stresses on the subgrade by multiplying the deflections


by the subgrade modulus is questionable.
The problem of including analysis of the subgrade conditions as well as the
tensile stresses in the slab for precast slabs was first addressed by Bull [46,47].
In his analysis, plate elements were used to model the slab, and brick elements
were used to model the underlying material. In this way both the stresses in the
slab and subgrade could be evaluated. Three-dimensional finite element
analysis of this type is quite expensive to conduct, and the results of this work
for 2 m square rafts has been converted into simplified design charts and a
microcomputer program [48,49].
12 PRECAST CONCRETE RAFT UNITS

The calculated stresses in the concrete pavement are related to pave-


ment performance through a fatigue relationship, and there are two basic
approaches to developing these relationships. The relationships can be devel-
oped on the basis of laboratory beam tests or using accelerated full-scale
traffic tests.

1.4.3 Concrete fatigue relationships


Beams may be cast in the laboratory and tested repetitively at loads less than
the concrete's ultimate flexural strength to failure at some number of load
repetitions. These results can then be used to develop fatigue relations such as
those in Figure 1.3. There is considerable scatter in concrete fatigue data so the
probability of failure is often shown as in Figure 1.3. During testing a beam
cannot be totally unloaded without causing rebound problems, and the ratio
of minimum to maximum stress during the test has considerable impact on the
fatigue relation as can be seen in the figure. Laboratory fatigue tests have
shown generally that concrete fatigue is the same in compression, tension, and
flexure, that frequency of loading can affect concrete fatigue depending on the
applied stress ratio, and that rest periods between loadings affect the concrete
fatigue strength.
Several concrete pavement design fatigue relationships are plotted in
Figure 1.4 with some of the beam test results from Figure 1.3. Note that the
ordinate in Figure 1.4 is the design factor which is simply the inverse of the

0.0 r-----,-----,----,------r---,----,-------,

--- ---
0.8

---- ........ --. ...... -.."

0.2

0'---_ _-L ....L. l...-_ _-L ...L- . l -_ _- l


o 10 10J 10'"
CYCLES TO FAILURE. N

Figure 1.3 Sample laboratory fatigue curves SMIN = minimum stress, SMAX = maximum stress,
P = probability of failure American Concrete Institute [52].
A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 13
3.0 r - - - r - - - - - - r - - - - - - - , - - - - - . . - - - - - - ,

2.5

2.0

1.5

1.0

0.5

a
1.000 10.000 100.000 1.000.000 10.000.000
COVERAGES OR LOAD CYCLES

Figure 1.4 Comparison of different concrete pavement design criteria (Rollings 1987).

stress ratio in Figure 1.3. The Portland Cement Association relation in


Figure 1.4 is probably the most widely used beam fatigue relationship and can
be seen to be a conservative interpretation of the laboratory results. The two
relationships marked CE in Figure 1.4 were developed by the US Corps of
Engineers from full-scale accelerated traffic tests of pavement slabs under
aircraft sized loads. Field tests of this sort include, to some extent, the stresses
developed in the pavement by temperature and moisture gradients and
include other effects such as non-uniform subgrade support. The two CE
relationships were developed from the same data, but one used the layered
elastic analytical model to calculate stresses while the other used the
Westergaard edge-loaded model. Since different models can calculate different
magnitudes ofstress, the fatigue relationship for field tests should be calibrated
to the specific model to be used in design. A similar effect can be observed in the
relationships in Figure 1.4 based on the AASHO road test results where
Treybig et al. [50] used the layered elastic model and Vesic and Saxena [51]
used the Westergaard model. The AASHO relationships both show a distinct
curved shape as opposed to the linear relationships exhibited by the
laboratory beam tests and the Corps of Engineers field tests. This discrepancy
is caused by the AASHO road test concrete pavements failing due to pumping
rather than concrete fatigue. Consequenctly, the AASHO road test relation-
ships have a fairly limited applicability to other soil and moisture regimes.
As noted earlier, the layered elastic model is probably not usable with
precast pavements, and therefore the two fatigue relationships in Figure 1.4
that are based on the layered elastic model are similarly inappropriate for
precast pavements. The Westergaard model is only applicable to larger
precast slabs that meet the required radius of relative stiffness requirements
discussed earlier. Therefore the CE-Westergaard relation in Figure 1.4 might
14 PRECAST CONCRETE RAFT UNITS

be used with appropriate sized slabs, but some questions remain about its
validity for these conditions since joint construction and subgrade support
might be appreciably different for precast slabs compared to the original test
slabs. The AASHO relationships are probably not appropriate owing to their
failure conditions and different construction details. The Portland Cement
Association, or some other similar laboratory beam fatigue relation, probably
represents the best fatigue relation to use with precast pavements at the
present time. The small precast slabs, such as the 2 m rafts, are sufficiently
small that thermal and moisture gradients in the slab should not cause any
problems.

1.4.4 Subgrade loading


In the past, subgrade conditions under rigid pavements might have posed a
pumping problem and required special pumping protection measures, but the
subgrade itself was not generally in any danger of being overstressed.
However, as noted earlier this is not necessarily true with precast pavements.
There is little data available at present to allow adequate analysis of this
problem. The procedure adopted by Bull [46,47J, using the compressive
vertical compressive subgrade strain criteria from flexible pavement design
methods, is a reasonable approach. Rocking of small precast slabs has been
reported to be a problem [19]. This may require a stabilized base or joint
connectors for the slabs. At present there is no clear method of analysing the
rocking problem at the design stage.

1.4.5 Slab handling stresses


All precast products must be designed for handling stresses, and the use of an
appropriate impact factor to account for rough handling would be prudent.

Table 1.4 Effect of pickup points on slab moments [23]

Slab Maximum M n Maximum M p %of maximum


length (m) Type of pickup L(m) (Nm) (Nm) moment

6.1 End 0.0 0 15494 100.0


1/4 and 3/4 points 1.52 3883 4368 28.1
Balanced moments 4.08 1940 1940 12.5
4.0 End 0.0 0 6947 44.8
1/4 and 3/4 points 1.0 1227 1942 28.0
Balanced moments 0.84 870 870 7.7

Noles:
1. Analysis assumes four pickup points, simply supported beam analysis, and both slabs 2 m wide
weighing 4181 kg and 2818kg respectively.
2. L denotes distance from end of slab to pickup point.
3. M n denotes negative moment.
4. M p denotes positive moment.
5. For balanced moments, negative and positive moments are equal.
A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 15
Design for handling must also consider whether the products will be handled
before they gain full strength and must consider any special loading that may
be applied such as stacking or transportation. It is quite feasible that the most
severe stress conditions that must be considered in design may come from
handling rather than traffic loads.
The pickup points for the precast slab must be adequate to withstand the
loads. They should also be located to minimize the stresses. Table 1.4 shows
how the location of the pickup points can be adjusted to reduce the moments
in the slab.

1.5 Joints
If precast pavement slabs are joined together they will provide structural
support to one another, reduce faulting between slabs, and may reduce rocking
problems that have been reported for some of the smaller slabs. However, by
joining the slabs together some flexibility will be lost. It will no longer be
possible to remove the slabs and re-use them after correcting settlement,
obtaining access to utilities, or to relocate the pavement. Whether connections
are made across the joints or not, depends upon the intended pavement use.
A conventional concrete pavement slab that uses dowels or keys in
construction joints, or that develops good aggregate interlock from contrac-
tion joints on short spacings, will be able to get significant structural support
from the adjacent slabs. Tests have found that for the purpose of design, it is
reasonable to reduce the free edge stress, calculated with the Westergaard
analytical model, by 25% to account for this load transfer across joints when
designing airfield pavements or parking areas where the free edge of the
pavement will not be trafficked [25]. A similar structural advantage is
recognized in the design of concrete highways if a tied concrete shoulder is
used adjacent to the concrete pavement rather than the conventional flexible
shoulder [42].
Considerable effort will be required for precast pavements to develop load
transfer levels equivalent to those of conventional pavements. There are some
patented load transfer devices such as were used in the San Diego airport
repair which are inserted into holes cored at the slab joints. Similarly, kerfs
could be sawn or cast in the slabs so that dowels could be placed across the
joints and then grouted into place [23]. Keys could be cast into the slabs, but
field assembly would be difficult, and it would be difficult to obtain the tight fit
necessary for an effective keyed joint. For some of the Michigan repairs,
dowels were inserted into drilled holes in the adjacent slab and then welded
to a steel plate cast into the slab [27].
Efforts to achieve high levels of load transfer, such as noted above, will add
significantly to the cost of the pavement and require more research and testing.
Some investigators have elected to use simpler connectors to maintain slab
alignment. This typically has been done by welding protruding reinforcing
16 PRECAST CONCRETE RAFT UNITS

bars together at a few points along the edge of the slab. The Brookings road
project used a grouted shear key between slabs but, as was seen in Table 1.2,
this added significantly to the expense of the project. Other investigators have
elected to leave the slabs unconnected. The space between slabs has been filled
with a variety of materials including sand, sand slurry, cement and sand
mixtures, fibre board and epoxy.

1.6 Construction
Precast pavement slabs are often placed on a 20-50 mm thick bed of sand or
sand cement and then rolled with a vibratory roller to seat and level them.
Another alternative approach that has been used is to inject grout under the
slab to raise it to the desired elevation. Precast pavements can have
smoothness problems due to the construction procedures and casting
tolerances used. Patterson [21] found 10% of the precast slabs in newly
constructed container terminal pavements had elevation differences of 5 mm
from surrounding slabs. This quality of construction may pose no particular
problem in storage areas but would be unacceptable for high-speed traffic.
Modern construction equipment such as fork-lifts or cranes can handle
sizable loads, thus large precast slabs can be handled in the field. Large slabs
can speed construction in open areas but can be awkward when transporting.
Equipment that employs a vacuum to lift precast elements is used in some
precast yards. Such equipment needs no pickup anchors in the precast slab
and simplifies field construction.
Construction equipment and procedures are adequate for the construction
of precast pavements for many applications, but improved construction
procedures are needed if the pavement is to take high-speed traffic. Construc-

Table 1.5 Suggested casting tolerances

Van der Wal and Walker (1976) Transportation Research


Measurement Flat wall panels Board [28] Repair slabs

Length and width ±3mm ~3.1 m ±6mm


+ 3 mm to - 5 mm for
3.1-6.2m
+ 3 mm to 6 mm for
6.2-9.2m
±6mm >9.2m
Thickness +6mm to -3mm ±3mm
Deviation of edge 1.6 mm per 3.1 m
from a straight line 6mm maximum
Squareness 6 mm difference in diagonals ± 13 mm in side length
Warpage 6mm
Surface deviation 6mm in 3.1 m ±3mm in 3.1m
A REVIEW OF PRECAST CONCRETE PAVEMENTS AND RAFTS 17

tion in the field can be greatly hampered by variable dimensions of the slab.
Table 1.5 shows some suggested casting tolerances available from several
sources.

1.7 Summary
Precast concrete construction has been used to build new pavements and to
repair existing pavements. Precasting pavement slabs allows the production of
a strong, durable product that will generally be of a higher quality than cast-in-
place pavements. Also, these precast slabs can be prestressed or can include
fibre reinforcing to significantly improve their strength and load-carrying
capacity. When precast slabs are used for pavement repairs, traffic is only
interrupted during the installation of the repair, and there is no delay
necessary for curing the concrete or for strength gain. The modular nature of
precast concrete pavements allows them to be taken up and re-used so that
they have particular value for temporary construction or where access to
underlying materials or utilities will be needed in the future.
Past experience has found precast pavement construction to be more costly
than conventional construction, and the final pavement is usually not as
smooth. Consequently, precasting has been used primarily for special
problems or situations rather than for general pavement construction.
However, there are distinct potential advantages for precast pavements that
may lead to their future wider use.
2 Analytical procedures for
concrete pavements
A.M. IOANNIDES

2.1 Introduction
Concrete slab-on-grade type pavements belong to that broad category, at the
interface of structural and geotechnical engineering, commonly referred to as
'soil-structure interaction problems'. As in numerous other engineering
applications, the pavement slab is treated as an elastic plate, but it is the
response of the supporting soil medium that is the governing consideration.
For an accurate evaluation of this response, the complete stress-strain
characteristics of the soil are required. The extreme variety of soil conditions
encountered in engineering practice all but precludes the development of
generalized stress-strain relationships applicable to any large group of soils.
The stress-strain behaviour exhibited by most natural soils is markedly non-
linear, irreversible and time dependent and these soils are, more often than not,
anisotropic and inhomogeneous.
The inherent complexity of real soils has led to the development of a number
of idealized models, which attempt to provide a useful description of certain
aspects of soil response under a specific set of loading and boundary
conditions. A major issue to be addressed, therefore, is how to determine which
model or idealization is best suited for any given application. The answer to
this question need not be unique. If fact, comparisons between responses
obtained using different models, might shed light on an area much larger than
the sum of individual portions illumined by each separate analysis. Several
factors must be considered. These include soil and pavement type, loading and
boundary conditions. Also included are the economics in terms of money and
time required for the analysis, design and construction ofthe proposed facility.
The two major subgrade idealizations commonly employed in concrete
pavement slab-on-grade analysis are reviewed in this chapter. Currently
available computer programs are also introduced, fully cognizant that the
tools presented here are becoming increasingly available to engineers.

2.2 Concrete pavement subgrade support characterization


2.2.1 The dense liquid foundation
In the determination of the structural response of concrete slab-on-grade
pavements, one of two fundamentally different hypotheses has been tradition-
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 19
ally used to idealize the properties of the subgrade. In the simplest of these
theories, the supporting soil medium is considered as a bed of closely spaced,
independent, linear springs. Each spring deforms in response to the stress
applied directly to it, while neighbouring springs remain unaffected. It is thus
assumed that the vertical stress q(x, y) occurring at any point on the
foundation surface is directly proportional to the deflection w(x, y) at that
point, i.e.
q(x,y) = kw(x,y) (2.1)
In this expression, k is the modulus of subgrade reaction, assumed to be
spatially independent. This idealization is commonly termed as a 'dense
liquid', and is almost universally ascribed to Winkler [1]. The first application
of the dense liquid model, by Hertz was, indeed, one involving a liquid rather
than a soil foundation [2]. In the analysis performed by Hertz, a floating ice
sheet is modelled as an infinite elastic plate and the Winkler assumption
represents a simple consequence of Archimedes' principle. The use of the
Winkler idealization in the study of beams and plates on an elastic foundation
was highlighted by the classical treatises of Zimmermann [3], Schleicher [4]
and Hetenyi [5].

2.2.2 The elastic solid foundation


In the second commonly employed support characterization theory the soil is
regarded as a linearly elastic, isotropic, homogeneous solid, of semi-infinite
extent. The terms 'elastic solid', 'elastic continuum', or 'Boussinesq's half-
space' are often applied to this idealization. It is regarded as a more realistic
representation of actual subgrade behaviour than the dense liquid model, in as
much as it takes into account the effect of shear interaction between adjacent
subgrade support elements. Consequently, the distribution of surface displace-
ment remains continuous. The deflection at any point occurs not just as a
result of the stress acting at that particular point alone, but is influenced to a
progressively decreasing extent by stresses at points further away.
Widespread use of the elastic solid foundation has been inhibited, however,
by its mathematical complexity. Unlike the Winkler model, where the
governing equations are of a differential form, problems associated with the
elastic continuum generally require the solution of integral or integro-
differential equations [6].

2.2.3 Can the dense liquid and the elastic solid be compared?
Analytical results obtained using a dense liquid (Winkler) foundation may
differ substantially from those based on elastic solid (Boussinesq) idealization,
especially in the case of subgrade stresses [7]. Vesic and Saxena [8], among
others, point out that 'no single value of k can yield agreement between the two
analyses of all statical influences'. A theoretical explanation of this pheno-
20 PRECAST CONCRETE RAFT UNITS

menon is provided by considering a non-homogeneous half-space. In situ soil


masses typically exhibit considerable divergence from the assumed homogene-
ity. As a result of depositional processes, especially the effect of overburden
pressure, soil stiffness generally increases with depth, with the possible
exception of a desiccated crust of higher stiffness near the surface.
The analysis of a non-homogeneous half-space presented by Gibson [9] is
particularly interesting, since it provides (as an unexpected bonus) a rigorous
theoretical relation between the dense liquid and elastic solid models. Gibson
examined the case of an incompressible medium (Poisson ratio, Ils = 0.5)
whose shear modulus Gs increases linearly with depth, z, i.e.
Gs(z) = G.(O) + mz. (2.2)
In the special case when the modulus is zero at the surface (Gs(O) = 0), he
showed that a rectangular strip loading causes a uniform surface settlement of
W o within the loaded area. This settlement is proportional to the applied
pressure and can, therefore, be written as:
(2.3)
where
ks = 2m. (2.4)
The factor ks is similar in nature to the modulus of subgrade reaction, and it is
shown in this case to be independent of the size or shape of the loaded area.
Outside the loaded area, the surface does not settle. This finding validates
Winkler's concept for a semi-infinite elastic medium, provided that Ils = 0.5
and its Young's modulus is given by Es(z) = 3mz. It was not established that
these conditions are both necessary and sufficient, only that they are sufficient.
It is conceivable that other combinations of Iliz) and Es(z) might lead to
similar results.

2.3 Closed-form solutions for concrete pavement responses

2.3.1 Slab resting on a dense liquid foundation


Several closed-form solutions for the response of a concrete slab supported by
a Winkler subgrade can be found in the literature. The most prominent among
them are those derived by Professor Harald Malcolm Westergaard (1888-
1950), whose pioneering analytical work has been at the heart ofconcrete slab-
on-grade pavement design since the 1920s. The 'Westergaard solutions' [to,
11, 12, 13] employ the theory of a medium-thick plate on a dense liquid
foundation, but are only available for three particular loading conditions: a
uniformly loaded circular, semicircular, or elliptical load applied at the
interior, edge, or corner of the slab. Furthermore, they assume a slab of infinite
or semi-infinite dimensions.
Since their first appearance in the early 1920s, the Westergaard equations
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 21
have often been misquoted or misapplied in subsequent publications. To
remedy this situation, a re-examination of these solutions using the finite
element method (FEM) was conducted by Ioannides et al. [14]. This exercise
yielded a number of significant results, among them the following:
(a) Several equations ascribed to Westergaard in the literature are erro-
neous, usually through a series of typographical errors or misapplic-
ations. The correct forms of these equations have now been conclusively
established, and are presented in Table 2.1.
(b) Westergaard's original equation for the edge stress is theoretically
incorrect. The long ignored equation given in his 1948 paper [12]
should be used instead.
(c) Using the FEM, improved expressions for maximum corner loading
responses have been developed, to replace Westergaard's semi-empirical
formulae. These are also included in Table 2.1.
A graphical adaptation of Westergaard's theory (which should be distingu-
ished from the Westergaard equations) was developed by Pickett and his
Table 2.1 Maximum response equations for slab on dense liquid foundation

Interior loading
Maximum bending stress, CT;

Ordinary theory:
BSIOT = {[3P(1 + /l) ]/2nh 2 } [In(21Ia) + 0.5 - y] + BSl20T
Special theory:
BSIST = {[3P(1 + /l)]/2nh 2 } [In(21Ib) + 0.5 - y] + BSl2ST
For square:
BSISQ = {[3P(1 + /l)]/2nh 2 } [In(21Ic') + 0.5 - y] + BSl2SQ
Supplementary, CT 2 (ordinary theory):
BSl20T = {[3P(1 + /l)]/64h 2 } [(aW]
Supplementary, CT 2 (special theory):
BSl2ST = {[3P(1 + /l)]/64h 2 } [(bW]
Supplementary, CT 2 (for square):
BSI2SQ = {[3P(1 + ,u)]/64h 2 } [(c'W]
Maximum deflection, 0;
Circle:
DEFlC = (PI8kI 2 ) {I + (1/2n)[ln(aI21) + y - 5/4)(aW}
where P = total applied load
E = Young's modulus of slab
/l = Poisson's ratio of slab
h = slab thickness
k = modulus of subgrade reaction;
a = radius of circular load
c = side length of square load
14 = {Eh 3 /[12(1 - /ll)k]}
b = [J (1.6a 2 + h2 )] - 0.675h if a < 1.724h
=a ifa>1.724h
c' = {e("/4)-IIJ2}c
Y = Euler's constant ( = 0.577 21566490)

Table 2.1 cont'd over


22 PRECAST CONCRETE RAFT UNITS

Table 2.1 (Contd.)

Edge loading
Maximum bending stress, CT.

Ordinary theory (semicircle):


BSEWOT = 0.529(1 + 0.54Jl) (P/h 2) [log! o(Eh 3 /ka1) - 0.71]
Special theory (semicircle):
BSEWST = 0.529(1 + 0.54Jl)(P/h 2)[log!o(Eh 3 /kbi) - 0.71]
'New' formula (circle):
BSEIC = [3(1 + Jl)P/n(3 + Jl)h 2] {In (Eh 3 /IOOka 4 ) + 1.84 - 4Jl/3 + [(I - Jl)/2]
+ 1.18(1 + 2Jl)(a!0}
'New' formula (semicircle):
BSEIS = [3(1 + p.)P/n(3 + p.)h 2] [In (Eh 3/IOOka1) + 3.84 - 4p./3 + 0.5(1 + 2p.) (a 2/Q]
Simplified 'new' formula (semicircle):
BSELS = ( - 6PW)(1 + 0.5p.) [0.48910g lO (a2/l) - 0.091 - 0.027(a 2/l)]
Simplified 'new' formula (circle):
BSELC = ( - 6P/h 2)(1 + 0.5Jl) [0.48910g lO (a/l) - 0.012 -0.063(a/Q]
Maximum deflection, c\
Original formula:
DEFEW = (1/-/6)(1 + 0.4p.)(P/kI 2)
'New' formula (circle):
DEFEIC = ({ P[(2 + I.2Jl)t]}/[(Eh 3 k)t])[1 - (0.76 + 0.4p.)(a/I)]
'New' formula (semicircle):
DEFEIS = ({ P[(2 + 1.2p.)tJ}/[(Eh 3 k}t]) [I - (0.323 + 0.17 p.)(a 2/Q]
Simplified 'new' formula (semicircle):
DEFELS = (I/J6)(1 + 0.4p.)(P/kI 2)[1 - 0.323(1 + 0.5p.)(a 2/Q]
Simplified 'new' formula (circle):
DEFELS = (I/J6)(1 + 0.4p.)(P/kI 2)[1 - 0.760(1 + 0.5p.)(a/Q]
where a 2 = radius of semicircle
b2 = [J(1.6ai + h2)] - 0.675h if a2 < 1.724h
= a2 if a2 > 1.724h

Corner loading
Deflection
b. = (P/kJ1)[1.1-0.88(atfl)] Westergaard
b. = (P/kJ1) [1.205 - 0.69(c/Q] Ioannides
Stress
CT. = (3P/h 2) Goldbeck, Older
CT. = (3P/h 2)[I - (atfl)°·6] Westergaard
CT. = (3P/h 2) [I _ (a/QO.6] Bradbury
CT. = (3P/h 2) [I - (atfl)1.2] Kelley, Teller and Sutherland
2
CT. = (3.2P/h ) [1- (atfQ] Spangler
2
CT. = (4.2P/h )(1 - {[(a/l)t]/[0.925 + 0.22(a/l)J}) Pickett
2
CT. = (3P/h )[1.0 - (C/QO.72] Ioannides
Distance to point of maximum stress
X 1= 2[(a l l)t] Westergaard
XI = 1.80CO.32Io.59 Ioannides
where a = radius of circular load tangent to both edges at corner
al = distance to point of action of resultant along corner angle bisector
= (j2)a
c = side length of square loaded area.
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 23
coworkers [15, 16]. These investigators presented influence charts which can
be used to determine the maximum responses obtained under an arbitrary
single- or multiple-wheel load, acting at the interior or at the edge of an infinite
or semi-infinite slab. It is interesting to note that, for a single-wheel load, the
edge moment chart agrees with Westergaard's 1948 equation, which supports
conclusion (b), above. Computerized implementations of the bending moment
influence charts have been prepared by Packard [17J for interior loading
(program PDLIB) and by Kreger [18] for the edge loading condition
(program H51).

2.3.2 Slab resting on an elastic solid foundation


Until very recently, analyses involving the elastic solid foundation had been at
a comparatively elementary stage, despite the general consensus among
engineers that the Boussinesq foundation is a much more realistic represent-
ation of the subgrade than the Winkler medium. Closed-form solutions to the
problem of a plate on an elastic solid have been limited. The special case of an
axisymmetric slab of infinite dimensions was treated by Hogg [19] and Holl
[20,21]. These researchers carried the numerical calculations to such a stage
that, for the first time, their results could be applied with relative ease to the
calculation of stresses and displacements in pavement slabs. Thus, Hogg
obtained expressions for the displacement and curvature of a plate under a
concentrated load, and for the curvature at the centre of the same plate loaded
with a circular load. Graphs of these quantities were also presented. Later,
Hogg also considered the case of a finite layer with a rough base [22] and
presented the exact mathematical solution for the plate displacement under a
point load in the form of an infinite integral and tabulated results.
On the basis of the work by Hogg and Holl, Losberg [23] developed a set of
equations for maximum responses of an infinite plate on an elastic solid under
a circular interior load. These are reproduced in Table 2.2. They are similar in
form to those presented earlier by Westergaard for the corresponding dense
liquid solution. In both cases, the equations consider only the first one or two
terms of an infinite series. A similar set offormulae had been obtained by Arora
and Khanna [24] by comparison to the dense liquid solutions.
Influence charts for the determination of the maximum deflection and
maximum bending moment in a slab of infinite size, loaded at its interior by an
arbitrary single- or multi-wheel load have been presented by Pickett et al.
[15,16]. For the case of edge loading, Pickett and his coworkers [25, 26] later
presented a similar chart for the maximum bending stress, which has not been
as extensively used as their previous charts. A computerized version of the edge
stress chart, called H51ES, has been developed by Ioannides [27]. No chart
has been presented to date for the maximum deflection occurring under edge
loading. The reason for this is that the theoretical treatment of this case
presented by Pickett et al. [25] contains significant errors.
24 PRECAST CONCRETE RAFT UNITS

Table 2.2 Maximum response equations for slab on elastic solid foundation

Interior loading (Losberg)


Maximum deflection, D;:
PI 2
D; = et: [I - (a/ly {0.1413 - 0.1034In(a/I,)}]
D3y3
Maximum subgrade stress, q;:
P
q; = - - [1 - 0.5513(all e ) + 0.1257(all,)2]
(3.j3) 1/
Maximum bending stress, u;:
-6P(1 + Il)
Ui = [0.1833 !oglo(alle ) - 0.0490 - 0.0 120(ally]
2
h
where Ie = J..j(2DIC)
D = flexural stiffness of the slab [ = Eh 3I {12( I - 1l 2 )}]
C = soil constant [ = EJ(I -Il;)]
Es = Young's modulus of soil
Ils = Poisson's ratio of soil
Edge and corner loading (loannides)
General form of best-fit equations:
R* = A loglo(c/I,) + B
except for:
Corner X I = A J (elle) + B
R* A B

Edge deflection -0.185 0.251 0.95


Edge subgrade stress -2.040 1.587 0.91
Edge bending stress - 2.148 1.042 0.98
Corner deflection -0.593 0.495 0.94
Corner subgrade stress -43.973 26.087 0.81
Corner bending stress -2.044 0.879 0.81
Corner XI 1.961 -0.412 0.86

where R* = [DD/PI:J, or R* = [bCI,/2P] for deflection


R* = [ql;IP] for subgrade stress
R* = [uh 2 /P] for bending stress.

More recent investigations by Ioannides and his coworkers have resulted in


major advances towards bringing the solution of the problem of a slab on an
elastic half-space 'to the same stage ofcompletion as the Westergaard solution'
[25J, applicable to the corresponding dense liquid case. At the heart of their
effort has been a systematic approach to the interpretation of numerical
analysis data (as obtained, for example, from the FEM), employing the
principles of dimensional analysis. As a result, predictive formulae for the
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 25
response of a long plate resting on an elastic solid foundation and subjected to
an edge or corner loading are now available. They were derived using results
from finite element (f.e.) computer program ILLI-SLAB, and apply to small
loaded areas of finite size [28]. These equations are also reproduced in
Table 2.2.

2.3.3 Limitations of the closed-form solutions


Although closed-form theoretical solutions are very desirable and are usually
fairly easy to use, they are not always available for the wide variety of practical
problems that arise in the field. Several assumptions made, although essential
to the theoretical formulation on which such solutions are based, are often
satisfied only partially, if at all, by real-life conditions.
Westergaard's dense liquid formulae are a case in point. As has already been
noted, they assume a single slab of infinite dimensions, loaded by a single tyre-
print at one of three specific locations. Thus, they are neither directly
applicable to a slab of finite dimensions nor to a slab loaded by a multiple-
wheel gear acting at an arbitrary location on the slab's surface. A third serious
shortcoming is their inability to handle joints, cracks or other weakness zones
in the slab. Such discontinuities are of paramount importance in determining
concrete pavement response.
Furthermore, to obtain closed-form solutions, the supporting medium must
be represented in a highly idealized fashion. Westergaard, for example, used
the Winkler foundation, and assumed that full contact existed between
pavement and subgrade. Clearly, several conditions of practical interest arise
in the field, which render such assumptions totally inadequate. Phenomena
such as local soil yielding, curling, warping, erosion and pumping, all lead to
significant departures from the assumed uniform and continuous subgrade
support. Soil variability and stress-dependent moduli (stress softening or stress
stiffening) add to the divergence between theory and reality.
The complexity and mathematical rigor of theoretical closed-form
approaches render a realistic simulation of such non-idealized conditions
intractable. Hence the popularity of numerical, approximate or iterative
techniques. These procedures have become increasingly feasible given the
progress achieved in computer technology in the last decade. An enormous
volume of computations may now be performed in a reasonable time,
permitting an adequately accurate solution to such problems to be obtained.
The main computerized techniques for this purpose available today are
discussed in the next section.

2.4 Computerized numerical analysis methods for concrete


slabs-on-grade
With the advent of high-speed digital computers, solution of complex
structural problems has been greatly facilitated. Among them is the case of a
26 PRECAST CONCRETE RAFT UNITS

concrete pavement, consisting ofjointed or cracked slabs of finite dimensions,


with or without load transfer systems at the joints and cracks. Such a
pavement may be under the action of an arbitrary single- or multiple-wheel
load, applied anywhere on the slab surface. A temperature differential between
the top and bottom surfaces of the slab may be acting on the pavement alone
or in combination with external wheel loads. The support of the subgrade may
be continuous or intermittent, while the subgrade itself may be represented by
any of the idealizations discussed above, or even more complex variants
thereof (e.g. non-linear, inhomogeneous, anisotropic, etc.).
A number of powerful numerical analysis methods have evolved in the last
few decades. These methods are applicable to a wide range of complex,
boundary value problems in engineering. For concrete pavements, in parti-
cular, the following three methods appear to be the most prominent:
(a) the finite element method (FEM);
(b) the finite difference method (FDM); and
(c) numerical integration techniques.
The particulars of the first two methods are discussed in considerable detail
in several standard textbooks. According to the FEM, the structural system to
be analysed is first subdivided into a number of discrete bodies ('finite
elements'). A set of simultaneous equations is then generated, describing the
response of the structure in terms of the applied loads and its total ('global')
stiffness. The latter reflects the contribution of all individual element
stiffnesses. The basic concept in a finite difference solution is to replace the
governing differential equation and the boundary conditions by corresponding
equations describing the variation of the primary variable (deflection) over
small but finite spatial increments into which the structure is again subdivided.
The third category of computerized numerical precedures often used for
concrete slabs-on-grade includes solutions involving integrals of Bessel,
elliptic or other functions over infinite or finite ranges. Use of digital
computers reduces the evaluation of such integrals to a routine task, requiring
only a small effort on the part of the user. Furthermore, computer results can
be very accurate since a large number of smaller subranges may be used
without any penalty in terms of user effort. This approach is fundamentally
different from the FDM or the FEM. In the FDM or the FEM, the numerical
procedure begins with the governing differential equations, which is an
essential part of the final solution. The use of numerical integration
techniques, on the other hand, is only an incidental choice of how to evaluate
the integrals in an expression derived after considerable manipulation of the
governing differential equation and the boundary conditions. In this sense,
this third category of solutions is more akin to an analytical (closed-form)
approach, than to the FDM and the FEM. Comparisons between these three
techniques can, therefore, be particularly enlightening.
The most important criterion governing the adequacy of all three of these
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 27
numerical approaches is the fineness of the subdivisions made during the
solution process. The requirement for a fine grid has to be tempered, by the
amount of computer memory available, as well as to the required execution
time. User guidelines for the design of adequate grids should be developed for
each new program, otherwise these powerful tools can be abused and result in
misleading conclusions.

2.4.1 Computer programs for concrete pavements


Numerous computer programs have been written for the analysis of pavement
systems using the three methods outlined above. The state-of-the-art in
computerized numerical analysis procedures for concrete pavements has been
presented in considerable detail in two recent comprehensive reports [27, 29].
The most important of these procedures are:
(a) ILU-SLAB This is a finite element (f.e.) program developed originally
for the analysis of jointed, one- or two-layer concrete pavements with
load transfer at the joints [30]. It was later expanded significantly to
incorporate a variety of subgrade support characterization models,
including the dense liquid (Winkler), elastic solid (Boussinesq), two-
parameter (Vlasov) and stress dependent (Resilient) idealizations [31].
More recently, the ability to accommodate a linear temperature
variation through the thickness of the slab has also been added [32].
The ILU-SLAB model is based on the classical theory of a medium-
thick plate, and employs the 4-noded, 12-degrees-of-freedom plate
bending finite element [33]. The Winkler type subgrade is modelled as a
uniform, distributed subgrade through an equivalent mass formulation
[34]. This is a more realistic representation than the four concentrated
spring elements used in other programs. The elastic solid option in ILU-
SLAB is based upon a procedure described by Cheung and Zienkiewicz
[35]. They assumed a piecewise uniform approximation to the subgrade
reaction, and formed the subgrade stiffness matrix by inverting the
flexibility matrix obtained using Boussinesq's theory. In the formulation
for the two-parameter foundation, the concept of strain energy was used
to obtain the additional stiffness matrices required [31]. These are more
accurate than those presented by Severn [36], which contained several
numerical as well as typographical errors.
ILU-SLAB uses a work equivalent load vector [33]. Various types of
load transfer systems, such as dowel bars, aggregate interlock, keyways,
or a combination of these can be considered at pavement joints. The
model can also accommodate the effect ofa stabilized base or an overlay,
either with perfect bond or no bond.
(b) J-SLAB This f.e. program is very similar to ILU-SLAB, and was
developed at the Portland Cement Association [37]. It includes the
capability for calculating stresses due to a linear temperature gradient
28 PRECAST CONCRETE RAFT UNITS

through the slab thickness, but is restricted to the Winkler subgrade


type.
(c) WESUQID and WESLAYER Written at the U.S. Army Engineer
Waterways Experiment Station, these two f.e. programs analyse a
concrete pavement resting on a dense liquid and an elastic solid
foundation, respectively [38]. They are based on a model developed by
Huang and Wang [39,40], and are very similar to ILU-SLAB and J-
SLAB. They permit the analysis of the effect of a linear temperature
gradient through the slab thickness and account for loss of support
(partial contact or initial gaps) between the slab and the subgrade.
(d) GEOSYS This f.e. program was developed for analysing rock-
structure interaction [41] , but was recently adapted to three-
dimensional concrete slab-on-grade analysis [29,42]. The modified
GEOSYS code accounts for the non-linear stress-strain behaviour of
the subgrade (stress dependence) through an iterative scheme.
(e) CFES Analysis ofaxisymmetric slabs offinite extent on an elastic solid
subgrade is performed in this program [43] using the method of
concordant deflections as proposed by Bergstrom et ai. [44].
(f) FIDIES This is a finite difference analysis program coded for
rectangular slabs on elastic solid [45].

2.5 Principles of dimensional analysis applied to concrete pavement data


interpretation
An obvious recourse to the lack of a closed-form solution to many practical
problems involving concrete pavements would be to use data obtained from
f.e. or other numerical procedures, and verify such predictions by comparisons
to actual field observations. The almost routine use of available sophisticated
computer programs, and the relative abundance of a large variety of carefully
collected in situ data, have transformed the problem confronting the
profession from one of data availability to one of data interpretation. The
principles of dimensio!1al analysis can be utilized very effectively in this
respect. These have been employed fruitfully in other branches of engineering,
but have largely been ignored in transportation facilities studies, particularly
since the introduction of computers in the early 1960s. Despite occasional and
admirable exceptions, the general trend in the last three decades has been to
show an overwhelming preference for - and an unlimited confidence in - the
results of sophisticated statistical analyses, without much consideration of the
underlying engineering interactions between the host of input parameters
involved. Although in highly empirical fields regression techniques will always
be an invaluable tool, the profession can benefit immensely by employing
dimensional analysis to determine the engineering dependent and indepen-
dent variables to be analysed. Without such exercise ofengineeringjudgement,
regression is lamentably bound to remain just that.
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 29
2.5.1 Dimensional analysis and data interpretation
The need for dimensional analysis is well recognized in those areas where
available analytical tools are not capable of yielding exact solutions. For
example, fluid mechanics is heavily dependent on numerical and empirical
work. Consider, for instance, the comments of fluid mechanics investigators
Roberson and Crowe [46]. They stress that in such areas, 'it is essential that
researchers employ dimensionless parameters [for] analysing model studies
and for correlating the results of experimental research'. For example, 'by
considering a non-dimensional form of Bernoulli's equation we will have made
a tremendous reduction in experimental work from that required before
considering the non-dimensional form. The process of non-dimensionalizing
the equation reduces the correlating parameters from five to two.' As a result,
considerable saving of time is effected with respect to data collecting, since the
non-dimensional factorial is much smaller than its dimensional counterpart.
Note, that as with the Bernoulli equation, it is often possible to ,[have] a clue
about the governing equation' from previous theoretical investigations, which
may themselves be incomplete. Nonetheless, 'by considering the dimensionless
form of that equation, we [are] able to obtain a set of dimensionless
parameters with which to correlate our data' [46].
Dimensional analysis was not unknown in transportation facilities studies
of the pre-computer era. Burmister's comments at the 1962 International
Conference on the Structural Design of Asphalt Pavements are a case in point.
From that forum, he advocated that the 'principles of dimensional analysis
should be rigorously followed, involving fundamental dimensionless ratios
which have physical significance'. This approach not only provides a useful
way of presenting theoretical and analytical data, but is also 'a more basic
approach in a comprehensive evaluation of field data, leading to dimension-
ally correct empirical relations' [47].

2.5.2 Dimensional analysis applied to pavement systems


The first step in applying the principles of dimensional analysis to pavement
systems, is to distinguish between 'input parameters' and 'independent
variables' entering the analysis, as well as between 'output values' and
'dependent variables'. It is often assumed that these pairs of terms have
identical meanings, thus resulting in extremely long factorials and incomplete
(often misleading) data interpretation. Regression algorithms obtained in this
way cannot be applied to data other than those for which they were developed.
In contrast, establishing independent and dependent variables by combining a
number of input parameters and output values into non-dimensional forms,
merely recognizes the fundamental engineering interactions between the
factors involved. This is much preferable to delegating this cardinal engineer-
ing task to the statistician, or more commonly the 'black box' of sophisticated
and complex statistical computer packages.
30 PRECAST CONCRETE RAFT UNITS

Through application of the principles of dimensional analysis, it can be


shown that the problem of an infinite slab-on-grade can be reduced to a non-
dimensional equation of the form [27,48,49]
R* = J(iog [a/I]) (2.5)
where R* is a dimensionless response; J is the logarithmic function of (a/Q
sought; a is the radius of the applied load; and I is the radius of relative stiffness
of the slab-subgrade system. The latter is expressed as a linear dimension, for
the dense liquid foundation, as
1= lk = 'V [Eh /{12(l-1t
3 2
)k} ] (2.6a)
or, for the elastic solid case, as
(2.6b)
in which E is the slab modulus of elasticity, It the Poisson ratio of the slab,
h the slab thickness, Es the Young modulus of the soil, Its the Poisson ratio of
the soil, and k the modulus of subgrade reaction.
The equations by Westergaard and Losberg, discussed earlier, present the
functional forms of J for the particular cases of the three primary maximum
responses, namely deflection, (j, bending stress, a, and subgrade stress, q, for
each of the three fundamental loading conditions, i.e. interior, edge and corner.
The non-dimensional responses, R*, can be extracted from these equations, as
follows [45]:
R* = [(jD/PI 2 ], or R* = [(jkF/P], or R* = [bCI/2P] for deflection;
R* = [qF/p] for subgrade stress; and (2.7)
R* = [ah 2 /p] for bending stress.
In these, D is the flexural stiffness of the slab [ = Eh 3/{12(1 - 1t 2 )} ]; C is the soil
constant [= EJ(l - Jl;)]; and P is the total applied load.
Thus, five of the six input parameters are lumped into a single non-
dimensional ratio (a/Q, which defines uniquely each of the non-dimensional
responses, independent of the values of the particular parameters. As in
Bernoulli's case, this problem is reduced to one of a single independent
variable, (a/I), and three dependent variables (the three non-dimensional
responses). A number of instances where progress has recently been achieved
through the application ofthe principles of dimensional analysis are discussed
below.

2.5.3 Implications oj dimensional analysis: some examples

2.5.3.1 Effect oj slab size Interpretation of numerical data using dimen-


sional analysis often leads to an extension of an available theoretical solution.
Adjustment factors, preferably based on a theory similar to that used for the
original equations, may be developed to account for digressions of in situ slab
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 31
behaviour from the assumed idealized conditions. For example, in order to
account for the finite extent of concrete pavement slabs, Ioannides et al. [31]
introduced the normalized length term, (Ljl), assuming that the width, J.v, of
the slab was equal to its length, 1. Using data from f.e. studies, they established
the minimum slab size requirements for the development of an infinite-slab
condition. These vary slightly depending on the subgrade idealization adopted
and the loading condition considered, but it is generally found that a
uniformly supported slab with an (Ljl) value in excess of about 5 behaves much
like an infinite slab. Ioannides, et al. [31] also presented plots illustrating the
effect of shorter slab lengths on the predicted responses. Slab size effects persist
significantly longer in curled slabs, where an (Ljl) ratio in excess of 15 is usually
necessary for infinite-slab response [50].

2.5.3.2 Consideration ofmultiple-wheel loads The effect of dual-wheel loads


may be quantified by (Sja), where S is the spacing of the two loads [51]. The
chart in Figure 2.1 can be used to convert a dual-wheel gear to a single tyre-
print, as a function of the wheel spacing. In this figure, P denotes the total
applied load on both wheels, while a is the radius ofeach tyre-print. From this,
a 'correction factor' may be derived, to be applied to the Westergaard single-
wheel interior load equation for the maximum stress, a i . This approach can
also be useful in investigating more complex loading gear assemblies.

