Structural Engineers Design Manual

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 304

GHD Pty Ltd

GHD Structural Design Manual


Revision 2
July 2010
Contents

Preface i

1. Introduction 1
1.1 General 1
1.2 Structural Design 2
1.3 Role Descriptions 3

2. Structural Design Documentation 4


2.1 Introduction 4
2.2 Concept Design Phase 4
2.3 Preliminary Design Phase 6
2.4 Developed Design Phase 10
2.5 Detailed Design Phase 13
2.6 Tender / Preconstruction Phase 15
2.7 Summary 15
2.8 References and Further Reading 16

3. Legible / Traceable Calculations 17


3.1 Introduction 17
3.2 General 17
3.3 Design Information 18
3.4 Component Design 18
3.5 References and Conclusions 19
3.6 Summary 19
3.7 Set-out of Calculations 19
3.8 Checking and QA 20
3.9 References and Further Reading 20

4. Software 21
4.1 Introduction 21
4.2 Finite Element Analysis (FEA) 21

5. Design Standards and References 24


5.1 Introduction 24
5.2 GHD Standard Structural Details 27
5.3 References and Further Reading 27

Used as a guide only, QA procedures must be followed 2


42/01012/07/44062 GHD Structural Design Manual
Revision 2
6. Structural Systems 28
6.1 Introduction 28
6.2 Overall Structure 28
6.3 Design Requirements 28
6.4 Braced Frames/ Trusses 29
6.5 Sway verses Braced Frames 29
6.6 Diaphragms 30
6.7 Floor Structure Systems 30
6.8 Structure Robustness and Redundancy 32
6.9 Structural System Selection 32
6.10 References and Further Reading 32

7. Serviceability 33
7.1 Introduction 33
7.2 Deflections 33
7.3 Durability 37
7.4 Vibrations 38
7.5 References and Further Reading 39

8. Design Loadings 40
8.1 Introduction 40
8.2 Design Standards 40
8.3 General Principals 41
8.4 Permanent / Dead Loads 41
8.5 Imposed / Live Loads 42
8.6 Wind Actions 43
8.7 Earthquake Load 44
8.8 Loading Calculations 46
8.9 References and Further Reading 47

9. Standard Technical Specifications 48


9.1 Introduction 48
9.2 Technical Specifications 48
9.3 Performance Specifications 48

10. Foundations 49
10.1 Introduction 49
10.2 General 49
10.3 Ground Investigation - Geotechnical Investigation Report 49

Used as a guide only, QA procedures must be followed 3


42/01012/07/44062 GHD Structural Design Manual
Revision 2
10.2 General 49
10.3 Ground Investigation - Geotechnical Investigation Report 49
10.4 General Foundation Design Approach 50
10.5 Foundation Bearing Pressure and Net Bearing Pressure 50
10.6 Bearing Capacity Estimates 51
10.7 Foundation design 58
10.8 Other Issues 62
10.9 References and Further Reading 62

11. Retaining Walls 63


11.1 General 63
11.2 Retaining Wall Types 63
11.3 Presumed Geotechnical Design Parameters for Concept Design 64
11.4 Surcharge Loads 66
11.5 Design 68
11.6 References and Further Reading 70

12. Dynamic Foundations 71


12.1 Introduction 71
12.2 Analysis of Dynamic Foundations 72
12.3 Design 74
12.4 Foundations for Reciprocating Machinery 75
12.5 References and Further Reading 76

13. Concrete Columns and Walls 77


13.1 Introduction 77
13.2 Preliminary Design 77
13.3 Detailed Design 78
13.4 Walls 79
13.5 Standard details 79
13.6 References and Further Reading 79

14. Reinforced Concrete In-situ - Elements 80


14.1 Introduction 80
14.2 General 80
14.3 Preliminary Sizing 80
14.4 Preliminary Reinforcement 88
14.5 Detailed design (Strength – moment and shear; Deflection –
short-term & long-term; Crack – flexural & shrinkage) 89

Used as a guide only, QA procedures must be followed 4


42/01012/07/44062 GHD Structural Design Manual
Revision 2
14.6 Standard Details 91
14.7 References and Further Reading 92

15. Prestressed and Post Tensioned Concrete 93


15.1 Introduction 93
15.2 General 93
15.3 Materials 93
15.4 Tendon Profile 95
15.5 Preliminary Sizing 96
15.6 Standard Details 96

16. Structural Steel 97


16.1 Introduction 97
16.2 General 97
16.3 Steel Grades and Sizes 97
16.4 Frames 97
16.5 Interaction of Shear and Bending 98
16.6 Connections 99
16.7 Protective Coatings 100
16.8 Sizing 101
16.9 Capacity Check 102
16.10 Standard Details 106
16.11 References and Further Reading 106

17. Concrete Masonry 107


17.1 Introduction 107
17.2 General 107
17.3 Preliminary Sizing 107
17.4 Capacity Check 108
17.5 Standard Details 111
17.6 References and Further Reading 111

18. Timber Structure 112


18.1 Introduction 112
18.2 General 112
18.3 Sizing 113
18.4 Capacity Check 113
18.5 Standard Details 117
18.6 References and Further Reading 117

Used as a guide only, QA procedures must be followed 5


42/01012/07/44062 GHD Structural Design Manual
Revision 2
19. Composite Construction 118
19.1 Introduction 118
19.2 Structural Arrangement 118
19.3 Sizing 118
19.4 Capacity Check 119
19.5 Standard Details 125
19.6 References and Further Reading 125

20. Pavement Design 126


20.1 Introduction 126
20.2 Failure Mechanism 126
20.3 Preliminary Thickness Design 126
20.4 Detailed Design 130
20.5 Standard Details 130
20.6 References and Further Reading 130

21. Precast Concrete Structure 131


21.1 Introduction 131
21.2 Hollowcore Slab (Plank) 131
21.3 Precast Concrete Wall 138
21.4 Standard Details 139
21.5 References and Further Reading 139

22. Water Retaining Structures 140


22.1 Introduction 140
22.2 Key Issues 140
22.3 Information Required Before We Start Structural Design 141
22.4 Basic design 141
22.5 Concrete Quality and Mix Design 141
22.6 Estimation of Peak Temperatures 142
22.7 Design for shrinkage and swelling 143
22.8 Intent of AS 3735 143
22.9 Design for Strength 143
22.10 Design for Serviceability 144
22.11 Waterstops 147
22.12 Joints 147
22.13 Design for earthquake 148
22.14 References and Further Reading 149

Used as a guide only, QA procedures must be followed 6


42/01012/07/44062 GHD Structural Design Manual
Revision 2
23. Bin Design 150
23.1 Introduction 150
23.2 General 150
23.3 Steps in the design 150
23.4 Selection Criteria 150
23.5 Material Properties 150
23.6 Design Loadings and Loads Combinations 151
23.7 Structural Design 156
23.8 Material of Construction 157
23.9 Corrosion 158
23.10 Maintenance 158
23.11 References 158

24. Durability in Design, Construction and Maintenance 159


24.1 Introduction 159
24.2 Project Durability Requirements and Asset Design Service Life 163
24.3 Environmental Exposure Classifications 165
24.4 Deterioration Mechanisms 168
24.5 Reinforced Concrete Typical Potential Durability Issues 171
24.6 Metal Elements Typical Potential Durability Issues 175
24.7 Rubbers/Elastomers, Protective Coatings, Sealants and
Adhesives Typical Potential Durability Issues 179
24.8 Waterproofing Reinforced Concrete Typical Potential Durability
Issues 181
24.9 Composites and Plastics Typical Potential Durability Issues 182

Table Index
Table 1 Typical Structural Property of Materials 7
Table 2 Indicative Reinforcement Rates for Concrete
Building Structures (kg/m3) 8
Table 3 Indicative Unit Weight for One-storey Steel Building
Structures (kg/m2) 9
Table 4 Typical grid dimensions 11
Table 5 Typical sizes for vertical coordination dimensions of
space 11
Table 6 Typical floor-to-floor configurations 11
Table 7 List of design documentation for a large-scale
project 15

Used as a guide only, QA procedures must be followed 7


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 8 List of design documentation for a small-scale
project 15
Table 9 Design codes and references 24
Table 10 Common floor structure systems 30
Table 11 Common limits for calculated deflection of beam
and slab 33
Table 12 Common limits for calculated deflection of steel roof
rafters 35
Table 13 Common limits for calculated horizontal drift of steel
frames 36
Table 14 Common limits for calculated deflection of timber
structure 36
Table 15 Structural Design Standards and Guidelines 40
Table 16 Typical Dead Loads for Structures 41
Table 17 Annual Probability of Exceedance - Earthquake 44
Table 18 Types of loading calculation 46
Table 19 List of standard specifications 48
Table 20 Presumptive Bearing Capacities to AS 1726 Soil
Descriptions 52
Table 21 BS 8002 Rock Groups for Use with 54
Figure 7 Design Values for Foundations on Sandstone and
Shale in Sydney Region (after Pells, 1978) 57
Figure 8 Pad on rocks 59
Table 22 Typical Presumed Unit weights of granular soil
(after AS 4678 and BS 8002) 64
Table 23 Typical Presumed Unit weights of cohesive soil 65
Table 24 Presumed Effective Strength Parameters (after AS
4678) 65
Table 25 Common surcharges for retaining walls 66
Table 26 Machine Vibration Level Classes 75
Table 27 Ultimate Compressive Strength 77
Table 28 Specification for Prestressed Strand 93
Table 29 Typical Span/depth for multi-span prestressed
floors 96
Table 30 Typical Connection Types 99
Table 31 Suggested Thickness Slab for Typical Loading
Conditions 126
Table 32 Fatigue Strength of Concrete in Flexure 127
Table 33 Recommended Thickness Slab (mm) for Typical
Loadings from Forklift Trucks 128

Used as a guide only, QA procedures must be followed 8


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 34 Maximum Recommended Joint Spacings for JUCP 129
Table 35 Deformed Bar Reinforcement for CRCP 129
Table 36 Reinforcement for RCP up to 13m Long 129
Table 37 Plank weight based on manufacturer’s design data 137
Table 38 Design Life Period in Australian Standards 165
Table 39 Environment Descriptions Example 166
Table 40 Deterioration Mechanisms Summary 169

Figure Index
Figure 1 Organisation Chart 2
Figure 2 Concept Design Phase 5
Figure 3 Preliminary Design Phase 9
Figure 4 Design Development Phase 12
Figure 5 Detailed Design Phase 14
Figure 6 Rock allowable bearing capacity for square pad
foundations with settlement not exceeding 0.5% of
foundation width (taken from BS 8004) 56
Figure 7 Design Values for Foundations on Sandstone and
Shale in Sydney Region (after Pells, 1978) 57
Figure 8 Pad on rocks 59
Figure 9 Surcharge calculations 67
Figure 10 Dynamic Foundations – General Guidelines 73
Figure 11 Span/d ratio for one-way slabs 82
Figure 12 Span/d ratio for two-way slabs (two adjacent edges
discontinuous) 83
Figure 13 Span/d ratio for two-way slabs (three edges
discontinuous, one long edge continuous) 84
Figure 14 Span/d ratio for two-way slabs (four edges
discontinuous) 85
Figure 15 Span/d ratio flat plate 86
Figure 16 Span/d ratio for simply supported beams 87
Figure 17 Span/d ratio for Continuous beams 88
Figure 18 Cross-sectional area for bars 90
Figure 19 Cross-sectional area per metre width for bar
diameters 90
Figure 20 Mesh specification 91
Figure 21 Standard Post-tensioning Duct Sizes 94
Figure 22 Tendon Profiles 95

Used as a guide only, QA procedures must be followed 9


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 23 Portal design charts (cross wind knee moment) 103
Figure 24 Portal design charts (longitudinal wind knee
moment: 5 m frame spacing) 104
Figure 25 Portal design charts (longitudinal wind knee
moment: 9 m frame spacing) 105
Figure 26 Wall properties and wall compressive load capacity 109
Figure 27 Distributed-load Capacities for Continuous-Span
Lintels 110
Figure 28 Distributed-load Capacities for Continuous-Span
Lintels 111
Figure 29 Timber post supporting roof or floor loads 114
Figure 30 Bearers 114
Figure 31 Joists 115
Figure 32 Strap brace – Speedbrace Type A Unit 116
Figure 33 Strap brace – Speedbrace Type B Unit 117
Figure 34 Span tables for simply supported composite beams
(office floors) 121
Figure 35 Span tables for simply supported composite beams
(retail floors) 122
Figure 36 Span tables for simply supported composite beams
(plant rooms) 123
Figure 37 Span tables for simply supported composite beams
(car parks) 124
Figure 38 Load capacity table for 150 mm thick hollowcore
plank 132
Figure 39 Load capacity table for 200 mm thick hollowcore
plank 132
Figure 40 Load capacity table for 250 mm thick hollowcore
plank 133
Figure 41 Load capacity table for 300 mm thick hollowcore
plank 133
Figure 42 Load capacity table for 350 mm thick hollowcore
plank 134
Figure 43 Load capacity table for 400 mm thick hollowcore
plank 134
Figure 44 Linear load distribution of point load (PCI) 135
Figure 45 Load distribution coefficients – centrally-placed line
load (FIP) 135
Figure 46 Load distribution coefficients – point load (FIP) 136
Figure 47 Load distribution coefficients – point load at edge
(FIP) 137