2Dr------..-----r---.....,....---r---......---,---.......,

1.8
(S/a)
1.6 0.0 (Westergaard)
1.3
1.4
2.5
3.8
1.2
5.0
7.1
~IQ. 1.0 10.6
0.8
0.6
04
02
°O.L------I--~:-----I..--~--"'O'-'---::l'::----'
02 0.4 06
(all)
Figure 2.1 Nondimensional plot for effect of dual-wheel spacing.
32 PRECAST CONCRETE RAFT UNITS

3 ,..--"T"""--,----,--r---r---,-----r--,.--..,

DenSe LIquId (New Eqn )


2 -0-
_
_
-0---_ j
..
_
0

- "" ,
--.-.. - -0 - ELASTIC SOLID

o
""""" ........ -
~
' "'
'/' -'0- ...

Dense Liquid (Old Eqn.; circle), ' ~,


".~, ""'-. ~
"'~ Uq.,,,rn. Eq•.; "mi-d~"" '
..
-5L.-_....L-_........_ _......._ ........._......L_ _....... _.....L.._---'~_...J

o 0.50 1.00 1.50 2.00


o 0
Tor R;
Figure 2.2 Nondimensional plot for the determination of LPEF.

2.5.3.3 Determination of load placement effect factor Figure 2.2 shows the
variation of the ratio of the maximum bending stress, a e, developing in a
concrete pavement slab loaded by an edge load, divided by the corresponding
stress developing under interior loading, ai' as a function of the load size ratio
(a/f) [49]. The need to determine this load placement effect factor (LPEF)
stems from the fact that a e is generally more critical for design purposes than
ai' and that the latter is relatively easier to determine analytically. It should be
noted that the curves in Figure 2.2 apply to the case of an infinite slab, while
smaller slabs may be expected to give lower LPEF values. The value of LPEF
of 1.4 to 1.5, recommended by Thompson and Dempsey [52] and other
investigators, probably refers to such shorter (or cracked) slabs.

2.5.3.4 Analysis of combined temperature curling and external loads It has


long been recognized that merely summing up the stresses due to the applied
wheel loads and those induced by curling [53,54] is inadequate, and often
even erroneous [55,32]. The pertinent independent variable driving the
system in this case is the non-dimensional product of (aAT), where IX is the
coefficient of thermal expansion of concrete, and AT the temperature
differential between the top and bottom surfaces of the slab. Using data from a
f.e. study, Ioannides and Salsilli-Murua [50] developed an equation for the
maximum tensile stress in a concrete slab under the combined action of an
external wheel load and a temperature differential. For this purpose, they
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 33
introduced the ratio p, defined as:
(2.8)
where a, max is the maximum combined tensile stress under curling and load,
and at Wes is the maximum tensile stress predicted by Westergaard (LlT = 0).
For the particular case of an infinite slab loaded by a single edge load, and a
day-time (curled down) linear temperature distribution through the slab
thickness, the following formula was derived for p:
p = A + B{a/I} + C {log 1o(a/0} (2.9)
where A, Band C are functions of Ll T only, as follows:
A = 1.0-0.9152 Ll T
B=1.6215LlT (2.10)
C= -0.8713LlT
Note that equations (2.10) are presented in terms of Ll T (in OF) for clarity,
assuming a = 5.5 x 10- 6 etF. The fundamental relation involves (all T), and it
would be very easy to redefine A, Band C for any other a-value. Also note that
for an infinite slab, the unit weight of concrete has no effect.

2.5.3.5 Effect of load radius The form of the independent variable (a/0
implies that the sensitivity of the pavement system response to changes in load
radius, a, is just as pronounced as the effect of variations in its radius of relative
stiffness, I. Yet, in the majority of concrete pavement studies reported in the
literature the effect of the size of the loaded area is completely disregarded.
This may be relatively unimportant in the case of full contact (e.g. zero
temperature differential), but it is tremendously more pronounced under
curling or other partial support conditions, particularly for (a/I) values
between 0.05 and 0.1. This is the range in which a large number of actual
pavements and loads fall.
Such considerations suggest that the equivalent single-axle load (ESAL)
concept [56], according to which all traffic loads are reduced to an equivalent
single-axle load of a standard weight, is flawed, since it most often implicitly
assumes a constant value ofload radius. A much more fundamental reduction
would have been to express mixed traffic in terms ofan equivalent radius of the
applied load, leading to an ESAR concept. The major reason, of course, for the
preference given to the ESAL concept is that axle loads are much easier to
determine and control than are tyre contact radii. Regrettably, the system
response is naturally oblivious to matters of practical expediency.

2.5.3.6 Selection of slab dimensions The prevalent recommendation in the


United States with respect to the maximum slab length is
(L/h):::;21 (2.11 )
34 PRECAST CONCRETE RAFT UNITS

while recent European experience suggests


(Llh)";; 25. (2.12)
Application of the principles of dimensional analysis indicates that the
pertinent fundamental engineering independent variables are not the slab
length, L, or width, W, but the non-dimensional ratios (LIl) and (WIl). A re-
examination of the f.e. results by Darter [57] suggests that limiting the (LIl)
ratio to about 4 will ensure that the combined stresses under a temperature
differential of + 30°F (or less) will not exceed a value equal to twice that
predicted by Westergaard. Note that this applies for a single slab with
unprotected free edges. If load transfer is provided, it may be expected that
longer panels may be allowed. A comparison between the (Llh) criteria most
commonly used in the USA and in Europe and the proposed (LIl) criterion
suggested here shows that problems may arise in a pavement designed using
an (Llh) criterion as E decreases or k increases. In addition, slab lengths
determined according to the two (Llh) criteria lie for the most part within the
range defined by (LIl) = 4 and (LIl) = 6. Thus, the former is shown to be a
fairly conservative choice, while (LIl) = 5 appears to be a very promising
alternative. Finally, the f.e. study conducted by Ioannides and Salsilli-Murua
[50] suggested selecting a lane width, W, so that (WIl) is between 3 and 6.

2.5.3.7 Determination offrictional stresses The expression commonly used


in computing the maximum tensile stress, O"r, arising in a concrete slab as a
result of frictional restraint assumes that a purely frictional phenomenon is
observed [54,58]. Applying the principles of dimensional analysis, and
considering a unit width of the slab, this can be restated in the following
generalized form:
(2.13)
where O"r is the maximum stress due to subgrade friction, {3 is a dimensionless
reduction factor accounting for the non-uniform friction factor developing
under the slab, /1 is the maximum friction factor as obtained from the sliding
test, Ye is the unit weight of concrete and t\ is the maximum distance of the
'fixed point' (or point of zero displacement) from any free edge of the slab.
The form of the Bradbury-Kelley equation given above is preferred over
more commonly encountered versions of it, because it is dimensionally
homogeneous. It also shows that O"r is directly proportional to the friction
factor, /1, the unit weight ofthe concrete, Ye' and the slab length, L. Conversely,
for the purely frictional conditions implicit in Equation (2.13), O"r is independ-
ent of the slab thickness, h, except that this influences the friction factor and the
dimensionless reduction factor, {3.
The value of the dimensionless reduction factor, {3, may be determined using
an approximate method presented by Kelley [58]. Interpreting the Kelley
method using the principles of dimensional analysis indicates that the
dimensionless reduction factor, {3, depends on an assumed displacement, D,
ANALYTICAL PROCEDURES FOR CONCRETE PAVEMENTS 35
1.0....===::::=::,....,...,....-r---.-..........,.....,......""T'T"!----,.---.,r--T'.........""T"T"rl

08

as
(3
04

02

0.1 LO 10
0/ = (~A) /aLlT
Figure 2.3 Nondimensional plot for the determination of fJ as a function of t/J.

necessary for the development of the maximum friction factor, the slab length,
L, the temperature drop experienced by the slab, /)'T, and the thermal
coefficient of concrete upon contraction, IX. Figure 2.3 presents the variation of
p as a function of the dimensionless number, 1jJ, defined as:
IjJ = (DI2A)/(IX /). T). (2.14)
The curve in Figure 2.3 can easily be adapted to account for enhancements
in the method of calculating (1r, such as have been proposed by Iwama [59].

2.5.3.8 Back-calculation ofpavement parameters A consistent and theoreti-


cally sound approach, utilizing the principles of dimensional analysis, and
leading to a closed-form back-calculation procedure for a two-layer slab-on-
grade pavement system, has been presented by Ioannides [48]. This simplifies
considerably the effort required in interpreting non-destructive testing data. A
unique feature of this approach is that in addition to yielding the required
back-calculated parameters (i.e. soil and slab moduli or slab thickness), it also
allows an evaluation of the degree to which the in situ system behaves as
idealized by theory, and provides an indication of possible equipment
shortcomings. The method is extremely powerful and versatile, and may be
extended to include other loading conditions, as well as multi-layer systems
(flexible pavements). When the back-calculation is performed on a personal
computer (program ILLI-BACK), execution time per deflection basin is
trivial (~/s) [60].

2.6 Conclusions
In this chapter the analytical procedures for concrete pavement analysis are
presented and discussed within the following four headings:
(a) concrete pavement subgrade support characterization;
36 PRECAST CONCRETE RAFT UNITS

(b) closed-form solutions for concrete pavement responses;


(c) computerized numerical analysis methods for concrete slabs-on-grade;
and
(d) principles of dimensional analysis applied to concrete pavement data
interpretation.
The chapter shows that, in the last ten years, considerable effort has been
expended in identifying areas of analytical inadequacy and in providing some
of the required solutions. Considerable further research, development and
data interpretation are required before sophisticated concrete pavement
analysis can be incorporated into practical design guides. Dimensional
analysis can be of immense value in this direction.
3 The design of precast concrete
raft-type pavements
J.W. BULL

3.1 Introduction

3.1.1 Raft production and use


The use of precast concrete is a well-established, economic, manufacturing
technique used by the construction industry for structural and semistructural
members. The technique has also reached a high level of efficiency in the
production of concrete raft-type pavements which are manufactured in the
quality-controlled environment of permanent factories.
Large numbers of concrete raft-type pavement units are produced in
glassfibre moulds. Into these moulds is accurately placed the top and bottom
reinforcing steel. The concrete is then poured into the mould. After twenty-
four hours the raft units are taken from the moulds using lifting keys and
stored prior to their transportation to site.
The use of raft pavement units has developed rapidly since the early 1960s
[1]. Raft pavement units have been used for temporary roads, highway repairs,
pavements subjected to heavy industrial traffic, airfield construction and port
container terminals. For example, it may be necessary for a highway engineer
to close ofTpart of a road with the minimum of traffic interruption, while at the
same time maintaining safety at lane change over points. It may also be
necessary to use the hard shoulder or land adjacent to the existing highway,
together with the central reservation, none of which, on its own, is suitable for
large volumes of traffic. However, if these areas are used in conjunction with
concrete raft-type pavements, they would provide a suitable pavement surface.
Consequently, in these conditions, raft-type pavements have distinct advan-
tages over other forms of pavements as they can easily be laid, moved and
re-used.
The use of raft pavement units for specialist solutions, such as container
terminals which require a very high load capacity and good durability, are
common. These terminals are usually built on fill areas and are subjected to
large subgrade settlements. The pavements are exposed to petroleum products
and other chemicals leaking from the container-handling equipment, which
will soften bituminous pavements and possibly lead to their complete
disintegration. In situ concrete pavements are more resistant to most
38 PRECAST CONCRETE RAFT UNITS

chemicals, but they are prone to cracking and will break up when differential
settlement occurs.
The early port pavements were designed for much smaller loads. For
example, in the mid-1950s the maximum gross weight of a four-axle rigid lorry
was 22.35 t [2]. Today, in container areas, axle loads of up to 800 kN and
450 kN container corner casting loads are now common [3].
The shape of raft pavement units is predominantly rectangular, although
hexagonal raft pavement units have been cast. The size of the rectangular raft
pavement units has ranged from 1 m square to 3.2 m by 5.3 m and even 2.29 m
by to.Om. Aspect ratios have been between 1 and 18 with thickness of between
75 mm and 220 mm. Raft pavement units weighing in excess of 9 t have been
produced. More recently work concerning the use of raft units in the USA [4J
and UK [5] has been published.
Plain concrete, unreinforced raft pavement units have been manufactured,
but they are thick and difficult to handle due to the low tensile strength of
concrete. In order to reduce the thickness and weight of raft pavement units,
various forms of reinforcement have been used. Prestressed rafts are able to
carry increased loads owing to the compressive prestress and its ability to close
cracks when the loads have been removed. However, the manufacturing
considerations relating to prestressing over the length of the short raft
pavement units limits its use. The inclusion of short lengths of steel or other
fibres in the concrete can increase the flexural strength, tensile strength,
ductility and toughness of the concrete and reducing spalling. The greatest
effect of adding fibre reinforcement is found in the concrete's ability to absorb
high levels of impact loading and to limit crack growth. In limiting crack
growth, water infiltration is reduced and load transfer across the cracks is
increased. The most usual form of reinforcement in raft pavement units is steel
bars or steel mesh. This form of reinforcement is used in the following design
method.

3.1.2 Raft design and laying


In the UK a major research and development programme [6J, at the
University of Newcastle upon Tyne, produced two design methods [5,7J for
2 m square, steel reinforced, high-strength concrete raft pavement units. One
design method is described in this chapter with associated later work being
referenced at the end of the book. The research identified the need for a
square raft as it reduced concrete bending stresses and allowed thinner raft
pavement uni ts. The thickness ranged from 120 mm to 200 mm.
Each raft pavement unit normally has, around the top edge, a steel angle
frame. Within the raft is a double layer of reinforcing steel, with one layer at the
top and the other layer at the bottom. The concrete base has an octagonal
recess to prevent raft creep during service.
The raft pavement units are laid using a fork-lift truck. The subgrade or soil
THE DESIGN OF PRECAST CONCRETE RAFT-TYPE PAVEMENTS 39
is graded to the required profiles. Then a granular subbase, preferably in excess
of 250 mm thick, is laid and compacted. A lean concrete subbase can also be
used.
A bedding course of sand 50 mm thick is placed and compacted on the
subbase. This bedding layer is used to level the rafts and to ensure that there is
support underneath all of the raft pavement unit rather than just part of it, as
may happen with in situ concrete pavements.
Any loss of subgrade support due to settlement can be accommodated by
lifting out the raft pavement unit, refilling and relevelling the subbase and then
refitting the raft pavement unit.
The joints between the rafts are maintained by the use of steel spacers. These
joints are then filled using dry sand or a sand slurry.
One unusual feature ofthe raft pavement unit is the lack of permanent load
transferring joints connecting the rafts. By not using permanent joints, the
rafts can be recovered, retained and re-used.

3.2 Previous methods of analysis

3.2.1 Pre-computer methods of analysis


The earliest pavements were designed using empirical knowledge. Before 1901
test roads [8] were being used to determine the most suitable pavement types.
By 1940 [9J, highway engineers were able to correlate subgrade types with
pavement performance. The move towards the use of analytical pavement
design began in the early 1940s. There already existed an analytical method
[10], but it was not widely used. The method used the concept of a dense liquid
supporting an elastic beam and was extended [11] for an elastic slab on a
Winkler subgrade.
Many design methods based on either Westergaard's analysis [II] or
multilayer elastic analysis [12] are available. The original theoretical analysis
of semi-infinite rigid pavements [l1J, was found to give erroneous results for
the raft-type pavements [7]. The analysis assumed that the reaction of the
subgrade was purely vertical and that there was only one layer beneath the
pavement. The value of the modulus of subgrade reaction, k, used in the
analysis was found to vary according to whether the base-bearing pressure, the
pavement flexural stress or the vertical displacement were required. Work by
Ackroyd and Bull [7J has shown that Westergaard's analysis gives inaccurate
results for 2 m square rafts and that the actual wheel load must be considered
[13].
The method of multilayer elastic analysis [14] was used for flexible
pavements and later formed the basis of the multilayer elastic pavement
analysis computer programs [15]. This method of analysis could not be used
without modification for concrete raft-type pavements, as it assumes the
pavement to be semi-infinite and uses the concept of a single-wheel load
40 PRECAST CONCRETE RAFT UNITS

equivalent to a multiple-wheel load, to simplify the analysis. This assumption


has been shown to be invalid for raft-type pavements [13].
In the late 1950s extensive testing offull-scale concrete pavements indicated
that the behaviour of many subgrades was closer to that of an elastic-isotropic
solid, rather than a Winkler subgrade. The elastic model uses Young's
modulus of elasticity E and Poisson's ratio v rather than k, the modulus of
subgrade reaction, as its principal variables. A single layer analysis of the
subgrade was found to be too simple to represent the pavement and multilayer
solutions were proposed [14]. With the development of computers, it was
natural that the multilayer elastic solution should be combined with the finite
element method and used to research concrete raft-type pavement structures,
as has been done in this chapter.

3.3 Repair criteria


For raft-type pavements neither the deflection criteria of flexible pavements
nor the cracking criteria for rigid pavements are adequate repair criteria [16].
For example, in flexible pavement design, a pavement would be considered to
have failed if the surface deformation exceeds 25mm [16]. For raft-type
pavements, edge movements in excess of 50mm have little effect on the larger-
diameter wheels of container-handling equipment. Any differential settlement
between rafts is overcome by lifting and relevelling the raft pavement units.
The repair of raft-type units is directly related to the combined maximum
tensile stress developed in the concrete. Concrete can be subjected to stresses of
between 36% and 78% of its ultimate stress without any signs of distress [17].
Above this value microcracks and fissures form. As compressive loading will
close these cracks, concrete will fail in tension at a lower stress than in
compression. It has also been noted in laboratory tests carried out by the
author that raft-type pavements, which would be considered as requiring
repair under rigid pavement cracking criteria, will successfully carry repeated
loads up to 850 kN.
In order to specify a condition at which raft-type pavements need either
maintenance (the serviceability limit state) or replacement (the ultimate limit
state), it was found that the critical parameters were the maximum combined
concrete tensile stress in the raft pavement unit [18J and the maximum vertical
compressive stress on the subgrade [5].
The serviceability limit state can then be expressed as:

(a) the number of load applications N, related to


(b) the maximum combined concrete tensile stress Q and
(c) the tensile stress of the concrete C, as follows:

N= 225000[~J (3.1)
THE DESIGN OF PRECAST CONCRETE RAFT-TYPE PAVEMENTS 41
From the multi-axial design stress equations [19] that match the concrete
normally found in raft-type pavements, a maximum concrete tensile stress of
5.20 MPa was obtained. Laboratory testing by the author gave maximum
concrete tensile stresses of between 4.90 MPa and 6.00 MPa and a tensile stress
of 5.55 MPa was used.
The ultimate limit state, that is the point where the raft-type pavement unit
would need to be replaced, required further research work which developed
[20] a relationship between:
(a) the number of load applications N to the serviceability limit state.
(b) the loaded area contact pressure p (in MPa) and
(c) the number of load applications 'Null'
to the point where the raft pavement unit would need replacing, as follows:

I - x [contact pressureJC (3.2)


Nu t - N 0.875

The value of 0.875 MPa was the basic tyre contact pressure assumed for
Equation (3.2). The value of C is related to the applied contact pressure and is
given in Figure 3.1. In the original research work it was assumed that the tyre
contact pressure would not exceed 2.00 MPa and the same value for C is used
for all contact pressures below 2.00 MPa.
The pavement life can also be related to the load induced maximum vertical
compressive stress on the subgrade [21] and expressed as the number ofload

o 4 6 8 10 12
ConI act Pressure in MPa

Figure 3.1 The value of 'C' related 10 the contact pressure.


42 PRECAST CONCRETE RAFT UNITS

applications to the point at which pavement relevelling is required. The


relationship between

(a) Young's modulus, E, as used in the finite element analysis, and


(b) the California Bearing Ratio (CBR) in %(A) of the subgrade

can be approximated by the equation [16]:


E = 10.0 x A (MPa) (3.3)
A further development [21J was to relate:
(a) the allowable soil vertical compressive stress, B (in kPa), to
(b) the number, Np, of load repetitions,
by using Equation (3.1) together with the allowable soil vertical compressive
microstrain [22]. The result is the equation

Np= [
280 XAJ4 (3.4)
B
The use of Equation (3.4) allows the prediction of the number of load
repetitions Np at which the pavement will require relevelling due to excessive
subgrade deflection.

3.4 Loading

3.4.1 Standard axle loadings


In observing the number ofload applications on a raft-type pavement, it is
easier to appreciate the physical size and gross weight of the vehicle, rather
than the individual wheel or axle load. For highway vehicles, it is usual for the
number of axles to be increased as the gross weight of the vehicle is increased.
This reduces the individual axle load that is limited by law; for example:
(a) In Austria, Denmark, West Germany, the Netherlands and Sweden it is
10.00t per axle (equivalent to 2.26 standard 80kN axles).
(b) Belgium, France, Greece and Luxembourg have a limit of 13.00 t per
axle (6.46 standard axles), while
(c) in Italy the limit is 12.00t (4.69 standard axles) and
(d) in the United Kingdom the axle loading is limited to 10.17 t (2.42
standard axles).
To assess the different loaded axle effects on raft-type pavements, the mixed
axle loadings can be related through the use of the 80 kN standard axle load
[16]. For example:
(a) the 16 t gross weight vehicle may be represented by 2.6 standard axles;
THE DESIGN OF PRECAST CONCRETE RAFT-TYPE PAVEMENTS 43
(b) the 32 t vehicle may be represented by 5.4 standard axles; and
(c) the 38 t gross weight vehicle by 5.1 standard axles.

3.4.2 Non-standard axle loading


The wheel loads on raft-type pavements are usually very much higher [3] than
the 80 kN standard axle loading. Consequently, a different relationship is used
to relate the maximum applied wheel load, as used in the raft-type pavement
design calculations, to the variety of applied wheel loads. The relationship is as
follows [23]:

WJ3.75 [P J1.25
N =Nx -
m [Wm xPm- (3.5)

where N m is the equivalent number of the maximum load applications, N is the


number ofload applications of the load W, W m is the maximum applied load,
Pm is the contact pressure of the maximum applied load, W is applied load
and P is the contact pressure of the applied load W.
The usual assumption made for infinite pavements, that a single-wheel load,
equivalent to a multiple-wheel load [24] can be used to simplify the analysis,
cannot be made for raft-type pavements. Bull and Salmo [13] found that the
load position and the number ofloaded points affected the concrete raft tensile
stress and the subgrade stress, as shown in Table 3.1.
The finite element results shown in Table 3.1 are for the same total load
using a variety of wheel configurations, which are then related to the concrete
and subgrade stresses. The first line of the table refers to a single centrally
placed wheel load. The second line is for the same wheel moved across the raft-
type pavement until the maximum concrete and subgrade stresses are found.
The maximum value of2.03 MPa occurred when the wheel was at the centre of
one edge, but the maximum subgrade stress of 26.7 kPa occurred when the
wheel was at one corner. In an infinite pavement, the two maximum stresses
would occur at the same wheel position. Table 3.1 also shows that for two
wheels between 1.0m and 2.8 m apart, the concrete stress varies between

Table 3.1 Concrete tension and subgrade stress related to load


position

Concrete
tension Soil stress
Load configuration (MPa) (kPa)

Single central load 1.00 10.0


Movable single load 2.03 26.7
Two wheels 1.0 m apart 1.14 21.2
Two wheels 1.5 m apart 1.12 19.2
Two wheels 2.0 m apart 1.14 17.2
Two wheels 2.8 m apart 1.12 11.4
44 PRECAST CONCRETE RAFT UNITS

1.12 MPa and 1.14 MPa, a difference of 2%, while the subgrade stress varies
between 11.4 kPa and 21.2 kPa a difference of 86%.
By taking into account the actual applied wheel loads and their precise
positions, the raft-type pavement units are found to carry significantly more
repeated loadings than some pavement design manuals would suggest [25].

3.5 Computer modelling


Unlike highway pavement design where a large amount of empirical data is
available, the design method for raft type pavements relies heavily upon
computer modelling, with increasing verification by laboratory and on-site
testing.
To investigate the complexities of the raft-type pavement, the size and shape
of the loaded area plus the stresses in the raft and the subgrade, the finite
element package PAFEC was used [26]. The PAFEC system has available
many finite element types including three-dimensional brick elements. Two of
the brick elements, one a 30-node element and the other, an 8-node element,
both with three translational degrees of freedom, U x' Uy' and Uz' were used.
The elements gave deflections and stresses at each node. A check was also
made on the nodal stress discontinuities to assess the output accuracy.
The finite element numerical analysis had to be related to two practical
considerations. Firstly, the computer model must accurately model the
practical raft-type pavement. Secondly, the laboratory experimental results
must be related to the finite element results. For these reasons, the details given
in Figure 3.2 and Table 3.2 formed the parameters from which the datum

150

50 Bedding Layer

300 Sub-Base

600
Subgrade

-/'--------:1"'00"'0-----+----,1=500-------1

Figure 3.2 The finite element model of one quarter of the concrete raft type pavement.
THE DESIGN OF PRECAST CONCRETE RAFT-TYPE PAVEMENTS 45
Table 3.2 Standard raft-type pavement parameters

Pavement layer Basic value

Load 10 kN, centrally placed


Raft-type pavement 2m by 2m by 150mm thick
50 MPa concrete
314mm 2 reinforcement, each way top and bottom
Bedding layer 50mm thick
7.5% CBR
0.25 Poisson's ratio
Subbase 300mm thick
20% CBR
0.25 Poisson's ratio (drained)
Soil/Subgrade 600mm thick
0.3% CBR
0.3 Poisson's ratio (drained)

values of the concrete stress and subgrade vertical compressive stress were
determined.
The initial finite element model used 104 elements, 197 nodes and 458
degrees of freedom, but models with up to 1224 elements, 2030 nodes and 5285
degrees of freedom were used to refine the element mesh and provide
asymptotic values of the stresses. The applied load was initially a single 10kN
central point load, but over 600 computer runs were made with single- and
multiple-wheel loads at a variety of positions on the raft and the stress values
recorded. From the computer runs, a series of graphs and tables were
constructed to quantify the effect each variable had on the concrete tensile
stress and the vertical subgrade compressive stress. The results were then
transferred to a microcomputer and a spread sheet programme.

3.6 Raft pavement design method


In Figure 3.3 part of the flow chart for the microcomputer-based raft-type
pavement design method is given. The program is interactive and the raft
pavement designer can enter and alter data at any of the data input positions.
The program will automatically recalculate the design using the amended data
and a list of the final input data and output data is printed at the end of the
calculations.
The computer assumes the pavement parameters shown in Table 3.2 are the
basis from which to calculate the loaded pavement requirements. Any
variation from the basic values is automatically calculated and compared with
the required raft-type pavement life expectancy.
The program requests the number of the heaviest wheel loadings the
pavement must sustain over its design life and the subgrade CBR. The
46 PRECAST CONCRETE RAFT UNITS

Output:
Applied raft pavement stress number lAPS]
Design subgrade stress number [ASS]

~----NO
L..-_.-- -...J

YES

Input:
Subgrade CSR
Subgrade depth
Subgrade saturation

I-------<~NO

YES

Input:
Granular Sub-base thickness
Granular Sub-base CSR
Granular Sub·base saturation

~~~~ _ _r----""'NO

YES

~~~ _ _. . J - - - - NO
YES

END
Figure 3.3 Microcomputer based now chart.
THE DESIGN OF PRECAST CONCRETE RAFT-TYPE PAVEMENTS 47
program calculates the design raft pavement stress number (DPS) and the
design subgrade stress number (DSS) to give the required pavement life.
The program now requests the concrete compressive strength, the heaviest
wheel load and expected wheel configuration. The program then calculates the
applied raft pavement stress number [APS] and the applied subgrade stress
number (ASS) and subtracts the relevant design raft pavement stress number
(DPS) or the design subgrade stress number (OSS). If the outputs have a zero
or a negative value, then the basic pavement model of Table 3.2 will give a life
equal to, or in excess of, respectively, the required pavement life. If the results
are positive numbers, then the raft pavement designer knows by how much the
design must be altered to achieve the required pavement life.
The program now requires data concerning the subgrade, its saturation
condition, CBR value and depth to the next stronger layer. The program
calculates the resulting APS and ASS and compares them with the design raft
pavement stress number (DPS) and the design subgrade stress number (DSS),
respectively. The pavement designer can then determine by how much the
design must be altered to achieve the required pavement life.
The designer is now free to optimize the subbase and raft-type pavement
unit design. A decision must be taken to use either a granular subbase (GSB)
material or a lean concrete subbase. The program allows both options to be
used, although only the granular subbase option is shown here. If a granular
subbase is to be used, the CBR value, the thickness and the degree of saturation
is required. In general a subbase with a minimum CBR of 20%, a minimum
thickness of 250 mm and adequately drained would be used. If a lean concrete
subbase is preferred, its thickness and strength will be requested. The program
will again show the relationship between the required and achieved values.
As it is not always possible to achieve the pavement life using the standard
raft-type pavement unit, the program will allow the amount of steel
reinforcement in the raft pavement unit and the raft-type pavement unit
thickness to be altered. After this comes the final stage of the program.
Calculations take place to determine the pavement life both in terms of when
the raft-type pavement will require maintenance, its serviceability limit state
and the point at which the subgrade settlement will make relevelling of the raft
pavement necessary.
In some locations, where the soil CBR values are very low and the applied
wheel loading is high, the critical design parameter will be an inadequate
subgrade. It will be necessary, in that case, to rerun the program to determine
the improvement required in the soil CBR.
A permanent record of the data input and output is printed. This allows
possible alterations to be made to the design at a later date.

3.7 Design example


An industrial client wishes to have built a raft-type pavement using 2 m square
raft-type pavement units. The pavement is to have a serviceability limit state of
48 PRECAST CONCRETE RAFT UNITS

105000 vehicle movements. Due to the low soil CBR and high settlement rate,
the client is prepared to accept relevelling of the pavement after such 40000
vehicle movements.
The vehicle has an axle load of 400 kN spread equally between two wheels
1.5 m apart. The soil has an 8% CBR, is 5 m thick and is neither completely
saturated nor completely dry. The pavement design engineer has to determine
the pavement design.
The client also wishes to know the point at which the raft pavement units
will require replacement (the ultimate limit state) if the vehicle load were
replaced by the same load due to container stacking, but with no relevelling
taking place. The maximum container contact pressure would be 12 MPa.
Table 3.3 gives the stages of the pavement design. Table 3.4 gives the
computer input and output.

Stage 1 The required data input in Table 3.3 is the subgrade CBR and the
number of vehicle movements that the raft-type pavement will be required to
sustain up to its serviceability limit state. The output lists the maximum raft
design stress number (DPS) as 120.99 and the maximum design subgrade
stress number (DSS) as 158.39. For the raft-type pavement units and the soil to
satisfy the client loading requirements neither DPS nor the DSS numbers must
be exceeded.

Stage 2 Using the standard 2 m square raft pavement unit as described in


Table 3.1, together with the details of the 400 kN axle load spread equally

Table 3.3 The seven stages of the raft-type pavement design example

Stage no. Data input Data output

Vehicle movements 105000 DPS=120.99


Relevelling after 40000 DSS= 158.39
Subgrade CBR 8%
2 Concrete 50 MPa APS-DPS=448.41-120.99=327.42
Total load 400 kN ASS- DSS = 83.92-158.39= -74.47
Two wheels 1.5 m apart
3 Subgrade CBR 8% APS-DPS=383.84-120.99=262.85
Subgrade 5 m deep ASS-DSS=204.09-158.39=45.70
4 GSB40% CBR APS-DPS = 308.51-120.99= 187.5~
GSB 600 mm thick ASS- DSS= 161.96-158.39=3.57
5 Raft 200 mm thick APS-DPS= 119.81-120.99= - 1.18
Raft steel I81Omm 2 ASS - DSS = 155.58 - 158.39 = - 2.81
Final values APS= 119.81 ASS = 155.58
6 Pavement maintenance required after 109 196 vehicle movements
Relevelling required after 42971 vehicle movements
7 Replace pavement units after 778199 load applications
THE DESIGN OF PRECAST CONCRETE RAFT-TYPE PAVEMENTS 49
Table 3.4 Printout of design example data input and data output

Number of vehicle movements before


(a) maintenance required 105000
(b) relevelling required 40000
Wheel load 400kN
Wheel configuration Two wheels l.5m apart
Subgrade eBR 8%
Subgrade thickness 5m
Subgrade saturation Normal
Granular subbase eBR 40%
Granular subbase thickness 600mm
Raft unit thickness 200mm
Raft unit steel reinforcing 181Omm 2 each way
Raft unit concrete strength 50MPa
Achieved number of vehicle movements before
(a) maintenance required 109196
(b) relevelling required 42971
(c) raft replacement required 778199

between wheels 1.5 m apart, the program calculates the loading stresses on the
raft and on the subgrade. The axle configuration that produces the maximum
stress in the raft does not usually produce the maximum subgrade stress. The
program determines the difference between the applied raft stress number
(APS) of 448.41, the applied soil stress number (ASS) of 83.92, both due to
Stage 2 and the DPS and the DSS numbers respectively. Table 3.3 shows that
the APS number is too high by 327.42, but that the ASS number is 74.47 below
the required value, which indicates that the normal subgrade strength is
adequate.

Stage 3 The data input now requires the subgrade CBR of 8%, depth of 5 m
and the degree of saturation. The program determines the difference between
the applied raft stress number (APS) of 383.84, the applied soil stress number
(ASS) of 204.09, both due to Stages 2 and 3 and the DPS and the DSS numbers
respectively. The table shows that the APS number is too high by 262.85 and
the ASS number is too high by 45.70, indicating that the actual subgrade
strength is not adequate.

Stage 4 The data input requires details of the granular subbase (GSB). The
GSB has a CBR of 40% and a thickness of 600 mm. The program again
determines the difference between the applied raft stress number (APS) of
308.51, the applied soil stress number (ASS) of 161.96 due to Stages 2, 3 and 4
and the DPS and the DSS numbers respectively. The table shows that the APS
number is still too high by 187.52 and that the ASS number is 3.57 too high.
50 PRECAST CONCRETE RAFT UNITS

Stage 5 By using a 200mm thick raft-type pavement unit with 181Omm 2 of


steel reinforcement each way in the top and in the bottom, the program
calculates the further reduction in the APS and ASS. The final values of 119.81
for the APS and 155.58 for the ASS are printed.

Stage 6 The program calculates the number of vehicle movements the rafts
can sustain before raft maintenance is required as 109196. A similar
calculation is made for the subgrade which is found to be able to sustain 42971
vehicle movements before settlement will require relevelling of the rafts. Thus
the client's serviceability limit state design requirements have been satisfied.

Stage 7 The program takes the number of vehicle movements the rafts can
sustain before raft maintenance is required of 109196 and, using the contact
pressure of 12 MPa related to Equation (3.2) and Figure 3.1, gives an ultimate
limit state of 778 199 load applications.
Table 3.4 lists the program printout of the input and output for the above
design example.

3.8 Conclusions
In this chapter a design method for 2 m square precast concrete raft-type
pavements has been described. The method has been successfully used in the
design of heavy duty and industrial pavements since 1986. This 'on-site' work
has resulted in the validation of the design method. Further research is
proceeding into the use of alternative shapes and sizes of raft-type pavement
units and details of this work may be obtained from the author of this chapter.

Acknowledgement
This research work has been supported by the Science and Engineering Research Council (SERC),
Grant No. GR/B 86101 and Redland Aggregates Ltd, Redland Precast Division, Barrow-upon-
Soar, Loughborough, Leics., UK.
4 The behaviour of precast
concrete raft pavements under
fatigue loading with special
reference to their use in port
areas
H. AL-KHALID

List of symbols
D = pavement damage ai = tensile stress at the bottom of a
W = wheel load slab due to interior loading
p = tyre or contact pressure ae = tensile stress at the bottom of a
ft = ultimate tensile strength of slab due to corner loading
concrete (Jc = tensile stress at the top of a

E5 = secant modulus of elasticity slab due to corner loading


F = percentage of fatigue life h = slab thickness
ace = compressive stress of concrete r 1 = radius of area of contact be-
under pulsating load tween load and slab
feee = static compressive stress of r 2 = radius of equivalent distri-
concrete bution of pressure
N = number of cycles to failure E = modulus of elasticity of
nr = number of cycles applied at a concrete
particular stress condition y = Poisson's ratio
N r = number of cycles which will /5 = radius of relative stiffness
cause fatigue failure at that C 1 , C 2 = correction factors to allow for
same stress condition redistribution of subgrade
a = stress due to a load reactions
MR = modulus ofrupture of concrete

4.1 Introduction
In the early 1970s, considerable growth in container trade occurred. Many
ports that handled containers suffered substantial pavement problems, as their
pavement design was often based on extrapolation from empirical procedures
applicable to highway pavements. It was then determined that port pavements
52 PRECAST CONCRETE RAFT UNITS

should be treated as a separate entity from conventional paving design and


construction.
With the realization that port pavements needed specific research the
development of analytical methods which took into account the heavy wheel
loads and site conditions normally found in ports was started. The analytical
procedures incorporate the development of a layered system structural
analysis and materials research. The procedures predict specific modes of
distress resulting from traffic and environmentally related causes, with the
structure being designed to ensure that critical values of stress will not be
exceeded.
Modern container terminals require extensive areas of paving for stacking
and moving containers. The provision of such areas often proves to be one of
the most expensive capital cost items that a port has to bear. The cost of
terminal surfacing works may be as high as 25% of the total investment of a
container terminal [1].
Several types of pavement can be used where heavy moving and stationary
loads are expected. Among these are bituminous, in situ concrete, concrete
block and precast concrete raft pavements.
Precast concrete raft pavements have the advantage of a hard concrete
surface, ideal in heavily loaded areas. The necessary flexibility is achieved as
the precast raft can be removed and relaid if settlement occurs in the
underlayers. The rafts are generally 2 m square with a protective steel angle
around the top edge to prevent the concrete spalling under stress con-
centrations due to impact loading. Raft units are laid on compacted sand to
give a uniform bedding. The subbase is granular and must be free draining to
prevent saturation and subsequent development of pumping after heavy rain.
Precast concrete pavements have a number of advantages:

• good quality control in manufacturing


• full strength achieved in off-site curing
• little plant needed for laying
• immediate trafficking
• they can be easily lifted and relaid to accommodate site settlement.
However, the disadvantages include:

• high cost of the raft which is aggravated by haulage as a typical raft unit
weighs 1.25 t
• if the units are larger than the track width of the handling equipment or
the distance between the corner castings supporting the containers, large
hogging bending moments can be induced in the rafts
• differential settlement between rafts must be controlled, as excessive steps
may be dangerous for moving handling equipment and may also cause
problems for surface drainage.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 53
The following sections deal with the loadings that precast concrete rafts are
subjected to when used for port pavements, the behaviour of concrete under
fatigue loading and the pavement models that have been developed to
represent such behaviour. In addition, the methods of analysis used for
pavements are reviewed and particular emphasis is placed on the use of the
finite element method in the analysis and design of precast concrete raft
pavements.

4.2 Port loading


A port is a dynamic and constantly changing entity. Changes are necessary to
achieve or maintain efficient operation and to ensure that any particular port
remains competitive. In the past two decades, for instance, with the change
from bulk to containerized cargo, many ports operations have had to conform
to meet the needs of their clients.
Containerized cargo has been accompanied by the use of fast mechanical
handling equipment and large open parking areas for the specialized container
handling trailer systems. As a result the maintenance of the large areas of
paving necessary for parking and storage has become an important aspect of a
port's operation.
These pavements are subjected to large numbers of relatively high loads
with the speeds being predominantly low and the riding quality being of
secondary importance. An added factor is the construction of ports on
reclaimed land which is prone to significant settlements.
This section reviews the types of loading experienced by port pavements,
considers their evaluation and examines the use of these evaluations for design.
The loading on port pavements can be divided into two types:
1. Dynamic loading due to vehicular movements
2. Static loading due to the load transferred by stacked containers,
materials and some types of vehicle.

4.2.1 Dynamic loading


Experience with highway pavements indicates that for vehicular loading the
combination of frequency and magnitude of the wheel load to give the
maximum effect is the critical factor for design, this is also the case for port
pavements.
There are essentially four types of vehicle being used extensively in ports:

1. Straddle carriers (Figures 4.1 and 4.2)


2. Front-lift trucks (Figure 4.4)
3. Sideloaders (Figure 4.6)
4. Tractor/trailer units (Figure 4.7).
54 PRECAST CONCRETE RAFT UNITS

Rail-mounted cranes or pneumatic-tyred cranes are also widely used but these
travel on specially constructed beams or well-defined strips which are specially
designed.

4.2. t. t Range andfrequency ofloading The axle or wheel load of container-


carrying vehicles is obtained by simple statics and the frequency of specific
wheel loads is related to the distribution of gross weights of containers
handled. The distribution of container weight will depend on the country
being considered and the types of goods handled by the port. Also, although
containers tend to be either 6.1 m or t 2.2 m long, or given by 20 ft equivalent
(TEU) units, the different types (aluminium, refrigerated, half height, liquid)
still introduce difficulties in the estimation of weight distributions [2].
In the UK Barber and Knapton have produced a gross weight distribution
based on surveys carried out for the period 1967-76 [2]. Further work by
these authors which is presented in Table 4.1 has resulted in distributions for
different percentages of 6. t m to t 2.2 m long containers [3]. Table 4.1 is used
in the British Ports Association Design Manual [4] and its updated version
published in t 988 by the British Ports Federation.

4.2.1.2 Damage to pavements Pavement damage is an arbitrary quantity


related to the stresses and strains throughout the pavement structure. The
relationship between pavement damage and wheel load is assumed to follow a
fourth power law, although estimates of the power vary from 3.5 to 6.0 [5].
In port pavements the loads are considerably larger than highway loads
and it is impracticable to extrapolate highway design recommendations.
Hence the following relationship has been used [4]:
(4.1)
where D is the pavement damage, W is the wheel load, and p is the contact or
tyre pressure.
A standard unit called the Port Area Wheel Load (PAWL) has been used
[2]. This is defined as a wheel load of 12.0 t and a contact pressure of
0.8 Njmm 2 . Comparison of the PAWL with the standard axle for highway
design defined as an 80.0 kN axle load and a tyre pressure of 0.5 Njmm 2
indicates that one PAWL is equivalent to approximately 100 standard axle
units.
The contact pressure is normally assumed equal to the tyre pressure and the
'area of contact' is assumed to be circular. This is not entirely accurate for some
of the larger machines which have heavy treads designed to penetrate the
surface, a factor particularly important when considering the stresses applied
to the pavement surfacing. However, at the subbasejsubgrade interface the
effects are dissipated thus reducing the differences.
In general, if the equation for damage is written as
(4.2)
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 55
Table 4.1 Distribution of container weight by gross weight

Container Proportion of 40 ftf20 ft containers


weight
(kg) 100/0 60/40 50/50 40/60 0/100

0 0.00 0.00 0.00 0.00 0.00


1000 0.00 0.00 0.00 0.00 0.00
2000 0.00 0.18 0.23 0.28 0.46
3000 0.00 0.60 0.74 0.89 1.49
4000 0.18 1.29 1.57 1.84 2.95
5000 0.53 1.90 2.25 2.59 3.96
6000 0.98 2.17 2.46 2.76 3.94
7000 1.37 2.41 2.67 2.93 3.97
8000 2.60 3.05 3.16 3.27 3.72
9000 2.82 3.05 3.11 3.17 3.41
10000 3.30 3.44 3.48 3.52 3.66
11 000 4.43 4.28 4.24 4.20 4.04
12000 5.73 5.24 5.12 4.99 4.50
13000 5.12 4.83 4.76 4.69 4.41
14000 5.85 5.38 5.26 5.14 4.67
15000 4.78 5.12 5.21 5.29 5.63
16000 5.22 5.58 5.67 5.76 6.13
17000 5.45 5.75 5.83 5.91 6.21
18000 5.55 5.91 6.00 6.10 6.46
19000 6.08 6.68 6.83 6.98 7.58
20000 7.67 8.28 8.43 8.58 9.19
21000 10.40 8.93 8.56 8.19 6.72
22000 9.95 7.60 7.02 6.43 4.08
23000 5.53 4.31 4.00 3.69 2.47
24000 2.75 1.75 1.50 1.25 0.24
25000 0.95 0.63 0.55 0.47 0.15
26000 0.67 0.40 0.33 0.27 0.00
27000 0.72 0.43 0.36 0.29 0.00
28000 0.53 0.32 0.27 0.21 0.00
29000 0.43 0.26 0.22 0.17 0.00
30000 0.28 0.17 0.14 O.ll 0.00
31000 0.03 0.02 0.02 0.01 0.00
32000 0.03 0.02 0.02 0.01 0.00
33000 0.00 0.00 0.00 0.00 0.00
34000 0.05 0.03 0.02 0.02 0.00

where k is a function of tyre or contact pressure, and if the empirical


probability density function (pdt) of wheel load is peW), then the average
damage per pass E(D) has been given as [5]
n
E(D) = k L Wf X P(Wi)~W
i= 1
(4.3)

where ~ W is the step length in Wand n is the number of classes in the pdf. The
total expected damage D T for N wheel passages and m wheels on the vehicle is
m n

DT = kN L L Wji
j= 1 i = 1
X P(Wji)~W (4.4)
56 PRECAST CONCRETE RAFT UNITS

In practice the damage is estimated by comparison against the standard wheel


load (PAWL).

4.2.1.3 Specific wheel loads


Straddle carriers The straddle carrier is a specialized piece of equipment
designed to lift and transport a container to a desired location. Many types of
straddle carrier are currently available, varying from the low profile models to
gantry cranes spanning two or three rows of stacked containers, as can be seen
in Figures 4.1 and 4.2.
Early models were based on a variable height portal-lift structure with the

Figure 4.1 Straddle carrier.

Figure 4.2 Gantry crane.


THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 57
'0r------.-----,-----.,...----,
- - Karricon 30'3
- - Valmet '097
"
.! - - - - - Clarke VC830
"
c:
~ 30
~
cv
c:
·0
c:o
u
o
o 20
:»a.
III
oJ
~
<I:
ll.

~ 10
cv
.D
E
;)
.z

° L.-L.-_ _---l'-- - - I_ _--l:~_ _____J

° 10 20
Gross Weight of Container ITonnesl
30

Figure 4.3 Relative damage to pavement by weight of container - Straddle carrier.

engine and transmission located at one end on an intermediate fixed portal


frame. This type can travel and lift containers over a two-stack-high
arrangement. More recent models tend to have a fixed portal structure. The
method of hoist and travel varies within each of these two types.
Of particular importance to pavement loading is the number of wheels that
are used to transfer the load. Either four- or eight-wheeled versions are
available with the eight-wheeled versions generally being 15-20 t heavier.
However, at the top of the lifting scale for four-wheeled versions more damage
may be inflicted on the pavement. The relative damage to the pavement by
weight of container for a number of different models can be seen in Figure 4.3
[3].
Front-lift trucks Front-lift trucks (Figure 4.4) have lifting capacities varying
from 1 t to more than 45 t, making them very versatile. Large container
handling versions from 23 to 45 t have been developed using fixed or telescopic
spreaders to accommodate varying container lengths. Lighter versions in the
0.82-1.1 t range are widely used to stack empty containers in blocks.
Since the load acts outside the wheel base the front axle load and subsequent
damaging power can be very high. Furthermore, the variation in damaging
power is considerable corresponding to the wide spectrum of vehicle and
container weights that can be used. It is therefore essential to have specific
58 PRECAST CONCRETE RAFT UNITS

Figure 4.4 Front lift truck.

o Vehicles
analysed

~
3 ............- . . I . -......._..I.-.....L_...L.........L_...L.........L_...L.........L~
50 60 70 80 90 100
Maximum Gross Weight of Vehicle
I unladen weight lifting capacity J (Tonnes)

Figure 4.5 Average damaging power of front lift trucks by maximum gross weight.

knowledge of the type of plant expected to be operating on the pavement if


correct design values are to be chosen.
By plotting maximum gross weight unladen, plus lifting capacity, against a
logarithmic scale of the relative damage per pass in PAWLs, a straight line
relationship is obtained which can be used to estimate the damage ifthe weight
of the vehicle is known (Figure 4.5) [2l
Side loaders Side loaders (Figure 4.6) are particularly useful where other
types of container-handling equipment are unsuitable. They are capable of
stacking containers up to three high and require wider access lanes than
straddle carriers while their turning radii are smaller than laden front-lift
trucks. The load is carried evenly on two axles, thus reducing the damaging
effect compared to front-lift trucks. While lifting, the machine cannot move
due to the use of stabilizers which produce large stresses in the pavement.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 59

Figure 4.6 Sideloader lift truck.

Repeated placement of these jacks in the same position can result in damage to
the pavement surface and structure.
Tractor/trailer units There are many types of trailer systems (Figure 4.7)
currently in use in ports, the use of any particular trailer being dependent on
the load it has to carry. Trailer systems are very convenient for ro-ro operation
allowing rapid and easy ramp loading. The types of trailer used include:
• Normal road haulage trailer, maximum gross weight 27 t
• fifth-wheel trailers for on-site handling only, with carrying capacities
varying from 40 to over 100 t
• drawbar trailers with steerable front axle or both front and rear axles,
used on site only, with a maximum capacity of 120 t
• special ro-ro trailer system (MAFI), the overall length being kept to a
minimum by using a 'gooseneck' connecting the front of the trailer to the
fifth-wheel tractor unit during towing.
As the axle loads of the tractors and trailers are of the same magnitude as
heavy highway traffic, highway pavement design criteria can be used. There
may be some difficulty in estimating the average value ofthe damaging power
per pass as the flat terminal trailers are designed for general cargo, the weight
and quality of which can vary considerably. Moreover, trailers are sometimes
used to carry containers which further complicates the problem.
The results of other container-handling equipment have been used to
estimate the average value of the damaging power [2]. From the results, the
true payload, when the damaging power of that pass was equal to the average
value, was approximately 85% ofthe maximum carrying capacity for the 6.1 m
units. For the larger 12.2 m units, indications were that this was nearer 65%.
The average value for the relative damage per pass in standard axles (80 kN,
0.5 N/mm 2 ) was then obtained by applying these proportions to the trailers.
60 PRECAST CONCRETE RAFT UNITS

WI WT W2

MOBILE {UNLADEN} CRANES

Figure 4.7 Tractor and trailer system.

4.2.1.4 The effect of channelization An important consideration in the


evaluation of dynamic loading is lane channelization. On highways where
restrictions are imposed due to lane marking, channelling is significant and
each axle can be considered as one load application. In the case of ports,
channelization is rare except in specifically designated access areas such as
container stacking areas. Even in this case, care is required as the layout of
storage and manoeuvring areas can be changed [5].
The distribution of wheels across a pavement can be defined by the standard
deviation. Figure 4.8 shows the transition from a very narrow lane, where
confined rutting would occur, to a wide pavement where the maximum passes
occur at the centrehne.
Two factors were considered in finding the actual distribution across a
pavement [2,3]:
• the standard deviation defined by the effective lane width
• the width of pavement and its position relative to the total width
available.
By varying the effective lane width and calculating the number of movements
over a chosen element in the pavement, the curves shown in Figure 4.9 were
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 61
track width (t)

E
<Il
E very narrow lane
<Il
>
'"
Cl

21 1.51 0.51 o 0.51 1.51 21

Dislance from cenlre of lane

Figure 4.8 Effect of increased lane width on pavement damage.

produced for a straddle carrier of width 4.0 m. The figure shows the curves
obtained for two points on the pavement, namely the centreline ofthe lane and
at a distance of half the track width from the centreline. It can be seen that the
maximum effect changes from under the wheel to the lane's centreline when the
lane width is just under three times the track width. A design curve for wide
pavements is also shown, giving the proposed reduction factors. The lane
width is given in terms of the track of the vehicle (t). Indications are that the
reduction factors vary from 1.0 for narrow lanes to 0.2 for very wide lanes. In
any port there are a number of different lane widths being used and the values
chosen will depend on the layout of the handling area being considered.

4.2.1.5 Braking. accelerating. cornering and bouncing Sudden changes in


motion can impose loads on the pavement which should be accounted for in
design. The damage caused by accelerating, braking, cornering and surface
62 PRECAST CONCRETE RAFT UNITS

Percenlage 01 lotal loads


occurring within 2 m wide Irack width
elemenl of pavement
-4 .....1
t t
22
.f-,I..-+

80
I.
~
I
ELW

ELW<31
1 = 1.0

60
Standard
devialion (m)
ELW 3t -5.5t
f = 0.50
5

40 4

20 2

o-f-----r----r---...,...-----r---,----..L.o
4 8 12 16 20 24 metres
(I) (2t) (3tl (41) (5!l (61)

Effective lane width (ELW)

Figure 4.9 Reduction factors for wide pavements.

unevenness are taken into account in the BPA design manual by a number of
modification factors, as shown in Table 4.2 [4]. Modification factors are also
given for proximity effects where two or more wheels are close together. These
factors are applied to the wheel load, W.

4.2.2 Static loading


In general, pavements designed to withstand the repetitive wheel loads are
capable of carrying the static loads. However, areas in which high contact
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 63
Table 4.2 Dynamic load factors'

fD fD
A. Braking c. Accelerations
Front-lift truck 1.3 Front-lift truck 1.1
Straddle carrier 1.5 Straddle carrier 1.1
Side-lift truck 1.2 Side-lift truck 1.1
Tractor and trailer 1.1 Tractor and trailer 1.1
B. Cornering D. Uneven surface
Front-lift truck 1.4 Front-lift truck 1.2
Straddle carrier 1.6 Straddle carrier 1.2
Side-lift truck 1.3 Side-lift truck 1.2
Tractor and trailer 1.3 Tractor and trailer 1.2

• Where two or three of these conditions apply simultaneously, the io factor should be multiplied
together.

stresses are expected also need consideration. In hotter climates than the UK,
where flexible pavements are used, the designer must consider the stiffness or
elastic modulus of bituminous mixes as a function of both temperature and
time of loading. Precast concrete rafts are relatively less affected by high
concrete stress due to static loads than conventional flexible pavements. The
principal loads to be considered are:
• container corner pads
• road trailer dolly wheels
• stabilizers of sideloaders.

4.2.2.1 Container stacking [n a container park, containers may be stacked


in rows or blocks of three to five or even six containers high. The load of the
container is transferred to the pavement by its corner castings which protrude
12.5 mm below the bottom of the unit and measure 178 by 162 mm in plan
producing a high concentration of stress. Since all the containers in a stack are
unlikely to be full the maximum gross weights may be reduced as indicated in
Table 4.3, according to the BPA design manual [4].

Table 4.3 Pavement loads from container stacking (after the BPA
Manual 1982)

Load on pavement (kg) for each


Reduction Contact stacking arrangement
Stacking in gross stress
height weight (MPa) Singly Rows Blocks

1 0 2.59 7620 15240 30480


2 10% 4.67 13 720 27430 54860
3 20% 6.23 18290 36580 73150
4 30% 7.27 21340 42670 85340
5 40% 7.78 22860 45720 91440
64 PRECAST CONCRETE RAFT UNITS

4.2.2.2 Dolly wheels oftrailers The front of parked trailers is supported by


two posts, each having 225 mm diameter by 88 mm wide steel wheels. If the
wheels do not penetrate the surface, a contact length of approximately 10mm
gives an applied stress of the order of 40 N/mm 2 . Some trailers have pivoted
plates usually 150 mm by 225 mm, with a turned up lap at each end, instead of
wheels, which reduces the pressure to about 2.0 N/mm 2 for the same load.
The special MAFl trailers have no front axle and when parked rest on a bar
made up of a 130 mm by 60 mm by 2145 mm long inverted steel channel
section with a wooden insert, the stresses applied are even smaller in this
case.

4.2.2.3 Stabilizing jacks When lifting or placing a loaded container,


sideloaders must be stabilized by two or four jacks. These carry 70-95% ofthe
combined vehicle and container weight when the mast is fully extended.
Although the overall loads carried are large the applied stresses vary between
0.4 and 1.3 N/mm 2 , the higher values occurring on the smaller machines where
only two jacks are provided.
The stresses involved are low, but some pavements have been damaged [3].
This may be aggravated by the jacks being placed in exactly the same position,
on the pavement thus promoting deterioration.

4.3 Fatigue behaviour of concrete


Fatigue is the process of progressive permanent change occurring in a material
which is subject to time fluctuating stress and strain. This results in the
development of cracks or complete fracture after a sufficient number of load
repetitions.
Considerable investigation has been made into the behaviour of concrete
under repeated loading. While much of the early work has been inspired by
failures in highway pavements, more recent work has been directed to
satisfying the needs of structural engineers designing bridges and other types
of structures subjected to repeated loading. This means that although most of
the properties of concrete under repeated loading can be reasonably assessed,
these properties will not necessarily be directly applicable to concrete
pavement rafts where the behaviour of a slab and the supporting subgrade
must be considered.
This section briefly reviews the present state of knowledge of concrete
behaviour under repeated loading, as work in this area forms the basis of a
number of fatigue models used in concrete pavement design.
Concrete behaves differently depending on whether it is reinforced or not;
hence, for discussion purposes the following categories will be considered:

• Plain concrete
• Reinforced concrete
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 65
Prestressed concrete is a further category; however, as precast concrete raft
units are not usually prestressed, this is not dealt with here.

4.3.1 Plain concrete


Extensive investigations have been carried out to identify and quantify the
behaviour of plain concrete under fatigue loading from as early as 1903 [8].
However, the limited nature of each investigation, the different methods of
testing and the forms of test specimen used, have made it difficult to correlate
the results of the investigations [7]. Thus, in many cases only general
correlations can be made.
As for static loading many factors have been found to influence the
behaviour of concrete under cyclic loading. Some of these are:
1. Age and strength of concrete
2. Rest periods
3. Rate of load application
4. Range of applied stress or sequence of load

4.3.1.1 Age and strength Early investigations on concrete in flexure and


compression under repeated loading indicated a fatigue limit of approximately
55% of the ultimate static strength [8,9, to, 11]. In 1903 tests were carried out
on cement cubes in compression at a frequency of 4 cycles/min while in 1923
flexural fatigue tests on 152mm by 152mm by 914mm specimens at
40 cycles/min cantilevered from a central support were carried out [8, 10]. The
main aim was to duplicate actual conditions occurring at the corner of
pavement slabs. Further tests were carried out on plain concrete and mortar
beams, 102 mm by 102 mm by 762 mm [11, 12]. Stresses above the fatigue limit
caused progressive deformation and final failure. In the tests carried out in
1924, which used cantilever beams and allowed stress reversal, age of the
specimen was an added factor [12]. The tests indicated that an endurance limit
of 54-55% It could be expected for mortar specimens over six months old and
50-55% It for specimens four months old. However, an endurance limit could
not be established between 40 and 60% It for 28-day specimens. Later
experiments with lightweight concrete found no 'fatigue limit' but established
fatigue strengths of 40-50% of the static ultimate strength at one million cycles
[13]. A modulus of rupture similar to the Purdue tests was used and complete
reversals of stress allowed. The age of the specimens varied from 23 to 84 days.
Further tests investigating the effect of speed oftesting on the flexural fatigue
strength indicated that no fatigue limit existed up to 10 million cycles [14].
Medium, 25 N/mm 2 , to high strength, 30 N/mm 2 , concrete was used. The
fatigue strength established for all beams tested was approximately 64% of the
ultimate static strength. Further work established that plain concrete exhibits
no fatigue limit at least up to 10 million cycles and concluded that the repeated
load which plain concrete may sustain for a finite number of repetitions
66 PRECAST CONCRETE RAFT UNITS

without failure is a critical percentage of the static ultimate flexural strength


[14]. Most specimens were aged a minimum of90 days so that concrete would
not continue to gain strength during the tests. The compressive strength was
approximately 30 N/mm 2 .
Concrete prisms tested in axial compression were investigated to determine
the effect of static strength [15]. Concrete grades of 40 and 60 N/mm 2 were
used and the fatigue strength at one million cycles varied between 66 and 71 %
of the static strength. Lower percentages were noticed for the higher grades
and for the smaller maximum aggregate size. The elastic strain was not
sensibly affected by the strength ofthe concrete or by the size of aggregate, but
the remaining strain for a given fluctuating stress increased for the lower
strength.
Around 1969 attention started to be paid to the nature offatigue damage, in
particular the type of cracking that developed. Consideration was given to the
unidirectional compression stresses and it was concluded that under fatigue
loading a more extensive system of cracks developed than observed in static
tests [16]. Reference [16] also suggested that progressive changes in the
modulus ofelasticity and in the ultrasonic pulse velocity indicated the damage
caused by repeated compressive loading. Moreover, it attributed the marginal
increase in the static strength observed during the fatigue test to the
temperature rise due to the energy dissipation and the loss of gel moisture.
Reference [16] also stated that under fatigue loading at low stresses, the strain
which was recovered on removal of the load, i.e. the elastic strain, increases
continuously with increasing number of cycles.
An equation based on the observations to predict the remaining life of
concrete was formulated as follows:
F = 299 - 2.99 E s (4.5)
where F is the percentage of fatigue life and Es is the secant modulus of
elasticity.
Further investigation of the mode of cracking was carried out using
specimens subjected to cyclic and sustained loading [17]. Microscopic
observation was used which found progressive microcracking similar to the
cracking process observed in static tests. Large increases in both lateral and
longitudinal strains and decreases in pulse velocity were reported at approach-
ing failure. It was suggested that two stages ofcrack propagation occurred:
1. Both the lateral and volumetric strains increase at increasing rates until
failure. For stresses less than 70% of the ultimate, when failure did not
occur, crack propagation was relatively slow.
2. Crack growth associated only with failure and characterized by a faster
growth rate of microcracks, a sharp increase in volume dilation and rapid
changes in ultrasonic measurement.
It was also observed that when a specimen was loaded to failure, before it
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 67
reached the second stage, its strength could be up to 15% larger than that of an
unloaded companion specimen. Hence it was hypothesized that during
sustained loading a consolidation of the paste occurs in which strengthening
may offset the damaging effect of the load.
In 1977, the applicability of the Palmgren-Miner partial damage hypo-
thesis to the fatigue of concrete was studied [18]. Ninety concrete cylinders
were tested at high frequencies (150-200 Hz) and the results plotted and
compared with the equation given below by lakobson in the Wohler diagram
[18]:

G';emax = 1.0 - 0.0634(1.0 - R)loglO N (4.6)


J ccc

where G'eemax is the highest compressive stress in the concrete under pulsating
load, Jeee is the static strength of concrete, R = G'eemin/G'eemax and N is the
number of cycles to fatigue failure.
lakobson's regression line was in agreement with the above equation and
showed satisfactory correlation with previous fatigue tests. The results did not
contradict the Palmgren-Miner partial damage hypothesis.
Further prediction work was put forward in 1979, when an equation for the
determination of the fatigue strength of plain, ordinary and lightweight
concrete when subjected to compressive stresses was proposed [19]:

(.J':"') =
J~ 1.0685(1.0 - R)loglON. (4.7)

Equation (4.7) stipulated that the Wohler diagram for concrete must be drawn
for constant values of R. Fatigue properties in compression were compared for
ordinary concrete of 2250 kg/m 3 density and lightweight concretes of densities
ranging down to 1500 kg/m 3 . Uncertainties in the fatigue test began to appear
for:

(J';ax)
J~ ~0.8. (4.8)

This was attributed to difficulties in exactly determining the static compressive


strength of the individual test specimens and to incipient creep effects. The
work was further expanded to determine the tensile strength of plain concrete,
and showed that Equation (4.7) was also valid for concrete in fatigue under
tensile stresses [20].
Flexural fatigue tests were carried out in 1973 on small beams to investigate
the effects of moisture, age and rate of loading on plain concrete [21]. The
results indicated that the fatigue performance was strongly dependent on the
age and moisture state of the concrete. It was suggested that a reasonably good
prediction oflong-term performance might be obtained by extrapolation from
the results of short-term tests under appropriate conditions.
68 PRECAST CONCRETE RAFT UNITS

Predictive work was also carried out in 1964, when the strength and
deformation characteristics of concrete subjected to high compressive stresses
in the range of low cycle fatigue was investigated [22]. It was concluded that
the behaviour of concrete subjected to repeatedly applied high stresses could
be predicted using the static stress-strain relationship, as a unique failure
envelope. For a given maximum stress, the specimen failed as soon as its strain
reached the value given by the descending portion of the static stress-strain
curve at that stress. A load cycle caused incremental strains only if a critical
stress corresponding to the intersection of the loading and unloading stress-
strain curve of the fatigue curve, was exceeded. This point was referred to as the
shakedown limit. For stresses below this point the strains followed a closed
loop without further development of permanent strains. It was shown also that
the shakedown stress increased with a reduction of the stress range. Analytical
relations were developed for the envelope curve and for the unloading and
reloading stress-strain curve.

4.3.1.2 Effect of rest periods As early as 1898, work was carried out on a
number of repeated load tests on tension briquets which suggested that rest
periods permitted recovery from fatigue effects [23]. Work on concrete beams
in flexure also investigated beams subjected to different stress histories [10]
when it was noted that beams which had been subjected to previous stress
histories could resist a greater number of load applications at an increased
intensity as long as the first stages of loading were below some critical value.
Further investigation was carried out in this area in 1924 in which a number of
cantilever beams were subjected to repeated loads, permitting reversal of stress
in flexure and allowing rest periods overnight and during weekends. It was
found that rest periods had only a temporary effect but it was noticed that one
specimen recovered almost completely after five weeks of rest caused by an
accidental shutdown. Stressing the concrete below its endurance limit
increased its strength.
A somewhat more comprehensive investigation was carried out in 1960 in
which the length of rest period was varied between 1 and 27 minutes for a given
number of tests [24]. In the tests 185 beams were subjected to 4500 load cycles
in flexure and then allowed varying rest periods during which the specimen
sustained a constant load equal to the lower limit of the repeated load. The
result showed that periodic rest periods increased the fatigue life of plain
concrete subjected to flexural loading. The increase in fatigue strength was
more pronounced as the length of rest periods increased up to 5 min, after
which the rest period seemed to have little effect. This is illustrated in
Figure 4.10 where the stress level S is plotted against the duration of rest
period.
The results also showed that the fatigue strength and fatigue life of concrete
subjected to repeated loads of varying magnitude, was influenced by the
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 69
0·70
T T T I I
Vl

1/1
1/1
CII
.s:.
C,
c 0·68 - - - - -
.::
"'J ~ ~

Vl ...
CII

"tl If)
~ CII
a. 0
a.
0·66 H -
« E :;:;
E ::>

-Vi -
~
E u 0·64
'x0 c
:t:
0·62 I I I I I
0 5 10 15 20 25 30
Duration of Rest Period. mins.
Figure 4.10 Effect of rest period on Portland Cement Concrete fatigue life.

sequence in which these loads were applied. If the upper stress level in a fatigue
test was varied between two values continuously during the test, the lower
value being constant, the fatigue life decreased with increasing magnitude of
the higher stress level and also with increasing number of cycles under the
higher stress. Loads corresponding to stress levels less than that which would
normally cause failure, could contribute to the damage initiated by previously
applied higher loads. It was also observed that the behaviour ofconcrete under
varying repeated loads could be affected by variations of strength during a
fatigue test due to the relief of shrinkage stresses and cumulative micro-
cracking.

4.3.1.3 Rate ofload application The effect of rate ofload application on the
fatigue strength of concrete was investigated in 1953 [25]. In these tests, on
plain concrete, repeated loadings applied at between 70 and 440 cycles/min
were used. The main conclusion was that the loading frequency within the
range investigated had little or no effect on the fatigue strength.
High-speed tests on cylinders of plain concrete were carried out in 1958 at
speeds of 500-9000 cycles/min [26]. No effect of rate of loading on fatigue
strength was observed.
In a 1956 discussion on the then current state of knowledge on the fatigue
problem, it was concluded that the rate of load application did not
significantly affect the fatigue response [27].
Investigations into the strength and deformation characteristics of plain
concrete subjected to high repeated and sustained loading were carried out in
1972 [28]. The speed of testing was varied from 4 to 414 N/mm 2 per minute.
70 PRECAST CONCRETE RAFT UNITS

The results indicated that when concrete was subjected to high repeated
compressive stress, a decrease in either the maximum stress level or the stress
range resulted in an increase of the number of cycles to failure. The authors
noted that at a high stress, a reduction in the speed of testing resulted in a
significant reduction of the fatigue life of concrete. Under repeated loading the
failure strains of concrete increased with decreasing stress level or decreasing
range ofloading. Damage caused by high repeated loads depended both on the
number of applied cycles and the total time the concrete had to sustain high
stresses.
An investigation was carried out into the effect of rate of application of
steadily increasing loads upon the fatigue strength of plain concrete in
compression [29]. It was concluded that the fatigue strength of each of the
three types of concrete made with gravel limestone or lytag aggregate was
enhanced by increases in the rate of loading. Accordingly accelerated fatigue
testing ofconcrete structures may produce an overestimate of their true fatigue
life. Sensitivity of the static strength to the rate of loading appeared to be
related to the stiffness of the aggregate relative to the matrix.
Work in 1973 on the flexural fatigue of small concrete beams suggested that
at least up to 20 cycles/min the rate of loading was unimportant [21].

4.3.1.4 Effect ofrange ofstress One of the earlier and more comprehensive
investigations to determine the effect of range of stress on fatigue strength was
carried out in 1959 [14], when 175 beams were tested under flexural loading
with the range being varied throughout the investigation. The variation within
each series of tests was such as to predominantly maintain a constant ratio, R,
between the minimum and maximum applied stresses. The values of R ranged
from 0.13 to 0.75. Figure 4.11 shows the curves resulting from this work. From
the figure the influence of the range of stress, measured by the ratio R, can be
seen. There is a successive decrease in the fatigue strength as R varies from 0.75
to 0.13. The investigators concluded that the fatigue strength (repeated loads

-
~
o :l
- CI>
5 1· 0
_ ,.--....,---,----..,...---r--....,--.......,
:;: - Q.
o X :l
0::20:: R" 0.75
: LL '0
U.
'-~
oc.2 1ll
08
. R"0'50
_0 'E
- :l
'tl
... 0
oz ~
0 R "0'25
U. E 0 R" 0'1.3 f
CI>
:l g -
III 0·6 00'18
~·x ~
00':::
U.~Ul

10 10 2 10 3 10 4 10 5 10 6
Fatigue Life----Cycles to Failure
Figure 4.11 Effect of range of stress on the behaviour of plain concrete under fatigue loading.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 71
sustained without failure) is a critical percentage of the static ultimate strength
and this percentage is a function of the range of stress to which the concrete is
subjected.
Further work in 1966 reported that concrete subjected to five cycles of a
maximum stress of 83-88% of the static ultimate strength showed no
subsequent effect as far as the static strength of the specimens was concerned
[30]. When the maximum stress was increased to 95-100%, seven of the eleven
specimens tested failed before reaching the fifth cycle. Thus it was concluded
that the loading capacity of the concrete decreased with increasing cycles up to
the first five cycles. Specimens which did not fail after five cycles at a maximum
of 90-100% showed either a decrease or an increase in their static strength. It
has been observed that the difference between the strain at the maximum and
minimum load levels of one cycle give an indication of the integrity of concrete.
Investigations into the effect ofcompressive stress gradient on the fatigue life
of plain concrete were carried out in 1966 [31]. The main test variables
considered were stress gradient and the maximum stress level. Concrete
composition, specimen size, frequency of loading and minimum stress level
were held constant. The main conclusions from the work were:

1. Stress gradients had a significant effect on the fatigue strength of plain


concrete in compression.
2. Concrete fatigue life was highly sensitive to small changes in maximum
stress level.
3. The current AASHO allowable concrete compressive strength of 0.40 f~
was a conservative estimate of the fatigue strength of concrete in
compressIOn.

The results of this investigation indicated that for a fatigue life of two million
cycles, a probability of failure, P = 0.00001, and a minimum top fibre stress of
0.10f~, a maximum compressive stress of 0.50f~ may be allowed at the top
fibres of prestressed flexural members as regards avoiding fatigue failure.

4.3.2 Reinforced concrete


The behaviour of reinforced concrete under cyclic loading is more complicated
than that of plain concrete. This is because rupture can be due to failure of the
steel or of the concrete. Failure in concrete can be in bond, diagonal tension or
flexure. However, the classification of the mode of failure is not always clear,
with beams appearing to have failed in diagonal fatigue probably having failed
due to a combination of bond and shear failure.
Fatigue load tests on reinforced concrete were carried out as early as 1898
[32]. However, only a few load repetitions were used. The first extension tests
into reinforced concrete beams were carried out in 1907 [9]. Fifty-nine beams
were tested at the age of one month and 23 beams at the age of one year. The
beams failed through the development of a tension crack and sometimes a
72 PRECAST CONCRETE RAFT UNITS

diagonal tension crack which seemed to appear gradually and increase in size
during the fatigue test.
Tests on both T-beams and rectangular beams were carried out in the 1930s
[33-35]. The reinforcement used a variety of different bent bars and/or
stirrups. As in the plain concrete tests attention was paid to the remaining
strains and deflections. It was noted that aged beams did not permanently
deform as rapidly as young beams. It was concluded that repeated loads below
some critical value (fatigue strength) do not affect the ultimate carrying
capacity of the beam and that cracks breathe (open and close) as long as the
elastic limit of the steel is not exceeded.
In general, the importance of creep and shrinkage was emphasized and it
was shown that they are inseparably involved in the fatigue phenomenon.
Other research on a series of eight beams reinforced with three 12 mm
diameter smooth bars tested to a limit of 11 million cycles found that the
fatigue strength was 70% of the ultimate load. The method of failure was
typically a crack progressing towards the top of the beam resulting in excessive
elongation of the steel and crushing of the concrete.
Various investigations followed where reinforced beams were designed to
fail in flexure [36,37]. It was concluded that the magnitude of repeated load
determined the failure mode. Generally, it was found that a low-magnitude
load resulted in the fatigue of the steel, i.e. flexural failure, while a high-
magnitude repeated load caused shear failure. This produced S-N (stress vs
number of cycles to failure) diagrams for diagonal cracking and failure in
shear. The fatigue strength for cracking at 10 million cycles was 57% of the
static cracking load and, for failure, 63% of the ultimate static strength.
Some work concentrated on the characteristics of bond under fatigue
loading [38,39]. Pull-out specimens were used and the conclusions were
somewhat different. On application of a repeated load of 50% of the static pull-
out strength, it was concluded that bars continued to slip with increasing
repetition ofload until failure [38]. The ratio, static pull-out strength (after 5
million cycles) to the original pull-out strength, was found to be about OS
However, using eccentric pull-out specimens, no evidence that failure of
specimens would occur by increasing the number of cycles was found, unless
the applied load was at least 80% of the ultimate bond strength [40]. There was
some difficulty in predicting the number of cycles to failure when the applied
load was near to the ultimate.
The mode offailure and interaction was further defined, by investigating the
behaviour of 60 beams under repeated loading [41]. Data were presented with
parameters for nominal shear stress, bond stress, concrete compression and
steel tension stress. The tests indicated that bond was the mode of failure most
susceptible to fatigue damage and that shear and diagonal tension failures
were likely to occur when the specimen was weak in bond. As with Kesler [36]
it was also indicated that the mode of fatigue failure depended upon the load
level, with the static mode also being important. The fatigue strength in bond
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 73
was difficult to ascertain as most specimens failed at less than 75 000 cycles.
Some work was carried out with beams subjected to simultaneous cyclic
and sustained load [29]. These tests indicated that a load with a fluctuating
and sustained component caused the deflection of the beam to increase with
continued application. A linear relationship was found when the logarithm of
the increase in deflection was plotted against the logarithm of the number of
cycles or length of time. It was also found that for all service load levels and
rates ofloading used in the tests, a beam subjected to a load with a fluctuating
component increased its deflection by about the same amount as a similar
beam subjected to a sustained load of magnitude eq ual to the maximum load
reached during the application of the fluctuating load.
The previous paragraphs show that most of the investigative work to date
has concentrated on the behaviour of beams under various modes of loading.
However, some experiments have involved reinforced slabs subjected to
repeated, concentrated loading and have indicated that slabs have great
reserves of strength [42]. The 'endurance limit' was found to be slightly over
50% of the static capacity and it was found possible to predict the fatigue
behaviour of slabs by means of the fatigue properties of the reinforcing steel.
Further work on slabs was carried out in 1974. Nine one-way spanning slabs
centrally loaded were tested [43]. The applied load induced stresses between
0.3 and 0.46 of the tensile strength of the steel fabric. References [42] and [43]
agreed and concluded that the fatigue characteristics of the slab were
controlled by the fatigue characteristics of the welded fabric. Reference [43]
also found that conservative estimates of the fatigue life value for collapse
could be determined using a similar procedure to that for determining first
wire fracture and using Miner's theory to predict cumulative damage effects.

4.3.3 Fatigue models


Many fatigue models have been developed to predict the fatigue life of concrete
pavements. In these models failure is defined either by the percentage
probability of rupture of the concrete or by the development of cracks to an
acceptable level. The main models in use today are discussed in this section
and are shown graphically in Figure 4.12.

4.3.3.1PCAjatigue model The mathematical relationship for the Portland


Cement Association (PCA) fatigue model [44] is:

log N = 11.78 - 12.110" for 0.5 < O"/MR


MR
and
10gN = 00 for O"/MR < 0.5 (4.9)
where N is the number of load applications to failure, 0" is the loading stress,
and MR is the concrete modulus of rupture. The mathematical formula
74 PRECAST CONCRETE RAFT UNITS

1· 00 ~-""'-""'-~-T""I""''T''""'''''~-~-~-'''''I

cr
.r;'
0.
c
~ 0·70
Vi
":::
III
0·60
L-

Vi
0·50

o· /,0 '--......I_......I_.....&_--I._--I._~..lo--.J._ ........_ ......


1 10' 10 2 10 3 10 4 lO S 10 6 10' 10 8
Load Applications to Failure. N
Figure 4.12 Summary of fatigue models for pavements.

describes a linear relationship for log N plotted against a/MR, and an


endurance limit of 0.5 is assumed.

4.3.3.2 Darter fatigue model Darter's model [45] was developed for use on
plain-jointed concrete pavements. He analysed the results of 140 concrete
beam fatigue tests from three separate studies. The curve fitted to the data
(Figure 4.13) was represented by the equation
17.61a
logN=17.61- MR . (4.10)

This equation is a mean regression curve in that it represents a failure


probability of 50%. To reduce the probability of failure to about 45%, the
equation becomes

10gN = 1661- 17.61a


. MR'

This was used for the design of plain-jointed concrete pavements deemed to
require 'zero maintenance'.

4.3.3.3 ARE fatigue model The ARE model [46J is based on the AASHO
road test sections which developed class 3 or 4 cracking [47]. The actual
numbers ofload applications experienced were converted to equivalent 80 kN
axle loads using AASHO equivalency factors for a terminal serviceability
index of 2.5 [48]. Representative mid-slab stresses in the direction parallel to
the traffic were obtained using elastic layer theory, and the resulting regression
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 75
1·00 ~-T"""'-"T"""'--r--...,..-""'T"---"---'--"-
__

0·90

a:

-
.c. 0·80
DI
C
Go
~ 0·70
"-
III
III KEY
~ 0·60
Vi t:. Kesler t:J. t:.
o Raithby and Galloway
0·50
o Ballinger

0·40 _ _~_~_~_-'-_-'-_-I._......l
_ _I.....-J
10 2 10 3 10 4 10 5 10' 10 7 10' 10 9
Load Applications to Failure. Nt
Figure 4.13 Darter's fatigue model.

equation was

N =23440(~R} (4.11 )

Figure 4.12 shows that the ARE model predicts a significantly longer fatigue
life than the laboratory models for high concrete stresses. The reverse is true
for applied stresses below 60% of the ultimate strength, with the ARE model
predicting a shorter fatigue life. This discrepancy is possibly due to the use of
mid-slab stresses with no consideration for voids, partial contact, curling and
other factors which tend to increase the actual stresses [49]. Reference [49]
also indicated that the most likely cause was the difference in triaxial and
uniaxial state of stress between the field and laboratory tests.

4.3.3.4 USAF distress model The US Air Force distress model, which is
used by the Army Corps of Engineers is based on observations of airfield
pavements [50]. The failure criterion in this design is the development of one
crack per slab for k-values (modulus of subgrade reaction) up to 54 kPa/mm
and two cracks per slabs for k-values in excess of this figure. Westergaard's
plate theory for edge stresses was used to compute the stresses [51].
Comparison of this model with the ARE model in Figure 4.12 shows that at
high stress levels the predicted fatigue life is considerably shorter. However, at
lower stress levels it predicts slightly higher fatigue life.
The model has been developed from a wide variety of field data but its
application to highway pavements is open to debate since the performance
data have been gathered from airport pavements with a small number (5000) of
total load applications but much greater loads.
76 PRECAST CONCRETE RAFT UNITS

4.3.3.5 Vesic distress model Vesic and Saxena analysed the AASHO road
test data but used Westergaard's plate theory to find the slab stresses [52].
However, the tensile stress caused by a load in the anticipated average wheel
path position was used instead of the mid-slab or edge stresses. By developing
a relationship relating the subgrade reaction k to the elastic layer properties of
concrete and subgrade, and using k and the number of load applications to
reach a present serviceability index (PSI) of 2.5, the formula

MR)4
N r = 225000 ( ----;;- (4.12)

was obtained.
As can be seen from Figure 4.12, the prediction of fatigue life using Vesic's
model and the ARE model, which both used the same performance data, is
different by a factor of 10-20 depending on the stress level. However, the
failure criteria used in these models are different, the ARE model using the
formation of a class 3 or 4 crack while Vesic defined failure by a PSI of 2.5.

4.3.3.6 Risc distress function The equation representing the Risc distress
function was also used to analyse the AASHO road test data [49]. The analysis
was based on plate theory resting on a multilayered elastic solid subgrade and
included effects due to actual load placement, slab geometry, load transfer
devices, and variation of material properties. The mathematical relationship is
MR)4.29
N r = 22 209 ( ----;;- (4.13)

where N r is the number of equivalent 80 kN axle loadings (equivalency by


AASHO method) required to produce a terminal serviceability index of 2.5.
Figure 4.12 shows that for high stresses the Risc function and the ARE
model predict similar lives. At lower stresses the Risc function is closer to the
USAF model giving a slightly higher fatigue life than the ARE model.

4.3.4 Port pavement model


Any model intended to represent the behaviour of precast concrete pavement
units (pepU) or rafts would depend on the failure criteria adopted. Models
used in highway pavements tend to adopt a higher level of conservatism the
more stringent the serviceability requirements. The serviceability require-
ments for highway pavements and port pavements using precast concrete raft
units are, however, quite different. On highways, where a rigid concrete
pavement structure is used, intolerable faulting (the difference in elavation of
two adjacent slabs at joints or cracks) values range from 3.0 mm to a maximum
of 6.2 mm dependent on the investigator [53]. These values are relevant where
relatively high-speed vehicles operate and where it is necessary to reduce tyre
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 77
noise to a minimum. The operations in a port, in comparison, result in low
vehicle speeds and investigators have reported that differences in elevation
between adjacent slabs of25-50 mm do not cause problems to the large wheels
of straddle carriers and other port equipment [54]. Hence indications are that
even though cracks may develop in the units, they do not have to be replaced
immediately.
None of the pavement models so far discussed has been developed from data
acquired using PCPUs subjected to heavy wheel loads, and design criteria for
this type of pavement are not readily available. As a result, although the
accuracy of any model chosen would need to be verified by experimental and
field evidence, the Vesic model using a PSI of 2.5 gives the highest number of
load repetitions at higher stresses and is therefore considered here to be the
most appropriate to represent the behaviour of the unit [54].

4.4 Structural analysis


Many, if not most of the methods of analysis so far developed relate to
highways. For highway pavements, the many modes of pavement distress
require that any model must have the capability of investigating some or all of
the following:
(1) finite slab analysis;
(2) the effects of load transfer at joints, taking into account different load
transfer devices;
(3) curling and warping;
(4) the effects of voids and partial contact;
(5) treatment of the slab as two rigid layers so that cement-treated bases can
be considered;
(6) representation of the subgrade as a three-layer elastic continuum;
(7) different traffic loading geometries; and
(8) computation of stresses and strains in the slab as well as in the layered
continuum
In the case of port pavements where precast concrete rafts are used, some of the
above features have been considered to be less important. For example, load
transfer at joints should be considered but load transfer devices such as dowel
bars, which are used in rigid pavements, are generally not used with precast
units. Also, although the normal constraints on pavement design due to
drainage, frost susceptibility and constructional requirements apply, thermal
warping stresses may be ignored because they can be considered negligible in
comparison to the traffic stresses [55].
The models used for representing concrete pavements in general are
reviewed, with particular emphasis being placed on those methods applicable
to concrete raft pavements.
78 PRECAST CONCRETE RAFT UNITS

4.4.1 Analysis models


The theories developed over the years to determine the stresses and deflections
in the concrete slabs may be divided into four major groups [49]:
1. Models of continuously supported slabs.
2. Layered systems.
3. Finite element and discrete element models.
4. Coupled models.