Used as a guide only, QA procedures must be followed 10


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 48 Structural adequacy, fire-rated wall 138
Figure 49 The Boundaries between Mass Flow and Funnel
Flow 151
Figure 50 Classification and Combination of Loads 152
Figure 51 Load Factors for Ultimate Strength Design (Ref.
Table 4.2, AS 3774-1996) 153
Figure 52 Examples of Loads from Bulk Solids during
Symmetric Filling and Discharging Conditions. 155

Appendices
A Standard Calculation Summary Sheet
B Standard Details in AutoCad
C Template for Design Information
D Analysis Formulae
E Useful Design Data
F Soil and Rock Description
G Prestress Design Notes
H Water Retaining Structures – Details

Used as a guide only, QA procedures must be followed 11


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Preface

This manual has been prepared to outline design procedures in order to standardise and improve the
quality of:
Design documentation;
Design approach; and
Format of structural calculations.
This manual provides direction and guidance for structural engineers on “how to design” and “where to
find useful and necessary information to carry-out design efficiently.
Refer to other GHD standard manuals for issues with regard to project management, job management,
quality management, document control, etc.
This manual will be updated periodically and suggestions for improvements are encouraged. There is a
large store of scanned references held in the N:\AU\Adelaide\Service\Structural\Technical Information,
Publications and Papers.

Used as a guide only, QA procedures must be followed i


42/01012/07/44062 GHD Structural Design Manual
Revision 2
1. Introduction

1.1 General
Structural engineering deals with the design of any structural system(s) for the purpose to support and
resist various loads applied to the structure directly or indirectly
Structural engineers are typically involved in the design of building and non-building structures, which
comprise of in-ground supporting systems typically called footings, main supporting structural frame,
super structure, cladding and supporting members. In addition the design of fixings and fittings
supporting non structural components may apply significant loads (gravitational or pressure induced
forces) that can form part of the design brief.
In building construction, the structural engineering field is a subset of civil engineering. In a practical
sense, structural engineering is largely the application of Newtonian mechanics to the design of structural
elements and systems that support buildings, bridges, walls (including retaining walls), dams, tunnels,
etc.
The analysis and design of structures involves the ability to assess the imposed actions on a structure
and design the components to withstand these imposed actions with a sufficient margin of safety against
the applicable limit states of design. Once this has been successfully completed, verified and
documented the construction team can safely fabricate and erect the structure and ultimately satisfy the
requirements of the Client.

Structures cover a large and varied field involving both marcro and micro design issues from concept and
scheme design through to detailed design and construction. Macro issues involve overall structure
stability and gross frame analysis whilst micro issues deal with individual component sizes and how they
are connected together.
Other areas that structural engineers can be responsible for include:
Whole of life costing i.e. durability of materials and serviceability design;
Guidance on value management involving improvement in existing design solutions for either cost,
time or functionality;
Assessment and review of existing structural systems and their ability to satisfy current limit states;
Consideration of construction/erection loads;
Inspection and verification of structural components during erection;
Reviewing fabrication and erection drawings; and
Pre-demolition inspection and guidance on appropriate demolition sequences.
The above list is by no means exhaustive. Good design practice dictates the early involvement of an
experienced structural design practitioner (5 –10 years minimum) who can guide less experienced
designers towards appropriate design objectives.

Used as a guide only, QA procedures must be followed 1


42/01012/07/44062 GHD Structural Design Manual
Revision 2
1.2 Structural Design
Structural design is an interactive process that creates a functional, economic and safe structure.
Equipped with engineering analytical skill and past experience, structural engineers are able to develop a
solution, conforming to the technical specifications, based on:
Client’s requirements;
Site conditions and constraints;
Design intention sketches for relevant parties;
Local regulations and design standards; and
Conditions of consents.
Generally in accordance with GHD procedures a thorough design process consists of 5 key phases:
1. Concept design phase (5%);
2. Preliminary design phase (15%);
3. Developed design phase (50%);
4. Detailed design phase (85%);
5. Tender / preconstruction phase (100%);
6. For a relatively small project, phases 1~3 can group into a preliminary design phase; and
7. Coordinated by the Job Manager (structural), all members of a design team (see Figure 1) must work
together to:
Ensure that applicable design criteria have been identified and applied to the design process;
Present a complete set of verified design documentation (refer Section 2) to GHD standard formats
shown in this manual or other relevant GHD manuals.
The standard design process is illustrated in Figure 1. The diagram describes the flow of responsibility
and tasks required to complete a structural design process.

Figure 1 Organisation Chart


Design Documents

Used as a guide only, QA procedures must be followed 2


42/01012/07/44062 GHD Structural Design Manual
Revision 2
1.3 Role Descriptions
Structural Engineer – a professional person who is involved in the design and supervision of the
construction of structures and structural components in: houses, theatres, sports stadia, hospitals,
bridges, oilrigs and office blocks.
Job Manager (Structural) (JM) – a senior structural engineer in charge of providing structural engineering
services (design report, design brief, technical specification and general notes).
Design Engineer (Designer) – a structural engineer undertaking detailed design for all or part of the
structure (design sketches, structural analysis and calculations).
Structural Drafter (Drafter) – a draftsperson preparing drawings for projects (standard and specific
details).
Design Checker (Checker) – a senior structural engineer with appropriate experience verifying the
Designer’s calculations.
Project Director (PD) – a service group manager/principal professional approving project deliverables.
Queensland: a Registered Professional Engineer of Queensland (RPEQ, civil); Victoria: A Registered
Building Practitioner; other states: a Chartered Professional Engineer (CPEng)

Other disciplines – disciplines (such as architectural, civil, hydraulic, electrical, mechanical, quantity
surveyor, etc.) other than the structural discipline involved with the design process.
The Client – the person or company responsible for engaging and paying for the structural engineering
service.