4.4.1.1 Models oj' continuously supported slabs Westergaard is considered


to be one of the first to treat the design of a highway pavement as a problem
requiring structural analysis [51]. The slab was assumed to be homogeneous,
have uniform properties and to be infinite in extent. The reaction of the
subgrade was assumed to be vertical only and to be proportional to the
deflection, i.e. the support is similar to that provided by a dense fluid (Winkler
foundation), and hence the subgrade has no shear strength [51].
Formulae were produced for the calculation of stresses due to a load placed
at the corner, at the edge and in the interior of the slab. Since all subgrades
possess some shear strength, the assumption regarding the subgrade contains
some error. Also the uniformity of contact between the slab and the subgrade
is frequently modified by warping ofthe slab. Non-uniform contact conditions
have been considered and the original functions for stress calculations have
been modified as follows [51, 56]
for interior loading,

(Tj = 0.275(1 + y)~ [IOglO( ~;;) - 54.54( 2YC 1


Z
]; (4.14)

for edge loading,

= 0.529(1 + 0.54y) W
ll z [ loglo (EI1
3
z
(Te kri ) + loglo ( 1 _r yZ ) - 1.0792 ] ;

(4.15)
for corner loading

(4.16)

where (Tj = tensile stress at the bottom of the slab due to interior loading,
(Te = tensile stress at the bottom of the slab due to corner loading
(Njmm Z ),
(Tc = tensile stress at the top of the slab due to corner loading (Njmm ),
Z

W = wheel load (N),


h = slab thickness (mm),
r 1 = radius of area of contact between load and slab (mm),
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 79
rz = radius of equivalent distribution of pressure (mm),
rz = r l where, r l > 0.724h,
rz = (1.6ri + hZ)O.5 - 0.675 h, where r < 0.724 h,
E = modulus of elasticity of concrete (N/mm Z),
y = Poisson's ratio for concrete,
k = modulus of subgrade reaction (N/mmz/mm),
is = radius of relative stiffness (mm),
is = [Eh 3 /{12(1- O'Z)k}] 1/4 and
C l , C z = correction factors to allow for a redistribution of subgrade
reactions.
Where depressions in the subgrade or warping of the slab leaves the corner
unsupported, it acts as a cantilever and the stress can be calculated from the
formula

(4.17)

The subgrade was considered as a semi-infinite elastic solid and using the
elastic properties of the subgrade (E and Poisson's ratio y) a mathematical
model was developed for the maximum stress and deflection in a semi-infinite
concrete slab under a single load at the interior [57,58]. Using influence
charts, the work of Westergaard was extended to include any loading
configuration. These models were based on the assumption that the slab
extended a long distance from the loaded area and are thus subject to a
number of shortcomings as follows [59]:

• the slab models are limited to pavements consisting of two layers (rigid
slab on subgrade), multiple-layered slabs cannot be analysed
• concrete slabs of finite dimension, slabs with more than one joint or
cracked slabs cannot be considered
• it is assumed that the concrete slab is continuously supported by the
subgrade, therefore the effect of voids or partial subgrade support cannot
be analysed
• slabs of non-uniform thickness, non-uniform material properties or non-
uniform subgrade support cannot be analysed.

4.4.1.2 M odefs of fa yered systems In layered systems, each pavement layer


is represented by its modulus of elasticity, its thickness and its Poisson ratio.
All layers are assumed to extend horizontally to infinity with the bottom layer
also extending vertically to infinity. The stress analysis of a single layer was
first formulated and solved in 1885 [60]. The applied load was assumed to be
concentrated and acting normal to a semi-infinite elastic and homogeneous
solid system.
The first solution for a generalized elastic multiple-layered system was
presented in 1943 [61]. This was later extended by other investigators to
80 PRECAST CONCRETE RAFT UNITS

include more generalized loading conditions and for viscoelastic multiple-


layered systems. Modern computers allow analysis of multiple-layered
systems subjected to multiple loads.
Despite the capabilities of these models they have a number of deficiences in
the analysis of rigid pavements [49]:
• layered system models are not capable of analysing pavement slabs of
finite dimensions
• various concrete pavement features such as joints, cracks, reinforcement,
voids under the slab, curling and warping and loss offoundation support
cannot be analysed.

4.4.1.3 Finite element and discrete element models The use of the discrete
element method was pioneered in 1949, when the slab was considered as an
assemblage ofdeformable hinges, rigid links and coil springs [62]. The first use
of the discrete element method for concrete pavement slabs had to wait until
1966 [63]. In the analysis the subgrade was idealized as a Winkler foundation
and the effects of joints and shrinkage cracks were taken into account by
reducing the original bending stiffness of the slab at those locations where
joints or cracks existed.
The model was later modified and improved by other investigators to
include elements of different sizes, anisotropic skew slabs and the idealization
of the subgrade as a semi-infinite elastic solid. The major disadvantages of the
discrete element formulations are that elements of varying sizes are not easily
incorporated and special treatment is needed at the free edge where stresses
cannot be uniquely determined.
The finite element method for the analysis of concrete pavements has been
used by a number of investigators. The concrete slab was modelled in a similar
manner to the discrete elements and the solution based on the principle of
minimum potential energy. Finite element models for the analysis of concrete
pavements may be grouped into the following major classes:

• slab models
• plane strain models
• prismatic models
• axisymmetric models
• three-dimensional models

Slab models consider concrete pavement slabs as medium thick plates


supported by an idealized subgrade [64-67]. These models all represent the
subgrade as a Winkler foundation.
In the model in [64J, a concrete system with joints is represented by up to six
slabs, with nine slabs being used for non-symmetric loading. Rectangular plate
elements with three degrees of freedom at each node (two translations and
one rotation) are used to represent the slab and stabilized base, or the slab and
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 81
overlay. Dowel bars are represented by beam elements, spring elements being
used to model aggregate interlock. The overall stiffness matrix is formulated
by superimposing the effects of individual element stiffnesses using topological
or element connective properties. Slabs with varying thickness OJ pavements
with tied shoulders having a thickness different from the slab can also be
analysed. Since beams and springs are used, the load transfer efficiency of
various joint designs can be determined, together with the effect ofjoint design
on critical slab stresses. The model does not consider the effects of curling,
warping, voids and partial contact.
The model used in [65] represents the concrete slab as being composed of a
single rigid layer. Two or four slabs are used where load transfer effects are
being considered and only one slab when no load transfer at joints is assumed.
Rectangular plate elements with three degrees of freedom per node are used.
Load transfer devices are not used, instead load transfer efficiency is assumed.
This type of model is restricted to slabs of uniform thickness, but does consider
the effects of curling, warping, voids and partial contact. Despite the
advantages of being capable of analysing slabs with all types of loading, slabs
of finite dimensions and different pavement features (cracks,joints, etc.), multi-
layered pavements cannot be handled.
The plain-strain idealization represents the pavement as a transverse slice of
the pavement having a unit thickness. Because of the simplifying assumptions
these models are not capable of evaluating various pavement features such as
joints, cracks, slab action of the concrete pavement slab and multiple loads.
In the prismatic idealization the pavement system is represented by a
constant two-dimensional geometric shape with respect to an infinite third
dimension. The first prismatic model was developed in 1969 and later extended
to the analysis of pavement structures [68,69]. The greatest limitation for
concrete pavements is that no variation of geometrical configuration is
allowed along the longitudinal axis of the pavement system. Therefore
prismatic models cannot handle any transverse discontinuities such as joints
or cracks, nor do they use a realistic representation of loads.
The axisymmetric model idealization is an almost general three-
dimensional system. The pavement is modelled as a multilayered cylindrical
system loaded symmetrically at the centreline. Despite the advantages in the
analysis oflayered systems, axisymmetric models are not capable ofevaluating
various concrete pavement features such as joints, cracks or various loading
conditions such as edge, corner or joint loads.
The three-dimensional models, where the actual geometrical configuration
of the entire system can be taken into consideration, are thought to be the most
desirable [54]. Many computer programs (SAP, PAFEC, ANSYS) have been
developed which employ three-dimensional finite elements and with the speed
and efficiency of modern computers, these are being used extensively.

4.4.1.4 Coupled models Coupled models can be described as structural


82 PRECAST CONCRETE RAFT UNITS

models obtained by coupling two or more of the following models:

• finite element or discrete element


• multiple-layered system
• analytical (closed form)

The formulation of coupled models has been carried out by a number of


investigators for the analysis of pavement systems [64,66,70]. Saxena [70J
coupled the discrete element slab model developed by Hudson and Matlock
[63] with the closed form solution of Boussinesq [60]. The discrete element
technique was used to idealize the concrete pavement slab and the Boussinesq
equation used to represent the behaviour of the subgrade as a semi-infinite
elastic solid. Saxena [70] also formulated a coupled model consisting of a
discrete element slab model and a three-layered elastic system, by using an
approximate solution technique [71]. The model did not consider the effects of
voids, partial contact, curling and warping and is only applicable to single
slabs with free edges. Load transfer effects were not considered.

4.4.2 Precast concrete pavement models


Many of the parameters required for the analysis of highway pavements are
not important for precast concrete pavement units. The two important criteria
for the determination of failure are the maximum horizontal tensile stress in
the concrete and the vertical subgrade stress. Early investigations into raft
unit analysis involved the use of a coupled model where a thin-plate finite
element was used to represent the slab and a Winkler foundation used to
represent the subgrade [55]. To determine the most suitable combination of
elements, comparative work was carried out based on a simply supported
plate and a plate resting directly on a Winkler foundation. The foundation was
composed of either spring elements or, in later models, a thin layer of brick
elements overlaying a rigid base. Provided that the elastic layer is thin enough
it was found that the latter model can accurately represent a Winkler
foundation. The loading was by means of a single point load or a pressure load
simulated by four equal point loads applied at the four corners of the central
element ofthe plate. The results were compared with theoretical solutions and
also Westergaard's solution. In the plate theory the stresses under the point
load are given as infinite, hence only the deflections could be compared for
point loading, pressure loadings were used to compare the stresses.
Further research has found that Westergaard's analysis gave increasingly
erroneous results for raft pavements as their side length reduced below about
8 m [54]. Hence, finite element analysis using an eight-note, three-dimensional
brick element was used to develop the numerical analysis. The element had
three orthogonal translation displacements at each node and was found to
give good results in situations where direct forces were predominant. A similar
model is used in this work.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 83
4.5 Finite element analysis
The finite element analysis described later was carried out using the finite
element package ANSYS and used the eight-node, isoparametric three-
dimensional element to represent the concrete units, the sand bedding layer,
the subbase and the subgrade. Each node of the element has three orthogonal
translational displacements.
For precast concrete pavements there are a number of parameters which
can influence the behaviour of the pavements, these include:
• the material properties of the pavement layers
• the load magnitude and number of repetitions
• tyre contact pressure
• environmental conditions.
These parameters were all considered and are discussed further below.

4.5.1 Previous work


A number of investigations on precast concrete pavements have been carried
out using the finite element package PAFEC [54,72-75]. These investigations
include:
• determination of the load position which causes maximum stress in the
concrete raft or in the subgrade
• effect of concrete raft size and placement of reinforcement
• effect of the properties of the pavement structure.

4.5.1.1 Determination ofload position Research has suggested that for large
flat slabs the free edge loading is the critical loading [76]. The slab curls
upward at the perimeter and considerably increases the tensile stress. The
corner stress is therefore critical for tensile stresses in the concrete. Critical
tensile stresses are also found at the top of the slab due to cantilever action
caused by loss of support under the corner.
Work on slabs 2.0m square found that corner loading resulted in rotation
and produces maximum stresses at the subgrade level [55]. Further research,
aided by on-site tests, indicated that the maximum concrete tensile stresses
were experienced when the load was placed at the centre of an edge. This
indicates that Westergaard's corner formula is not applicable to raft-type
pavements since it predicts a lower failure load than the edge formula.

4.5.1.2 Size ofraft Concrete raft paving units vary in size from 10.Om by
2.29 m down to OJ m square [74]. In analysing the effect of size of raft on
pavement behaviour, the concrete tensile strength was treated as critical since
it was considered that overloading of the subgrade could be corrected by
relevelling and relaying the raft unit [74]. A single 100 kN load placed at the
centre of the unit was used with the raft size being varied from 2.44 m square to
84 PRECAST CONCRETE RAFT UNITS

Table 4.4 Raft size related to load applications, concrete stress and subgrade stress (after Bull
and Luheshi 1989)

Raft size
(mm xmm) Plan Raft Concrete Subgrade
150mm area mass Load cycles stress stress
thick (m 2 ) (kg) LIB to maintenance (%) (%)

2440 x 2440 5.954 2143 1.000 291282 92.27 83.32


1824 x 2440 4.451 1602 1.3377 220727 97.95 95.42
2000 x 2000 4.000 1440 1.000 211 158 100.00 100.00
1220 x 2440 2.976 1072 2.000 164981 106.36 114.95
1300 x 2000 2.600 936 1.5385 175 233 104.77 122.08
916 x 2440 2.235 805 2.6638 81636 126.82 130.21
608 x 2440 1.484 534 4.0132 30283 162.50 149.54

0.608 m by 2.44 m. Table 4.4 shows the effect of change of raft size on the
concrete and subgrade stress. From these results a large square raft was
indicated as the most efficient. In practice the weight of the unit must also be
considered as this will have implications on the cost of construction and
transportation.

4.5.1.3 Raft reinforcement Several workers have investigated the effect of


reinforcement amount, position and orientation on the load-carrying capacity
of concrete raft units. Early workers used simply supported slabs and their
main conclusions were [77]:
• altering the angle of the reinforcement to the edge of the raft changed the
ultimate load
• spanning the reinforcing bars in accordance with the elastic bending
moment distribution would increase raft stiffness, reduce deflections,
reduce crack widths, but would only minimally reduce the amount of
reinforcing steel
• stopping-off bars would reduce the amount of reinforcing steel.
Later researchers used PAFEC with a three-dimensional brick element and
reinforcement running along the bottom or both the top and bottom of the
elements [77]. The reinforcing steel was arranged to be parallel to or at 45° to
the edges of the 2 m square raft. Using a series of computer runs with varying
areas of reinforcement, the number of applications of a 100 kN load at the
serviceability and ultimate limit states was determined. The results are
illustrated in Table 4.5.
From the results it is seen that the ultimate limit state, the point at which the
raft will need to be replaced, is over twelve times more than that of the
serviceability load repetitions. For both limit states it was found that placing
the reinforcement in both the top and bottom of the raft increases its life more
than the reinforcement in the bottom. Laboratory work was in progress at the
time to corroborate the numerical analysis.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 85
Table 4.5 Relationship between serviceability limit state, ultimate limit state and the amount of
steel reinforcement in a raft (after Bull and Luheshi 1989)

Load repetitions to Load repetitions to


serviceability limit state ultimate limit state
Area
reinforcement Reinforcement Reinforcement
(mm2)
per metre width Top and Top and
of raft Bottom bottom Bottom bottom

0 211158 211158 2565049 2565049


85 215365 218228 2616154 2650932
212 224403 230177 2725943 2796083
425 235657 249672 2862651 3032899
850 261089 292144 3171587 3548465
2125 347821 448057 4225167 5442787
4250 511 725 795415 6216197 9662330

4.5.2 Current analysis


The structural design of a pavement subjected to heavy wheel loads is
thought to be a function of the number and magnitude of the imposed wheel
loads, contact pressure, pavement material properties and environmental
conditions.
The numerical analysis carried out by the author of this chapter is an
attempt to provide design charts for raft pavements adopting the PAWL (i.e.
12 t load and a contact pressure of 0.8 MPa) as the possible means of loading
the raft surface. Design charts are developed based on variation of the material
properties of the pavement layers and design based on these charts are
compared to those developed using a 10 kN point load [54]. The finite element
program, ANSYS, allowed a maximum wavefront of 200, which restricted the
number of elements that could be used and this placed some limitations on the
analysis.

4.5.2.1 Parametric study In the analysis the effect of changes in the


following was considered:
1. Load position
2. Subgrade Young's modulus
3. Subbase Young's modulus and thickness
4. Concrete thickness.
The influence on both the concrete and subgrade stresses of the Poisson ratio
of the pavement layers, the thickness and modulus ofthe bedding sand and the
subgrade thickness beyond 1.2 m, has been investigated and found to have
predominantly little effect [54]. These parameters were therefore held
constant throughout the investigation.
A basic full finite element model, as detailed below, was used to determine
86 PRECAST CONCRETE RAFT UNITS

the stresses as the parameters were varied:


Precast concrete raft
2000 x 2000 x 150 mm thick
Cube strength 50 N/mm 2
Modulus of elasticity 34 kN/mm 2
Poisson's ratio 0.15
25 finite elements
Sand bedding layer
2000 x 2000 x 150 mm thick
Modulus of elasticity 75 N/mm 2
Poisson's ratio 0.25
25 finite elements
Subbase layer
5000 x 5000 x 300 mm thick
Modulus of elasticity 200N/mm 2
CBR value 20%
Poisson's ratio 0.25
49 finite elements

A A
L ---'
m Load position

PLAN

2
CD Precast concrete
ralt

3 - fi\Sand bedding
\V layer
I.-
Q) Sub-base layer
~ f7;\ Sub-grade
\::,) layer

SECTION A- A

Figure 4.14 Finite element model-load at centre.


THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 87
Subgrade layer
5000 x 5000 x 1200mm thick
Modulus of elasticity 3 N/mm 2
Poisson's ratio OJ
98 finite elements

The load was placed on 380 mm square area positioned at the centre, the
corner or centre of an edge dependent on the stress condition being
investigated. The subgrade CBR was varied from 2%, representing a relatively
soft clay, to 30%, a stiffclay or medium compacted granular material. Similarly
the subbase CBR was changed from 20 to 100% representing a medium- to
well-compacted materiaL Parameters used in practice were also taken into
account in varying the subbase thickness from 200 to 700 mm and the concrete
thickness from 100 to 350 mm.
Figures 4.14-4.16 show the basic model used for the three load positions.
Symmetry was used to reduce the computer storage space required when the
central load position was being analysed and also in assessing the accuracy of

~ Load position

PLAN

1 I'f\ Precast concrete


'-' raft
CD.
®- o
(;\ Sand bedding
layer
v @ Sub - base layer

? r- fi:\ Sub- grade


\.::,) layer

SECTION A-A

Figure 4.15 Finite element model-load at centre of edge.


88 PRECAST CONCRETE RAFT UNITS

L
m Load position

PLAN

Precost concrete
2 CD ralt
Sand beddi ng
3 (3) layer

4 @ Sub-base layer

Sub-grade
@ layer

SECTION A- A

Figure 4.16 Finite element model -load at corner.

the method. To determine the relative accuracy the mesh size was reduced
continuously to the limit of the computer program with the maximum stress
being checked in each case. One quarter of the model was discretized for this
aspect of the analysis and the results are shown in Figure 4.17. From this it
appears that, as the degrees of freedom are increased, the concrete and
subgrade stresses tend towards unique values.
The maximum horizontal tensile stress in the concrete and the maximum
vertical compressive stress in the subgrade were both noted. Similar to
previous investigations, the concrete tensile stress was considered the more
critical as relevelling of the PCPUs can take place when subgrade subsidence
occurs. Also the ultimate tensile strength of concrete was taken as 5.5 MPa.
This was based on laboratory work, which indicated a strength between
4.9 MPa and 6.0 MPa [54]. The value chosen allows for differences due to
manufacture of each individual unit, and is that given by other researchers
[47,52]. Values for Young's modulus (E) of the subgrade are related to the
CBR through the approximate relationship, E = 10 CBR (MPa) [78].
Stresses in the concrete and subgrade are expressed as a percentage of the
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 89
CIMPal

51kPal 0.25

12·0 0·20
_ _- - 5
o
10·0 0·15

8'0 0·10

6·0 0·05

L·O O·OOL-_.......L._ _..L...._--L_ _....I.-_.......I_ _•


o 800
Degrees of Freedom

Figure 4.17 Concrete (C) and soil (5) stresses due to a load applied centrally as a function of
degrees of freedom.

180

160
/ ----
~140
....... )1'" I~SUbgr'lCone.
~ 120
II!
1;) 100 /
.~
Q)

80
)
!Il
II!Q. 60 /
E )~
oo 40
'-
.!:!
'iii
20
cI
c:
~ 0

-20
~ . - ..
-----'
_--~
~----~ ~---~
~-----
50 100 150 200 250 300
Subgrade Modulus (MPa)
Figure 4.18 Concrete and subgrade stresses related to subgrade Young's modulus.
90 PRECAST CONCRETE RAFT UNITS

2~\~
4 \ \. ~SUbg'l
Cone.
0'1
~, \~
<
,
0'1

'".,.""
~ 6
1ii
\
Q)
> -8 ,,
'iii
0'1
,
~
,
"-("-
Q. -1 0
E
o
~-1 2 ,

"
~
.~ -1 4

""
~ "
" ~
-1 6
...... ...... _.
-1
,~
~OO 300 400 500 600 700 800 900 1C o
~I
Subbase Modulus (MPo)
Figure 4.19 Concrete and subgrade stresses related to subbase Young's modulus.

stresses in the basic model to allow the effect of the changes in the material
properties to be easily compared.

4.5.2.2 Results Figures 4.18-4.21 show the vanatlOn of concrete and


subgrade stresses with changes in material properties. In Figure 4.18 the
subgrade Young's modulus is varied and, as it increases, the vertical
compressive stress in the subgrade increases while the horizontal tensile stress
in the concrete decreases. However, the concrete and subgrade stresses both
decrease as the subbase thickness and/or modulus is increased, as shown in
Figures 4.19 and 4.20 respectively. A similar trend is evident from Figure 4.21
as the concrete thickness increases. As in earlier investigations, changes in the
concrete thickness have the most significant effect on the concrete stress [54].
Table 4.6 shows the stresses in the concrete and the subgrade in the basic
model, which have been adjusted using the factors determined by reducing the
mesh size down to 50 mm.

4.5.2.3 Design example Consider a container terminal laid on a 2% CBR


subgrade. The pavement is to be constructed using a 2 m by 2 m unreinforced
PCPU and must carry 500000 load repetitions of a 400kN wheel load.
The single PAWL placed anywhere on the raft produces a maximum
concrete tensile stress of 2.278 MPa and a maximum subgrade compressive
stress of 107 kPa (Table 4.6) in the basic model. If the concrete stress is
expressed as a percentage of the ultimate stress (5.5 MPa), PUS, then this load
produces a PUS of 41.4%.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 91
15
~
10 ~SUbgr.r-
"'~ "Cone.

III
5
..\
III
o
....III~

o
~
'(ij
III
~
Q.
E -15
-5

-10
"" \, ~,
o ~. ~
~-20
'(ii
~ -25
K ' .....
I- ........ ~,. . ~
-30
~
-3~00 250 300 350 400 450 500 550 600 650 700
Subbase Thickness (mm)
Figure 4.20 Concrete and subgrade stresses related to subbase thickness.

40

30
...
..
I ~SUbgr-l
"Cone. ---
~
..........
20
.
'.'
,

--
III 10

---r-. .. .
III

~ 0
--., ~
III "\
~ -10 ...,
.(ij
III -20
\.
~
Q.
E -30 \
",
o
~ -40
'~
(J)

~ -50
'.
c: ~.
~ -60
' .... '.
-70
'" ,,~

-8~00 120 140 160 180200220240260280300320340360


Concrete Thickness (mm)
Figure 4.21 Concrete and subgrade stresses related to PCPU thickness.
92 PRECAST CONCRETE RAFT UNITS

Table 4.6 Concrete and subgrade stresses in basic model

Maximum horizontal Maximum vertical


concrete stress subgrade stress
Load position (MPa) (kPa)

Centre of slab 1.542 18.75


Centre of edge 2.278 55.30
At corner 1.128 107.0

Using the fatigue model of reference [52J, that is

N = 225000[~RJ (4.18)

where N = 500 000, load repititions of the 400 kN (138 PUS) wheel load, then
2
CT, the allowable concrete stress, is 89.1% ofMR (4.5 N/mm ), or 89.1 PUS. The
designed raft pavement must have a PUS less than or equal to 89.1. This is
achieved as shown in Table 4.7.
Hence the design objective is achieved by using a subbase thickness of
450 mm with a CBR of 50% over the given subgrade. As an alternative the
concrete thickness could be increased leading to a similar reduction in the
concrete stress and achievement of the design objective. Further charts for
concrete strength and raft size could be produced which would permit a
relatively easy design procedure. Although the finite element analysis has been
carried out as accurately as possible given the limitations of the program, the
results cannot be considered to be accurate without experimental evidence to
substantiate them.

Table 4.7 Reduction of load PUS to design PUS

Alteration in
subgrade bearing
Variable to Alteration of pressure
be considered PUS (kN/m 2 )

Basic design load: 120kN 41.4 PUS 107.0


A. Design Load: 400 kN 138PUS 356.67
(400/120 x 41.4) (400/120 x 107)
B. Site conditions: 2% CBR (-10%) (+25%)
(Figure 4.18) -13.8PUS + 89.0
C. Increase subbase thickness to (-15%) ( -17.0%)
450 mm (Figure 4.20) -20.7PUS -60.63
D. Assume subbase CBR increased (-11%) (-8%)
to 50%, (Figure 4.19) -15.l8PUS -28.53
Adding A, B, C, and D 88.32 PUS 356.51
TIIE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 93
5

o
1-- Bull et 01
• Current Concrete stress
,.-....
~
-5 1\\
\'\
'-J

(f)
(f)

....~
\'
-10
(f) ,,
,
2Q) -15 "
------- --- _--
~

\ ....

--"'-- ...
-

'"'"'"
()

g -20
o
E -25
:::l
E
.§ -30
~
~~
-35
~
50 100 150 200 250 300
Subgrade Modulus (MPa)
Figure 4.22 Comparison of concrete stresses related to subgrade Young's modulus.

-e- Current
I _. concrete stress I

2
i'>'
\
... ...
Bull et 01

4
1'.
, ...
(f)
(f) (\ , ...
....
~ - 6
(f)

.... -8 \D .... ........


"'
---
Q)

~
()

g -1 0 1\ ---
o ~
E -1 2 \
\bi
:::l
E
.§ -1 4
~

-1 6
"\~
-1 8
~
r-- I---
200 300 400 500 600 700 800 900 1<1 ~o
Subbase Modulus (MPa)
Figure 4.23 Comparison of concrete stresses related to subbase Young's modulus.
94 PRECAST CONCRETE RAFT UNITS

5
I
I - Bull et 01
I -@o Current Concrete Stress
II
5 I~
(fl
(fl

~
0 I"
+-'
~ ............ -.........