Used as a guide only, QA procedures must be followed 3


42/01012/07/44062 GHD Structural Design Manual
Revision 2
2. Structural Design Documentation

2.1 Introduction
Design documentation is a compilation of documents (reports, calculations, certificates, etc.) and
drawings (sketches, detailed drawings etc). It is a comprehensive written document prepared by the
design team showing how engineering theories are applied with the consideration of local regulations,
functional, economic, constructible and safe structural principals. The actual content of design
documentation varies to suit the nature of different projects and design phases. The aim of this section is
to provide engineers with an understanding of the design process and the requirements for structural
design documentation.

2.2 Concept Design Phase


In the Concept Design Phase, a viable structural system will be developed in collaboration with the Client
to set the direction of a project. typically during this phase the Job Manager (JM) should:
1. Make the Client aware of structural implications regarding the Client’s requirements;
2. Identify all the necessary inputs (note: site visit(s) must be undertaken for projects involving
modification to existing structures);
3. Study viable options;
4. Choose preferable options after Project Director’s review (15% review);
5. Provide Design Report with sketches for the client’s approval; and
6. Generally the Concept Design Report consists of:
Introduction;
– Details of the Client
– Brief description of the project
– Scope of work
Design inputs;
– Client brief (Budget, time schedule, design life, special features concepts, etc.)
– Site location (Geotechnical information, terrain category, site constraint, etc.)
– Key risks and assumptions
Critical evaluation of different structural options based on;
– Structural type & form (refer Section 6)
– Main gravity and lateral load resisting systems (refer Section 6)
– Floor system (refer Section 6)
– Ground retention system (refer Section 11)
– Foundation system (refer Section 10)
– Facade system
– Roof system
– Cost (short-term and long-term)

Used as a guide only, QA procedures must be followed 4


42/01012/07/44062 GHD Structural Design Manual
Revision 2
– Constructability
– Feasibility for future extension or modification
Conclusions/Recommendations;
Attachments:
– Sketches
– Correspondence
– Record of discussions
– Information of recommended proprietary products / systems
Examples of Concept Design Reports for previous GHD projects are available. It is recommended that
designers seek previous examples from the project director.

Figure 2 Concept Design Phase

P repa re C o ncept D esign R ep ort,


conce ptu al ske tche s

P D : ap proval of C oncep t D esign R e port

JM : Issua nce of C on cept D e sign R epo rt

C lie nt’s: appro val o f C on cept D e sign R epo rt

Used as a guide only, QA procedures must be followed 5


42/01012/07/44062 GHD Structural Design Manual
Revision 2
2.3 Preliminary Design Phase
In this phase the JM is required to fine-tune the design concept with the output being the Design Basis
Report and preliminary drawings, which are of sufficient quality for a quantity surveyor to carry out an
initial cost estimate.

Design Basis Report


For a large-scale project, the preparation of Design Basis Report is very important as it serves as a
comprehensive document representing the design requirements for a design team. The development of
an up-to-date Design Basis Report will assure the Client and the Project Director that the designers have
considered in detail the requirements of the project. The JM (structural) should produce a Design Basis
Report that typically consists of:
Introduction:
– General project information (project title, site location, project component, name of owner &
employer, etc.)
– Meteorological and topographic information of proposed site
– Purpose of basic Design Basis Report
Structural systems for every component (gravity and lateral resistance, refer Section 6);
Design codes, reference, aids and guidelines (inclusive of guideline from the Client):
– Local design codes, standards and commentaries (refer Section 5)
– Design reference and guidelines (refer Section 5)
– Documentation reference (specified by the Client)
– Technical specifications (for key elements at this stage, refer Section 9)
– Computer software (refer Section 4)
Load criteria (refer Section 8):
– Material dead loads
– Imposed loads
– Wind loads
– Earthquake loads
– Geotechnical loads
– Dynamic loads
Materials Properties (refer Table 1):
– Concrete
– Steel
– Masonry
– Timber
– Others
Geotechnical (refer Sections 10 & 11):
– Basic of geotechnical design
– Pile capacity

Used as a guide only, QA procedures must be followed 6


42/01012/07/44062 GHD Structural Design Manual
Revision 2
– Pad capacity
– Lateral earth pressure
– Surcharge
– Hydraulic pressure
Design limit states:
– Ultimate strength limit state
– Serviceability limit state (limits of deflection, vibration, etc. refer Section 7)
– Stability limit state
– Durability (for aggressive soil condition)
– Crack control (watertightness design)
– Fire or acoustic engineering requirement
– Construction methodology governing the design
Check and reviews (levels of checking based on design phase).
Attachments.

Table 1 Typical Structural Property of Materials

Material Modulus of Shear Poisson’s Thermal Density


elasticity, E modulus, G ratio expansion (kN/m3)
(GPa) (x10-6 /oC)
Concrete at 28 21~341 0.42E 0.2 10 24
days
Steel 205 0.38E 0.3 12 78.5
Aluminium alloy 70 0.37E 0.33 23 27.2
Stainless steel 200 0.3845E 15 79
304L
Aluminium 105 0.42E 0.3 16-19
bronze
Cast iron 65-95 0.4E 0.25 11-13 70.7
Wrought iron 150-220 0.4E 0.25 11-12 75.4
Timber – 14.2 7.3
hardwood
Timber – 9 5
softwood
Water 60 9.8

Note:
1. Long term Modulus of Elasticity approximately 50% of stated value.

Used as a guide only, QA procedures must be followed 7


42/01012/07/44062 GHD Structural Design Manual
Revision 2
At the end of the Preliminary Design Phase the JM should produce a structural system that is:
Clear and straightforward (i.e. load paths); and
Based on rational design and detailing.

Preliminary Drawings
Preliminary drawings in this stage shall include:
1. All key plans (smaller scale) with adequate information for initial cost estimates (reinforcement rates
& steel unit weights).
2. Necessary sections showing the configuration and sizes of proposed primary frames.
3. Preliminary foundation layout with adequate information for the initial cost estimate (reinforcement as
kg/m 3, steel as kg/m2, etc).
4. Critical details that have significant cost implications (special structural connections, strengthening to
existing structures, etc.).
5. Typical values of reinforcement rates (kg/m3) and steelwork usage (kg/m2) based on the types of
development are shown in Table 2 and Figure 3 as guidelines. The actual value to be adopted can
vary and depends very much on experience and engineering judgement. It is advised that for any
large-scale project where a variation in rate will create a significant error in job costing, that the rates
be built up from a preliminary design.

Table 2 Indicative Reinforcement Rates for Concrete Building Structures (kg/m3)

Column Beam PT Slab Slab Wall


Residential 150 ~300 180 ~ 250 PT = 5.0 120 100
development
6. Rebar =
40
Commercial 200 ~ 350 200 ~ 300 PT = 6.5 150 100
development
Rebar = 45
Light industrial 250 ~ 300 250 ~ 300 PT = 7.5 175 125
development Rebar = 55
Light warehouse 300 ~ 350 250 ~ 300 PT = 7.5 175 125
(imposed live Rebar = 55
load less than
10kPa)

Used as a guide only, QA procedures must be followed 8


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 3 Indicative Unit Weight for One-storey Steel Building Structures (kg/m2)

Commercial Light industrial Medium industrial Heavy industrial


development development (with development (with development (with
(braced frame) light duty cranage) medium duty cranage) heavy duty cranage)

20-25 20 for portal frame 40 for portal frame 80 for portal frame
2.7 for purlin and 2.7 for purlin and girt 2.7 for purlin and girt
girt

Figure 3 Preliminary Design Phase

Preparation of Design Basis Report

Preparation of prelim drawings

PD: approval of Design Basis Report

JM: issue of prelim drawings

Used as a guide only, QA procedures must be followed 9


42/01012/07/44062 GHD Structural Design Manual
Revision 2
2.4 Developed Design Phase
With Preliminary Design approval from the Client (i.e. in written form) and final geotechnical report,
wind report and schematic architectural details (showing building geometry and anatomy, refer Table
4, Table 5, Table 6) are available, design development can be undertaken by the design team to:
Finalise the set-out of critical vertical members (columns and walls);
Determine size of all primary and most secondary structural members; however there may be some
architectural, services and secondary support members not defined at this stage;
Design generic connection details;
Agree serviceability performance criteria with the Client and update the Design Basis Report
accordingly;
Confirm assumptions with relevant parties and update the Design Basis Report accordingly;
Identify input by other disciplines (floor openings, service ducts, beam penetrations, etc.); and
Incorporate likely erection / construction requirements.
A design review (50%) with the Project Director and the checker shall be arranged by the JM before
he/she passes design tasks to designers.
The JM shall make sure all design criteria to be applied to design tasks are described on Job Briefing
Sheets (see 3.3).
Designers shall not proceed with calculations without the following:
Design Brief (including latest revision of the Design Basis Report);
Preliminary sketches/drawings;
Architectural details; and
Geotechnical investigation report.
Designers shall study and understand the given information before they start their calculations.
Refer to Section 3 for more details on how to prepare legible and traceable calculations.
Following the development of the design the preliminary drawings prepared during the preliminary stage
should be revised to include:
1. Key plans with information for cost estimates (reinforcement as kg/m 3, steel as kg/m2, etc)
2. Layout and size of secondary framing members (e.g. lift, stairs, canopies and platforms)
3. Preliminary foundation layout with adequate information for cost estimates (reinforcement as kg/m 3,
steel as kg/m2, etc)
4. Generic reinforcing details for typical primary elements
5. Typical connection details for primary elements
6. Define elements covered by proprietary design (precast floor, roof panel, piling, etc.)

Used as a guide only, QA procedures must be followed 10


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 4 Typical grid dimensions

Type of building Preferred dimensions (m)

Offices and retail 6.0, 7.2, 9.0, 10.5, 12, 15 grids

Retail outlets 5.5, 9 or 11 grids

Car parks 8.4 grids

One-storey industrial building 6.0, 7.2, 9 between portal frames

Table 5 Typical sizes for vertical coordination dimensions of space

Dimension / spaces Range of space (mm) Multiples of size (mm)

Floor to ceiling, floor to floor 3600 100


(and roof)
3600 ~ 4800 300

> 4800 600

Zones for floors and roofs 100 ~ 600 100

> 600 300

Changes of floor and roof levels 300 ~ 2400 300

> 2400 600

Openings in walls 300 ~ 3000 300 or 100

Table 6 Typical floor-to-floor configurations

Zone name Dimension reference Requirement Common values

Structural Zone Height of structural Specified by Based on selected


member structural engineers grid dimensions and
for strength structure system
requirements (refer design
guidelines)

Service Zone Maximum deflection of Specified by 50 mm maximum


structural member structural engineers
for deflection
requirements

HVAC ducts or terminal Specified by Approx 500 mm


device mechanical
engineers

Support for HVAC ducts Proprietary products Approx 50 mm

Sprinkler zone Specified by 50 ~ 150 mm


hydraulic engineers

Used as a guide only, QA procedures must be followed 11


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Zone name Dimension reference Requirement Common values

Light and ceiling zone Specified by electrical 150 mm


engineers

Headroom Clear headroom Specified by architect 2.4 m for office;


/ client
2.1 m for toilet,
2.1 m for car parks,
etc

Floor zone Raised floor Specified by electrical 200 mm ~ 450 mm


engineers

Figure 4 Design Development Phase

JM: update Design Basis Report,


prepare Job Briefing Sheets

JM : collate all inputs for


Detailed D esign

Designer:
Is design info, arch details
& geotechnical report
available?

Designer: calculation of m ajor


elem ents & connections

Drafter: CAD dwg preparation


from designer’s sketches

Used as a guide only, QA procedures must be followed 12


42/01012/07/44062 GHD Structural Design Manual
Revision 2
2.5 Detailed Design Phase
In this phase, the design team creates documents to clearly define the design of all structural elements.
Design details should be co-ordinated with other disciplines and the design sketches should be updated
with CAD details. The complete content of structural calculations is indicated in Section 3.
The design team shall be given final architectural layouts + typical details and coordinate with
drafters to make sure the following is shown on structural drawings:
Dimensions of the main building grids, critical structural elements, and other elements that are the
direct responsibility of the structural engineer;
Reference the architectural plans or other disciplines for other dimensions (unless agreed otherwise)
Design and documentation of secondary architectural elements where requested by the Architect
Reinforcing details and connection details
Note: The level of design details shown on drawings in this phase, particularly for concrete and masonry
elements, varies in industry between regions, building type and procurement methodology. A major
factor is the capability of the local building industry to efficiently provide the construction phase
documentation. The level of details outlined in these guidelines is appropriate where the contractor has
the skills and resources to efficiently provide construction phase documentation. For some projects the
design consultant may require a greater level of detailing. The appropriate level of design details required
should be agreed with the client prior to the commencement of the project. Refer to GHD Standard
Details (path to be advised) for more details.