'"
(f)
5
- --- ...
~
Q)
..........
~~~
+-'
Q)
b -1 0
c
o
() -1 5
~
E
E-
:J 20 "'- ~
"-
'x
~ -2 5
i'...
-30 """- ..... """'4
-3~00 250 300 350 400 450 500 550 600 650 700
Subbase Thickness (mm)
Figure 4.24 Comparison of concrete stresses related to subbase thickness.

4.5.2.4 Discussion The analysis used here was compared to that provided
by reference [54] who used a IOkN load placed centrally on a pepu.
Comparable levels of discretization and material properties were used in the
two analyses. Results from the two investigations are shown in Figures 4.22-
4.29 and are discussed below.
Concrete stresses Comparison ofthe changes in maximum concrete stress
as the parameters of the pavement layers are varied, is illustrated in
Figures 4.22-4.25. Generally the two studies appear to predict similar trends,
but there is a pronounced departure when the subgrade modulus is increased
from 20 to 300 MPa (Figure 4.22). This work is seen to give a somewhat greater
reduction in the concrete stresses than that of reference [54]. A similar trend
is evident as the subbase thickness and modulus are varied with closer
agreement at lower values of the parameters. Relatively good agreement is
evident as the concrete thickness is varied. These differences are probably a
result of the different load positions used, indicating that the variation of
stresses as parameters are changed is influenced by the position of the load.
This could be further complicated by the tensile forces which are introduced in
the model when the load is placed at the centre of an edge. It must also be
remembered that there are differences between the two models in the way the
concrete tensile stresses are defined. In practice the soil has no tensile strength
and the PCPU will tend to tilt upwards under corner or edge loading.
However, for central loading, particularly under a small magnitude load, these
forces may not be as important. The effect of these tensile forces appears to be
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 95
60
I
:I
50
4
\ e-- Bull et 01
Current concrete stress
........
~ 30 \
\ \
\
gJ 20
Q)
b(/) 10
I\..'
~
'l'-.
Q) 0
~
b -1 0
c
8 -2 0
I~
E -30 ~
::l (~
E -40
'xo -50
~
~ ~ ....
-60
"R ~
-7 0
:-...:: ~
-8~00 120 140 160 180200 220 240 260 280 300 320 340 360
Concrete Thickness (mm)
Figure 4.25 Comparison of concrete stresses related to concrete thickness.

less significant as the thickness of the concrete layer, which is in direct contact
with the toad, is changed.
To determine the effect of the tensile forces, further modelling could be
carried out on a trial basis. After each trial those nodes under vertical tensile
forces at the sand/concrete interface could be progressively released until a
point was reached when the slab deflected upwards and the sand/concrete
interface is in compression only. The stresses resulting could then be compared
with those as determined in this study.
Subgrade stresses Changes to the maximum subgrade stress, as the
parameters of the pavement layers are varied for the two investigations, are
illustrated in Figures 4.26-4.29. As for the concrete stress, the general
behaviour indicated by the two investigations is similar. Differences in the two
methods are greater as the subgrade and subbase parameters are changed.
Under corner loading the results indicate that although the magnitude of the
subgrade stresses may be greater than those due to central loading, the
increase in stress as the subgrade modulus is increased is less than that caused
by a central load. There also appears to be a smaller reduction in subgrade
stress as the subbase modulus is increased. A similar investigation to that
described above could be carried out to determine the effect of the parametric
changes when no vertical tensile soil forces are allowed in the model.
Further considerations The precast concrete raft units when used on site
will be subjected to a vast number of loading configurations. The use of a
distributed load in positions which are expected to cause maximum stresses is
96 PRECAST CONCRETE RAFT UNITS

400

350 H ~ Current Subgrode stress I


-- Bull et 01 I ##
##
###
##
##

,-... ##
##
~ ##
'-" 300 ##
(/)
(/) ,,#fII"
....~ 250 ,

""
If)
,
Q)
"0 200
,,
,,
~
.g
0> 1D--
150
,, V
If)
,, /
E 100
E
::J

'xo 50
,I
I
V
~

a V
-50
o 50 100 150 200 250 300
Subgrade Modulus (MPa)
Figure 4.26 Comparison of subgrade stresses related to subgrade Young's modulus.

~ II~c;rent;Ubgro~e' st;.E!ss~L-
I. -:~~ull. et
-2
\ .....
"-
0.11
,-... \
~ -4

"" ~
'---' \
(/) \
(/) -6
....... \
Q)
\
If) -8
\
~

"" "
Q)
\
] -10
...0> 1\,
.D
::J
-12 ,,
,,
If)

""
E -14
::J ,
E -16 ' ..... .....
'x ~
~ -18
- ...
... ......
-20 .........
-2~00 300 400 500 600 700 800 900 1000
Subbase Modulus (MPa)
Figure 4.27 Comparison of subgrade stresses related to subbase Young's modulus.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 97
15
)

10 ~
"~~ I ~ Current subgrode stress
'\.." ••• Bull et 01

(II
(II
~
0 '\
+'
(II ~,'.
K '.
Ql -5
"0 '.
o
......
"
0,-10 '
.J:J (~ ,
:J '.'.
..... .........
'"
(II -15
E
E-20 (~
..........
'x
~ -25

-30
"l ~
~)
-3~00 250 300 350 400 450 500 550 600 650 700
Subbase Thickness (mm)
Figure 4.28 Comparison of subgrade stresses related to subbase thickness.

12
10
... ~ Current subgrode stress I
,
8
.. ••• Bull et 01

""
~
......... 6
\'
(/)
\ \,
(/) 4
~
Ql

V;
L-
2
~ 0
1\
"\
~ -2
0>
.J:J .\
:J -4
(f) ..,~
E -6
":' ........

-
:J
E -8

. r---..
..,.."
'g -10 ",
r-- -e:>
::i: ".
-12
......
-14 '. ........
....... .
-1 ~ 00 120 140 160 180 200 220 240 260 280 300 320 340 360
Concrete Thickness (mm)
Figure 4.29 Comparison of subgrade stresses related to PCPU thickness.
98 PRECAST CONCRETE RAFT UNITS

thought to be more desirable than the use of a central load only, as the worst
conditions are considered for design. In addition, it is believed that a single
point load applied to this type of model results in inaccuracies directly under
the load and some difficulty in accurately assessing the maximum stresses may
be encountered [79]. A distributed load is applied to the model as four equal
point loads at the corners of the element and usually gives a more accurate
result. The accurate assessment of both investigations, however, requires
experimental and/or field evidence for comparison.
Other factors that could cause error in representing the raft pavement in
ports include the non-linear behaviour ofconcrete after it has cracked, the load
transfer mechanism of the pavements and the effect of multiple-wheel loads.
The finite element analysis as carried out here assumes linear behaviour
throughout, but this will not be the case as the concrete is subjected to higher
stresses. At these stresses, cracks may appear in the concrete and the concrete
stresses may then vary in a non-linear manner. This would be expected to have
an effect on the distribution of stresses and therefore cause a change in the
charts produced. Further experimental and/or site investigations would
therefore be required to provide a basis for comparison.
In service the PCPUs are thought to transfer load by interlock and this
probably results in a reduction of their deflections and subgrade stresses [54].
While these effects could be represented in a finite element model by using a
suitable element type to connect the slabs, the assessment of the method would
still be dependent on experimental evidence. Further work is required in this
area. Additionally the effect of multiple-wheel loads on the pavement structure
could be investigated. Data relating to the dimensions and loads of vehicles
used in ports was collected during this work in an effort to determine the
stresses imposed by these. However, the current limitations of the program did
not permit this analysis to be carried out, and there are now plans for it to be
dealt with in a separate research program.

4.6 Conclusions
From the literature and analytical work carried out and reported in the
preceding sections, the following remarks can be made:
1. Precast concrete raft pavements provide a hard concrete surface ideal for
heavily loaded areas and the necessary flexibility required for weak
foundations that are encountered in practice. Their main disadvantage is
their cost, which is further increased due to haulage. A typical 2.0 m by
2.0 m unit weighs 1.25 t.
2. The relationship between highway pavement damage and wheel load is
assumed to follow a fourth power law. Since the loads in port areas are
considerably higher than highway loads, it is impracticable to extrapolate
highway design recommendations, and the relationship given as equation
(4.1) has been used.
THE BEHAVIOUR OF PRECAST CONCRETE RAFT PAVEMENTS 99
3. A standard unit called the Port Area Wheel Load (PA WL), and defined as
a wheel load of 12.0 t with a tyre pressure of 0.8 MPa, has been
recommended for use in port pavement design [2l
4. In port pavements, channelization is an important consideration in the
evaluation of dynamic loading.
5. The fatigue strength of plain concrete in compression, tension or flexure
for a life of 10 million cycles has been found to be approximately 55% ofthe
ultimate static strength. Concrete seems not to possess an endurance limit
up to about 10 million cycles. The implication here is that there is no
limiting value of stress below which the fatigue life will be infinite.
6. For the effect of range of stress, it has been shown that a decrease of range
between maximum and minimum load results in increased fatigue
strength for a given number of cycles.
7. Data are not currently available to show the effect of randomly varying
loads on the fatigue behaviour of concrete.
8. The effect of rest periods and sustained loading has not been sufficiently
explored. Laboratory tests have shown that rest periods and sustained
loading between repeated load cycles tend to increase the fatigue strength
of concrete; however, if the sustained stress level is above 75% of the static
strength, then sustained loading may have detrimental effects on fatigue
life.
9. Frequency of load between 70 and 900 cycles/min has little effect on the
fatigue strength, provided the maximum stress level is less than about 75%
of the static strength.
10. Fatigue failure appears to be due to progressive internal microcracking.
Loading is also likely to cause changes in the core structure of the
hardened paste. Creep effects must also be considered as they become
more significant as the rate of loading decreases.
11. Reinforced concrete beams may fail in either the steel or the concrete. The
concrete failure may be in bond, diagonal tension or flexure.
12. The Vesic model, which allows a higher number of load repetitions at
higher stress ranges, is considered to be the most suitable to represent the
fatigue behaviour of precast concrete raft units.
13. Westergaard's analysis has been found to give increasingly erroneous
results for concrete raft pavements as their side length is reduced below
about 8m.
14. The use ofthe finite element method, and the use of the three-dimensional
brick element to represent the pavement layers, has been found to give
good results for precast concrete rafts, where direct forces are
predominant.
15. Two parameters have been identified for precast concrete raft pavement
design - namely, the concrete tensile stress and the subgrade compressive
stress. The concrete tensile stress is the more important as settlement of the
subgrade is thought to be easily corrected.
100 PRECAST CONCRETE RAFT UNITS

16. Maximum tensile stresses in the PCPU occur when the load is placed at
the centre of an edge, while placing the load at a corner causes maximum
compressive stress in the subgrade.
17. Finite element analysis indicates that a large square raft is the most
efficient as far as reduction of the concrete tensile stress is concerned.
18. Research has indicated that placing reinforcement in the top and bottom
of raft units increases the life of the unit more than if reinforcement is
placed in the bottom only.
19. The analysis carried out here and investigations by Bull indicate that, as
the parameters of the pavement layers are varied, the thickness of the
PCPU has the maximum effect on the tensile stresses induced in the raft.
20. A comparative assessment between models using concentrated point
loads and models using distributed pressure loads revealed that a more
accurate prediction of stresses can be obtained using the latter. The two
investigations indicated similar trends but predominantly predicted
different stress variations.
21. Further research is required to assess the effect of vertical tensile stresses
induced in the foundation when the raft is loaded at an edge, and to assess
the load transfer mechanism in a group of rafts under fatigue loading.

Acknowledgements
The author wishes to thank the British Port Federation and Dr John Knapton for granting
permission to use the copyright material of Tables 4.1 and 4.2 and Figures 4.1-4.9. The effort of
Mr R.A.O. Bryan, a student of the author, expended in producing essential material for the
chapter is gratefully acknowledged.
5 Performance improvement
of precast, reinforced and
prestressed concrete raft
units (beam and slab) under
impulsive loading
M. FUJII and MIYAMOTO

5.1 Introduction
In recent years, there has been an increasing use of concrete in various fields of
construction activity. Not only is concrete being used for offshore structures,
nuclear power plants and barges, but there is a growing possibility that it could
also be used for structures in space [1]. Additionally, the application of high
strength concrete and prestressed concrete is becoming increasingly common.
The study of the behaviour of structures under impulsive loading is still in its
infancy. There is a need to study quantitatively the mechanical behaviour of
those structures, especially in the stages prior to failure. Furthermore,
performance improvement indices and concepts of performance improve-
ment have to be determined in order to design impact resisting structures. The
design codes in most countries adopt an equivalent static load in representing
impact loads [2,3]. But even though it is effective up to the maximum stresses,
it would not be able to withstand the effects of excitation of the higher modes of
vibration, a change in failure mode due to propagating stress waves, scabbing
at the rear face of the impacted structure, etc., which are peculiar to structures
under impulsive loads. Therefore, there is a necessity for a dynamic approach
in designing such structures.

5.2. Load characteristics

5.2.1 Types of impulsive loads


The problem of impulsive loads can generally be divided into the two
following groups, depending upon the type of striking body and the mode of
failure of the impacted structure [4-6]:
(a) Soft impacts. Causes deformation to the striking body. Propagation of
stress waves is negligible and the failure mechanism is quite similar to
102 PRECAST CONCRETE RAFT UNITS

that of static failure. The ratio between the mass of the impacted
structure and the striking body, failure region and amount of scabbing
are the main concerns. Typical failure modes are shown in Figure 5.1a.
(b) Hard impacts. Barely any defomation forms on the striking body.
Impacting velocity is high in this case, thus complicated stress waves can
be expected to be the main cause offailure. The shape and dimensions of
the striking body together with the perforation depth are the main
concerns. Typical failure modes are shown in Figure 5.1 b.
Generally, impulsive loads confronting structures in civil engineering are
soft impacts with a loading period (until maximum loading is achieved) of
about 10- 1 -10- 3 s.
The effect of an impulsive load acting on a structure can bring about serious
dc:mage not only to the structure itself but also to the environment around, as
in the case of an aircraft crashing into a nuclear reactor or if a ship were to
crash into an offshore oil platform. But in all these cases, the possibility of such

(a)

Penetration and Spalling Penetration, SpaJling and Scabbing Perforation

Shear failre

(b)
Fipre 5.1 Types of failure mechanism during impulsive loading; (a) failure mechanism for
soft impacts. (b) failure mechanism for hard impacts.
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 103
Yma< (t)

Or-------------~Y

,--------------------,
I G1 m3~.;:'j~1
I ~' J c, -.~ \mN I Target
I m, 9, 9 N-, mN-l C :
L
I C, m, G, N-,

-
~

Missile velocity Va
x
Figure 5.2 Schematic representation of multimass model from Bignon and Riera [10].

an occurrence is very low. The occurrence of such an incident, together with


the expected ultimate state of the structure, has to be properly considered and
an introduction of probabilistic approach for safety evaluations is becoming
common in these cases [7]. There are already a number of countries where
such steps are introduced in the design of nuclear reactors [8,9].

5.2.2 Load modelling


Structures in civil engineering are usually confronted with impulsive loads
which have a rather slow loading rate (in the order of microseconds) but with a
long duration of action. The 'Riera model' [10] has a rather wide application
in the simulation of soft impulsive loads, especially in cases where the striking
body is deformable. In this model, the impacting body is considered as a
bidimensional assemblage of 'in-line' masses. The masses are interconnected
by springs which are non-linear visco-elastoplastic, represented by Cj in
Figure 5.2. The straight bar elements are joined at each mass point by a visco-
elastoplastic rotational spring, indicated by Gj • A change in the impacting
mass can be considered in the simulation ofload characteristics by splitting the
impacting body.
From the material behaviour point of view, a difference in behaviour under
various loading conditions or loading rates is expected. It is reported that due
to the effect of inertia an increase in strength followed by a decrease in
deformation capability can be expected during higher strain rates or loading
rates [6].

5.3 Analytical studies

5.3.1 Analytical methods


In most cases, concrete structures undergoing soft impacts are assumed to
undergo a transient phenomenon during analysis. Rotary inertia as well as
104 PRECAST CONCRETE RAFT UNITS

damping is included in the equations of motion under forced vibrations and


the effects of characteristics peculiar to the impulsive phenomenon - namely,
the dependence of material characteristics as opposed to the loading rate, local
deformation at point of impact and excitation of higher modes of virbration -
are usually considered. Analysis for a perfectly elastic body, taking into
account local deformation at the point of impact and excitation of higher
modes of vibration, can be achieved by either direct integration of the
equations of motion [11] or application ofthe Hertz contact law together with
the Fourier series [12]. Both of these methods bring about an exact solution to
the problem. It should be noted that these methods can only provide reliable
solutions within the elastic regime.
When considering material non-Iinearities and their dependence on time,
the finite difference method [13-16] or the finite element method [6,17-22]
prove to be more efficient. For problems in the field of civil engineering, a
complete solution is usually required, i.e. not only must the final states of the
impacted structure be known, but also the entire history of motion. Excitation
of the higher modes of vibration caused by inertia can be studied quantita-
tively using these methods. Moreover, bond characteristics between concrete
and reinforcement and shear transfer mechanism after the formation of cracks
[19] can be included. The state of stress in the concrete section, cracking
conditions, degree of perforation or penetration and amount of scabbing can
be successfully simulated by means of the non-linear dynamic finite element
method. However, the finite element process has the disadvantage of being
both costly and time consuming if the structure is not properly idealized.

5.3.2 Non-linear model


The constitutive model for beha viour in the elastoplastic region is based on the
model proposed by Isobata [23]. Concrete is considered as an orthotropic
body as cracks form in various directions. The body is assumed to have a
modulus of elasticity of Ep , and Poisson's ratio of vp , as opposed to (E and v) for
an isotropic elastic body. With the help of conversion parameters ('7, ~)
obtained from uniaxial material characteristics, the elastoplastic body can
then be expressed in terms of E, v, '7, ~. The axes x, yare taken along the
direction of principal stresses. The conversion parameters can thus be defined
as follows:

(5.1)

The following equation can be derived from the Maxwell-Betti reciprocal


theorem:
v·E
y x
=v·E
x y
=!' .(",.", )1/2' v 'E
<"xy 'Ix 'Iy • (5.2)
The modulus of rigidity Gxy can be expressed as a function of the modulus of
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 105
elasticity and Poisson's ratio of the various axes as follows:
1]~/2.1];/2. E
(5.3)
2(1+~xy·v)
j: = j:1/2.j:1/2 (5.4)
"xy "x "y .
From equations (5.1) and (5.3), the following relation is obtained:
1+ v
I] xy = 1+ .1]1 / 2.1]1 / 2
j: .v x y.
(5.5)
~xy

The stress-strain at the plane stress field of the orthotropic elastoplastic body
is thus:
(5.6)
(5.7)

(5.8)

(5.9)

(5.10)

vp =(j:~x ·v·j:~y ·v)1/2 =;;"='xy ·v . (5.11)

Q)
-<Jl
. - <Jl
<JlQ)
compressive tensile c: ...
a,.·..
strain strain ..... <Jl

compressive tensi Ie
stra in strain

Q)

·en>
...'"c.",
Q)
",
11,=0
11,
o=:
EQ)
0",

a) Concrete b) Steel

Figure 5.3 Idealized stress-strain curve for material.


106 PRECAST CONCRETE RAFT UNITS

The conversion parameter 1] for the modulus of elasticity in the elastoplastic


body as used in equation (5.9) is obtained from the idealized stress-strain
curve for concrete as illustrated in Figure 5.3a. The stress-strain relation
below the tensile stress limit is considered to be linear. On reaching this limit,
tensile rupture occurs and the elastic strain energy in the element is dissipated
to the surrounding nodes in the form of equivalent nodal forces. As for
Poisson's ratio, it is assumed that ~ = 1.0. The same method is also used to
define the elastoplastic model for reinforcement.

5.3.3 Finite element model for beams


Figure 5.4 shows the finite element assembly used for the reinforced concrete
beams that were studied. The dimensions of the beams are as follows: total
depth, 15 em; width, 15 em; total length, 130 em; span, 120 cm. The beams are
doubly reinforced, with concrete being modelled by isoparametric rectangular
4-node elements, while reinforcement is represented by line elements. The
isoparametric rectangular element is used because the effects of rotary inertia
can be considered and this is found to give a higher degree of accuracy [19].

-80 C.L.

<ll
I
I

50 5x 20
I--
7 x 50=350 8 x 25=200 =100
650

(a)
C.L.
2Dl0 I
~ rrrr

{ 3Dl0

I\.
I~
<31

<31

~r
D.

5 x 100 =500 90 60

650
(Unit: mm)
(b)
Figure 5.4 Finite element meshes for RC beam; (a) finite element meshes for concrete, (b) finite
element meshes for reinforcement.
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 107
The consistent mass matrix is also applied here. The impulsive force-time
relation obtained from accelerometers placed on the striking weight during
tests (Figure 5.5), stress-strain relation of materials acquired from static
uniaxial tests (see Figure 5.6), coefficient of shear transfer after cracking and
the spring constants (K h and K v) that represent concrete-reinforcement bond
behaviour, were fed into the analysis. Various time increments (~t) for the
integration process were at first applied to determine a stable time increment.
As shown in Figure 5.7, if the effects ofloading rates on the static uniaxial tests
are considered to be due to rotary inertia and damping, and since damping is
small enough to be ignored during soft impacts, this analysis in itself indirectly
involves the effect of loading rates on material behaviour. The effect of shear
transfer at crack surfaces due to aggregate interlock or dowel action, etc., is
taken into account by introducing a coefficient of shear transfer (shear
retention factor) which reduces the shear stiffness in the element once cracks
occur [19]. The bond link element [24J is used in modelling the concrete-
reinforcement bond behaviour [25].

'""
z 1.2
lo'
CD
0;
II
~ 0.8
Q)
0
'-
.E
t>aJ 0.4
a.
E

5 10 20
time (ms)
(a)

failure
6
z ,/

""
lo'
CD

"
0;
II ,/
't;
~
4 ,/
Q)
,/
0 ,/
'-
....
0 ,/

t><II 2
a.
E

0 2 3 4 5 6 7
time (ms)
(b)
Figure 5.5 Impulsive load - time functions and definition of loading rate; (a) in elastic, (b) at
failure.
108 PRECAST CONCRETE RAFT UNITS

a'k=390kgf/cm 2 (38.2MPa)

-a) concrete - ~ - - - - -...,


/ I

/ :
~,=3.33X105Kgf/Cm2
E .u=35oo IJ
:
x (3.26x 10·MPa) I

2 3 4
\! strain (x 1()31J )
50kgf/cm 2 (4.9MPa)

b) steel a .u=47kgf/ mm 2 (460.0MPa)

a .y=36kgf/mm 2 (352.8MP~
--
- - - - - - ---..-.----,
--- I
I
I
I
I E .. =50000 11
x E.=2.1 x Wkgf/ cm 2 (2.06x 105MPa) I
'-'

o 2 3 4 5
strain (x 10·11)
Figure 5.6 Idealized stress-strain diagrams for steel and concrete.

5.3.4 Finite element model for slabs and handrails


Slabs and handrails were modelled using the non-linear layered finite
element analysis [26,27] as shown in Figure 5.8a (1/4 part) and 5.8b
(1/2 part), respectively. The 4-node rectangular plate-bending element was
applied here. The element nodal-point degrees of freedom are the transverse
displacement Wi and sectional rotations 0xi and Oy;. The impulsive force-time
relation and material properties from static uniaxial tests were the input data
for this analysis. The slabs and handrails had doubly reinforced sections and
were represented by rectangular elements, which were eight layers in depth
(two layers of reinforcement and six layers of concrete). The modulus of
elasticity for reinforcement was calculated from the amount of reinforcement
in both the x and y axes. Strain in each layer is assumed to be as shown in
Figure 5.9 and the neutral axis can be given by the following equations:

(5.12)
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 109
600 VL=10 tf/ms

~E 500
0

"-.....
C> 400
~ VL=2.5tf/ms
..._----:-= ;~~=.----' ...

0
.,'" o--t(

z'"
" .; _ .. I

:110 .... .ot... ' :


.,'" , \
'stress-strain curve of
I
rrnn q
""',," "',, ",,,,,1 TP'='OO
Cll I

''-.,"
>
200
'"
I I

:(300mmX 150mmx50mm)
1000 2000 :1100 4000
average strain (II)

(a)

(tf)

Q)
o - - experiment (RC beam)
'-
.E - - - calculation (RC beam, [C) =0)
-·-calculation (RC beam, [C) ~O)
oQ)
Q.
E
* [C) : damping matrix

o1l
"0
o'"
...J

o 2 4 6
deflection (mm)
(b)

Figure 5.7 Effects of rotary inertia and viscous damping on stress-strain relation and impulsive
load - displacement relation; (a) concrete column (h = 10/;',), (b) RC beam (h = 10/;',).

e tEc ·t 2 + Es
= -------==;=;----'-----'-
LA sy( Zsyi (5.13)
y Ec·t + Es L:A sy ;
where ex, ey denote the centre of the elastic region in the x and y directions,
respectively; Ec> Es are the concrete and reinforcement moduli of elasticity,
respectively; A sxi , A syi are the average cross-sections per unit length in the x
and y directions for the ith layer of reinforcement, respectively; Zsxi' Zsyi are the
distances in the x and y directions from the middle of the ith layer to the top
surface of the slab; and t is the slab thickness.
o
-

650 ,Concrete layer y Concrete layer


10 x 60=600 flli
:8 ·m ~ CI ;g
C C CI tTl
CI n
L
CI Clr- ~
CI CI
lilt)
~ICI
"lit)
X
y ClCO ----
CI ~
co n
X
CI
j~
A -- x100=800 14 x 150=6QO
'~"
~X J 3 X2OOj60 OX \ Steel layer
C.L. _: o 2000 ;I>-
iL..8 Steel layer C.L.
'"
::J
C.L.
(Plan) (Section B-B) (Plan) (Section 0-0) ~
==l
en

(a) RC slab (1/4 part) (b) RC handrail (1/2 part)

Figure 5.8 Layered finite element meshes for slab and handrail (units are in mm).
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 111

Zoxi
t

~ !
Z z
(a) (b)

Figure 5.9 Strain in (a) concrete and (b) reinforcement layers.

Table 5.1 Time increments used in various analyses

Type of analysis

Type of structure Elastic (ms) Failure (ms)

RC slabs 0.2 0.05


RC handrails: Prior to maximum impulsive load 0.1 0.025
After maximum impulsive load 0.1 0.1

The slabs were simply supported at both sides while the handrails had fixed
supports on the lower side. The consistent mass matrix was applied together
with the non-conforming plate-bending element. Different time increments
(Lit) were used in the integration process and the values indicated in Table 5.1
were found to provide quite stable values for this analysis.

5.3.5 Dynamic solution of equilibrium equation


The semidiscrete equation of motion for a finite element assembly can be
written as follows:
[M]'{ O}I + [C]· {U}I + [K]'{U}, = {R}, (5.14)
where [M], [C], [K] represent the mass, damping and stiffness matrices
respectively while {R}, is the external force vector. {O}I' {U}I' {U}I are the
acceleration, velocity and displacement vectors, respectively.
The discretized equation for equation (5.14) during a time step of M is
shown by

[Ml {LiOL-+,Ht + [Cl {LiUL-+,+M + [Kl {LiUL-+t+M = {LiRL-+'Ht' (5.15)


As equation (5.15) is simply an approximative equation of motion, it is
solved using the Newmark-,8 method, which consists of the following
112 PRECAST CONCRETE RAFT UNITS

equations [17,18]:

{U}I+t" = {U}I + L\t· {UL + L\t2. [(1/2 - fJH O}r + fJ{ O}'Hr] (5.16)
{U}IHr = {U}r + (1/2)' L\t·( {O}r + {O}'H')' (5.17)
The discretized equation for acceleration can be obtained from
equations (5.16) and (5.17) which describe the evolution of the approximative
solution as follows:
.. I 1 . I ..
{L\U},_'H' = - - 2 {L\U},_rH'-- {U}I - - {UL· (5.18)
fJ·L\t fJ·L\t 2·fJ
Substituting equation (5.18) into equation (5.15) brings about the following
equation:

([K] + 2.;.L\t [C] + fJ.~t2 [M] }{L\U},_'HI


1 . (4' fJ - 1) ..)
= { L\R},_IHr+[C] ( 2'fJ{U},- 4.fJ ·L\t·{U}1

1 . I .. )
+ [M] ( fJ. L\t {U L+ 2. fJ {U} I .
(5.19)

In the above equations, the parameter fJ and the time step L\t are closely related
to the accuracy of the integration and also the stability of the dynamic
solution. In this study, the parameter fJ =t
is used because it satisfies the
necessary stability conditions. Based on various calculations, the most
appropriate value for the time step L\t was selected. The integration of the
equation of motion (equation (5.15)) with respect to time can be obtained by
solving equation (5.19). In order to improve the accuracy, the following
iterative method was also applied:

[M]
. - fJ.L\tI.{UL - 2.fJI..)
(fJ.L\t12 {L\U};_I+t" {U}t

I { } (4' fJ - 1)
+[C] ( - 1- {L\U }':-I+tu-- U 1+ ..)
'L\t'{U}r
2·fJ·L\t 2·fJ 4·fJ
+ [K]rHt·{L\U}:_rHr = {L\RL-tHt - [K]tHt' {L\U}::t\~t

- -1- [Cl {}i-I i-I


I [Ml{L\U},_tHr
L\U l-tHr---2 (5.20)
2·fJ·L\t fJ·L\t
where {L\U}::I\~I' {L\U}:_t+tu are the increase in displacement vector during
the iterative time of i-I and i. Generally during soft impacts, the effects
of viscous damping are small enough to be ignored. In the procedure here,
it is assumed that [C] = o.
The flow of the iterative process for equation (5.20) is as follows:
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 113
(1) Formation of stiffness matrix [K*]
1
[K*] = [K] + --·[M]. (5.21)
P·l!.t
2

(2) Calculation of the constant parts of the load vector

1 . 1 .. )
{l!.R*}t_t +.1.1 = {l!.R}t-tHt + [M] ( p'l!.t {U}t + 2P {U}r . (5.22)

(3) Setting of the initial values for the variable parts of the load vector
{RErr}i=O=O. (5.2:3)
(4) Formation of the load vector
{l!.R*}:_rHr = {l!.R*}._rHt + {RErr}i-l (5.24)
(5) Calculation of the discrete increment in displacement {l!.U};-tHI using
the band matrix method.
(6) Calculation of the variable parts of the load vector

{M,,,}: ~ - [K~,:,~ (AU):H: - p_ ~t' [M] {AU}:"'H' 1 (5.25)


{R Er .} = {RErrl + {l!.R Err } J
(7) Decision on degree of convergence
(5.26)
where the abbreviation 'tol' stands for the amount of tolerance while II II is
the Euclidean norm.
If the above equation is satisfied, the proceeding step would be step (8).
Otherwise, the flow will return to step (4).
(8) Calculation of the displacement, acceleration and velocity;
{U}tHI = {U}l + {l!.U}:_t+M
.. 1 . t. 1.. ..
{U}tHt = p. l!.t 2 {l!.U}:_t+M - p. M fUll - 2P {U}. + {UL (5.27)

. . 1 .. ..
{U}t+t>t = {U}t + 2l!.t( {U}l + {U}l+t>l)

5.4 Experimental studies

5.4.1 Test program for beams


In order to verify the validity ofcalculations using the model mentioned above,
RC beams similar to the analytical model were subjected to both static and
114 PRECAST CONCRETE RAFT UNITS

weight

bolt bolt
I
load cell

o : meas. point of acceleration


X: meas. point of steel strain
lSI : meas. poi nt of deflection
(unit: mm)

Figure 5.10 Test apparatus and measuring point of deflection, acceleration and strain.

impulsive tests. The static and impulsive tests were carried out in different test
frames. A 200 tf universal testing machine was used for the former. The beams,
which were simply supported, were loaded through a square steel plate (15 em
by 15 em by 1 em) placed at the mid-span. Elastic tests as well as failure tests
under continuous loading were carried out. On the other hand, impulsive tests
were carried out using the apparatus shown in Figure 5.10. Flat cages were
placed on the friction plane while load cells with ball bearings were fixed to the
bottom of the beam ends to make them freely rotatory and thus give the effect
of a simple support. Moreover, the ends were bolted to prevent them from
lifting off their supports. A mass of 100 kgf was dropped from various heights
at the mid-span to induce impulsive force. A rubber pad was placed on the
loading plate to provide soft impacts. The beam, rubber pad and mass were
considered to behave as one single body during the impulsive action. Besides
that, it is assumed here that the rubber pad causes no damping. In the
experiments, h = 1cm was used for the elastic test while h = 60 cm was used for
the failure tests. Measurements for strain, deflection, acceleration, crack
conditions and failure conditions were taken at points indicated on
Figure 5.10.

5.4.2 Test program for slabs and handrails


The dimensions of the slabs used for the verification tests are as shown in
Figure 5.11. For the static tests, load was applied through a 50 tf capacity
hydraulic jack, and was gradually increased by increments of 1.0 tf. At each
Rei nforcement
C.L ..,
LDwer face Upper face
~ i <:7 / / ,..,
7 / \ ,~
/ ~ §
C'I ("'J
::=
tTl
C'I I - 0= c tri
('I)
0=
,Strain gaug~ ~
('I) C'I
C'I 1-(1-
1-+----1-+- 3!:: ~
- II
II C'I
3-1,,If ~
(Q
+--+--=+--+-,,-1 t ~ 4 :
c::
-..,.
2 Prestressing x
C'I x C'I I
~
X
-- 4 ' tendon 4
It') - 2 4 _ , Cil
~
('I) o
'"
"'-'\
\ \ "'-
~.- ~
~J~-iT §2
~ ~l "tf
c::
I•• 01: ... cj>~1 .I~ fa
: :
8 x 154 = 1232 <:
tTl
3~-Ii :~:1:=I1il I:!=
(unit: mm) ~ 7x162 1134 1t- (unit: mm) 5;p-
(PC3 - 1 ) O
(a) (b) ~
Figure 5.11 Details of slabs; (a) reinforced concrete slab, (b) prestressed concrete slab (full
prestressing).

.-
.-
Vl
116 PRECAST CONCRETE RAFT UNITS

1.92m

Rubber pad
Loading plate

-----Slab speciment
c:::::::::===,;----
Supporti ng frame
Figure 5.12 Testing apparatus for impulsive tests (slabs).

loading step, the load was held constant to measure automatically the
deflection, concrete and reinforcement strain, crack width and force of
prestressing tendons (for PC slabs only) [28]. The displacement at the centre of
the loaded steel plate for all slabs were recorded by an X- Y recorder.
On the other hand, the apparatus used for impulsive loading test was a
pendulum-type impulse load testing machine which was specially designed to
derive a half sine-wave load with a mass of 500 kgf weight (see Figures 5.12 and
5.13). A rubber pad was placed on a square steel loading plate (15 em by 15 em

Figure 5.13 Testing apparatus for slabs.


3950 250

250
Wire rope \l
o
n Z
\l
I~
p ::>:I
tTl
....,
\ ~o~
8co tTl
'- ~,~0 Falling
It) ::>:I
) (Detail of handr ail)
height. h ) >
>~--? ~~ :::l
- ~ ..... " ..... ' - . '
'-~ I'
Discontinuouslline ~
....,
Vl
plate
Loading plate C
- Z
(i;
'" "' I tTl
::>:I
"
~
Steel girder ::§
t""'
~
Figure 5.14 Details of impact loadmg test for concrete handrails. <
tTl

5>
"Z
Cl

- .l
--
118 PRECAST CONCRETE RAFT UNITS

Figure 5.15 Impact loading system for concrete handrail.

by 1 cm) to obtain soft impact (impulsive load). The acting force was measured
with the help of an accelerometer attached to the falling mass.
For the impulsive load tests, two sets of tests were carried out. For the first
[28], the height of the falling mass was gradually increased in increments of
2.5 cm until the appearance of the first visible crack. A crack gauge was then
attached and the height of fall was increased in 5.0 cm increments until
ultimate failure. Impulsive loading was repeated ten times for each height of
fall. The outputs from strain gauges, crack gauges, load cells and acceler-
ometers were amplified and then recorded by an analog data recorder. In
the second set of tests [29], the height at which failure would occur was
calculated using the analysis mentioned in the previous section. Elastic tests
were carried out for a height of fall of 1.5 cm followed by tests for failure under
one single impulsive load using the calculated height of fall. Measurements for
deflection and acceleration response were carried out. The measuring system
consisted of an eddy current type non-contact displacement transducer, ultra-
small high-capacity accelerometers and an analog data recorder.
There is the possibility of collisions occurring when traffic accidents take
place, such as a vehicle crashing into a concrete handrail. A full-scale test to
simulate such collisions was performed. For this experiment, three cranes were
used to induce impulsive force. Two cranes were used to support the falling
mass while a third crane kept it at a specified falling height (see Figures 5.14
and 5.15). An impact force of 2.0 tf was directly induced through a loading
plate to the handrails with the applied point being equivalent to the height of
the centre of gravity of a truck. Measurements for deflection and acceleration
response were taken at points over a wide range.

5.4.3 Verification of analytical results for beams and slabs


The accuracy of the analysis can be verified by test results. Figure 5.16a shows
the impulsive load-midspan displacement function for the elastic tests and the
analysis. It can be seen that there is no distinct difference in stiffness for both
cases, even though the curved shapes are different. This is assumed to be
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 119

z
co 1.0
~
.,;

-II

IV
...
.e"
to., 0.5
c.
E

deflection (mm)

0 0.1 0.2
(a)

6
z
~
co
5
.,;
II
4
~

-".,
IV
...0 3
0
c. 2
E

deflection (mm)

0 2 3 4 5
(b)

Figure 5.16 Impulsive load - midspan deflection functions (RC beam); (a) elastic (hysteresis
curve), (b) up to failure.

-
30
~

-- 4 '-"
'U
'U
'" 20
'"
~
~ 3 Ql
CI> >
> I 'u;
~ 2 I :;
c. 10
"c. , I
- - Experiment E --Experiment
E ,, ----Calculation - - - -Calculation

0 0.1 0.2 3.0 0.4 0.5 o 2 3


Deflection (mm) Deflection (mm)

(a) (b)

Figure 5.17 Impulsive load - midspan deflection functions (RC slab); (a) elastic (hysteresis
curve), (h) up to failure.
120 PRECAST CONCRETE RAFT UNITS

caused by support conditions for the beam or 'noise' effects during the
measurement of the discrete deflections. The results of the failure tests are
given in Figure 5.16b which shows the load-midspan displacement curves for
the elastic and until failure loading, respectively.
The verification of the analysis on slabs is given in figure 5.17, which shows
the impulsive load-midspan deflection for the elastic as well as the failure tests.
There is a slight difference in the curves towards the failure point, but the
stiffness in the early stages are quite similar. From these figures it can be
concluded that the analyses for both the beams and the slabs give quite
accurate results.

5.5 Concepts of performance improvement for impulsive loading

5.5.1 Concepts of per:formance improvement for beams


As the analytical method has been shown to be quite accurate, calculations on
various materials were carried out and the contributions towards performance
improvement were studied. It was found that an evaluation method on
concrete beams under impulsive loading can be carried out based on the
following items [30]:
(1) Degree of improvement in load capacity
(2) Degree of suppression of drop in deformation capability at failure
(3) Degree of suppression of drop in stiffness
(4) Degree of suppression of critical local deformation
(5) Degree of suppression of increase in failure region.
A few factors considered to have effects on the impulsive behaviour of RC
beams were analytically studied. They were beam depth (all beams had a depth
of 15 cm except for the h-18 beam which was 18 cm in depth), concrete strength
(RC: normal strength concrete; a-860: high-strength concrete), usage of high
yield strength steel (HT) and the addition of steel fibre (SF-133). A model
derived from Hannant's theory [31J was employed for the material character-
istic of steel fibre.
The failure load-loading rate relation is shown in Figure 5.18. Under high
loading rates, the load capacity shows an increase mainly because of inertia.
The effects of inertia depend upon the sensitivity of the failure load-loading
rate relation. In other words, comparison can be carried out based on the
curve gradient. The SF-133 beam shows a larger degree of sensitivity
compared to the other beams and it can be concluded that the introduction of
steel fibre causes an improvement in load capacity. When high yield strength
steel is used, an improvement in load capacity can be expected at the higher
loading rates.
Figure 5.19 shows the relation between the deflection at failure and the
loading rate. Except for the HT beam, the deflection at failure decreases with
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 121

15 /'
4
'"'
CJ
10 3 .;;

z 3 ~
10
"
~
~

::.<:
<Xl CIl
a. E
a; 8 CIl
II E Q)
..c
E ", 2 2
~

Q) u
0:
6
~
~
Q) ~

3 4
~
<tl
0
~ '>-
'>-
CIl
'>- a; Q)
~

a; 4
~

" CIl
..!:!
CIl
'>-
"..!:!III a;
2
loading rate (tf/ms=9.8KN/ms) "..!:!
III

0
0 2 3 4 5

Figure 5.18 Effect of loading rate on load at failure.

3.5
~
CJ
15 §
SF-133
- - --- -
<I)

3.0
- - -- - - - - -
~

6 E
III
2.0 t?
~
Q)
E loading rate (tf/ms=9.8KN) 1.0
..c
E III
10 Q. U
0:
Gl
E
~
0 2 3 4 5 ~ ~
~ Q) 0
'>-
4
~
III
'>- RC .-
~
~
HT 0.8
m
""r-............. ------
~
~

------- --===--= a;
-
CIl
c: h-18 ..- - -
/ -.
'>-
.;0; c: <tl
CJ
;;: 2
Q)
"..L--" - - - - - - ___ 0
.;:; c:
0.4 0
.2

"
0-860
Q) Ol

" ;;:
Ol Gl
;;:
" Q)

loading rate Ctf/ms=9.8KN/ms) "


0
0 2 3 4 5

Figure 5.19 Effect of loading rate on deflection at failure.

the increase in loading rate. The 0'-860 and SF-133 beams do not show a drop
in deformation capability but for HT beams, deformation ability increases
with the loading rate. Deflection at failure for the SF-133 beam is about three
times that of the RC beam under static loading, thus making it the most
effective. For the HT beam, it can be seen that, at high loading rates, the
deflection at failure is greater than that of RC beams, therefore showing its
effectiveness.
122 PRECAST CONCRETE RAFT UNITS

14

/ 12
.!:!
..., 2.5
10 S
0 2.0 8 r-. ~
"'
cO 1:> E
0> OJ OJ
II
2.0
" 6 E
Q.
1l
...
E
50
, 4 1.5 5 >-0:
C,)

"''"" ' h-18___ ?", ~ 5


1.5
-- --- .,:'! ...

-- 1---- --
~
.. .. ~
x
~
/ 0-810,......-
,......- 1.0 III
>-
...III
0>

-.'
>-
.,...c:
0> 1.0 ;§ .,
c:

.,
2l 0.5 / 0.5 ~
2 loading rate (tf/ms=9.8KN/ms)

0
0 2 3 4 5
Figure 5.20 Effect of loading rate on total energy.

Figure 5.20 shows the relation between the total energy and the loading
rate. It is clear that the total energy increases as the loading rate is increased.
Since total energy can be defined as the amount of energy required for beam
failure under a single impact, it is approximately equivalent to the energy
absorbed by the member. Comparatively, the total energy for the SF-133 beam
is largely due to the addition of steel fibre, as it causes a very large increase in
ductility. Under soft impacts (low loading rates), an increase in beam depth, or
in concrete strength, has little effect on beam ductility.
Local deformation tends to occur at failure when subjected to impulsive
loads. Curvature at failure (1IR) will be used here to quantitatively study local
deformations. In order to eliminate the effects ofdeflections from the curvature
at failure, the following index will be used to evaluate local deformations:
. Curvature at failure (II R)
In dex 0 f Ioca I deformatIon = . (5.28)
Deflection at failure (15.)
Figure 5.21 shows the relation between the index of local deformation and
loading rate. Increasing concrete strength results in a large index. Increased
depth gives a value between that of the 0'-860 and RC beams, with local
deformation becoming more sensitive to loading rate. A stable low index
value is obtained in the case of steel fibre reinforced beams, showing that local
deformation is greatly improved. For high loading rates, the usage of high
yield strength steel gives results quite similar to that of the SF-B3 beams.
The failure region is affected by inertia and the area tends to be smaller
during impulsive loading. The failure region is the area where internal energy
is released, and when the area is small the cracks will be large and the failure
condition deteriorates. Sketches of crack conditions at failure were examined
CONCRETE RAfT UNITS UNDER IMPULSIVE LOADING 123

0-860 __ -7 4
/-- /'
" E
./
3 '
'E" ..c'"
C. <Il

'\ h-IV / 3
..
c
.~
iii .-/'
E
.E 2
'y'
..,
Q) /. " RC
2

.2'" HT
b
..,x
<Il

C
loading rate (tf/ms=9.8KN/ms)
OL..-_ _--'-_ _-----'. -'---_ _--'-_ _- . J
o 2 3 4

Figure 5.21 Effect of loading rate on index of local deformation.

.!!
5
SF-I33 ~
"E
~

-E
----- ----
- -- - --~~----.-
~ ~

-b 1.0
3
4
~_._.-
HT '" '"
E .8
<Il
l-
0.8 -; ~
I-
~ 3 , h 18
....'"
"
.El-
~'-'
....0 0.6 ~
--- -- -~-=---
Q)
I-

c 2
__ 0-860 -. b "
.~ 0.4§ ~
'"
E ~ "E
l- C
o
0.2 .0,
loading rate (tf/ms=9.8KN/ms) E
0
0 2 3 4 5

Figure 5.22 Effect of loading rate on region of failure.

and the area of failure was calculated. The relation, with regard to loading rate,
is shown in Figure 5.22. For high yield strength concrete, the failure region is
smaller than that of RC beams at the lower loading rates but, at the higher
rates, the failure condition is markedly improved. On the whole, SF-133 beams
have large values. Therefore, it can be concluded that an addition of steel fibre
reinforcement is effective in improving the failure conditions.

5.5.2 Concepts of performance improvement for slabs and handrails


Performance improvement of concrete slabs [26,32] with regard to material
type was studied for the following four material types; normal strength
124 PRECAST CONCRETE RAFT UNITS

high loading rate


50 ,-7-
40 § 40
"0
't:
:;; 30 ~Ql
30
>
'"
2 ';;;
:;
~ 20 ~ 20
';;;
:; - R C slab
Q.
E - - - HRC slab
- 10 --- - FRC slab

o 123 o 1 2 3
Deflection (mm) Deflection (mm)
(a) (b)

Figure 5.23 Impulsive load - midspan deflection functions for various types of slabs;
(a) experimental results, (b) calculations.

reinforced concrete (RC), high strength reinforced concrete (HRC), steel fibre
reinforced concrete (FRC) and lightweight reinforced concrete (LRC).
Figure 5.23 shows the impulsive load-midspan deflection relation for the slabs.
The characteristics of each slab are clearly simulated in the calculations, when
compared to the test results. The HRC as well as FRC slabs show a high initial
stiffness and, as for the latter, only a slight drop in stiffness can be noticed after
cracking. For the LRC slab, the stiffness throughout is quite low.
Figures 5.24 and 5.25 show the distribution of deflection at midspan
(transverse direction) and the crack progression within the slab section,
obtained from calculations respectively. The LRC slab is found to have a
relatively spread-out distribution of deflection, but failure occurs at an early
stage. An almost equal amount of deflection throughout the section can be
seen in the FRC slab and it can be assumed that yield lines occur at failure. For
the HRC slab, local deformation is assumed to occur at the failure region and,
as seen in the crack progression, the failure region is small. On the whole, as
loading rate increases, the distribution of deflection and crack progression
becomes localized.
The distribution of the direction normal to the principal stress (during
maximum impulsive load) for the RC and FRC slabs (1/4 size) is shown in
Figure 5.26. This is the direction in which cracking occurs. This result is in
good agreement with the experimental results. The direction of tensile
cracking at the rear face of the FRC slab (shown by solid lines) is parallel to the
direction of yielding lines.
Figure 5.27 shows sketches of the rear face of the slabs at ultimate failure,
obtained after the experiments had been completed while Figure 5.28 shows
the actual photographs taken after the tests. For the RC and the HRC slabs,
flexural cracks form under repeated impulsive loading. As loading is increased,
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 125
~ C.L. ~ C~.L_. _ C.L.
EO~---­
EO l E O
E
~; 1 ~=19tf ~; 1 ~ 1
g P=26tf g .,E
r
(J (J (J

rY
aJ aJ aJ

o~ 2 ~2
Y ~2
0
Z rYe
(a) RC slab (b)LRC slab (e) HRC slab

r-.C.;:::.L..:....- - - - - ~C.L.

~ 0 P=17tf
'E
~ 1
EO~
~
; 1
P=31tf

., g P=35tf

rY
(J (J
aJ

~2 P=34tf %2
r
e Y o Z

(d) FRC slab (e) RC slab


(high loading rate)
Figure 5.24 Distribution ofdeflection for various types ofslabs - transverse direction (analytical
results).

C.L. C.L.

rYiiiii~.riiiii~s
(a) RC slab (P=22 tf) (d) FRC slab (P=21 tf)

JVE.ms
C.L. C:1L.

[Y'II• •
r I
(b) LRC slab (P=22 tf) (e) RC slab (high loading rate (P=22tf»)
C.L.
t:8J Tensional crack
~ Digits show the amount
of drop in rigidity

(c) HRC slab (P=23tf)

Figure 5.25 Crack progression for various slabs (analysis).


126 PRECAST CONCRETE RAFT UNITS

C.L.

" ,'-/.. 1'/.'/.'-/'/1;(,


.f. '-1/ I/. 'x '/ 'xi/.',(
· . '1It/) >!..X'. /.X
· ' 77. 7. 'x '7)! XI/.
,r7. '/7J<,RV)rx
.' ,F!]'79RV'xt;l
'.. ,I'I-,'I'j /Ixl/ -xt/.
. .1/./ ";( 'xJ/.I/.IX'lx