The primary content of structural drawings should include:


General notes
Drawings defining all structural elements, including, plans, elevations, sections and details, with
adequate cross-referencing;
Define all connections by either defining specific connection details or referencing to GHD Standard
Details or specifying forces for a propriety connection system
Construction sequences and positions of control/construction joints
Include stairs, plant platforms and glazing support
Pre-camber/set established for members
Based on the actual requirements of the project, technical specifications (refer Section 9) shall be
arranged by the JM for:
Each structural trade;
Performance requirements;
Performance criteria for proprietary design;
Method statement for critical construction processes governing design.
As shown in Figure 1, the JM is the main coordinator and supervisor for the design process they shall
ensure that calculations and detailed drawings satisfy GHD QA processes prior to seeking Project
Director’s approval.

Used as a guide only, QA procedures must be followed 13


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 5 Detailed Design Phase

Designer: calculation of major


elements & connections

Drafter: CAD dwg preparation


from designer’s sketches

Compile documentation

Used as a guide only, QA procedures must be followed 14


42/01012/07/44062 GHD Structural Design Manual
Revision 2
2.6 Tender / Preconstruction Phase
A Structural Design Certificate is usually prepared by the JM and signed by the Project Director before
the issue of tender/construction drawings.
The JM should have one complete set of design documentation in order to undertake:
Feedback to Requests for Information (RFIs)
Preparation of necessary certificates
– Form 15 (QLD) – Compliance Certificate for Building Design or Specification
– Form 55 (TS) – Certificate of Others (Buildings)
– Forms for other states

2.7 Summary
The tables below show the structural design documentation required for each design phase:

Table 7 List of design documentation for a large-scale project

Concept Preliminary Design Detailed Tender


Design Design Development Design

Design Report

Design Basis
Report

Marked-up
Plans

Sketches

CAD drawings

Specifications

Calculations

QA sheets

Design
Certificates

Table 8 List of design documentation for a small-scale project

Preliminary Detailed Tender


Design Design

Design Basis
Report

Marked-up
Plans

Sketches

Used as a guide only, QA procedures must be followed 15


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Preliminary Detailed Tender
Design Design

CAD drawings

Specifications

Calculations

QA sheets

Design
Certificates

2.8 References and Further Reading

Used as a guide only, QA procedures must be followed 16


42/01012/07/44062 GHD Structural Design Manual
Revision 2
3. Legible / Traceable Calculations

3.1 Introduction
The purpose of this section is to assist designers in presenting structured legible calculations and design
documentation.

3.2 General
Structural calculations prepared after the 50% stage (preliminary design phase) are critical documents
required to demonstrate safety and structural integrity.
A set of calculations should be separated into individual categories for each structural component.
Typically this would include the following:
Standard calculation summary sheet (see Appendix A)
QA Sheets
Contents page
Design information (see Appendix C)
Component design
– Foundation
– Substructure
– Columns and walls
– Component design shall consist of following essential portions:
Design intention/subject
Summary sketches showing component configuration and actions
Analysis results
Conclusion
Summary
– Floors and superstructures
– Stairs
– Connections
– Facades
Note: be consistent with units and clearly identify span assumptions, supports, reactions, shear
forces, bending moments and whether the load is a serviceability limit state (SLS) or ultimate limit
state (ULS). Include the ULS load factors and the combinations that are being used.
Attachments
– Design brief
– Computer outputs for a group of structures
– List of latest architectural and structural drawings

Used as a guide only, QA procedures must be followed 17


42/01012/07/44062 GHD Structural Design Manual
Revision 2
It is compulsory in GHD for designers to:
Prepare an empty file with category folders before commencing calculations
Always complete the calculation page header
Ensure references on drawings match those in calculations
Present calculations in accordance with the format set out in this section

3.3 Design Information


Design engineers shall review and understand the Job Briefing Sheet (GHD Standard form: QA020A)
together with a copy of design information from the JM before commencing design work. The designer
shall study and comprehend all design inputs prior to commencing calculations.
Reasonable assumptions should be made and shown in design criteria if some of the inputs are not
confirmed. Correspondence or records of discussion must be printed out and included as attachments of
Design Basis Report.
As relevant parties should confirm assumptions during the detailed design stage, structural calculation
shall be updated accordingly.

3.4 Component Design


Prior to commencing the component design, it is recommended that a discussion with the design team
be convened to:
Understand the structural system
Discuss analysis methods
Identify all the components (i.e. floor structure, columns, walls, foundations etc)
Identify all interfaces (beam and slab, beam and column, column and base etc)
The designer should undertake a hand calculation of the structural component to develop the end result
before launching into the computer model. This can help identify potential errors in the model. Do not rely
on the software to do the member design unless you can competently set the correct design parameters.
Designers shall seek the advice of senior engineers if unsure.
All calculations should be neatly set out and legible. Always fill out the calculation page header and
number on every page. A typical set out of calculation components is shown in Section 3.7.
It is very common that structural analysis always covers several components. Computer print-outs could
be shown as “sketches” for the component design with following contents:
Configuration of component with necessary dimensions
Loading diagrams for all primary load cases (dead, live, earth pressure, operation loads, etc.)
Results of analysis showing deflection, moment diagrams, shear diagram, axial loads, etc.
Results of members designed to the requirements of serviceability, stability, strength, etc.

Used as a guide only, QA procedures must be followed 18


42/01012/07/44062 GHD Structural Design Manual
Revision 2
3.5 References and Conclusions
The calculation shall clearly indicate references quoted and conclusions drawn:
When a formula reference is quoted, always write it down in full and place it in the left hand margin of
the calculation page.
When a clause from a design standard is quoted, always quote the clause number and standard
When a computer result is used, always print out all the necessary “sketches” (refer 3.3)

3.6 Summary
It is recommended that a summary of all the subjects covered by the calculations be included to:
Make sure the design outputs meet the requirements of design inputs,
List out the results together for self-certification and design check,
Highlight all the outstanding assumptions and have them indicated on tender / construction drawings
accordingly.

3.7 Set-out of Calculations


A proper structural calculation set should be prepared with GHD standard QA sheets and presented with
the following contents:
1. Design information (criteria)

1.1 Structural System


1.2 Loading Criteria

1.3 Limit States & Limits


1.4 Material Specifications
1.5 Design Reference & Software
1.6 Assumptions
2. Foundation/shoring/retaining wall Design
2.1 General arrangement
2.2 Column / Wall Load Chase-down
2.3 Design of foundations
3. Substructure (pile cap, slab on ground, etc.) Design

3.1 General arrangement


3.2 Design of pile caps
3.3 Design of slab on ground
4. Superstructure Design
4.1 Key plans and elevations

Used as a guide only, QA procedures must be followed 19


42/01012/07/44062 GHD Structural Design Manual
Revision 2
4.2 Structural Models
4.3 Loading Diagrams
4.4 Overall structural appraisal (deflection)
5. Element (columns, walls, floors, stairs) design

5.1 Sketches
5.2 Member design
7.3 Summary
6. Connection design
6.1 Input from output of structural analysis
6.2 Sketches of connection
6.3 Summary
7. Outstanding issues

Appendix A – Computer Analysis Results

Appendix B – Design Certificate

3.8 Checking and QA


The appointed Checker should not proceed with checking unless the calculations have been prepared
with the format set out in this section. Checking the final calculations is to be carried out in accordance
with GHD Quality Assurance Procedures.

3.9 References and Further Reading

Used as a guide only, QA procedures must be followed 20


42/01012/07/44062 GHD Structural Design Manual
Revision 2
4. Software

4.1 Introduction
Application of commercial software is very common in the design development phase and detailed
design phase because it is fast, accurate and easy to check. Pre-ordination of softwares according to
their functions will make design checkers concentrate more on the design inputs and outputs without
spending too much time on the analysis process. Designers shall familiarise themselves with software in
common use through internal and external training sessions.
All current available structural software can be found on Compass (Lotus Notes) in the GALEXE
database. Software requests are managed through this database.

4.2 Finite Element Analysis (FEA)


The use of FEA has historically been associated with the mechanical, mining and bridge design
industries. Increasingly FEA analysis is becoming more widely available and accepted into mainstream
structural design practices. The analysis packages have been adapted to cater for building and civil
structures and FEA enables detailed visualization of where structures bend or twist, and indicates the
distribution of stresses and displacements.
The modern FEA packages include specific components such as vibration, thermal, wind and fluid
pressures working environments and some these packages can be integrated in BIMM systems. In a
structural simulation, FEA can help in producing stiffness and strength visualizations and in minimising
weight, materials, and costs.

Common FEA applications include:


Algor
ANSYS
Etabs
Ram Concept
Robot
SAP
SAM
Staad Pro
Strand 7
These applications may also be used to model soil structure interaction.
The decision to proceed with FEA analysis should costed in at the original proposal stage and the
development of the model should take place under the watchful eye of an experienced practitioner.

Used as a guide only, QA procedures must be followed 21


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Software such as Ram Concept are specific structural packages and are designed to work for structural
systems. The general purpose packages (eg. ANSYS, Strand 7 etc.) by their nature are potentially much
more open to inappropriate use in that they’re much more flexible in how models are connected,
constrained, loaded, material properties established, analysed etc. It should be noted that the general
packages should not be used unless persons are suitably trained and/or supervised.

4.2.1 Points to note and common pitfalls


Modern FEA packages provide impressive 3D solid models of a structure. The results look good, but
could be completely incorrect. The following is a brief commentary on common errors encountered in
modelling.
Boundary conditions: Incorrect boundary assumptions are probably the most common form of
errors in FEA analysis results. Not only modelling of elastic constraints but also basic assumptions
on the interaction of the structure with the boundary. Basic constraints, such as symmetry boundary
conditions or interaction between different element types can cause significant errors.
Loading: Loading is a complex topic in most FEA packages. For design standards, loading tends to
be an envelope definition of extreme applied load over space or time, whereas FEA analysis requires
the use of a number of actual or consistent loading conditions to estimate an envelope of extreme
load effects.
Verification: All analysis should include a check that the total load applied in the model matches the
design requirement and that the loading is sensibly pathed through the boundary. Many FEA
packages do not easily provide this report. Review of the result should be done by someone
experienced on other jobs and has some idea of what stress magnitudes, stress patterns and
deflections are expected. If the results do not match the expected conditions, then an explanation is
needed to why the results are in contrast to initial expectations.
Advanced Modelling: All engineering systems have some performance which is not exactly
described by a linear model i.e. performance is non-linear. However the most appropriate solution
may not be to try and model the non-linear effects but to estimate the impact of ignoring the non-
linearity in the FEA solution. Using a non-linear FEA model is only applying a different numerical
model to the analysis. This may not provide a better answer for design. If a non-linear analysis is
required, the initial analysis is usually done as a linear model. For instance, carrying out a response
spectrum earthquake analysis as a pre-cursor to a non-linear full time history analysis. If the
response spectrum analysis indicates that stresses are well within the design limits, then there is no
need to undertake a full time history analysis.
Materials: Most materials used in construction are not of uniform properties. This must be
considered in the design and also the estimation of numerical performance. Variability can have a
significant effect on how the load is transferred through the structure. Common examples are
– Variability of sub-grade stiffness,
– Cracking in RC concrete and also in the sub-grade,
– Variability of steel performance beyond nominal yield,
– Creep and shrinkage of concrete, and
– Creep of most plastics
Numerical modelling of this material performance is often not the best design approach.

Used as a guide only, QA procedures must be followed 22


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Design Criteria: Most design standards are empirical, with the design load and analysis method
matched to the design standard and acceptance criteria. With the adoption of FEA analysis
techniques, it is often that the design criteria need to be reviewed.
A good example is the suitability of detailed FEA modelling of welds and corners in steel design
which can produce estimates of local stress (strain) concentrations which can be difficult to manage
for stress limitations and fatigue. Regions of high stress should not be readily ignored, and must be
checked to see if they are load or displacement controlled. SOME regions of high stress identified in
FEA modelling are not identified by the analysis method that was assumed when the associated
design criteria were developed.
Dynamic Performance: Most structural design considers extreme load and strength. Dynamic
performance is more related to expected load history and stiffness.
FEA uses a sophisticated numerical model to describe a real system. The complexity of the
numerical model is no guarantee that it better describes the real structural performance, especially
when considering the design life of the structure and the materials used.

In applying any numerical model approximation to a real structure it is essential to understand how
well the modelling describes the real performance over time. You should be aware of what the
modelling has not adequately considered and what is the best way to take account of this in the
design.

Used as a guide only, QA procedures must be followed 23


42/01012/07/44062 GHD Structural Design Manual
Revision 2
5. Design Standards and References

5.1 Introduction
The main design reference for a project shall be the Structural Design Brief prepared by the JM. Design
standards and references in common use are shown in Table 9.