· 'HI;'I';' IJX:IXlx
· ' ,if.. 'I. 'f.I/J/r/t/tJ
C.L. __ -+=.L'1Lf7"",Fl£-,-"Y- J7
L 'J..L'liU ·F7.LL:'r/ ·I f~l..L.
:u...:J.'.w
i I:>

(a) Reinforced concrete slab

C.L.
·H, " tx'
. -',' .Ix
.' ".Ix
" . ,t/.
. " " " ,1/
, , - ' / ,t/.
. _ ,i:lt/rx
. - -. - 1/..1/1."
,', ,fit/Ix
- -. -.. - 'f-Hv
C. L.- _.j...LLLl..1..LO-...L-.l..Ll-...L-~'LJ...1'-A:J.£. ;--tFlxL
,
(b) Fiber reinforced concrete slab
----1st layer (compression)
--8th layer (tension)

Figure 5.26 Direction perpendicular to principal stress in slabs (analysis).

the slabs deteriorate with scabbing at the rear face occurring and finally the
slab fails due to punching shear. Comparatively, the HRC slab has a higher
deformation ability and resists high impulsive loads, flexuraIly, before finally
failing under shear. The failure region is also smaller in this case. The FRC slab
resists impulsive loads flexurally throughout and yield lines form before failing
due to flexure. The amount of scabbing is greatly improved with the FRC slab.
The effects of prestressing were also studied experimentally [28] by
comparing results from slabs under different prestressing states and also with
steel fibre in the prestressed slabs. Many studies have shown that the structural
damping due to load histories and ductility factors of PC structures were
inferior to that of RC structures under impulsive loads. Figure 5.29 shows an
example of a prestressed slab under impulsive load failure. From test results, it
was found that the position and direction of initial cracking was irregular
under impulsive loading and scabbing was observed at the point of intersecting
cracks. When compared to the other types of slabs, the prestressed slabs tend
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 127

(a) Reinforced concrete slab

(b) Fiber reinforced concrete slab

(c) High strength concrete slab

Figure 5.27 Crack pattern at failure (experiment).

to fail at short notice and the amount of scabbing of cover concrete is large,
mainly due to the prestressed forces. The degree of scabbing decreased
remarkably with the introduction of steel fibre. The residual crack width and
deflection were smaller when compared to those of the RC slabs. All the
prestressed slabs failed under punching shear except for the steel-fibre slab,
which failed under flexure due to the crushing of the concrete in the
compression zone. Thus, an improvement in ductility can be expected with the
addition of steel fibre in the prestressed slabs.
Performance improvement of concrete handrails was carried out by
128 PRECAST CONCRETE RAFT UNITS

(a)

(b)

(c)

Figure 5.28 Crack pattern at failure (slabs); (a) RC slab, (b) FRC slab, (c) HRC slab.

comparing calculations and results from full-scale tests [32]. Figure 5.30
shows a comparison of crack pattern between the experimental results and the
analytical results at failure, while Figure 5.31 shows an example of a handrail
that was tested. The curves showing the direction perpendicular to the
principal stress at an early stage of impact is given in Figure 5.30a. The
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 129

Figure 5.29 Crack pattern at failure (PC - full prestressing),

C.L. C.L.
, . I- 'x'.
IXI',( .' '" "';"-k x " /. '7 7·x. 'f .. y.. ;..<
. {\ ... :.. . L..X'K .\, -...;.......;.--~ '/. 'j.:r.- ' xX
.. ·1.. -...:. " ',I'XyJ';'-/ -) X . /, X
. . '/'/'" .·tIYI'-i.~·. xl'---"-. "-.:. .. ' ...... '. 7. y. " . xy
::;;:, ,"x' ' x ... /.'r---..::. y ;:: '.',', .R' 77 ' , ' f
. ".', . I'X'I'X ...:.- ' ,...>.-N '>( l<,' ;,.. '" " ' ',,','.' 'x 'x '. x
, :f"<I'Xr->:.'.-\--", xr--.::.. -....: \: .. ' '7
' .... >( '7
.' " " , 'I'X'~'I'" ....;- .. ' xf--'-K X "x '.' , ,'. y y. ' ;/ Y
", yt--+-, X ',"'" , .. , " ' )<)<
c;o
>( f--i. .' -:. 'x' I---'-

(a) at t=O.125msec (b) at t=5.0msec


- - 1st layer ----- llth layer (rear face)
(front face)
(Analysis)

Rear face

&L--_\..l...--__\LL/_--L..-I.L-I ITop face

(Experiment)

Figure 5.30 Crack pattern of reinforced concrete handrail at failure.


130 PRECAST CONCRETE RAFT UNITS

(a)

(b)

Figure 5.31 Crack pattern of reinforced concrete handrail at failure; (a) front face, (b) rear face.

analysis not only makes clear the estimation of the crack pattern at failure with
a high degree of accuracy, but also shows the disorder in stress due to stress
waves in the early stages of impact.

5.6 Conclusions
The results with regard to performance improvement under impulsive loading
can be summed up as follows:
(1) A logical assessment of impact resistance of concrete beams can be
carried out based on the following items:
• Degree of improvement in load capacity
• Degree of suppression of drop in deformation capability at failure
• Degree of suppression of drop in rigidity
• Degree of suppression of critical local deformation
• Degree of suppression of increase in failure region.
CONCRETE RAFT UNITS UNDER IMPULSIVE LOADING 131
(2) The following indices can be utilized for evaluating impact resistance:
load capacity, deformation capability, total energy, curvature at failure
and the failure region.
(3) The improvement of ultimate tensile strain in concrete by means of steel-
fibre reinforcement is the most effective solution for impact resistance.
(4) The usage of high yield strength steel as reinforcement is effective for
impacts with high loading rates.
(5) On the whole, increasing only the concrete strength could decrease
impact resistance. This can be overcome by improving the ultimate
tensile strain of high-strength concrete.
(6) The introduction of steel fibre guarantees a sufficient energy absorption
capacity, not only due to its contribution to the structural resistance, but
also due to an increase in energy distribution capacity.
6 The behaviour of plain
undowelled raft-type concrete
pavement
M. POBLETE

6.1 Introduction
A raft-type pavement is a flat structure of plain cement concrete slabs,
characteristically thin with respect to length and width, placed over a subgrade
as schematized in Figure 6.1. Each slab has its own boundary conditions
determined by the support medium and by contact with surrounding slabs.
The contacts between the slabs, which are called joints, are capable of
accommodating shear stresses when the wheels of traffic pass across one slab
and on to the next.
Transverse joints are usually formed by a saw cut or by a thin strip partially
inserted in the fresh concrete during compaction. The aim is to weaken the slab
section in such a way that the tension stresses associated with the shrinkage
can break the slab at predetermined locations. Thus the interactions between
slabs lie mainly with granular interlock between the irregularities ofthe crack,
which are formed around the coarse aggregate in the concrete; particularly in
the case of the high-hardness siliceous aggregates. On average, joint spacings
are between 3.5 m and 6 m and occasionally more, often with a slant between 0
and 10 degrees to improve serviceability as well as to provide a gradual
transfer of loading between the slabs.
The pavement is supported by a substructure of selected materials, placed in
layers over the natural soil to improve the bearing capacity, as shown in
Figure 6.2. The pavement layers are gradually improved in quality, from the
roadbed soil up to the subbase, the last layer being of high shear strength and
very low compressibility, with the aim of providing a uniform, stable and
permanent support to the pavement slabs. The last layer also prevents the
damaging effects of moisture and frost action in the fine-grained soils effecting
the pavement slab [1]. If the roadbed soils are permeable, well graded,
granular and sufficiently compacted, no additional subbase may be needed.
These ideal conditions are seldom encountered in practice and a subbase of
selected granular material is invariably justified, as is the necessity of
recompaction of the roadbed soil. When the expected dynamic loads are very
heavy, as in the traffic of large trucks or aircraft, non-erodable bases, bound
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 133

TRAFfle

PLAIN CEMENT
CONCRETE SLABS

Figure 6.1 Basic elements of a plain undowelled concrete pavement.

~CONCRETE PAVEMENT

~0
~ ~ ~
/OPTIONAL BASE OF STABILIZED
MATERIAL (Trealed with cement.
o 0 () 0 0 0 Cl Q <' ,0 asphalt, lime or ather booer)

~~O~:oo::o»'~ci a:~
~ " _ ,~ "_ . _ . _ SUBBASE OF SELECTED MATERIAL

:-- ~. : : - " ;",:":':...":... :. .~. ~ROAD BED SOIL IMPROVED BY


IN-SITU RECOMPACTION

.~
. rv
NATURAL UNDISTURBED ROAD BED SOil
. ~

.~ ..

Figure 6.2 Typical cross-section of a pavement structure.

with Portland cement or asphalt, are sometimes found to be necessary in


addition to the subbase.
The existence of free water into the pavement layers, whether in full or
partial saturation, is, in general, highly detrimental. Special precautions are
taken to prevent the infiltration of rainwater, by sealing joints, cracks and all
other discontinuities, as well as providing quick and positive drainage devices.
Having stated the principal aspects ofa plain cement concrete pavement, the
following sections will give a description of the general behaviour that
characterizes rigid-jointed pavements within a number of environmental
134 PRECAST CONCRETE RAFT UNITS

conditions. Until very recently, rigid pavements were idealized as simple rafts
resting on invariable supports and reaching the limit of their working life after
a very predictable number ofload repetitions. However, practical evidence of
premature distress in many in-service pavements suggested that other
important variables, such as thermal and hydraulic effects in the slabs
themselves, were being either neglected or insufficiently weighed. These effects
arise as the pavement is more exposed to solar radiation than any other
cement concrete structure. Satisfactory comprehension and evaluation of all
these effects are essential to define and implement strategies for pavement
maintenance, as well as for the design of new pavements intended to last for
many years.

6.2 Laying of a plain cement concrete pavement


Shortly after the pouring of fresh concrete, chemical reactions associated with
the hydration of cement begin. The process is exothermic and increases the
temperature of the concrete mass above its initial value. The higher the
ambient temperature, due to solar radiation, the warmer will be the concrete
and the subgrade surface. Consequently, the maximum temperature of the
hardening concrete may exceed the admissible recommended limit of 32°C
[2,3]. A high temperature within the fresh concrete, accelerates hydration and
reduces concrete strength. Furthermore, rapid evaporation may exceed the
capability of the water retention of the curing compounds. This may cause
plastic shrinkage, crazing and the subsequent cooling of the hardening
concrete would introduce tensile stresses. During the night following construc-
tion, the hardening mass of concrete cools and strong shrinkage takes place.
The magnitude of the shrinkage depends on the coefficient of thermal
expansion, Ct, and the temperature drop.
Tensile stresses so developed first produce cracks at the predetermined
joints locations, usually one crack at every two or three nominal slabs. This
process proceeds gradually during the following day and night cycles until all
joints are activated [4]. As a matter of illustration, Figure 6.3 shows the
variation of opening of ten consecutive transverse joints on a typical highway
pavement at the age of 16 days, when five out of ten consecutive joints are
activated. After a longer time period, all the joints are active but maintaining
the identity of those that cracked first, which stay relatively more open, even
after several years of service, as shown in Figure 6.3b.
The process of shrinkage is a complex phenomenon, clearly described by
Neville [2]. Neville distinguishes between the irreversible first drying
shrinkage of the concrete mass initially saturated, and the reversible moisture
movements caused by alternating storage under wet and dry conditions.
Concrete pavements in the field are directly exposed to solar radiation from
BEHAVIOUR OF PLAIN UNDOWELLED RAfT-TYPE CONCRETE PAVEMENT 135
JOINT OPENINCilS (mm)
2.--- ---..

1.5

0.5

2 3 4 5 6 8 9 10

JOINT NU ..IER
(a)
2 JOINT OPENINCilS (mm)

1.5

0.5

O-'--'''P"---- L..-...... -_.---'''P"----"'I '--......---."


2 3 4 5 6 7 6 9 10
JOINT NU ..IER
(b)
Figure 6.3 Evolution of joints openings with pavement age; (a) TS-03 (Chile) summer 7.45 am,
16 days old; (b) 4.16am, 2.3 years old.

the very moment of pavement laying, thus the drying out by evaporation starts
very early and from the top surface downwards. Considering that the
pavement base remains humid, reaching at most an equilibrium moisture
content, a differential shrinkage takes place simultaneously with the process of
concrete hardening and joints formation. As a result, the newly laid slabs
undergo a process of deformation called 'warping'. Warping lifts up the slab
edges as soon as the concrete reaches sufficient stiffness to carry its own weight.
The magnitude ofthis initial warping caused by the early shrinkage is difficult
to ascertain, but it is believed to depend on the cement content of the mix and
on the rate of concrete hardening, which in turn depends on the temperature of
the fresh concrete mass [2]. Curing under a permanent source of water, like
flooding or ponding, simply delays the development of initial warping, which
is resumed as soon as the pavement is allowed to dry out. If the fresh concrete is
allowed to crack freely, the joints spacing will depend primarily upon the
136 PRECAST CONCRETE RAFT UNITS

temperature of the mix, the type of aggregate controlling the coefficient of


thermal expansion iX, plus the type and content of cement controlling the
hydration heat. These would explain why, based upon local experience,
different joint spacings are used in different countries and climates.
In summary, the plain cement concrete pavement is laid as a system of
interacting slabs of predefined geometry, with some initial upward concavity
or warping. The warping is expected to be partly relieved by creep, though the
high strength of pavement quality concretes probably reduces its contribution.
The practical consequence of the initial upward concavity is that the edges of
slabs tend to become and remain unsupported, unless a rather soft base
accommodates a 'seat' under the bottom surface of the slab.

6.3 Effects of temperature change and reversible moisture movement


Every day the concrete pavement is exposed to solar heating. Consequently,
the temperature of the slab rises, starting with the top surface and moving
down to the bottom through a heat conduction process. When the heating is
suspended by sunset or cloudiness, a reverse process of cooling takes place,
reducing the internal temperatures of the slab. Temperature reversal happens
throughout the day. According to the results of systematic measurements
using thermal sensors installed at various depths in concrete pavement slabs,
Figure 6.4 shows the typical internal temperature distributions, which are
representative of different times during a typical hot sunny day [5]. From these

TOPSlWACE

7 AM ';"
0.06
....E
-
..... 0.09
; /
,J::
~
Q) 0.12
0
HEATING
0.15
COOLING
0.18

0.21
15 20 25 30 35 40 45 50

Temperature ('"c)

Figure 6.4 Temperature distributions through slab thickness during a hot sunny day cycle
(TS-05 (Chile) summer).
BEHAvrOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 137
results it can be seen that the temperatures in the pavement are at a minimum
in the early morning hours. The maximum occurs typically in the afternoon,
around 1500-1700 h. The top surface temperature may be four times greater
than the bottom surface temperature, with the temperature distribution
through the depth being non-linear. During cloudy winter days the thermal
conditions in the slab are more stable; nonetheless, there is a tendency for the
top surface to be slightly cooler. This is due to the thermal inertia of the whole
substructure and the surface being continuously exposed to wind and ambient
temperature changes.
Similar data for the temperatures measured at different depths are expressed
in Figure 6.5, in terms of the time during any typical daily cycle. It may be seen
that the condition in the top surface of being cooler than the bottom tends to
predominate from late afternoon, through the night and until the next
morning, this condition lasting about twice as long as the opposite condition
and being repeated every day, with the only difference lying in the amplitude of
temperature variations. Similar behaviour has been reported for the climates
of Spain, Florida, USA, and other places [6,7]. Thus it may be considered to
be generally correct for a major part of the world. If the complete data of many
days are integrated throughout the full year, it can be concluded that, for
approximately two-thirds of the time, the pavement is in a condition of being
cooler at the top surface, which is defined as a negative temperature
differential. The positive temperature differential is restricted to one-third of
the overall time, though it may be of high amplitude during the hotter hours of
a sunny summer day [8].
To illustrate the typical effects produced by temperature changes, Figure 6.5
shows the simultaneous vertical displacements of five selected, instrumented
points on two slabs, together with the full thermal cycle ofa hot sunny day. The
continuous measurements, obtained by means of an analog/digital measuring
system, show that the pavement slabs undergo a process of deformation called
'thermal curling', going from the uplifting of the corner points 1,2 and 5 during
the night cooling when the top surface gets cooler than the bottom
[Figure 6.6a], to the progressive downward movement of the same points
when the pavement is heated by the sun in the next morning to end up with a
slight sinking of the corners at the time of the highest temperatures
[Figure 6.6b]. Meanwhile, the centre point, 4, behaves in the opposite manner,
providing support to the slab during the night and in the early morning hours.
Interestingly enough, it is the behaviour of point 3 which is consistent with the
upward concavity ofthe cold hours, but tending to lift up again when the slabs
are expanded by the high temperatures. The transverse edges are restricted to
rotation and, as a result, a pseudo-cylindrical shape is produced in the slabs, as
in the tiles on a roof, providing some support along longitudinal edges
(Figure 6.6c). It can be determined that the condition of a plane or flat slab is
merely the transition from one condition to the other, and is therefore of little
practical significance due to its very short duration. In the case discussed, the
138 PRECAST CONCRETE RAFT UNITS

. \_·4·_-1-(0)
1.2 ~---,------,,---,-----~---~--

E
0.8 .. \ ,V'l+ 3~
-+3.4 +3.6
5
~--"'/"';:;'....- \ - - - - + - - - - - j

.§ 06 SLAB CONTROL POINTS


OJ
c
; 0.4
a.
:::>
0.2

-0.2 L- L.- ~ __

4 PM o 8 AM 4 PM
Day Hour

50
51

0°....'"' Th. SENSORS 52

...
Q)

iii
...
40 !3l]:
+-~ 54
Q)
a. ,
55
E
Q) !
I- 30

-
Qj
...c
Q)

.5 T5-03 (CHILE) (b)


20
4 PM 0 8 AM 4 PM
Day Hour

Figure 6.5 Simultaneous vertical displacements of five slab points during a complete day
thermal cycle.

change in shape is produced under a strong posItIve heating, which is


indicative of a pre-existing moisture warping. On the other hand, the
traditional downward concavity shown in Figure 6.6b may appear only if the
transverse edges are free to rotate, but this is infrequent, occurring only when
the mean temperature in the pavement remains low enough during a sunny
day.
The joint openings, resulting from the expansions and contraction of the
slabs, constitute an important parameter controlling joint efficiency in its
function to transfer load from one slab to another. This movement is in
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 139
(a)

(b)

(c)

Figure 6.6 Typical shapes of slabs deformed by temperature, (a) upward concave slabs,
(b) and (c) downward concave slabs.

addition to the influences already discussed on restricting the edge rotations. It


is worth noting that, in general, for any given pavement the joint openings
measured at the middle plane change linearly with the mean temperature
according to the coefficient of thermal expansion (tl) of the concrete
(Figure 6.7). On sunny spring or summer days, when the temperature increases
sufficiently and the lower edges of the slabs are in contact, compression is
built up and the openings exhibit a non-linear decrease that tends towards the
complete closure of all the joints. During the extreme situations of very high
mean temperatures, together with low absolute openings, the compression in
the joints can be so high that the undersirable effect of buckling of an entire
pavement strip can be produced. This is known as 'blow up' of the pavement
slabs. When comparing the experimental data given in Figure 6.7 - obtained
at different seasons and therefore under different moisture conditions in the
pavement - the vertical distance between the winter line of openings and the
autumn line, considered as a dry reference state, represents the swelling of the
moist concrete [9]. Similar effects are evidenced by the behaviour of pavement
140 PRECAST CONCRETE RAFT UNITS

32 TS - 03 (CHILE)

• WINTER

28 o

AUTUMN
SPRING

N
C SPRING

b IiJ. WINTER
V WINTER
24
E
E
E 20
w
z COEFF. OF
c(

...
..J THERMAL
EXPANSION
...w 16
0

-Htl-ID-J
~

I-
:i

c( 12

'"z
C)
MIDDLE PLANE
z
...
w
0
8
I-
Z
Q ,
o L . . - - - - - - " - - -.........---"'O"'--O_ _----'
5 15 2S 35 45
MIDDLE PLANE TEMPERATURE (oCI

Figure 6.7 Longitudinal thermal deformation of a PCC pavement measured at joints.

deflections under load [4], as well as in laboratory tests or in field trials [10]. It
is important to emphasize that in pavements the hydraulic effects are
essentially differential, with the upper part of slabs predominantly drier than
the bottom, resulting in a seasonal upward warping which adds to the upward
curling already discussed. The same kind of conclusions have for long been
made in the technical literature [10-12], though not properly included in the
rational design methods [13]. This is probably because field observations at
some particular location give the impression that warping and curling are
subtractive rather than additive.
Considering the overall effects of climatic changes of temperature and
moisture, it is possible to identify three main components for the upward
warping and curling, that appear to predominate in arid climates:
(1) 'permanent' warping, which comes during the first few hours after
construction, due mainly to irreversible drying shrinkage;
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 141
(2) 'hydraulic' warping of seasonal variation, due to reversible moisture
movements in the concrete; and
(3) 'thermal' curling, which is modified by daily temperature variations.

In wet climates, with rainfall well spread over the year - particularly if the
rainfall is during the warm months - a constant and uniform moisture content
through the slab thickness can be reasonably assumed, and in that case
hydraulic warping could be ignored. In general, at any given day, 'permanent'
and 'hydraulic' warping are coupled, being almost impossible to evaluate
separately as reliable moisture data through the concrete thickness are usually
very difficult to obtain. This is opposite to what occurs with temperature only,
and the only way to decouple the effects appears to be to take some residual
moist reference state, if this can be assessed. The curling component is much
easier to determine through slab temperature measurements, for which multi-
sensors can be implemented in the slab [5], or by using thermodynamic
equations to deduce the entire temperature distribution in the pavement
thickness using the surface temperature.
To summarize, notwithstanding that all concrete pavements are construc-
ted to similar specifications, differences arise due to the environment and
climate of the place of construction, particularly when the drying conditions
produce warping and the consequent uplifting of slab edges and corners. To
diminish these detrimental effects some solutions have been proposed, other
than that of only increasing the slab thickness [9]. The most relevant solutions
are (a) the use of open graded bases connected to appropriate drainage, with
the aim of creating the conditions for a more uniform drying of the entire slab
in addition to the exposed top surface, and (b) the avoidance of concreting
during high ambient temperature so as to minimize initial joint opening and
warping.

6.4 Structural response


If there were no hydraulic warping or thermal curling the pavement slab
would rest flat upon the base with full support under every point and,
consequently, with zero internal stresses. In such a condition the traffic loads
would be the only source of stresses, their magnitudes being evaluated with the
aid of elastic theory. In this regard recognition must be paid to Harold M.
Westergaard (1888-1950), who developed analytical solutions for slabs on
grade, by assuming the support as a bed of Winkler springs. Many other
researchers have revised the classical Westergaard equations, which are
considered to form the foundation of almost all concrete pavement design
methods, with a view to extending the range of applications and making
adjustment to reconcile theory with practice. However, the improved
understanding of actual pavement behaviour is constantly introducing
additional variables, such as warping and curling which reduce and modify the
142 PRECAST CONCRETE RAFT UNITS

support and restrict the edges of slabs in a way that seems to disregard
Westergaard's straightforward approach. Fortunately, the powerful tools of
modern computers make it possible to apply superposition of wheel loads and
slab body forces with the desired sophistication. It must be recognized that the
problem involved is indeed very complex and, consequently, more research is
needed to clarify the many aspects that control the behaviour of in-service
concrete pavements and ultimately the overall pavement life performance.

6.4.1 Deflection
Once the pavement is built, the most direct structural response is the deflection
or vertical displacement of the slab due to the moving load. Measuring this
deflection by means of high-performance equipment, such as the Falling
Weight Deflectometer, the French Lacroix Deflectometer and others is usual,
but they all give displacements that are relative, more or less, to some
uncertain reference plane [14,15]. It is convenient to have absolute displace-
ments measured with respect to fixed references, ideally deeply anchored in the
subsoil. It is recognized that this approach is not feasible for systematic and
long-term pavement evaluations but in research activities it can produce
invaluable insight into the actual pavement behaviour.
Upon the application of a wheel load, the pavement slab undergoes a
complex pattern of deformation, similar to a rigid plate restricted by a range of
boundary conditions. For example, in the case of an upward-curled slab, a
wheel load applied at a corner will produce a slab rigid body rotation around
the supported slab centre; plus the straining within the elastic range of the
concrets itself and of the substructure. Therefore, the actual deflection of any
point on the slab has three components which must be borne in mind for the
appropriate interpretation of measured results. To show this behaviour,
typical recordings of the simultaneous absolute deflections produced by a
moving, loaded truck axle, at two adjacent corners and at the centre ofthe slab
are presented in Figure 6.8 for two different thermal conditions on a summer
day. At the corner points 1 and 2, two maxima are observed, representing the
deflections due to the passage of, first, the front axle and then the rear axle. A
slight rocking is detectable by the negative deflections produced when the rear
axle passes across the far edge of the slab. During the afternoon hours, when
the top surface is warmer than the bottom surface, the deflections of the same
corners are substantially reduced owing to the increased support and to the
compressive stresses developed by the expanded slabs due to joint closure. The
different magnitudes of adjacent corner deflections can be attributed to
different levels of upward curling, as well as to joint interlocking. The assumed
influences of the geometrical differences between slabs has proved to be oflittle
significance [4]. At the centre point, the deflections are much smaller, because
either the upward-curled slab at that point is well supported, or, in the reversed
thermal condition, the wheels passing near the edges do not directly pass over
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 143

HOUR MEAN TEMP.: 22.7 oC HOUR MEAN TEMP.: 22.2oC


06:06 A.M. TEMP. DIFF.: _7°C 05:12 P.M. TEMP. DIFF.: +7°C

0.00

0.25
~
--,
i rv
0.50
: i

E JOINT, \ SLAB
s
THICKNESS
0.23 m
-
.
ttl
0.75 I---
E ~
1.00 I--- -
z 1.25
\. -.f-4.56~4.06 +
0 CORNE ~ 1
1.60

....
f.)
0.00
~ r.... /'
w j
0.25
..J

"- 0.50
if
V
w
0.75 CORNE R2
c

0.00

0.25
......... ~
- TS-Ol (CHILE)
0.50 CENTER 4
o 10 15 20 25 o 5 10 15 20 25
TIME (sec) TIME (sec)

Figure 6.8 Simultaneous deflections under travelling load for two temperature conditions in
summer.

the centre point. Consequently, the peak deflection of the upward-curled slab
centre point is an accurate indicator of the subgrade reaction. Looking at the
data plotted in Figure 6.9, a direct relationship between peak absolute
deflection and corner uplift is clearly apparent, for both the winter and
summer environmental conditions. The summer environmental conditions
show a narrow loop differentiating the heating from the cooling branch, which
is considered to be another manifestation of the compressive stresses built up
by the high temperatures that restrict the free movements of the slab edges.
The variables discussed above can be brought together in the following
general expression for peak absolute deflection <5 of the corner as follows:
(6.1)
where d is the upward curling of the corner, induced by the moisture and
temperature differentials through the slab thickness; 00 is the deflection of the
144 PRECAST CONCRETE RAFT UNITS

MAX. DEFLECTION (mm) J


1.6 r - - - - - - - - - - - - - - - - . , - - - - - - ,

{\ -3.6
I
"
~2
III
-+-3.6~
!
!
I
1.2 _ _ _ _; _ _ · · · _ . _ . ,....HH"_••

Slab thlel: ness' 0.235 m :


cement Stabilized Base I:
:
i
i m(~
I •
!.•
!

0.8 --------.---,---.--l ---~.- _....L_. !.._-_._ _-_ __ __.


I I
I
1

• I
0.4 ~----,-,.-fT. ._+--------+----_ -

• SUMMER
+ WINTER

OL-----..L------'--------'-------'
o 0.4 0.8 1.2 1.6
UPLIFT OF CORNER d I (mm)
Figure 6.9 Slab corner displacements and maximum deflections during winter and summer day
cycles.

full supported edge; and m is a coefficient, that is characteristic of each


pavement and is related to the initial joint openings.

6.4.2 Joint efficiency


In undowelled Portland Cement Concrete [PCC] pavements, the load
transfer capacity of the joints derives its importance from the empirical and
theoretically accepted fact that the traffic-induced stresses in the slabs depend
upon the capacity of structural collaboration of the neighbour slabs, through
the interlocking developed at transverse and longitudinal joints. The eventual
deterioration of the load transfer would imply a progressive increase in the
internal stresses and also the initiation of slab cracking at a reduced number of
load repetitions. Consequently, the load transfer is usually considered as a
pavement condition variable [16,17], although the value of this variable
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 145
would not be unique as it depends upon thermal conditions, as well as
upon load position, being either on the approach or on the leave edge of the
slab [18].
The joint efficiency, or simply the load transfer, is well defined by the ratio of
simultaneous deflections of the adjoining edges when the load is positioned
near the joint. Although there are several definitions of related joint deflection,
the expression used in Figure 6.10 is simple and can be applied to a load
approaching, as well as leaving, the slab [16]. Results obtained from absolute
deflection measurements during a complete thermal cycle, in a typical Chilean
pavement, are used to illustrate that the evaluated joint efficiency varies
according to the thermal conditions of the pavement and its daily variation

LEAVE SLAB

~cJ; (t~\1 -_ -..


I"i~
1:1
I' I ..___.-.
_-
,. I ____

i! /------cfp)
SIMULTANEOUS
DEFLECTION ~. './
RECORDINGS I:'

JOINT
JOINT EFFICIENCY ('I.)

80 I---,..."q!:.-~+----t-'---r-
I
60 i - - - - - i i - - .-+\-----L-----i----t----H-:±:-J"---i

40 I
AXLE LOAD
80 J::N
20 1S'-0I (CHILE)
:I YEARS OLD
oL..-_ _.l..-_ _..I.-_ _...1-_ _- ' -_ _-L-_ _---'--_ _- '
2 PM 10 PM 6 AM 2 PM
DAY HOUR

Figure 6.10 Typical variation of the joint efficiency during a summer day cycle.
146 PRECAST CONCRETE RAFT UNITS

can be satisfactorily explained in terms ofthe actual transverse joint openings,


with no apparent influence of the slab thickness, type of base or any other
variable. Furthermore, the joint efficiency computed from the deflections due
to the load leaving the slab can be very different when the load is approaching
the slab, unless the joints are so tightly closed that the pavement behaves as a
continuous strip. This difference has been attributed to an inclined angle in the
induced joint crack, which can transfer a different proportion of load
depending upon which way the load approaches the crack [4]. As the thermal
cycles and load repetitions accumulate in the pavement, a tendency for the
acute lower edges to break has been observed in medium age pavements, thus
causing a deterioration in joint efficiency. In the long term a common residual
value would be reached, though still varying with each day's thermal
conditions and with the seasonal cycles of moisture swelling and drying
shrinkage depending upon the particular climatic location [9].
In short, deflections produced by the traffic loads are a direct manifestation
of the structural capacity ofa pavement, both ofthe substructure as well as the
pavement itself; however, the absolute magnitude of the deflection at any
moment is strongly dependent on the environmental conditions affecting the
support of individual slabs and on the efficiency of the joints which connect the
slabs and make an articulated structure. For the purpose of the structural
evaluation of jointed pavements, due account must be taken of the thermal
and hydraulic conditions at any given time, which can be an extremely
complex task, and therefore these evaluation activities should be attempted
during seasons of relative climatic stability, as can often be encountered during
cloudy or foggy days.

6.5 Effects of pumping


In concrete pavements the term 'pumping' is associated with the movements of
infiltrated rainwater that can saturate the base material and the slab/base
interface. This interface can be open at slab edges due to the upward slab
warping and curling. Due to the action of the wheel loads crossing a transverse
joint, water is moved in circular patterns by a mechanism of'pumping out' and
'suction' each time the slab edges are loaded and unloaded, as illustrated in
Figure 6.11. At outer corners, where the deflections are higher, the water
moves essentially backwards and at high speed, with a tendency to erode the
base surface underneath the 'leave edge' and to accumulate material under the
'approach edge' [19].
One of the essential condition that prevents pumping is the sealing ofjoint
and crack discontinuities with flexible mastic or inorganic compounds such as
silicon, polysulphide, polyurethane and others. However, in pavements with
poor joint efficiency the deflections of slab edges can induce such a large shear
deformation that most joint sealants will quickly fail unless they are put in
wide grooves designed to withstand the shear deformation. In recent years
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 147
(a) TRANSVERSE JOINT

1 STEPPING OR
FAULTING

SANDY FINES ACCUMULATION

(b) CIRCULAR /' "


MOVEMENT / \
OF WATER

WHEEL LOAD
\( ~)

CONDITIONS FOR PUMPING


- Open Interface
- Water

- Deflections I Poor Joint Efficiency

Figure 6.11 Slab faulting and pumping mechanism; (a) longitudinal section, (b) plant view.

much argument has arisen on whether to seal or not to seal new pavements
[20J, primarily because it is thought to be difficult if not impossible in the long
term to keep the joints sealed. For example, joint effectiveness is reduced when
the joints are most open during the low temperatures normally associated with
rainy days. For these reasons, instead of sealing, a quick positive drainage is
promoted by providing open-graded permeable bases and slotted pipes at the
outer sides of the pavement slabs to readily conduct the water off the road
[21]. The joint sealing is still considered useful in that it prevents the entry of
incompressible debris that would cause spalling of the joint and a consequent
reduction in pavement serviceability [3,9].
Pumping is also caused by cavities beneath the slab edges. These cavities are
148 PRECAST CONCRETE RAFT UNITS

produced by the upward slab warping on a rigid base. Pumping effects are also
observed at initially well-supported slab edges due to the pore pressures
developed within the saturated voids of the base material, initiating the
movement of solid particles and creating a cavity. To prevent this action
cement-treated bases with 4% or more ofcement (by weight) have been used, to
make them truly unerodable although at the same time stiffer [22]. In highly
rigid upward concave slabs the elastic deflections ofthe loaded corners (15 - Do)
given in equation (6.1), are sometimes less than the uplift d and the slab/base
interface does not close to cause erosion. This has been observed in the case
illustrated in Figure 6.9. In other places, where the warping and curling
appears to be unimportant, the erosion by pumping is reported to be a serious
problem and sometimes it is necessary to fill the cavities with injections of
mortar to regain corner support. In these kinds of climates the use of dowels is
widely accepted as a way of improving joint efficiency and diminishing the
corner deflections. In places where upward warping is significant, the efficiency
of conventional steel dowels is considered doubtful [9].
The most visible and measurable effect of pumping is the slab faulting
developed at transverse joints, as shown in Figure 6.11, which can severely
affect the pavement serviceability. Faulting between slabs of as much as 6 to
10 mm or more has been noted [23]. In general, the development offaulting is
due to the superposition of erosion voids under the 'leave edge' of the slab and
of an accumulation of sandy fines under the 'approach edge' [19]. The sandy
fines can be due to the erosion of the base, and most importantly they can come
from the shoulder material which has been sucked out during the elastic
recovery of the 'approach edge'. Other effects are the erosion holes developed
at the shoulder, in front of the slab corners and also produced by the water
being pumped due to the action of the wheel loads. Erosion starts systemati-
cally at the 'leave corners', destroying the shoulder surface finish and later
proceeding towards the 'approach side' at a more advanced state of distress.
With respect to the magnitude of the shoulder holes and of the slab
faulting, some dependence on the erosion susceptibility of the shoulder fill
material has been observed, it being more important in sandy fills than in
coarse granular materials. The controlling factor undoubtedly is the pro-
longed permanence of water trapped by low permeable shoulder materials,
sometimes known as the 'bathtub' effect [24]. To prevent these undesirable
effects from developing in new pavements, or to maintain them below some
threshold value in existing pavements, the effective action appears to be the
installation of maintainable edge drains to avoid the water trap and to stop
fines mobilization from the shoulder [3].
With regard to the base material, two somewhat opposite options are
currently considered; that is, rigid cement-treated bases and flexible open-
graded granular bases. Tn fact, cement-treated bases with 4-4.5% cement are
considered non-erodable and unable to loose sandy fines, but at the same time
they are impermeable, therefore restricting the water flow to be horizontal
BEHAVIOUR OF PLAIN UNDOWELLED RAFT-TYPE CONCRETE PAVEMENT 149
only. The granular subbases also permit some downward water flow. On the
other hand, cement-treated bases beneath upward-curled slabs provides a
rigid support restricted to the central part of the slab, increasing the volume of
water to be pumped from the slab/base interface. From these points of view,
rigid and impermeable bases would develop faulting more easily than granular
subbases, thus favouring the use of open-graded bases to permit a quick
positive drainage. The granular subbase could be treated with a flexible
asphaltic binder to provide an improved support under the whole of slabs. The
binder would also make as it truly non-erodable under all possible situations
of slab warping and curling.

Acknowledgements

Many of the concepts discussed in this chapter are the result of research conducted in the
concrete paved highway network of Chile. Grateful acknowledgement is due to the National
Highway Administration for their permission to use some of the data and to all collegues
associated with the research project at the University of Chile.
7 Rapid pavement repair using
precast concrete rafts
L.J. KENNEDY

7.1 Introduction
Concrete rafts quickly spring to mind when considering the need to replace a
pavement surface as fast as possible. This chapter examines the possibilities of
such repairs. The development of repair systems has been driven by the need to
achieve the rapid reinstatement of airfield pavements in war, after they have
been cratered by aerial bombardment. The primary requirements of such
repairs will be considered, followed by design considerations. The use of
concrete rafts for rapid pavement repair has led to the development of
specialist equipment which will be described in general terms. Also, details of
practical experience will be reviewed. Finally, consideration will be given to
the use of rafts for road pavement repairs plus possible future developments.

7.2 Preliminary considerations


The use of concrete rafts in repairs involves two principal areas of consider-
ation; the strength and integrity of the raft itselfand the functioning of the rafts
within a complete repair system. The raft must be designed to cope with loads
that arise during manufacture, transportation, stockpiling, installation, and
normal trafficking. The main considerations of raft design are covered in
Chapter 3, but it is worth noting that aircraft wheel loads can be particularly
severe. Heavy loads and minimum size wheels lead to very high tyre pressures.
For fast jets of the fighter type, tyre pressures up to 2.4 MPa (350 psi) are
possible. This is not generally a problem for concrete rafts, which are
manufactured to a high concrete strength specification, but it does mean that
raft foundation conditions are critical, especially when larger aircraft with
heavier wheel load assemblies are to be carried. Also, it is important to
consider carefully the design detailing when making rafts for airfield pa vement
repair, e.g. the shape of the raft edge. This aspect will be covered later.
When contemplating the incorporation of rafts within a repair, it is
necessary to have a clear idea of the site conditions and the repair process.
Bomb craters come in many types and sizes but the most common is the open
bowl type as shown in Figure 7.1. The pavement is ruptured and blown
outwards leaving a lip of pushed up pavement known as 'heave', debris is
RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 151

Figure 7.1 Typical bomb crater (Crown copyright).

True Diameter

Apparent Diameter

,,
Appamt Oe,1lt
,

Figure 7.2 Section through crater.

strewn around and loose material falls back into the hole (called 'fall-back').
The condition of the foundation soil will vary depending upon the soil type
and how it reacts to explosive blast. In any event, it will be in a variable and
unknown condition. Figure 7.2 illustrates the crater characteristics.
The repair is tackled in stages; including debris clearance, backfilling of the
crater bowl and providing the running surface of the repair. A critical aspect of
preparing the repair is dealing with the heaved pavement surrounding the
crater bowl. Debris clearance is carried out using conventional wheeled
loaders. Loaders with 'four-in-one' buckets (including a dozer and clam-shell
capability) are the most suitable. The heave is dealt with either by removal,
which is the usual method, or by pushing it back down. The latter is a
specialized technique still under development, but it is a process that can leave
152 PRECAST CONCRETE RAFT UNITS

a cracked pavement surface at the crater edge which affects the use of repair
surfacing techniques. Some surfacings are laid, carpet-like, over the repair and
can cover the cracks, while other methods, like rafts, need a good edge to lie
against. When the site is tidy, with debris cleared and heave removed, the
crater bowl may be backfilled either with imported ballast rock that does not
need compaction, or with selected debris pushed back into the bowl. Debris
backfill may be compacted using a dynamic compaction technique.

7.3 Surfacing types


The surfacing generally comprises an unbound layer of graded crushed rock,
that is compacted in a conventional manner using a heavy vibrating roller, and
a final surfacing. The final surfacing has been the subject of much research and
development, especially as the essential requirement is to achieve an
operational repair in the minimum time. The main generic types of surfacing
are described below.

7.3.1 Mat systems


Many current methods of crater surfacing employ mats made up of expedient
metal surfacing formed from planks of material. The World War II pierced
steel planking, which still exists in quantity, has been used, but it has little
flexural strength. Most users employ extruded or welded aluminium panels
designed to cope with soft ground conditions (down to CBR 4%). Usually the
panels are built into rectangular patches which can be rolled or dragged over
repairs and then bolted down. The principal disadvantage is that a step is
formed on the running surface, typically 32-38 mm thick. A US variation is to
use a fibreglass mat that is much thinner but this needs a well-graded base
layer which cannot be laid and compacted quickly and reliably in temperate
climates.

7.3.2 Flush repair systems


Flush repair systems are required to avoid the step problem and have become
increasingly necessary as aircraft have become heavier and more complex.
Mat systems impose operational limitations. Flush repair systems can range
from conventional pavement construction to novel systems using expensive
or complex binder materials. Conventional techniques using concrete and
asphalt gain strength too slowly. Many alternative options using resins,
polymers, modified bitumens and cements have been tested. Placing methods
have ranged between the use of premixed materials and the use of flood
grouting which has been put onto previously placed stone layers. Many of the
options have proved unsuitable because of problems with handling, finishing,
environmental limitations and excessive cost. One of the methods that
overcomes many of these problems is the use of precast concrete raft units (or
RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 153
'slabs', as they have become known in military parlance). This is not the only
favoured solution to the problem but it has been adopted by a number of
nations as a practical and safe method.
There are limited published data on crater repair methods because the
majority of research and development has been carried out by defence agencies
and recorded in unpublished reports. However, Kennedy [1] has covered the
principal elements of the repair process, as well as including a survey of
research and development activities in this field. Given the techniques that
have been evolved, the use of precast concrete rafts is a viable and accepted
method for crater repair surfacing. The remainder of this chapter is devoted to
the employment of precast concrete rafts.

7.4 Key factors in repair design


The key factors in the design of rafts used in rapid airfield repairs are loading,
durability and specific military needs. Other key factors, when set in the
context of the complete repair, are riding quality, installation speed, plus the
maintenance and repair of the repair system itself. All these factors are
considered below:

7.4.1 Loading
Rafts must be strong enough to sustain repeated loading by aircraft with very
heavy wheel loads and high tyre pressures. A minimum design requirement of
200-1000 passes by the design aircraft load over the bomb crater repair and
before maintenance is required might be specified. The wheel loadings could
vary as shown in Table 7.1
Loading data must include the aircraft main gear layout (single, twin, twin
tandem, etc.) with appropriate bogie dimensions (track, wheelbase), together
with the proximity of other main gear assemblies which would affect the
loading at a depth beneath the pavement.
The other principal factor affecting the raft loading is the foundation
condition. Given the circumstances of the site, the foundation is likely to be
strong but variable. Good quality base materials will have been laid but the
work will have been done hastily and soft or uncompacted areas may occur.
However, for design purposes, it is necessary to have a foundation strength. A
typical value would be 30% CBR and certainly not less than 20% CBR. An

Table 7.1 Typical aircraft loadings on pavements

Aircraft type Main gear loads (kN) Tyre pressures (MPa)

Tactical (e.g. Tornado) 100-140 1.7-2.4


Strategic transport and civil 700-1100 0.75-1.5
(e.g. TriStar)
154 PRECAST CONCRETE RAFT UNITS

alternative measure of base strength is the modulus of subgrade reaction (k)


and values in the range 100-240 kNjm 2 jm have been measured depending on
the specification and compaction of the base layer. The use of available data
that give direct correlations between CBR and k values is not recommended.
The thickness and strength requirement for the rafts can then be decided
empirically from past experience or by analysis.

7.4.2 Durability
Durability is a critical factor in many expedient repair systems but precast
concrete rafts are less prone to difficulties. The rafts are inherently durable
given an effective system of quality control during their manufacture. They are
generally made with high-strength concrete and cured under controlled
conditions before being loaded. Their reinforcement system controls cracking

Figure 7.3 Cracked raft after loading (Crown copyright).


RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 155
and they have excellent freeze/thaw, chemical and fuel resistant qualities.
Rafts, like all pavement concrete, inevitably crack in service (see Figure 7.3)
but experience has shown that the cracks do not deteriorate significantly with
continued trafficking. The slabs thus remain serviceable. Less certain is the
durability of the overall repair, where one is dependent upon foundation
performance. The principal faults are settlement and rocking under load, but
damage can be caused to the slabs themselves at the butting edges. Spalling of
the concrete is unacceptable due to the danger of damage to aircraft,
particularly their engines, from loose fragments.

7.4.3 Specific military needs


Generally, the requirements of civil and military aircraft are similar and
certainly this is so for the larger strategic transport military aircraft. They are
often civil aircraft in military livery or they are specifically designed for low
quality airfields (e.g. Cl30 Hercules or C5 Galaxy aircraft), and can cope with
rough pavement surfaces. The tactical fighters and bombers are like race-
horses, which have become finely tuned for maximum performance. This is
generally interpreted as performance in the air and as carriers of weapons.
Concessions to good performance in the ground run are often compromised
during the development phase, so that, in the extreme, long runways of fine
smoothness are required. Heavy weapon loads are hung under wings and
fuselage, where runway undulations can cause excessive airframe loadings.
Aircraft undercarriages are made as light and as small as possible which limits
their shock-absorbing qualities. Finally, tactical aircraft often need to use
aircraft arrestor systems and trail a hook to pick up the arrestor cable
stretched across the runway. When it is deployed, the trailing hook can bounce
and touch the runway surface. There is then a particular risk that the hook
might catch on the edge of precast concrete rafts and this hazard must be
considered.

7.4.4 Ride quality


The need for a smooth runway repair surface has been outlined. It is necessary,
however, to establish the limits of roughness that can be accepted in emergency
situations. In this context 'roughness' does not refer to the surface texture of
the repair but to the undulations in the repair profile and any steps or other
discontinuities. Undulations will be measured in relation to a datum based
upon the original longitudinal gradient of the runway. When precast concrete
rafts are used in repairs, it is clear that the step limit might be the most critical
one. Setting roughness limits is far from straightforward. Each aircraft type
will have a different response to runway roughness, also they can have a range
ofloadings and mode of operations. In the end, one must rely on the experience
of those involved in aircraft testing to provide acceptable roughness limits to
cover all aircraft types. Typical limits are given in Table 7.2.
156 PRECAST CONCRETE RAFT UNITS

Table 7.2 Typical limits of acceptable runway roughness

Maximum vertical deviation from a straight line datum joining sound ±30mm
pavement levels each side of repair
2 Maximum vertical deviation from peak to trough in the level profile ±30mm
across the repair (related to the same datum as in I).
3 Maximum step between adjacent slabs or slab and sound pavement over ± ISmm
a distance up to 300mm

The roughness limits given in Table 7.2 are meant to be 'carefree' limits, such
that landing and taking-off aircraft need not take any special precautions
when trafficking the repair. Greater limits are possible, especially for aircraft
designed for rough field landings. Vertical profile deviations in the range 50-
60 mm might be acceptable with steps up to 30 mm. Care should be taken to
measure repair performance under load so that elastic deformations are
considered. In addition, note should be taken of any raft tendency to rock
under moving loads. Dynamic effects caused by aircraft moving at speed
should be appreciated. Raft bounce and translation under braking loads can
also become a hazard.

7.4.5 Installation speed


The best raft installations are those where care is taken to prepare the raft
foundation. However, the requirement of rapid pavement repair is for a
minimum time to be spent on each work phase. The net result is a compromise
where adequate short-term performance is achieved for minimum preparation
time. One advantage of raft repairs is that individual rafts can be lifted and
reset to accommodate settlement or raft tilting, either initially or after a period
in service. Raft bomb crater repairs might typically be expected to remain in
service up to 90 days before permanent repairs can be undertaken. The speed
of raft repairs may be increased by the use of special equipment and well-
practised procedures. These aspects will be developed further.

7.5 Repair design


The repair comprises a stable backfill, a base layer and a surfacing finish. With
precast concrete rafts as a surfacing, the repair designer must concentrate upon
the form and preparation of the base layer and on any special treatment of the
bulk fill crater filling. The principal factors affecting the foundation design are
the loading and the construction method appropriate to rapid repairs. These
aspects need some explanation.
The most straightforward loading is that caused by single wheel gear loads.
For tactical military aircraft with very high tyre pressures, the effects are
accommodated mainly in the pavement layer, i.e. the precast concrete rafts,
and the base layer below. Multiple wheel main gears on large commercial or
strategic military transport aircraft have lower tyre pressures but overall apply
Aircraft: Phantom F4J Max. Landin2 Weight: 46 OOOlb Aircraft: Tornado GR Mkl Max. landing Wei2ht: 37 500lb
Sin21e Main Gear Load: 43.9% [20861.67kg) Sin21e Main Gear Load: 47.7% 117006.80kl)
Max. AUW: 56 OOOlb Max. AUW: 63 OOOlb
125396.82kgl 12857t.42kg)
~
Wheel Arran2ement Wheel Arrangement >
'"d
, a
'"d
b I a. 23ft 4in [7.11m) I a. 17ft 3in 15.25m) >
:t:, b. 17ft llin 15.46mJ b b. 10ft 2in [3.09ml <
m
~
...,Z
;:Q
a m
I. · .j '"d
>
.~ ;;3
Aircraft : Boeing 747/200C Max Landing Wt : 629 860lb Aircraft: Tristar K Mkl/KC Mkl Max landing Wt : 368 OOOlb
Single MaiD Gear Load : 23.1 % [285 650.79kg] Single Main Gear Load: 45.7% 1166 893.42kgl 5i
Max AUW : 823 OOOlb Max AUW : 542 OOOlb z
Q
. '1 (373 242.63kg] 1245 804.98kgl '"d

Wheel Arran2ement Wheel Arrangement ~


':'~'c
.. ;"'T n
~
...,
a. 73ft 6iD 122.40m) ii a. 70ft Oin 121.33m]
b. 33ft 7in 110.23m) b. 36ft Oin 110.97m)
8
z
b e n
c. 3ft Sin 11.04m] c. 4ft 4in [1.32m) ;:Q
m
d. 4ft Sin [1.34m] r d. 5ft lOin 11.17m) ;j
e. 11ft 7in (3.53m]
-.-'.-
:t~± ;:Q
f. 9ft 4in 12.84m) >
~
a

Figure 7.4 Aircraft undercarriage arrangements. VI


-.J
158 PRECAST CONCRETE RAFT UNITS

much greater loadings. These are spread through complex undercarriages so


that the vertical stresses from the loading are felt at greater depths. Typical
undercarriage arrangements are shown in Figure 7.4. So for larger aircraft the
designer must take care to provide strength at depth.
The need for fast construction militates against the normal civil engineering
approach, where construction materials are carefully prepared, laid and
compacted with a quality control scheme to protect standards. As far as
possible, compaction is avoided or minimized and materials are selected to
optimize performance. High quality materials such as crushed rock can be
stockpiled for the emergency situation. Designers also need to consider new
materials, such as geotextiles, that will improve performance. These points
have to be borne in mind when considering construction of the repair.

7.5.1 Bulk fill


There are two basic approaches to bulk filling. The first is the clean bowl
technique. Here, the crater is cleared of fall-back material and filled to the
appropriate level with single-sized ballast rock (size range 40-100 mm but
usually 50 or 60 mm nominal size). A geotextile may be used as a separation
medium to line the hole, but should only be used for a good engineering
purpose, e.g. where poor subsoil conditions exist. Apart from cost, the
difficulty of handling geotextiles in windy conditions should be borne in mind,
and the time taken should be justifiable. The second approach, used by UK
military engineers, is where selected crater debris is pushed back into the crater
with any fine-grained subsoil material (i.e. clay) being removed. The debris is
compacted with dynamic compaction equipment and a regulating layer of
ballast rock is usually added to achieve an acceptable top level. It is well worth
applying some rolling compaction to the regulating layer to stabilize the top
surface.

7.5.2 Base layer


There are two basic options for constructing the base layer with a series of
variations within each option. One option is to use a graded material that is
compacted. The second option is to use a single-size material, smaller in size
than that used for bulk filling, which does not need compacting. The base
materials are usually specified as good quality crushed rock but naturally
occurring gravels may be satisfactory. Graded materials are sometimes
specified with no fines. There are a number of reasons for this: (a) the material
can be maintained in emergency stockpiles near the repair site for many years;
(b) it is free from the problems of migrating fines; (c) the material is easier to
screed level with a scraping blade or beam; and (d) the material can be
compacted in wet conditions, i.e. wet of optimum. The disadvantage is that the
final result is not as dense or as strong as a more fully graded material. A strong
surface layer is needed to distribute point loads, especially if a single-size
material is used.
RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 159
Single-size aggregates are particularly suitable for rapid repairs, where their
use leads to savings in time. They are easy to stockpile, place and screed level
and do not require compaction. They do need the containment of the existing
surrounding pavement and subbase layers to allow them to function. Often,
there is a case for laying a geotextile separation layer beneath the base layer.
Once the base material is chosen, a decision is needed on the layer thickness.
As with all pavement design, great reliance is placed upon past experience and
the guidelines derived from it. It is at least possible in repair design to carry out
testing trials to evaluate the performance of repairs using a range of base
thicknesses, with representative trafficking loads. Tn practice, base designs for
tactical aircraft have varied from 100 mm of 10 mm stone to 200 mm of 20 mm
stone and, for the larger aircraft, from 125 mm of 6 mm stone to 300 mm of a
well-graded and compacted material. More details of full repair systems are
given later.
The other component in repairs using precast concrete rafts is that of a sand
bedding layer between the base layer and the rafts. Raft manufacturers usually
specify a sand bedding layer based upon their experience with civil engineering

0I1111ll1 PlYlIIIII
CmltutliOi lall

----- ----- /",---


---
/
hlk nu / Basi laytr

.- .-
/

----- .../
Figure 7.5 Section through completed repair.

Figure 7.6 Partly completed repair (Crown copyright).


160 PRECAST CONCRETE RAFT UNITS

projects. With emergency airfield repairs the sand layer can be a source of
settlement. Some settlement might be acceptable where large areas of rafts are
laid for hardstandings but where rafts are set into an existing pavement the
differential settlement effect is deleterious. Sand bulking is also a problem.
Where a bedding layer is needed, e.g. to allow for an uneven surface of the base
layer, it should be kept as thin as possible. A practical minimum thickness is
20mm.
A typical section through a repair is shown at Figure 7.5. This summarizes
the features described above. A part completed repair is shown in Figure 7.6.

7.6 Raft design


The basic design of rafts to withstand loading is covered in Chapter 3. This
section will deal with aspects specifically related to rafts used in emergency
pavement repairs.

7.6.1 Raft size


Most practical work has concerned rafts that are 2 m square. This is because
many tests and trials have used the proprietary rafts marketed by Stelcon and
their licensees. Their standard design is for a raft 2 m square and 140 mm thick.
The rafts are factory made using high-strength concrete (typically, 55 N/mm 2
or more) with steel mesh reinforcement in the top and the bottom. Two lifting
holes are incorporated within the raft area where lifting keys may be inserted
and then removed after placing. The standard thickness of 140 mm has
satisfied most needs but UK military engineers have had 200 mm thick rafts
produced to cater for large wide-bodied aircraft.

Figure 7.7 Surface spalling at butt joint (Crown copyright).


RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 161
7.6.2 Raft edge detail
The main proprietary raft has steel angle edging to protect the upper arrises
during handling and when in use. Rafts for use in emergency repairs on
military airfields are specified without the angle edging. This is because test
experience has shown the danger of aircraft arrestor hooks catching on the
steel angle edging. In some respects, this is an unfortunate modification. Rafts
obviously become more susceptible to handling damage but there is also the
problem of edge spalling while in use. It is normal in standard civil engineering
situations (for hardstandings) to lay rafts with a 5 mm gap between them by
using suitable shims. In pavement repairs, the dynamics of trafficking causes
the rafts to move, both horizontally and vertically. It is usual practice,
therefore, to place rafts to butt against each other to minimize horizontal
movement. With rafts touching each other, stresses are caused because
pressure is developed between rafts as loads move across them. If the pressure
point is too near to the raft surface there is a likelihood of spalling. The point is
illustrated in Figures 7.7 and 7.8.

---Rollinr Lnad

(a)

(b)

--- --- .......


/'--- ................
Lonrer Shear Path

Figure 7.8 Edge details; (a) bad detail, (b) good detail.
162 PRECAST CONCRETE RAFT UNITS

The problem arose because a raft manufacturer had a draw angle on the raft
edge to ease the vertical lifting of the raft from the mould. To achieve the
desired edge, rafts must be cast using removable side shuttering, as was done
with the older 150 mm thick rafts. The recommended detail shown in
Figure 7.8 permits safe butting between rafts. This design of raft is covered by a
UK Patent (held by the UK Ministry of Defence).
The other aspect of edge detailing is the expectation a designer might have of
transferring vertical loads between rafts. Some form of positive interlock
design is possible but would increase the manufacturing process complexity
and also increase raft-laying time. Early proponents of rapid raft repairs
stressed what they called a 'keystone' effect where vertical loads between rafts
were transferred by means of sand filling in the joints. In part, sand was pushed
up into the joints during raft bedding activity. Sand was also brushed into the
joints. Loads would then be transferred through the sand, relying on the sand's
shear resistance and to some extent the inverted-vee shape of the joint. Sand in
the joints is a very effective load transfer system for small blocks (of brick size)
and is widely used even on airfield runways. However, it is doubtful if the
benefit extends to the larger raft systems. Dynamic effects can cause transient
vertical movements of 10 mm or more at the joints. It is also in the essence of a
rapid repair system that the work can be carried out rapidly without particular
care being taken. Careful sand filling of joints is time-consuming work. The
consensus of opinion for rapid repairs is to butt joint the rafts and use no sand
filling. Also, load transfer between rafts is assumed not to occur.
It is important to note that the inverted-vee shape at the bottom of the edge
detail (see Figure 7.8) has an important function. Rafts often tend to be laid at a
small nose-up angle (see Figure 7.9) and as they rotate into position, bedding

Figure 7.9 Raft laying using lifting beam (Crown copyright).


RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 163
material can be pushed up into the joint. The lower space in the joint
accommodates this material without forcing the rafts apart.

7.6.3 Raft reinforcement


Rafts can be reinforced in the bottom layer alone when only light loadings are
considered, e.g. warehouse floors, but for normal airfield applications top and
bottom reinforcement is necessary. Reinforcement design is usually under-
taken by the specialist raft manufacturers, often with advice from consultants.
Obviously over the years considerable experience has been gained, which has
led to optimal and economic designs. Consideration must be given to a range
of loading conditions, starting with the manufacturing process, going through
transportation, stockpiling and installation to the expected service loadings.
With rafts designed for airfield use, the heavy concentrated loads due to
aircraft undercarriages, are the predominant and ruling loading condition.
The raft designer must consider the worst case loadings, with the
undercarriages placed to produce the maximum effect for both the hogging
and sagging moments. These moments define the amount of reinforcement
needed in the top and in the bottom of the raft. Large heavy undercarriage
bogies cause significant hogging moments as they roll across the rafts so that
even strong well-made rafts may crack in the top surface (see Figure 7.3).
Typical bogie layouts are shown in Figure 7.4. The designer must also
consider the foundation support provided by the base and subgrade layers
before arriving at the reinforcement design.
It is normal to have equal weights and spacing of reinforcement in both
directions and in the top and bottom raft layers. Some designs have less
reinforcement in the bottom layer. Cracking is controlled by creating closely
spaced reinforcing mesh arrangements. Usually, cold drawn wire reinforce-
ment is used with welded joints. The density of the mesh is increased at the edge
zones usually by reduced spacing. This is because higher moments are
developed near edges and corners. The requirement is confirmed by experi-
ence. Reinforcement sizes range from 5 mm to 11.5 mm diameter with mesh
spacings from 50 mm to 250 mm centre to centre. A typical design for heavy
aircraft loadings is 8 mm wire at 150 mm spacing, increasing in the highly
stressed edge zones.

7.7 Repair boundary


It is possible to achieve very accurate levels on rapid raft repairs by a
combination of good technique and special equipment (to be discussed later).
The problem area is invariably where the repair abuts the original pavement.
This might ordinarily be a difficult problem but, in this case, it is compounded
due to the surrounding pavement having been subjected to explosive shock
and, perhaps, some form of heave reduction treatment. In any event, heave of
164 PRECAST CONCRETE RAFT UNITS

up to 10 mm is accepted in the original pavement to avoid cutting back even


further. Another area of variation arises because it is necessary to cut back the
original pavement to a square or oblong shape in plan to fit the raft format and
to allow the use of rapid screeding equipment (see Figure 7.6). Craters
generally have a circular plan shape and it is clear that uneven edge levels can
arise. It is not possible to completely overcome these difficulties. The aim is to
keep level variations within the step allowance and this needs particular care in
setting the top levels of the base layer and in using a good bedding-in technique
for the rafts. It also highlights the importance of selecting the original
pavement edge to be sound and reasonably constant in its level variation.
Another feature of the repair boundary is the gap between the rafts and the
original pavement. In general, this must be filled with a strong and durable
filler but, given the tendency for rafts to move under dynamic loading, some
flexible filler material must be incorporated. It is desirable to minimize the size
of the gap. However, to achieve a minimal gap all round, great care is needed
with the preparatory work and this does not accord with the emergency nature
of the task. Firstly, the cut lines must be measured and marked on the
pavement and then the pavement must be cut with high speed diamond-tipped
saws. An allowance must be made to account for setting out inaccuracies,
deviating cut lines and hasty raft laying. A normally accepted gap provision is
100mm all round. It is desirable to reduce this gap if possible. One solution is
to lay rafts close up to the pavement on two sides at least, providing the sides
are reasonably straight and limit gap-filling to two sides only.

7.8 Special equipment


Special equipment has been developed to optimize raft-laying as a rapidly
emplaced repair system. The principal equipment is a special subsurface screed
beam for achieving base and bedding layer levels. Another is the raft-lifting
beam that enables up to three rafts to be carried and placed from a front-end-
loader bucket. Finally, the system introduction has led to the development of
high-speed saws capable of cutting pavement quality concrete 300 mm thick at
a rate of 1mlmin, or even quicker.

7.8.1 Subsurface screed beam


The subsurface screed beam was developed by a Dutch company, Bruil-
Arnhem Groep BV, to meet the specific military requirement for the rapid
repair of airfield pavements using raft units. The purpose of the equipment is to
screed aggregate layers at a depth below the general pavement level within the
confined area of the repair. The overriding aim is to do this activity quickly,
accurately and to permit faster repair times than would be possible using
conventional equipment and methods. Bruil developed their Multilevel
Equalizer Type 82 APM 161M as shown in Figure 7.10 to meet the
specification of the United States and Dutch airforces.
RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 165

Figure 7.10 Screeding with sub-surface screed beam (Crown copyright).

The equipment has eight 2 m long screeding plates that can be individually
lowered to cover repair widths up to 16m. The length of the screeding plates
was selected to conform to 2 m square rafts.
The beam is supported on bogies at each end with castored wheels. It
is moved by towing or pushing the bogie top frame using tractors already
employed on site. The company can also provide a cantilever system to screed
6 m outside the bogie and al10w craters wider than 16 m across to be tackled.
However, a running strip of rafts must be laid to carry one bogie.
The beam is moved between crater sites by towing it 10ngitudinal1y from one
end. It is possible to wind up each screeding plate to give adequate ground
clearance.
The beam needs careful handling. The surface to be screeded must be laid to
a slight surcharge and the screed plates must maintain a small 'bow-wave' to
ensure that low spots do not occur. A good deal of hand work with shovels is
required to maintain an even surcharge. Care is needed in pushing the beam to
keep it square and to avoid it binding with the repair edge. Finally, the last part
of the surface, say about 1 m wide, must be levelled by hand.

7.8.2 Raft-lifting beam


The raft-lifting beam is an ingenious means of carrying up to three 2 m square
rafts from a stockpile or a fiat bed vehicle, such that the three rafts can be laid
directly into position at the same time. The lifting beam is illustrated in
Figure 7.9 and it was developed by Bruil of Arnhem to meet a Dutch airforce
requirement. The beam is carried on and fixed to the bucket of a medium-sized
front-end-Ioader. The rafts are stabilized during movement by steel pads
which hold them 'nose-up'. This attitude also assists laying. The angle of the
166 PRECAST CONCRETE RAFT UNITS

Figure 7.11 Saw cutting concrete pavement (Crown copyright).

rafts can be varied by crowding the bucket. The number of rafts carried
depends upon their individual weight, which for 2 m square rafts is related to
their thickness, and the capacity of the loader. The attachment chains allow
rafts to be carried in ones, twos or threes.

7.8.3 Concrete cutting saws


The repair system calls for a pavement cutting capability, so that craters can be
cut back to an oblong shape to suit the raft-laying and the subsurface screeding
operation. Pavements will be constructed in pavement quality concrete,
Marshall asphalt (asphaltic concrete) or a combination of the two. A typical
cutting target is to cut a concrete slab 300 mm thick at the rate of 1 m run per
minute. Under the impetus of the military requirement, a number of
manufacturers have produced suitable models. One is illustrated in
Figure 7.11. Such a saw would have a power source in the range 50-95 h.p.
(37.5-71.3 kW) and be capable of carrying a diamond-tipped saw blade of
1000mm diameter or more. The saws need a separate water supply but
incorporate their own pump.
Pavement cutting is a skilled operation if a straight line is to be maintained.
Also, it is important to match the saw blade to the material being cut. One
must not only consider the asphalt and the concrete, but also the aggregate
type in both materials plus the binder characteristics. It is necessary to seek
expert advice from saw and blade manufacturers and, if possible, to carry
out trial cuts on the pavement such that data are available before the
emergency arises.
RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 167
7.9 Practical experience
All practical experience to date has been gained in tests and trials involved in
developing rapid repair systems for military use, though some work on civil
road repairs using precast rafts has been carried out in Holland. The
procedures outlined in the preceding sections have evolved from these various
tests, which have ranged from small-scale field trials to full repairs on
explosively prepared craters. Repairs have been tested with special load carts
simulating aircraft wheel loadings as well as with actual aircraft. Aspects of the
tests are given in this section.

7.9.1 Early tests


During the late 1960s, tests were carried out at the Military Engineering
Experimental Establishment at Christchurch to examine the use of Stelcon
rafts for Rapid Runway Repair. This followed visits to the Southampton
Container Terminal where rafts were being used for container hardstanding.
The foundation used at Christchurch was 75 mm of 10mm Lytag laid on
shingle. A sand bedding layer was used and sand was brushed into joints. No
compaction was used. After trafficking with a load cart simulating a fighter
wheel load, there were settlements of up to 38 mm and raft tilting. Plate
bearing tests showed low bearing values. This inadequate system was not
developed further.

7.9.2 Recent developments


Rafts came back into consideration in the early 1980s, when the need for a
flush repair surface became important. The initiative was led by the US Air
Force in Europe. The German and Dutch airforces developed similar systems.
The equipment developments outlined earlier allowed a feasible repair
method to be evolved. The design comprises ballast rock bulk fill, with a
geotextile lining in the crater bowl, and a base layer offine gravel (10 mm single
size) 100-150mm thick, laid on a geotextile separation layer. Compaction is
used on the top ofthe ballast fill (two passes of a 10 t vibrating roller), but not
on the base layer, which isjust screeded level. The rafts are laid and bedded in
with several passes of the vibrating roller per row of rafts. Care is needed in
setting the base top level to achieve an acceptable final raft level. This is really a
matter of trial and error on a particular airfield with particular materials, but
bedding in settlements of up to 40 mm are typical. A British variant of the
design shows a more conservative approach, using a base layer of20 mm single
size stone 200 mm thick.

7.9.3 Heavy aircraft tests


The design described above is for tactical aircraft, but in 1985 both US and UK
research agencies conducted major trials to develop systems for larger aircraft.
168 PRECAST CONCRETE RAFT UNITS

- ,
~.
.

Figure 7.12 Raft repair during aircraft trials (US Air Force Engineering and Service
Laboratory).

The US tests concentrated on repairs suitable for military strategic transport


aircraft (such as the C5 Galaxy), that is, aircraft with high flotation
undercarriages and some 'rough field' capability. The UK considered civil
wide-bodied aircraft types (such as TriStar).
The US tests included a raft repair with only minor modifications from the
standard design already mentioned. Firstly, the top of the ballast fill was more
thoroughly rolled (six passes) and, secondly, the rafts were well bedded by
proof rolling with a load cart. A large number of aircraft movements were
made across the repair, which performed very well as shown in Figure 7.12.
Settlement and rocking movements were no greater than 25 mm with the most
critical area being at the junction between the repair and the original
pavement. Other points to note are that the crater was machine dug, so that
explosive effects were absent, and the base stone was a 6 mm pea gravel,
smaller than the normal 10 mm size.
The UK tests were carried out on a range where craters could be blown. It
was a major trial which included four raft repairs and the aim was to develop a
satisfactory design that could cope with wheel bogie loadings up to 100 t.
Evaluation of the special equipment described earlier was a key element. Only
load cart testing was possible but this allowed demanding trafficking of up to
100 passes with the heavily laden bogie. The repair design was developed, in a
conservative way, from a civil engineering standpoint. The principal require-
ment was for 200mm thick rafts which were specially manufactured. The 2 m
square format was retained and an edge detail was specified to facilitate side-
butting at the joints. The crater fill was dynamically compacted using a special
falling-weight technique but the top of the fill was dressed with a regulating
RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 169

Figure 7.13 Edge gap filling with fast curing mortar (Crown copyright).

layer of imported stone, which was rolled. No geotextiles were used as the site
and repair procedure did not need them. The base layer was 300mm thick and
consisted of a fully graded crushed rock (to an airfield pavement specification
for unbound base layers and of 37.5 mm maximum size). The base was fully
compacted and a sand bedding layer was added under the rafts. Finally, the
rafts were bedded in with a roller.
The slab repairs performed adequately. Most settlement was immediate, so
that thorough bedding-in paid dividends. The sand bedding layer was
considered to be a source of settlement and, ideally, it should be eliminated. All
the step problems occurred at the junctions between the repair and the cut
back pavement. Clearly, great care is needed in this area. Also, a variety of gap-
filling techniques were tried (see Figure 7.13). These included special fast-
curing cement mortars and asphalt mixes but the fibreboard strip was essential
to permit slab movement under load. The fault with casting the slab edges out
of vertical has already been mentioned and this led to surface spalling. On one
repair, the base layer was replaced by a 5 mm pea gravel that was laid without
compaction. It performed badly under the very heavy loadings and was not
examined further. Finally, special prestressed rafts were tried. They measured
5 m by 2 m in plan. They worked well enough, though there was a danger of
excess rocking movements on the long side.

7.9.4 Current experience


A number of countries have adopted raft repair methods for emergency
runway repairs and they train regularly. Proving tests have been made with
aircraft. One problem that has arisen in training is the wear and tear on the
rafts through repeated use. This is generally in the form of chips and spalls
170 PRECAST CONCRETE RAFT UNITS

around the edges, problems that would be avoided with the traditional Stelcon
raft with its angle iron protection. One solution proposed is the incorporation
of polypropylene fibres in the concrete mix to improve the surface toughness of
the raft, but this solution has not been tested so far.

7.10 Road repairs


The company which manufactures the special raft installation equipment,
Bruil of Arnhem, has marketed a system for rapid road repair. Their concept is
to use large rafts that would fill a road carriageway bay but making use ofthe
general procedures developed for military purposes. The concept is illustrated
in Figure 7.14. The raft sizes are 5.1 m by 3.6 m by 0.2 m thick. They have been
used not only for the rapid replacement of sections of roadway but also for new
roads where speed of laying is a critical factor. A load transfer system using
dowels has been developed, but this increases manufacturing complexity and
raft laying time. A thoroughly researched design method for 2 m rafts used
for roads and road repairs is also available from the UK raft manufacture
Redland Precast Ltd.

Figure 7.14 Road repair system (Bruil of Arnhem).


RAPID PAVEMENT REPAIR USING PRECAST CONCRETE RAFTS 171
7.11 Future developments
The main limitations on the use of precast concrete rafts for rapid pavement
repairs are the time taken to prepare the foundation layers and the dangers of
uneven settlement and tilting of individual rafts. While rafts are relatively
cheap, the special installation equipment is costly. Where considerable
concurrent activity is necessary, such as on an airfield in war, the capital
investment needed is high although very much less than the cost of any single
aircraft using the airfield. Another disadvantage is the need to cut back
pavement to an oblong shape. This removes much pavement that is otherwise
satisfactory. Future developments need to be aimed at removing these
disadvantages.
There is scope for studying the format of the raft itself. Most practical work
has depended upon the proprietary Stelcon raft that is 2 m square and 140 mm
thick. Rafts of different shape, size and thickness could be examined and
indeed research into this area is being conducted at the University of
Newcastle upon Tyne [2]. Load transfer devices should be considered, aiming
for those that are cheap and simple to incorporate and use. A particularly
difficult area is the junction between the repair and the original pavement.
Most unacceptable settlements or steps occur there. Some form of levelling or
load transfer device would be a considerable benefit.
There are potential advantages in avoiding the need to cut back good
pavement by setting rafts within the natural repair shape, which is rougWy
circular, using some form of filling or grouting to bridge the large uneven gaps
that would occur. This was tried by UK research engineers in 1985 but the idea
was not pursued as it did not allow the efficient use of the special screed beam
and required two different repair systems on site. The result was a combination
of difficulties rather than an overall improvement. However, further study of
the options might bring dividends.
In final summary, precast concrete rafts provide a good, practical method
for the rapid repair of pavements. The use of standard 2 m square rafts has
been optimized for rapid repairs by the development of special installation
equipment. Foundation designs have been developed using small single size
stone as a base, that can be easily handled and screeded. Numerous tests and
trials have led to solutions of areas of practical difficulty, such as the treatment
ofthe edge gap. It has been possible to prove the various repair methods by full
load testing using load carts and aircraft. Finally, numerous nations have
invested in full raft repair systems for military purposes. The scope for civil use
is there, where speed is critical.
8 A review of the analysis,
design, manufacture and use of
precast concrete raft pavement
units
J.W. BULL

8.1 Introduction
In this, the final chapter, it would be prudent for the editor to draw together the
main threads that have run through this book. With these threads it is possible
to develop a precis of good engineering practice for the analysis, design,
manufacture and use of precast concrete raft pavement units (raft units) and
also to suggest areas for future research and development. The fact that this
book is being published, indicates that the use of raft units for airfields,
highways and port pavements plus floor areas in heavy industrial situations is
steadily increasing. Raft units have distinct advantages over rigid pavements
where loading and load contact pressures are high and where subgrade CBR
values are low.
Raft units are more costly than conventional concrete pavement construc-
tion and the ride quality is inferior, but raft units have distinct cost advantages
when used for temporary roads and emergency repairs as they are easily and
quickly laid and moved in adverse weather conditions, plus being re-usable. In
areas of substantial subgrade settlement, the raft units can be lifted out and the
subbase relevelled. If the applied loading is increased, the existing raft units can
be replaced with thicker raft units. The removed raft units are then available
for use elsewhere. In terms of whole life costing, raft units become increasingly
competitive.
A major advantage of rafts is that they are of known quality. They are
normally manufactured in a factory and transported to the site, but they can
also be manufactured and stored on site. For example, a seaport area or an
airfield may store rafts for an emergency where a concrete paved area may
need reorganizing, relaying or urgently repairing. A raft pavement repair may
take as little as two hours and the pavement is available for immediate use.
Traffic is only interrupted during the repair and their is no delay necessary for
concrete curing.
The modular nature of the rafts gives them a particular value where access
A REVIEW 173
to underlying public utility equipment is needed. Raft units can be laid
knowing that they can be lifted out, the equipment attended to without the fear
of traditional trench reinstatement problems. Raft units do not transfer loads
between themselves by interlock. This is an advantage for pavement mainten-
ance, but it does mean that the ride quality will be lower and that the subgrade
stresses will be higher than for rigid pavements.

8.2 Manufacture of raft units


The factory manufacture of raft units brings with it the high quality control
that is free from the inclement weather conditions associated with on-site
concrete placement. The factory environment ensures high quality control of
the materials and production of the raft units. For example, raft units can be
inspected at the factory before being transported to site rather than the on-site
construction situation where an unsatisfactory piece of concrete has to be cut
out and replaced. Factory conditions ensure that large numbers of raft units
are rapidly produced, with concrete strengths of 60 MPa and above, thus
giving a highly durable product.
Following casting the raft units are vacuum lifted from the moulds and
stored, suitably stacked, up to eight high. Great care is exercised to ensure the
concrete has gained sufficient strength before using the two lifting holes cast
into the raft unit and attached to the reinforcement. Two layers of reinforce-
ment are used in the raft for crack control.
Raft unit manufacturers can produce raft units in sizes from 300 mm square
up to 3.2 by 5.3 m and even 10.0 by 2.29 m, but raft unit size has standardized at
2m square, with a thickness usually between 140 and 200mm.

8.3 Materials used in raft manufacture


Precast concrete is used primarily for problem areas or special situations
rather than for general pavement construction. This is partly related to the fact
that raft unit materials can be clearly specified, checked and placed to satisfy
specific requirements. For example, the inclusion of crushed rock aggregates
improves wheel skid resistance, while other additions to the mix design can
increase the concrete resistance to certain types of chemical attack. Increasing
the concrete compressive strength can reduce impact resistance, unless the
ultimate concrete tensile strain is improved.
The square is the most structurally efficient raft unit shape and facilitates the
use of symmetrical steel bar/mesh reinforcement. Bar or mesh reinforcement
should be placed in both the top and the bottom of the raft with a suitable
amount of concrete cover. The reinforcing bar spacing changes depending
upon its location in the raft unit, with more reinforcement being required near
to the raft edge owing to the increased load-induced stress. Increased
174 PRECAST CONCRETE RAFT UNITS

reinforcement is needed adjacent to a corner, with stopped-off bars giving


some advantages. In fact, the type of wheel load can affect the amount and
positioning of the bars. Prestressed raft units have been manufactured, but
their small size means that the prestress is not effectively used, especially at the
raft edge. For this reason, raft units are usually prestressed in one direction
only and are rectangular (non-square) in shape, thus reducing the efficiency of
a prestressed raft unit.
A major advantage of steel bar/mesh reinforcement is its ability to hold the
raft together after cracking. A cracked raft is acceptable as it is still able to
function as a pavement due partly to the subbase support and the fact that
cracking initially has little effect on ride quality. High-yield bars are used to
increase impact resistance at high loading rates. The use of fibre reinforcement
will improve the concrete's resistance to spalling, improve the flexural strength
and improve the impact resistance. The addition of the fibres increases energy
absorption and energy distribution, but once the concrete has cracked, bar or
mesh reinforcement is needed to hold the raft together.
The top edge of the raft is often subjected to impact loading. Normally angle
steel is placed along the top edge and welded to the upper reinforcing layer to
prevent impact induced concrete spalling and to improve the membrane and
bending strength of the raft. The angle steel must not detach from the raft, as
vehicle damage and personal injury may result.
In some rafts the edge angle steel is dispensed with and a redesigned concrete
edge detail introduced. In this case the whole raft side is redesigned and the
usual gap between the rafts is closed. The rafts abut each other and transfer
some load by shear interlock.

8.4 The analysis of raft unit pavements

8.4.1 Concrete stress and subgrade stress


Raft units are used to their best advantage where loads are large and where
subgrade CBR values are low. Consequently, the two most important raft unit
design parameters are the raft concrete flexural tensile stress and the vertical
bearing pressure at the top of the subgrade. The concrete stress is the more
important, as excessive concrete tensile stress will cause the raft to crack.
Excessive subgrade stress can be corrected by lifting out the raft unit and
relevelling the subbase. The use of small rafts reduces the concrete bending
stress, but increases the raft deflection and the subgrade stress. Some pavement
design methods do not consider subgrade stress as it is assumed to have a
negligible effect on pavement performance. For raft pavements, the subgrade
stress must be calculated as it is a serviceability state. Excessive concrete stress
causes concrete fatigue and cracking owing to progressive internal microcrac-
king, but further research is needed into the fatigue of concrete pavements. The
most suitable fatigue model for raft unit pavements appears to be the Vesic
A REVIEW 175
model used in Chapter 3. The Vesic model allows a higher number of load
repetitions at higher stresses than other fatigue models. Concrete fatigue life is
not infinite no matter how low the stress ratio, but decreasing the stress range
does increase the fatigue life.
For raft units the serviceability limit state is defined as the point where raft
relevelling or some other minor maintenance is required. As cracked rafts will
still carry very heavy loads of 850 kN successfully, the concept of a raft unit
ultimate limit state has been introduced. This concept is related to the
serviceability limit state, but modified by the wheel load contact pressure in
order to predict the load repetitions at which a raft unit pavement would need
replacing.

8.4.2 Analytical and numerical analysis


Westergaard's analysis of concrete slabs on grade was first published in the
mid 1920s and even today still forms the basis of many rigid pavement design
methods. However, Westergaard's analysis assumes an infinitely sized con-
crete slab and the analysis gives increasingly erroneous results as the slab side
length reduces below about 8 m. Normal sized rafts have a side length of
around 2m.
Westergaard's analysis can also only be used for three single wheel load
positions - namely, the interior, edge and corner positions. For raft units the
actual number of applied loaded wheels together and their positions must be
analysed. Thus the equivalent single-wheel load (ESWL) concept used in
conjunction with Westergaard's analysis cannot be used for raft units.
The finite element method, which uses an elastic solid as the raft foundation
and not the Winkler medium of a bed of springs, is more realistic and gives
more accurate results. Also raft separation from the subbase can be taken into
account. However, the use of the finite element method to model a loaded raft
unit pavement is in its infancy when compared to the sophistication of flexible
pavement analysis.

8.4.3 Loading
Owing to the high loading often applied to raft unit pavements, the use of the
standard highway axle load of 80 kN cannot be extrapolated. The port area
wheel load (PAWL) of 12000 kg with a contact pressure of 0.8 MPa is more
realistic for heavy duty pavements and can be used for raft pavements to relate
a series of alternative single-wheel loadings. However, for raft units, by taking
into account the actual applied wheel loads and their precise positions, as
suggested in Chapter 3, raft unit pavements are found to carry significantly
more repeated wheel loadings than the British Ports Federation (BPF)
design manual discussed in Chapter 4 would suggest for rigid concrete
pavements.
176 PRECAST CONCRETE RAFT UNITS

For a single-wheel load, the maximum concrete stress occurs when the load
is at the centre of an edge, while a load at the raft unit corner produces the
highest subgrade stress. For a multiple-wheel load the maximum concrete
stress may occur at almost any location on the top or bottom of the raft. For
low loads on high-pressure (2-3 MPa) tyres, raft pavement ride quality must
be high in order to reduce to a minimum the high resultant stress in the raft.
The resulting subgrade stress for high tyre pressures is low. In the case of
heavily loaded multiple wheels with low (0.8-1.0 MPa) tyre pressures, the
major stress is in the subgrade with lower stresses in the raft. The speed of
application of the load also affects the raft: an excessively high load very
quickly applied will cause a smaller failed area than if the same load is applied
slowly. Rest periods and sustained loading between repeated loads increase
the concrete fatigue strength, provided that the static load stress level is below
75% of the static concrete strength. Increasing the raft unit thickness is the
single most important way of increasing the raft fatigue life.

8.5 On-site laying of raft units

8.5.1 Subbase
The subbase layer is most highly stressed when the load and the load contact
pressures are high, and for this reason the minimum acceptable subbase CBR
and subbase thickness are 20% and 300 mm respectively. The subbase must
have a high shear strength, low compressibility and prevent the damaging
effect of moisture and frost action on the raft unit pavement layer. The subbase
must be permeable, well graded and granular to allow both vertical and
horizontal water flow - crushed rock being preferred. For very heavy
loadings, the subbase must be bound with Portland cement or asphalt, but this
does stop all vertical draining, forcing the runoff to flow either above the raft
units or through the bedding sand. Positive drainage is essential. The use of
geotextiles to separate layers and to increase layer strength is welcomed.
Dynamic compaction of the subbase also improves layer strength.
Because of the difficulty of ensuring full contact between the raft unit base
and the subbase, a sand bedding layer of 50 mm is used. This layer is a source of
settlement, which for large newly laid raft areas presents no problem. For small
repaired areas, the bedding layer thickness is reduced to 20 mm to prevent raft
roughness caused by differential settlement.

8.5.2 Joints
The use of discrete raft units means that the joints decrease ride quality. The
rafts are undowelled, but some load transfer due to shear interlock does take
place. The amount of load transfer is dependent upon the joint width, joint
A REVIEW 177
movement, joint filler, load type and whether the load is approaching or
leaving the joint. Also some form of joint is inevitable to reduce the step when
raft units abut an existing pavement.
Open joints between raft units are useful in that they allow water to
percolate to and drain through the subbase. In some cases, the raft unit joints
are sealed to prevent rainwater infiltration with these pavements being laid to
a fall. Joints reduce ride quality and many high-speed vehicles, such as aircraft,
require a high ride quality and a fine smooth surface to prevent unacceptable
vibration. Certain aircraft when landing or taking off require no unexpected
protrusions or steps between the raft units that could catch trailing arrestor
hooks. This effectively establishes a limiting runway roughness of 30 mm.
Sealing the joints reduces the roughness, reduces raft rocking and reduces the
steps between rafts. However, sealing the joints makes the recognition of
subgrade settlement more difficult.
Raft rocking can cause edges to become unsupported with perhaps concrete
spalling occurring, but rafts do not suffer from the solar radiation warping so
apparent with large concrete slabs.

8.6 Conclusions
Raft unit pavements have distinct advantages over in situ rigid concrete
pavements in that:
(1) there is high quality control in the factory;
(2) the units are laid rapidly and the pavement can be used immediately
irrespective of weather condition; and
(3) raft units are able to sustain very heavy loads and adapt to large
subgrade settlements.
To increase raft unit pavement life, the single most important contribution is
made by:
(J) increasing the raft unit thickness;
but further increases in raft unit pavement life can be obtained by:
(2) use of a square raft shape;
(3) use of fibre reinforcement;
(4) use of variably spaced and stopped-off reinforcing bars;
(5) increasing the raft unit size; and
(6) use of a permeable, well-graded crushed rock subbase, that has a bound
upper layer if very heavy loads are expected.
Further research, development, data acquisition and data interpretation are
required to bring raft unit pavement design to the same level of sophistication
as is found in flexible, pavement design. Specifically, research is required to:
178 PRECAST CONCRETE RAFT UNITS

(1) develop further analytical methods of raft unit pavement analysis;


(2) investigate further the numerical methods of raft unit pavement design;
(3) develop a flexible joint that transfers loads between raft units and will
limit surface roughness, but which will still allow speedy removal of
single raft units;
(4) reduce raft unit movement and spalling;
(5) reduce subgrade and subbase settlement; and
(6) find an alternative raft unit shape, size and thickness.
References

Chapter I
1. R.S. Rollings, 'Concrete Block Pavements', Technical Report, USAE Waterways
Experiment Station, Vicksburg, MS (1983).
2. K. Sharp, and P. Armstrong, 'Interlocking Concrete Block Pavements', Australian Road
Research Board, Melbourne, Australia (1986).
3. 1. Knapton, 'The Structural Design of Heavy Duty Pavements for Ports and Other
Industries', British Ports Association, London (1983).
4. B. Shackel, The Evolution and Application of Mechanistic Design Procedures for Concrete
Block Pavements', Proc. Third Internat. Conf on Concrete Block Pavements, Rome, Italy
(1988).
5. Proc. First Internat. Corif'. on Concrete Block Pavements, Newcastle upon Tyne, UK (1980).
6. Proc. Second Internat. Conf on Concrete Block Pavements, Delft, Netherlands (1984).
7. Proc. Third Internat. Conf on Concrete Block Pavements, Rome, Italy (1988).
8. S. Glushkov, and B. Royev-Bogoslovskii, Construction and Maintenance of Airfields,
Moscow, USSR (1970).
9. A. Hanna, et al. 'Technical Review of Prestressed Pavements', FHWA-RB-77-8, Federal
Highway Administration, Washington DC (1976).
10. AJ. Harris, 'Prestressed Concrete Runways: History Practice and Theory, Part 1', Airports
and Air Transportation, Vol. 10, No. 121, London, UK (1956).
11. D. Vandepitte, 'Prestressed Concrete Pavements - A review of European Practice', J.
Prestressed Concrete Inst. Vol. 6, no. 4, Chicago, Ill. (1961).
12. K. Sato, T. Fukute, and H. Inukai, 'Some New Construction Methods for Prestressed
Concrete Airport Pavements', Proc. Second. Internat. Conf on Concrete Pavement Design,
Purdue University, Indiana (1981).
13. V. Maidel, and A. Timofeev, •Precast Concrete Roads and Footways', Gorods-koe.
Khozayaisto, Moslevy, 26(3), Moscow (1962).
14. A. Birger, and A. Klopovskii, 'Prefabricated Roads Made from Vibra-Rolled Concrete
Slabs', Stroitelnye i Arkhitek Muskvy, 10(3), Moscow (1961).
15. S. Mikhovich, L. Tarasenko, and N. Tolmachev, 'Prefabricated Concrete Surfaces on
Industrial Roads in the Donbass', Automobil nyi Dorogoi, 24(2) (1961).
16. E. Dubrovin, et al. 'Precast Reinforced Concrete Slabs in Road Construction', Gorodskie
Khozayaisto Muskvy, 36(9) (1962).
17. B. Smolka, 'Prefabricated Prestressed Concrete Surfaces', Automobil nyi Dorogoi, 27(3)
(1964).
18. N.T. Stepuro, et al. 'The Construction of Sectional Concrete Pavements', Automobil nyi
Dorogoi, 27(3) (1964).
19. I. Mednikov, Y. Malchanov, and L. Gorodelskii, 'Analysis and Jointing of Polygonal
Road Slabs', Osnovaniya Fundamently i Mekhanika Gruntou, Vol. 11, NO.5 (1974).
20. L. Larson, and W. Hang, 'Construction of a Prefabricated Highway Test Section', Highways
Research Record, 389, Washington DC (1972).
21. W. Patterson, 'Functional Pavement Design for Container Terminals', Proceedings,
Australian Road Research Board, Vol. 8, Part 4 (1976).
22. R. Rollings, 'Corps of Engineers Design Criteria for Rigid Airfield Pavements', Proc. Second
Internat. Conf on Concrete Pavement Design, Purdue University, Indiana, USA (1981).
23. R. Rollings, and Y. Chou, 'Precast Concrete Pavements', Miscellaneous Paper GL-81-10,
USAE Waterways Experiment Station, Vicksburg, MS (1981).
24. R. Rollings, 'Field Performance of Fibre Reinforced Airfield Concrete Pavements',
DOT/FAA/PM-86/26, Federal Aviation Administration, Washington DC (1985).
180 PRECAST CONCRETE RAFT UNITS

25. R. Rollings, 'Developments in the Corps of Engineers Rigid Pavement Design Criteria',
Proc. Fourth Internal. Con! on Concrete Pavement Design, Purdue University, Indiana, USA
(1989).
26. 'Runway Repair Sets Fast Pace', Engineering News Record, Vol. 206, No. 10, New York
(1981).
27. S. Jones, and 1. Iverson, 'Use of Precast Slabs for the Repair of Faulted Joints in Concrete
Pavements', Special Report, Federal Highway Administration, Washington DC (1971).
28. 'Reconditioning High-Volume Freeways in Urban Areas', Synthesis of Highway Practice
No. 25, Transportation Research Board, Washington DC (1974).
29. L. Byrd, 'Precision Concrete Cutting and Repair System for Pavements', Roadways and
Airport Pavements, SP51, American Concrete Institute, Detroit, Mi. (1975).
30. 1. Overacker, 'Thruway Repairs Concrete Slabs Overnight', Public Works, Vol. 105, No.3
(1974).
31. P. Grimsley, and B. Morris, 'An Approach to Concrete Pavement Replacement That
Minimizes Disruption to Traffic', Special Report 153, Transportation Research Board,
Washington DC (1975).
32. 'Prefab Pavement Sections for PCL Repairs Slice Time and Cost for Caltrans', Better
Roads (September 1974).
33. B. Elkins, F. McCullough, and R. Hudson, 'Precast Repair of Continuously Reinforced
Concrete Pavement', Research Report 177-15, Centre for Transportation Research, Austin,
Tx. (1980).
34. 1. Rosenburg, et al. 'Rapid Runway Repair (RRR) In-house Test and Evaluation',
AFESC/ESL-TR-85-65, Air Force Engineering and Services Centre, Tyndall AFI5, Fl.
(1988).
35. W. Brabston, and T. Voller. 'Precast Concrete Slab Design for Repair and Restoration of
Paved Surfaces (REREPS)', Miscellaneous Paper GL-86-6, USAE Waterways Experiment
Station, Vicksburg, MS (1986).
36. A. Meyer, F. McCulloch, and D. Fowler, 'Polymer Concrete for Precast Repair of
Continuously Reinforced Concrete Pavement on 1-30, Near Mt. Pleasant', Research Report
No. 246-1, Federal Highway Administration, Washington DC (1981).
37. R. Packard, 'Computer Program for Airport Pavement Design', Portland Cement
Association, Skokie, 111. (1984).
38. H. Westergaard, 'Stresses in Concrete Pavements Computed by Theoretical Analysis',
Public Roads, Vol. 7, No.2 (1926).
39. H. Westergaard, 'New Formulas for Stresses in Concrete Pavements of Airfields', Trans.
Am. Soc. Civil Engineers, Vol. 113 (1948).
40. F. Parker, et at. 'Development of a Structural Design Procedure for Rigid Airfield
Pavements', Technical Report GL-79-4, USAE Waterways Experiment Station, Vicksburg
MS (1979).
41. R. Rollings, 'Design at Overlays for Rigid Airport Pavements', DOT/FAA/PM-87-19,
Federal Aviation Administration, Washington DC (1988).
42. S. Tayabyii, and P. Okamoto, 'Thickness Design of Concrete Resurfacing', Proc. Third
Internat. Conf. on Concrete Pavement Design and Rehabilitation, Purdue University,
Indiana (1985).
43. A. Ioannides, et al. 'Finite Element Analysis of Slabs-on-Grade Using a Variety of Support
Models', Proc. Third Internat. Con! on Concrete Pavement Design and Rehabilitation,
Purdue University, Indiana (1985).
44. R. Behrman, 'Model Tests of Rigid Pavements', USAE Ohio River Division Laboratories,
Mariemont, Oh. (1964).
45. A. Ioannides, et al. 'The Westergaard Solutions Reconsidered', Transportation Research
Board, Washington DC (1985).
46. 1. Bull, 'The Analysis of the Interaction Between Precast Concrete Pavement Units and
Soils Using Three Dimensional Elastic Analysis', App. Solid Mechanics Conference,
University of Strathclyde, Glasgow (1985).
47. 1. Bull, 'An Analytical Solution to the Design of Precast Concrete Pavements', Internat. J.
Numerical, Methods Geomech. 10 (1986).
48. 1. Bull, 'The Design Analysis of Raft Type Concrete Pavements Using Finite Elements',
Proc. Tenth Canadian Congress on Applied Mechanics, University of Western Ontario,
Canada (1985).
REFERENCES 181
49. 1. Bull, and S. Salmo, 'The Use of the Equivalent Single Wheel Load Concept for Discrete
Raft Type Pavements', Computers and Geotechnics, 3 (1987).
50. R Treybig, et al. 'Overlay Design and Reflection Cracking Analysis for Rigid Pavements'
Vol. 1, Development of New Design Criteria, Report No I. FHWA-RD-77-6, Vol. 1,
Federal Highway Administration, Washington DC (1977).
5t. A. Vesic, and S. Saxena, 'Analysis of the Structural Behaviour of Road Test Rigid
Pavements', Highway Research Record No. 291, Washington DC (1969).
52. American Concrete Institute, 'Considerations for Design of Concrete Structures Subjected
to Fatigue Loading', ACI 215R-74, Detroit, Mich, USA (1981).
53. F.M. Mellinger, 'Investigation of Prestressed Concrete Sectional Mats', Technical Report
No. 2-4, Ohio River Division Laboratories, US Army Engineer Division, Ohio River,
Cincinnati, Ohio, USA (1956).

Chapter 2
1. E. Winkler, Die Lehre von der Elaslizitiit und Festiykeit (The Theory of Elasticity and
Stiffness), H. Dominicus, Prague (llS67).
2. H. Hertz, 'Ober das Gleichgewicht schwimmender elastischer Platten' ('On the equilibrium
of floating elastic plates'), Ann. Physik und Chemie 22, Wiedemann's (1884).
3. H. Zimmermann, Die Berechnung des Eisenbahnoberbaues (Calculation of the Upper
Construction Surface for Railway Tracks), Ernst und Korn Verlag, Berlin, 1888; Second
edition (1930).
4. F. Schleicher, Kreisplatten au!, Elaslicher Unterlage (Circular Plates on Elastic Found-
ations), Julius Springer, Berlin (1926).
5. M. Hetenyi, Beams on Elastic Foundation Theory with Applications in the Fields of Civil and
Mechanical Engineering, Univ. of Michigan Press, Ann Arbor (1946).
6. A.P.S. Selvadurai, Elastic Analysis ofSoil-Foundation Interaction, Developments in Geotech-
nical Engineering, Vol. 17, Elsevier (1979).
7. A.M. loannides, Discussion of 'Subgrade contact pressures under rigid pavements' by YT.
Chou, J. Transport. Engng Ill, NO.3 (May 1985).
8. A.S. Vesic, and S.K. Saxena, 'Analysis of structural behavior of AASHO road test rigid
pavements', NCHRP Report No. 97, Highway Research Board (1970).
9. R.E. Gibson, 'Some results concerning displacements and stresses in a non-homogeneous
elastic half-space', Geotechnique 17, No.1 (1967). Correspondence: 18, No.2; 19, No.1
10. H.M. Westergaard, 'Stresses in concrete pavements computed by theoretical analysis', Public
Roads 7, No.2 (April 1926). Also in Highway Research Board, Proceedings, 5th Annual
Meeting (1925, published 1926), Part I, under title 'computation of stresses in concrete
roads'.
It. RM. Westergaard, 'Analytical tools for judging results of structural tests of concrete
pavements', Public Roads 14, No. 10 (December 1933).
12. RM. Westergaard, 'New formulas for stresses in concrete pavements of airfields', Trans.
ASCE 113 (1948). Also in Proc. ASCE 73, No.5 (May 1947).
13. H.M. Westergaard, 'Stresses in concrete runways of airports', Proc. 19th Annual Meeting,
Highway Research Board, Washington, D.C., 1939. Also in Stresses in Concrete Runways of
Airports, Portland Cement Association, Chicago, IL (December 1941).
14. A.M. Ioannides, M.R. Thompson, and EJ. Barenberg, 'Westergaard solutions reconsidered',
Transportation Research Record 1043, Transportation Research Board, National Research
Council, Washington, D.e. (1985).
15. G. Pickett, M.E. Raville, W.e. Janes, and FJ. McCormick, 'Deflections, moments and
reactive pressures for concrete pavements', Bulletin No. 65, Engineering Experiment
Station, Kansas State College (October 1951).
16. G. Pickett and G.K. Ray, 'Influence charts for concrete pavements', Trans ASCE 116(1951).
17. R.G. Packard, 'Computer program for airport pavement design', Special Report SR029.02P,
Portland Cement Association, Skokie, IL (1967).
18. W.e. Kreger, 'Computerized aircraft ground flotation analysis - Edge loaded rigid pave-
ment', Research Report No. ERR-FW-572, General Dynamics Corp., Fort Worth, TX
(January 1967).
19. A.RA. Hogg, 'Equilibrium of a thin plate, symmetrically loaded, resting on an elastic
182 PRECAST CONCRETE RAFT UNITS

foundation of infinite depth', London, Edinburgh and Dublin Phi/os. Mag. and J. Sci., Series 7,
25 (March 1938).
20. D.L. Holl, 'Equilibrium of a thin plate, symmetrically loaded, on a flexible subgrade', Iowa
State College, J. Sci., 12, No.4 (July 1938).
21. D.L. Holl, 'Thin plates on elastic foundations', Proc. 5th Internat. Congr. on Applied
Mechanics, Cambridge, Mass., 1938 (published 1939).
22. A.H.A. Hogg, 'Equilibrium of a thin slab on an elastic foundation of finite depth', London,
Edinburgh and Dublin Phi/os. Mag. and J. of Science, Series 7, 35 (April 1944).
23. A. Losberg, Structurally Reinforced Concrete Pavements, Doktorsavhandlingar Vid
Chalmers Tekniska Hogskola, Goteborg, Sweden, 1960.
24. M.G. Arora and S.K. Khanna, 'N umerical analysis of rigid pavement resting over Hookean
subgrade continuum', Numerical Methods in Geomechanics (C.S. Desai, ed.), ASCE, Vol. I
(1976).
25. G. Pickett, S. Badaruddin, and S.c. Ganguli, 'Semi-infinite pavement slab supported by an
elastic solid subgrade', Proc. First Congr. on Theoretical and Applied Mechanics, Indian
Institute of Technology (November 1955).
26. G. Pickett, and S. Badaruddin, 'Influence chart for bending of a semi-infinite pavement slab',
Proc. 9th Internat. Congr. on Applied Mechanics, Vol. 6, Universite de Bruxelles (1956)
(published 1957).
27. A.M. Ioannides, 'Analysis of slabs-on-grade for a variety ofloading and support conditions',
Ph.D Thesis, Univ. of Illinois, Urbana (1984). Also published by US Air Force Office of
Scientific Research, Report No. TR·85·0083, Air Force Systems Command, USAF, Bolling
AFB, D.C. 20332 (September 1984).
28. A.M. Ioannides, 'The problem ofa slab on an elastic solid foundation in the light of the finite
element method', Proc. Sixth Internat. Conf. on Numerical Methods in Geomechanics,
Innsbruck, Austria (1988).
29. A.M. Ioannides, J. Donnelly, M.R. Thompson and EJ. Barenberg, 'Three-dimensional finite
element analysis of a slab on stress dependent elastic solid foundation', U.S. Air Force Office
of Scientific Research, Report No. TR-86-0143, Air Force Systems Command, USAF,
Bolling AFB, D.C. (June 1986).
30. A.M. Tabatabaie and EJ. Barenberg, 'Structural analysis of concrete pavement systems', J.
Transport. Engng 106, No. TE5 (September 1980).
31. A.M. loannides, M.R. Thompson and EJ. Barenberg, 'Finite element analysis of slabs-on-
grade using a variety of support models', Pmc. Third Internal. Conlon Concrete Pavement
Design and Rehabilitation, Purdue University (April 1985).
32. G.T. Korovesis and A.M. Ioannides, Discussion of'EtTect of concrete overlay debonding on
pavement performance' by T. Van Dam, E. Blackmon, and M.Y. Shahin, Transportation
Research Record 1136, Transportation Research Board, National Research Council,
Washington, D.C. (1987).
33. O.c. Zienkiewicz, The Finite Element Method (3rd edn), McGraw-Hill (1977).
34. DJ. Dawe, 'A finite element approach to plate vibration problems', J. Mech. Engng Sci., 7,
No.1 (1965).
35. Y.K. Cheung and O.c. Zienkiewicz, 'Plates and tanks on elastic foundations: an application
of finite element method', Internat. J. Solids and Structures 1 (1965).
36. R.T. Severn, 'The solution of foundation mat problems by finite element methods', The
Structural Engineer, 44, No.6, London (June 1966).
37. S.D. Tayabji and B.E. Colley, 'Analysis ofjointed concrete pavements', Report FHWA/RD-
86/041, U.S. Department ofTransportation, Federal Highway Administration, Washington,
D.C. (1986).
38. Y.T. Chou, 'Structural analysis computer programs for rigid multicomponent pavement
structures with discontinuities - WESLlQID and WESLA YER; Report 1: Program
development and numerical presentations; Report 2: Manual for the WESLIQID
finite element program; Report 3: Manual for the WESLAYER finite element program',
Technical Report GL-81-6, U.S. Army Engineer Waterways Experiment Station (May
1981).
39. Y.H. Huang, 'Finite element analysis of slabs on elastic solids', J. Transport. Engng J. 100,
No. TE2 (May 1974).
40. Y.H. Huang and S.T. Wang, 'Finite-element analysis of rigid pavements with partial
subgrade contact', Transportation Research Record 485, Transportation Research Board,
National Research Council, Washington, D.C. (1974). -
REFERENCES 183
41. Agbabian Associates, Analytic Modeling of Rock-Structure Interaction, Vols 1-3, prepared
for Advanced Research Projects Agency, Bureau of Mines, Report No. AD-761-648, 649, 650
(1973).
42. A.M. Ioannides and J.P. Donnelly, 'Three-dimensional analysis of slab on stress dependent
foundation', accepted for presentation and publication, 67th Annual Meeting, Transport-
ation Research Board, National Research Council, Washington, D.C. (1988).
43. A.M. Ioannides, 'Axisymmetric slabs of finite extent on elastic solid', J. Transport. Engng
113, No.3 (May 1987).
44. S.G. Bergstrom, E. Fromen and S. Linderholm, Investigation of Wheel Load Stresses in
Concrete Pavements, Swedish Cement and Concrete Research Institute, Proc. No. 13, Royal
Institute of Technology, Stockholm (1949).
45. A.M. Ioannides, 'Finite difference solution for plate on elastic solid', J. Transport. Engng 114,
NO.1 (January 1988).
46. J.A. Roberson and C.T. Crowe, Engineering Fluid Mechanics (2nd edn), Houghton Miffiin
Company, Boston, MA (1980).
47. D.M. Burmister, Open Floor Discussion, Session 1, Proc. Univ. ofMichigan (First) Internal.
Can! on the Structural Design of Asphalt Pavements, Ann Arbor, MI (1962).
48. A.M. Ioannides, 'Dimensional analysis in NDT rigid pavement evaluation', accepted for
publication in J. Transport. Engng (June 1988).
49. A.M. Ioannides, Discussion of'Response and performance of alternate launch and recovery
surfaces that contain layers of stabilized material' by R.R. Costigan and M.R. Thompson,
Transportation Research Record 1095, Transportation Research Board, National Research
Council, Washington, D.C. (1986).
50. A.M. Ioannides and R.A. Salsilli-Murua, 'Temperature curling in rigid pavements: an
application of dimensional analysis', accepted for presentation and publication, 68th Annual
Meeting, Transportation Research Board, National Research Council, Washington, D.C.
(1989).
51. A.M. Ioannides, Discussion of 'Thickness design of roller-compacted concrete pavements'
by S.D. Tayabji and D. Halpenny, Transportation Research Record 1136, Transportation
Research Board, National Research Council, Washington, D.C. (1987).
52. M.R. Thompson, and B.I. Dempsey, 'Development of preliminary ALRS stabilized material
pavement analysis system (SPAS)', Technical Report ESL-TR-83-84, USAF Engineering
and Services Center, Tyndall AFB, Florida (March 1984).
53. H.M. Westergaard, 'Analysis of the stresses in concrete roads caused by variations in
temperature', Public Roads 8, No.3 (May 1927).
54. R.D. Bradbury, Reinforced Concrete Pavements, Wire Reinforcement Institute, Washington,
D.C. (1938).
55. L.W. Teller and E.C. Sutherland, 'The structural design on concrete pavements', Public
Roads 16, No.8 (October 1935); 16, No.9 (November 1935); 16, No. 10 (December 1935); 17,
No.7 (September 1936); 23, NO.8 (April, May, June 1943).
56. AASHTO Guide for Design ofPavement Structures, American Association of State Highway
and Transportation Officials, Washington, D.C. (1986).
57. M.l. Darter, 'Design of zero-maintenance plain jointed concrete pavement: Vol. 1-
Development of design procedures', Federal Highway Administration, Report No. FHWA-
RD-77-11I, Washington, D.C. (June 1977).
58. E.F. Kelley, 'Application of the results of research to the structural design of concrete
pavements', Public Roads 20, No.5 (July 1939); No.6 (August 1939).
59. S. Iwama, Experimental Studies on the Structural Design of Concrete Pavement, Pavement
Laboratory, Public Works Research Institute, Ministry of Construction, Japan (May 1964).
60. A.M. Ioannides, E.I. Barenberg and J.A. Lary, 'Interpretation offalling weight defiectometer
results using principles of dimensional analysis', Proc. Fourth Internal. Con! on Concrete
Pavement Design and Rehabilitation, Purdue University (April 1989).

Chapter 3
1. E.M. Dubrovin et aI., 'Precast reinforced concrete slabs in road construction', Gorodskie
Khozayaisto Moskvy, 36, No.9 (1962).
2. R.I. Williams, 'Maintaining the roads to carry the loads', J. Inst. Highways and Transport. 35,
No.4 (April 1988) 28-35.
184 PRECAST CONCRETE RAFT UNITS

3. S.D. Barber, 'Pavement design in port areas', Ph.D. Thesis, Univ. of Newcastle upon Tyne,
UK (1980).
4. R.S. Rollings and Y.T. Chou, 'Precast concrete pavements', Misc. Paper GL-SI-IO, US
Army, Engineers, Washington, USA (1981).
5. J.W. Bull, 'An analytical solution to the design of precast concrete pavements', Internat. J.
Numerical Methods Geomech. 10 (1986) 115-123.
6. J.W. Bull, 'Investigation into the possibility of producing design charts for raft type
pavements', SERC Grant No. GR/B 86101 (1982-1984). In association with Redland
Aggregates Ltd.
7. R.F. Ackroyd and J.W. Bull, The design of precast concrete pavements on low bearing
capacity subgrades', Comput. Geotech. 1 (1985) 279-291.
8. A.N. Hanna, 'Technological review of prestressed pavements', FHWA-RD-77-8, Federal
Highway Administration, Washington DC, USA (1976).
9. Boston Society of Civil Engineers, Contributions to Soil Mechanics. 1925-1940, Boston, USA
(1940).
10. E. Winkler, Study in Elasticity and Strength, H. Dominicus, Prague (1867).
11. H.M. Westergaard, 'Stresses in concrete pavements computed by theoretical analysis', Public
Roads, 7, No.2 (1926) 25-53.
12. D.M. Burmister, 'The theory of stresses and displacements in layered systems', Proc.
Highway Research Board (1943).
13. J.W. Bull and S.H. Salmo, 'The use of the equivalent single wheel load concept for discrete
raft type pavements', Compu/.. Geomechan. 3 (1987) 29-35.
14. D.M. Burmister, 'The general theory of stresses and displacements in layered soil systems', J.
App. Phys. 16, (1945) 89-94.
15. S.F. Brown and J.M. Brunton, 'Computer programmes for the analytical design of asphalt
pavements', J. Inst. Highways and Transport. 39, No. 8/9 (1982).
16. D. Croney, The Design and Performance of Road Pavements, Dept of the Environment,
TRRL, HMSO, London (1977).
17. R. Jones and M.F. Kaplan, 'The effects of course aggregate on the mode of failure of
concrete in compression and flexure', Mag. Cone. Res. 9, No. 26 (1957) 89-94.
18. A.S. Vesic and S.K. Saxena, 'Analysis of structural behaviour of road test rigid pavements',
Highway Research Board, No. 291 (1969), pp. 156~158.
19. D.W. Hobbs, 'The strength and deformation of concrete under short term loading', Cone.
Assoc. Tech. Report TRA/42.484 (September 1972).
20. J.W. Bull, R. Anang and M.H.H.M. Ismail, 'The conditions of fracture of heavy duty
pavement units and the relationship between the serviceability and ultimate limit state', 3rd
Internal. Conf. on Creep an.d Fracture ofEngng Materials and Structures, Swansea, UK (5-10
April 1987), pp. 1033-1043.
21. J.W. Bull and H. Al-Khalid, 'An analytical solution to the design of footway paving flags',
Comput. Geotech. 4 (1987) 85-96.
22. Shell International Petroleum Co. Ltd., Shell Pavement Design Manual, Asphalt Pavements
and Overlays for Road Traffic, London (1978).
23. W. Heukelom and AJ.G. Klomp, CO/lSidera!ion of Calculated Strai/ls at Various Depths ill
Connection with the Stability of Asphalt Pavemel1ls, Univ. of Michigan, Ann Arbor (1967),
pp.155-168.
24. M.A. Sargous and S.K. Wang, 'Equivalent single wheel edge loads on rigid pavements',
Research Report No. CE73-12, Univ. of Calgary (1973).
25. British Ports Association, The Structural Design of Heavy Duty Pavements for Ports and
Other Industries (enlarged edn), BPA, London (1983).
26. PAFEC, PAFEC Ltd, Data Preparation User Manual, Level 6.1, Nottingham, UK (1984).

Chapter 4
1. M.M. Meletiou, and 1. Knapton, 'Container terminal pavement management', UNCTAD
Monographs on Port Management, Monograph No.5, United Nations Publication ( 19821.
2. S.D. Barber and 1. Knapton, Port Pavement Loading, Dock and Harbour Authonty (Apnl
1979), pp. 379-356.
REFERENCES 185
3. S.D. Barber and J. Knapton, Port Pavement Loading, Dock and Harbour Authority (March
1980), pp. 362-367.
4. BPA, The Structural Design ofHeavy Duty Pavementsfor Ports and Other Industries, British
Ports Association (1982).
5. G. Tickell, Course Notes, Dept of Civil Engineering, Univ. of Liverpool (1988).
6. ACI Committee 215, 'Considerations for the design of concrete structures subjected to
fatigue loading', Pmc. ACI Journal, 71 (March 1974) 97-121.
7. G.M. Nordby, 'Fatigue of concrete - a review of research', ACI Journal, 55 (August 1958)
191-219.
8. J.L. van Ornum, 'Fatigue of cement products', Trans. ASCE 51 (1903) 443.
9. J.L. van Ornum, 'Fatigue of concrete', Trans. ASCE 58 (1907) 294-320.
10. H.F. Clemmer, 'Fatigue of concrete', Proc. ASTM 22, Part 2 (1923) 409-419
11. R.B. Crepps, 'Fatigue of mortar', Proc. ASTM 23 Part 2 (1923) 329-340.
12. W.K. Hatt and R.E. Mills, 'Physical and mechanical properties of Portland cements and
concretes', Bulletin No. 34, Purdue University (1928) pp 34-51,91-95
13. H.A. Williams, 'Fatigue tests oflightweight aggregate concrete beams', Proc. ACI Journal 39
(April 1943) 441-448.
14. J.W. Murdock and CE. Kesler, 'Effect of range of stress on fatigue of plain concrete beams',
Proc. ACI Journal 55, No.2 (August 1959) 221-232.
15. E.W. Bennett and S.E. Muir, 'Some fatigue tests on high-strength concrete in axial
compression', Cement and Concrete Ass., No. 59 (June 1967) 113-117.
16. E.W. Bennett and N.K. Raju, 'Cumulative damage of plain concrete in compression', Proc.
Southampton Civil Engineers Materials Corif'. (1969) pp. 1089-1102.
17. S.P. Shah and S. Chandra, 'Fracture of concrete subjected to cyclic and sustained loading',
Proc. ACI Journal 67 (October 1970) 816-825.
18. R. Tepfers, C Friden and L. Georgsson, 'A study of the applicability to the fatigue ofconcrete
of the Palmgren-Miner partial damage hypothesis, Mag. Caner. Res. 29, No. 100 (September
1977) 123-130.
19. R. Tepfers and T. Kutti, 'Fatigue strength of plain, ordinary and lightweight concrete', Proc.
ACI Journal 76, No.5 (May 1979) 635-652.
20. R. Tepfers, 'Tensile fatigue strength of plain concrete', Proc. ACI Journal 76, No.8 (August
1979) 913-933.
21. K.D. Raithby and J.W. Gallaway, 'Effects of moisture condition, age and rate of loading on
the fatigue life of plain concrete', Abeles Symp. on the Fatigue of Concrete, ACI Publication,
SP-41-14, Detroit (1973) pp. 315-330.
22. RP. Sinha, K.H. Gerslle and L.G. Tulin, 'The response of singly-reinforced concrete beams
to cyclic loading', Proc. ACI Journal 61, No.8 (August 1964) 1021-1028.
23. De Jo1y, 'The strength and elasticity of Portland cements' ('La resistance et l'elasitute des
ciments Portland'), Anna/es. Pants et Chaussees 16, Series 7 (1898) 198-244 (in French).
24. H. Hilsdorf and CE. Kesler, 'The behaviour of concrete in flexure under varying repeated
loads', T & A.M. Report No. 172, Univ. of Illinois, Urbana (1960) 147 pp.
25. CE. Kesler, 'Effects of speed of testing on the flexural fatigue strength of plain concrete',
Proc. Highway Research Board 32 (1953) 251-258.
26. B.M. Assimacopoulos, R.E. Warner and CE. Ekberg, 'High speed tests on small specimens
of plain concrete', J. Prestress Concrete lnst. 4, No.2 (September 1958) 53-70.
27. J.A. Neal and CE. Kesler, 'The fatigue of plain concrete', Proc. Internal. Con! on the
Structure ofConcrete, London (September 1965), Cement and Concrete Assoc., pp. 226-237.
28. E.M. Awad and H.K. Hilsdorf, 'Strength and deformation characteristics of plain concrete
subjected to high repeated and sustained loading', Civil Engineering Studies, Struct. Res.
Series No. 372, Univ. of Illinois (February 1972) p. 266.
29. P.R. Sparks and J.B. Menzies, 'The effect of rate of loading upon the static and fatigue
strength of plain concrete in compression', Mag. Concrete Res. 25, No. 83 (June 1973) 73-80.
30. S.P. Shah and G. Winter, 'Response ofconcrete to repeated loading', RI LEM Internal. Symp.
on the Effects of Repeated Loading on Materials and Structural Elements (September 1966)
Mexico D.I. p. 26.
31. F.S. Op1e and CL. Hulsbos, 'Probable fatigue life of plain concrete with stress gradient',
Proc. ACI Journal 63 (January 1966) 159-181.
32. M. Considere, 'Influence of reinforcement on the properties of mortars and concrete'
186 PRECAST CONCRETE RAFT UNITS

('Influence des armatures metalliques sur les propriHes des mortiers et hetons'), c.R. /'Acad.
Sci. 127 (1898) 992-995.
33. E. Probst, 'The influence of rapidly alternating loading on concrete and reinforced concrete',
The Structural Engineer, London 9 (1931) 419.
34. E. Probst, Principles of Plain and Reinforced Concrete Construction, Arnold, London (1936)
pp.64, 128,250.
35. E. Probst and F. Treiber, 'Reinforced concrete beams under frequently repeated loading',
Bouingenieur, Berlin 13, No. 21/22 (1932) 285-289.
36. T.S. Chang and C.E. Kesler, 'Static and fatigue strength in shear of beams with tensile
reinforcement', Proc. ACI Journal 54 (June 1958) 1033-1058.
37. T.S. Chang and CE. Kesler, 'Fatigue behaviour of reinforced concrete beams', Proc. ACI
Journal 30 (August 1958) 245-254.
38. Cw. Muhlenbruch, 'The effect of repeated loading on the bond strength of concrete', Proc.
ASTM 45 (1945) 814-845.
39. Cw. Muhlenbruch, 'The effect of repeated loading on the bond strength of concrete - 2',
Proc. ASTM 48 (1948) 977-985.
40. E.S. Perry and N. Jundi, 'Pull-out bond stress distribution under static and dynamic
repeated loading', Proc. ACI Journal 66, No.5 (May 1969) 377-381.
41. I.R. Verna and T.E. Stelson, 'Failure of small R/C beams under repeated loading', Proc. ACI
Journal 59, No. 10 (October 1962) 1489-1504.
42. C Loo, 'The behaviour of reinforced concrete slabs subjected to concentrated load repeated
many times', MSc. Thesis, Dept of Civil Eng., Queen's Univ., Kingston, Ontario, Canada
(August 1968) 52 pp.
43. N.M. Hawkins, 'Fatigue strength ofconcrete slabs reinforced with wire fabric', Abeles Symp.
on the Fatigue of Concrete, ACI Publication, SP-41-14, Detroit (1973) pp. 315-330.
44. Portland Cement Association, Thickness Designfor Concrete Pavements (1966).
45. M.1. Darter, 'Design of zero-maintenance plain jointed concrete pavement', Vol. 1,
Development of Design Procedures, Federal Highway Administration (1977), Report
No. FHWA-RD-77-111.
46. H.J. Treybig, B.F. McCullough, P. Smith and H. Von Quintus, 'Overlay design and reflection
cracking analysis for rigid pavement', Vol. I, Development of New Design Criteria (1977),
FHWA Report No. FHWA-RD-77-76.
47. Highway Research Board, The AASHO Road Test - Report 5, Pavement Research, High-
way Research Board (1962) Special Report 61E.
48. American Association of State Highway Officials, The AASHO Interim Guidefor Design of
Pavement Structures (1972).
49. K. Majidzadeh and G.J. lives, 'Evaluation of rigid pavement overlay design procedure,
development of the OAR procedure', (1983) Final Report DTFHII-9489.
50. G.M. Hammit, R.L. Hutchinson, J.L. Rice and 0.0. Thompson, Multiple-wheel Heavy Gear
Load Pavement Tests, Vol. IV, Air Force Weapons Laboratory (1971) Technical Report
No. 70-113.
51. H.M. Westergaard, 'Stresses in concrete pavements computed by theoretical analysis', Public
Roads 7 (1926).
52. A.S. Vesic and S.K. Saxena, 'Analysis of the structural behaviour of road test rigid
pavements', Highway Research Record, No. 291 (1969) pp. 156-158.
53. D.L. Spellman, J.R. Stoker and B.F. Neal, 'California pavement faulting study', California
Division of Highways (1970) Research Report 635167-1.
54. J.W. Bull, 'An analytical solution to the design of precast concrete pavements', Internat. J.
Numer. Analyt. Methods Geomech. 10 (1986) 115-123.
55. R.F. Ackroyd and J.W. Bull, 'The design of precast concrete pavements on low bearing
capacity subgrades', Comput. Geotech. (1985) 279-291.
56. L.W. Teller and E.C Sutherland, 'The structural design of concrete pavements', Public
Roads 3, No.8 (1943) 167-212.
57. A.H.A. Hogg, 'Equilibrium of a thin plate symmetrically loaded, resting on an elastic
subgrade of infinite depth', Philos. Mag., Ser. 7, 25 (1938).
58. D.L. Hall, 'Thin plates on elastic foundations', Proc. Internat. Can! on Applied Mechanics,
Cambridge, MA (1938).
59. G. Pickett and G,K, Ray, 'Influence charts for concrete pavements', Trans. ASCE 116(1951).
60, J. Boussinesq, Application des Potentials, Paris (1885).
REFERENCES 187
61. D.M. Burmister, 'The theory of stresses and displacements in layered systems and
application to the design of airport runways', HRB Proc. (1943).
62. N.M. Newmark, 'Numerical methods of analysis of bars, plates and elastic bodies', in L.E.
Grinter, (ed.), Numerical Methods of Analysis in Engineering, Macmillan, New York.
63. W.R. Hudson and H. Matlock, 'Analysis of discontinuous orthotropic pavement slabs
subjected to combined loads', Highway Research Record No. 131 (1966) pp. 1-48.
64. A.M. Tabatabaie and EJ. Barenberg, 'Structural analysis of concrete pavement system',
Paper presented at the 1979 Research Session of the American Society of Civil Engineers,
Boston, Massachusetts (April 1979).
65. YH. Huang and S.T. Wang, 'Finite-element analysis of concrete slab slabs and its
implications for rigid pavement design', Highway Research Record No. 466 (1973).
66. YH. Huang and S.T. Wang, 'Finite element analysis of rigid pavements with partial
subgrade contact', Highway Research Record, No. 485 (1974).
67. YH. Huang, 'Finite element analysis of slabs on elastic solids', J. Transport. Engng 100 TE2
(May 1979).
68. E.L. Wilson, 'Solid SAP, a static analysis program for three-dimensional solid structures',
Berkeley, Structural Engineering Laboratory, Univ. of California, Report SESM 71-19,
(1969).
69. J.E. Crawford and R. Pichumani, 'Finite-element analysis of pavement structures using
AFPAU code (nonlinear elastic analysis)', New Mexico, Air Force Weapons Laboratory,
Kirtland Air Force Base, Technical Report No. AFWL-TR-74-71 (1975).
70. S.K. Saxena, 'Pavement slabs resting on elastic foundation', Highway Research Record,
No. 466 (1973).
71. W.e. Steinbrenner, 'Tafelen zur Setzungsberechnung', Die Strasse I (1951). Also: Proc.
Internat. Conf. on Soil Mechanics. Cambridge. Massachu~etts, 1936, Vol. 2.
72. J.W. Bull, 'Designing precast concrete pavements using microcomputers', J. Inst. Highway
Transport. 5, No. 33 (1986) 21-24.
73. J.W. Bull and A.H. Khalid, 'An analytical solution to the design of footway paving flags',
Comput. Geotec~. 4 (1987) 85-96.
74. J.W. Bull and Y.B. Luheshi, 'The experimental and finite element analysis of non-square raft
type concrete pavements', 3rd Internat. Symp. on Numerical Models in Geomechanics
(NUMOG 3), (May 1989), Niagara Falls, Canada, pp. 707-715.
75. H.A. Khalid and J.W. Bull, 'Finite element analysis of precast concrete pavements with
reference to wheel load configuration and the implications on their design', Research
Mechanica (1989).
76. L.D. Childs and l.W. Kapernick, Tests of concrete pavements on gravel subbases', Proc.
ASCE, HW3, Paper 1800, Vol. 84 (October 1958).
77. M.W. Kweincinski, 'Some tests on the yield criterion for a reinforced concrete slab', Mag.
Cone. Res. 17, No. 52 (September 1965).
78. D. Croney, The Design and Performance of Road Pavements, HMSO, London (1977).
79. S.G. Millard, Private communication, Dept of Civil Engineering, Liverpool Univ. (1989).

Chapter 5
1. T.D. Lin, 'Concrete for lunar base construction', Concrete Int. (1987).
2. Japan Road Assoc., Specifications for Highway Bridges (1987).
3. AASHTO, Standard Specifications for Highway Bridges, 13th edn (1983).
4. RILEM, CEB, IABSE, lASS Interassoc. Symp., Proc. Concrete Structures under Impact and
Impulsive Loading, Introductory Report, Berlin (BAM) (1981).
5. M. Fujii and A. Miyamoto, A Review of Current Research on Concrete Structures under
Impulsive Loading, Japan Concrete J. 21, No.9 (1983).
6. RILEM, CEB, IABSE, lASS Interassoc. Symp., Proc. Concrete Structures under Impact and
Impulsive Loading, Berlin (BAM) (1982).
7. H. Hwang, B. Ellington and M. Shinozuka, 'Probability-based design criteria for nuclear
power structures', J. Struct. Engng. 113, NO.5 (1987).
8. A.K. Kar, 'Impact effects of Tornado missile and aircraft', J. Struct. Div. (ASCE) 105,
No. STII (1979).
9. W. Jonas, R. Meschkat, H. Reich and E. Rudiger, 'Experimental investigation to determine
188 PRECAST CONCRETE RAFT UNITS

the kinetic ultimate bearing capacity of reinforced concrete slabs subject to deformable
missiles', Trans. 5th SMirt Conference (1979).
10. P.G. Bignon and J.D. Riera, 'Verification of methods of analysis for soft missile impact
problems', Nuclear Engng and Design 60 (1980).
II. T. Tobe and M. Kato, 'A study on the laternal impact of beams', Trans. Japan Soc. Mech.
Engrs 38, No. 314 (1972).
12. G. Hughes and A.W. Beeby, 'Investigation of the effect of impact loading on concrete beams',
The Struct. Eng. 6OB, NO.3 (1982).
13. F. Stangenberg, 'Nonlinear dynamic analysis of reinforced concrete structures', Nuclear
Engng and Design 29 (1974).
14. R.W. Clough and J. Penzien, Dynamic of Structures, McGraw-Hili (1975).
15. J.P. Wolf, K.M. Bucher and P.E. Skrikerud, 'Response of equipment to aircraft impact',
Nuclear Engng and Design 47 (1978).
16. M. Abdel-Rohman and J. Sawan, 'Impact effect on R.e. slabs: analytical approach', J. Struct.
Engng 111, No.7 (1984).
17. M. Fujii, A. Miyamoto and H. Morikawa, 'An analytical study on the behaviour of
reinforced concrete beams under impulsive load', Proc. Japan Soc. ofCivil Engineers 360, V-3
(1985).
18. M. Fujii, A. Miyamoto and K. Sakai, 'Nonlinear finite element analysis of reinforced
concrete beams under impulsive load', Proc. JCI Colloquium on Finite Element Analysis of
Structures (1984).
19. M. Fujii, A. Miyamoto, K. Sakai and H. Fushita, 'Dynamic nonlinear modelling of
reinforced concrete beams under impulsive loads', Japan Concrete Inst. 7th Con[. (1985).
20. M. Fujii, A. Miyamoto, T. Nakamura and H. Fushita, 'Nonlinear dynamic analysis of
prestressed concrete beams under impulsive loads', Proc. 40th Annual Con! Japan Soc. Civil
Engineers (1985)
21. R. Houlston, J.E. Slater, N. Pegg and e.G. DecRochers, 'On analysis of structural response
of ship panels subjected to air blast loading', Comput. Struct. 21, No. 1/2 (1985).
22. J. Ghaboussi, W.A. Millavec and J. Isenberg, 'RIC structures under impulsive loading', J.
Struct. Engng 110, No.3 (1984).
23. O. Isobata, 'Two-dimensional elastic-plastic analysis of concrete structure by the finite
element method', Archil. Inst. Japan 189(11) (1971).
24. D. Ngo and A.e. Scordelis, 'Finite element analysis of reinforced concrete beams', J. of ACI
(1967).
25. A. Miyamoto and M.W. King, 'Non-linear analysis and bond modelling of RC beams
subjected to impulsive loads', Proc. Japan Concrete Inst. 10, No.3 (1988).
26. M. Fujii, A. Miyamoto, H. Fushita and K. Oka, 'Behavior of critical regions of concrete slabs
subjected to single impact', Japan Concrete Inst. 8th Con! (1986).
27. F.R. Hand, D.A. Pecknold and W.e. Schnobrich, 'Nonlinear layered analysis of RC plates
and shells', J. Struct. Div. (ASCE) 99, No. ST7 (1973).
28. M. Fujii and A. Miyamoto, 'Improvement of the impact resistance for prestressed concrete
slab', RILEM, CEB, IABSE, lASS Interassoc. Symp.: Concrete Structures under Impact and
Impulsive Loading, Berlin (BAM) (1982).
29. A. Miyamoto, M.W. King and H. Masui, 'Non-linear dynamic analysis and evaluation of
impact resistance of reinforced concrete slabs under impulsive load', Proc. Japan Concrete
Inst. 11 (1989).
30. A. Miyamoto, M.W. King and M. Fujii, 'Non-linear dynamic analysis and design concepts of
RC beams under impulsive load', Proc. Pacific Concrete Con[., Vol. 2, New Zealand (1988).
31. DJ. Hannant, Fibre Cements and Fibre Concrete, Wiley, England (1978).
32. A. Miyamoto, A. Nishimura, M. Fujii and Y. Kajitani, 'Field testing and measurement of
concrete structures', Second Internat. Colloquium on Concrete in Developing Countries, India
(1988).

Chapter 6
I. AASHTO Guide for Design of Pavement Structures, AASHTO, Washington, D.e. (1986).
2. A.M. Neville, Properties of Concrete, Pitman, (1975), pp. 213-216 and 321-337.
3. Standard Specifications, State of California, Dept of Transportation (January, 1988).
4. M. Poblete, R. Valenzuela and R. Salsilli, 'Load transfer in undoweled transverse joints of
REFERENCES 189
PCC pavement', Transportation Research Record, Washington, D.C., Paper No. 87-0739.
5. IDIEM, Field Instrumentation, Equipment, Measurement Methodology, Univ. of Chile,
Natl. HWY Adm. CHILE (1986). In Spanish.
6. P. Spralz, 'Concrete pavement construction', Publication AT 11, Chilean Cement and
Concrete Institute (1983). In Spanish.
7. 1. Armaghani, T. Larsen and L. Smith, 'Temperature response of concrete pavements',
Record No. 1121, Washington, D.C. (1987), pp. 23-33.
8. M. Poblete, R. Salsilli, R. Valenzuela, A. Bull and P. Spratz, 'Field evaluation of thermal
deformations in undoweled PCC pavement slabs', Transportation Research Record,
Washington, D.C., Paper 87-0738.
9. IDIEM, Concrete Pavement Research Project, Univ. de Chile, Natl. HWY Adm. CHILE,
Final Report (1989). In Spanish.
10. B. Tremper and L. Spellman, 'Shrinkage ofconcrete: comparison oflab & field performance',
Highway Research Record No.3, 1963.
11. N. Hveem, 'Slab Warping Affects Pavement Performance', J. ACI 47 (1951) 797-808.
12. R. Ytlcrbcrg, 'Shrinkage and curling of slabs on grade', ACI, Concrete Internal. Mag. Nos 4,
5,6(1987).
13. Thickness Designfor Concrete Highway and Street Pavements, Portland Cement Association,
USA (1984).
14. M. Darter, E. Barenberg and A. Yrjanson, 'Joint repair methods for PCC pavements',
Appendix C, Void Detection Procedures, Final Report, NCHRP 1-21, Univ. of Illinois (1984).
15. A. Baucheron, 1. Gramsammer, P. Keryell and M. Paillard, 'Le deflectographe 04', Bull.
Liaison LPC, No. 129 (1984).
16. M. Darter, Design ofZero-Maintenance Plain Jointed Concrete Pavements, Univ. of Illinois,
1976.
17. L. Ortega and P. Spratz, 'Performance measurcs on plain concrete pavements without
dowels', Proc. Internal. Conf on Roads and Development, Paris (1984), pp. 199-204.
18. P. Foxworthy, 'Concepts for development of nondestructive testing and evaluation system
for rigid airfield pavements', Ph.D. Thesis, Univ. of Illinois (1985).
19. F. Neal, 'California PCC pavement faulting studies: a summary', CALTRANS Laboratory
Report FHWA/CA/TL-85/06 (December 1985).
20. H. Cadergreen, Drainage of Highway and Airfield Pavements Wiley (1974).
21. G. Wells, 'Evaluation of edge drain performance', CALTRANS Laboratory Report
FHWA/CA/TL-85/15 (1985).
22. M. Ray, 'Raport general introductif: Drainage et erodabilite aux contacts entre la dalle de
Beton, la fondation et la bande d'arret d'urgence" Bull. Liaison LPC, No. 138 (1985).
23. M. Darter, 1. Becker, M. Snider and R. Smith, 'Portland cement concrete pavement
evaluation system COPES', NCHRP Report 277, Transportation Research Board,
Washington, D.C. (1985).
24. M. Poblete, R. Valenzuela, E. Clasing, 1. Salgado and P. Gutierrez, 'Faulting evolution in
undoweled PCC pavements with slabs curled upwards', Proc. 4th Internat. Conf on Concrete
Pavement Design & Rehahilitation, Purdue University, Indiana (1989).

Chapter 7
1. L.J. Kennedy, in Proc. 2nd Internat. Colif. on The Bearing Capacity ofRoads and Airfields (eds
C. Ward and C.K. Kennedy), WDM Ltd, Bristol, England (1986), p. 195.
2. 1.W. Bull, 'The experimental and finite element analysis of non-square raft type concrete
pa vements', 3rd Internat. Symp. on Numerical Model in Geomechanics (NUMOG Ill), Niagara
Falls, Canada (May 1989), pp. 707-715.
Index
AASHO road test 74 coupled models 81
acceleration 62 crack growth 38
aggregates 173 cracking 5
aircraft crater 150
tests 167 crazing 134
wheel loads 150 curling 140
airfield construction cutting
ARE fatigue model 74 machines 8
analysis models 78, 174 saws 166
analytical
methods 103, 175 damage 54
studies 103 Darter fatigue model 74
data interpretation 29
back-fill 151 deflection 142
base layer 158 dense liquid foundation 18, 19, 20
beams 106 design 9
bedding example 47, 90
layer 159 method 45
sand 39 dimensional analysis 24, 28, 29, 30
blow-up 139 discrete element models 80
bouncing 62 dolly wheels 64
Boussinesq 19, 27 durability 154
braking 62 dynamic
British Ports Association 54 compaction 152
bulk fill 158 loading 53
solution III
channelization 60 edge detail 161
clay bricks I elastic solid foundation 18, 19, 23
closed form solution 20, 25 energy distribution '131
computer experimental studies 113, 114, 118
modelling 44
programs 27 fatigue
concrete models 73
age 65 relationships 12, 64, 65, 68, 69, 70, 71,
blocks I 73, 74, 75, 76
cutting saws 166 faulting 148
fatigue relationships 12, 64 fibre reinforcement 6, 120, 174
pavement design 9 finite difference method 26, 104
strength 65 finite element models 80, 83, 104, 106, 108,
stress 94, 174 175
construction 16 finite sized slabs 10
containers 52 flexural strength 6
stacking 63 flush repair systems 152
terminals 52 foundation 18
cornering 62 frictional stress 34
cost 7 front lift trucks 53, 57
192 INDEX

geotextiles 158 PCA fatigue model 73


PCPU 76
handling stresses 14 performance improvement 120, 123
handrails 108 plank sections 3
hard impacts 102 plastic shrinkage 134
heavy aircraft tests 167 port
hexagons 2 loading 53
hydraulic effect 134 pavement model 76
practical tests 167
impact resistance 131
precast
impulsive loads 101, 120
concrete slab 4
installation speed 156
operations 6
jacks 64 prestressing 5
joints 15, 176 pumping 146
efficiency 144 punch-outs 9
sealant 146
raft
layered systems models 79 design 38, 45, 160
laying edge detail 161
concrete pavements 134 laying 38, 176
rafts 176 lifting beam 165
lifting materials 173
beam 165 production 37
holes 173 reinforcement 84, 163
liquid foundation 18 size 83, 160
load storage 173
application 69 use 37
characteristics 101 rapid pavement repair 150
modelling 103 reclaimed land 53
placement 32, 83 reinforcement 5, 6, 84, 163, 173
position 83 repairs 6, 170
radius 33 boundary 163
transfer 145, 162 design 153, 156
loading 14, 42, 43, 53, 56, 62, 153, 175 criteria 40
frequency 54 rest periods 68
range 54 ride quality 155
rate 69 RISC distress function 76
road
manufacturing rafts 173 construction 2
mat systems 152 repairs 170
materials 173 roughness 155
methods of analysis 39
military needs 153 sand bedding 159
missile mats 3 saws 166
modulus of subgrade reaction 19, 39, 154 screed beam 164
moisture movement 136 sealant 146
moving rafts 173 semi-infinite slabs 9
multiple wheel loads 31 serviceability limit state 40
non-linear models 104 shrinkage 134
numerical side loaders 53, 58
analysis 25, 175 skid resistance 173
m"odelling 44 slabs 9, 10, 153
dimensions 33
parametric study 85 handling stresses 14
pavement models 78, 108
damage 54 on grade 18, 25
design 9 prestressing 5
repairs 7, 150, 170 reinforcement 5
INDEX 193
small paving 6 temperature
soft impacts 101 change 136
soil-structure interaction 18 curling 32
solar heating 136 test program
Soviet Union 2 for beams 118
special equipment 164 for handrails !l4
stabilizing jacks 64 for slabs !l8, 167
stacked containers 53 thermal
standard axle loading 43 curling 137
static loading 62 effects 134, 137
steel fibre reinforcement 120 tractor units 53, 59
straddle carriers 53, 56 trailer units 53, 59
stress 14 transportation 6
range 70
ratio 12
structural ultimate limit state 40
analysis 77 undowelled pavements 132
efficiency 4 USAF distress model 75
response 141
subbase 176 Vesic distress model 76
subgrade 7, 10, 14
deflection 42
loading 14 warping 135, 140
stress 95, 174 Westergaard 20, 78, 141, 175
support 18 wheel load 56, 150
sub-surface screed beam 164 wheels 64
surface types 152 Winkler 19, 27, 78, 141, 175

S-ar putea să vă placă și