Table 9 Design codes and references

Title Latest Version Description Publisher Champions

AS 3600 2009 Concrete Standards To be updated by


Structures Australia each OC

AS 3735 2001 Concrete Standards To be updated by


Structures for Australia each OC
Retaining Liquids

AS 4678 2002 Earth-retaining Standards To be updated by


Structures Australia each OC

AS 2870 1996 Residential Standards To be updated by


foundation and Australia each OC
slabs

AS 3798 1996 Guidelines on Standards To be updated by


earthworks for Australia each OC
commercial and
residential
developments

AS 2159 1995 Piling Design and Standards To be updated by


Installation Australia each OC

AS 4100 1998 Steel Structures Standards To be updated by


Australia each OC

AS 4673 2001 Cold-formed Standards To be updated by


stainless steel Australia each OC
structures

AS 3995 1994 Design of steel Standards To be updated by


lattice towers and Australia each OC
masts

AS 2312 2002 Guide to the Standards To be updated by


protection of Australia each OC
structural steel
against
atmospheric
corrosion by the
use of protective
coatings

Used as a guide only, QA procedures must be followed 24


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Title Latest Version Description Publisher Champions

AS 1418.1 2002 Cranes, hoists Standards To be updated by


and winches Australia each OC

Part 1: General
requirements

AS 1657 1992 Fixed platforms, Standards To be updated by


walkways, Australia each OC
stairways and
ladders – Design,
construction and
installation

AS 3990 1993 Mechanical Standards To be updated by


equipment – Australia each OC
Steelwork

AS 2327.1 2003 Composite Standards To be updated by


structures – part Australia each OC
1: simply-
supported beams

AS 3700 2001 Masonry Standards To be updated by


Structures Australia each OC

AS 1684.1 1999 Residential Standards To be updated by


timber-framed Australia each OC
construction

AS 1720.1 1997 Timber structures Standards To be updated by


Part 1: Design Australia each OC
methods

AS 1720.1 2006 Timber structures Standards To be updated by


Part 2: Timber Australia each OC
properties

AS 1170.0 2002 Structural design Standards To be updated by


actions: General Australia each OC
principles

AS 1170.1 2002 Structural design Standards To be updated by


actions: Australia each OC
Permanent,
imposed and
other actions

AS 1170.2 2002 Structural design Standards To be updated by


actions: Wind Australia each OC
actions

Used as a guide only, QA procedures must be followed 25


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Title Latest Version Description Publisher Champions

AS 1170.3 2003 Structural design Standards To be updated by


actions: Snow Australia each OC
and ice actions

AS 1170.4 1993 Structural design Standards To be updated by


actions: Australia each OC
Earthquake loads

T48 1999 Industrial floors Cement and To be updated by


and pavements Concrete each OC
Association of
Australia

Portal frame 1991 Portal design Australia Institute To be updated by


design charts of Steel each OC
Construction

Masonry design 2005 Concrete Concrete To be updated by


retaining Masonry each OC
structures Association of
Australia Limited

HB 124 2000 Design of Standards To be updated by


concrete masonry Australia each OC
buildings

HB 237 2000 Detailing and Standards To be updated by


construction of Australia each OC
concrete masonry
buildings

HB 218 2004 Pavement Design Standards To be updated by


Australia each OC

DCT/V1/03 1999 Design capacity Australia Institute To be updated by


tables for of Steel each OC
structural steel – Construction
Volume 1: Open
Sections

DCT/V2/02 1999 Design capacity Australia Institute To be updated by


tables for of Steel each OC
structural steel – Construction
Volume 2: Hollow
Sections

DHSC/01 1996 Design of Australia Institute To be updated by


structural steel – of Steel each OC
Hollow section Construction
connections

Design reference 1999 Concrete Longman To be updated by


Structures each OC

Used as a guide only, QA procedures must be followed 26


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Title Latest Version Description Publisher Champions

Design reference 2005 Purlin & Girts Lysaght Structural To be updated by


Solutions each OC

Design reference 2005 Access Products Webforge To be updated by


Division each OC

Design reference 2005 Lysaght W-dek Lysaght To be updated by


Design Software each OC
User’s Guide

Design reference 2005 Using Bondek – Lysaght To be updated by


Design and each OC
Construction
Guide

Technical Manual 2003 Hollowcore National Precast To be updated by


Flooring Concrete each OC
Association
Australia
(NPCAA)

Design reference Steel Design


Guide Series 11 –
Due to Human
Activity

Design reference Floor Vibration in The American


Buildings Institute of Steel
Construction

5.2 GHD Standard Structural Details


GHD Structural Standard details are available in AutoCad and PDF format and can be inserted into
drawings (Appendix B).
Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks) for reference. There is a reference number for every standard detail;
Designers should quote the number when it is adopted.

5.3 References and Further Reading


Standards Online: http://www.saiglobal.com/online/autologin.asp

Used as a guide only, QA procedures must be followed 27


42/01012/07/44062 GHD Structural Design Manual
Revision 2
6. Structural Systems

6.1 Introduction
The structural engineer’s role is to select the appropriate structural system for a structure and ensure it is
detailed appropriately. The structural system will vary depending on various factors including:
End use/client requirements
Constructability requirements
Code requirements
Experience
Material(s) selected for construction
Durability requirements
The structural system selected will determine how the structure performs under the loadings encountered
during its service life.

6.2 Overall Structure


There are a wide variety of structural systems available to the designer. The main structural systems are
as follows.
Bearing Wall Systems: A building system with load bearing walls supporting essentially all vertical
loads with shear walls or braced frames providing the horizontal resistance.
Braced Frame System: A complete space frame system supporting vertical loads with shear walls or
braced frames providing horizontal resistance. Axial forces within the compression struts and tension ties
resist the deflections within these structures. Examples of braced frames are trusses.
Sway Frame System: A sway frame system supports vertical and horizontal loads through bending
stiffness of structural members and connections. A typical example is a portal frame structure.
Dual System: A hybrid structural system essentially comprising of a complete space frame to support
the vertical loads and at least a quarter of the horizontal forces. The total horizontal resistance is
provided by the combination of the moment frame, shear walls or braced frames in proportion of their
relative rigidities.
For the moment resisting frame system, three levels of ductility are recognised: ordinary, intermediate
and special. The design and detailing requirements for each level of ductility is provided in the
appropriate material standards, i.e. AS 3600 Concrete Structures Code and AS 4100 Steel Structures
Code.

6.3 Design Requirements


Design of structural systems demand that the engineer checks the following aspects:
Ultimate strength – ensuring the entire structure has the capacity to resist its ultimate design load
cases and load case combinations.

Used as a guide only, QA procedures must be followed 28


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Serviceability – ensuring the structure can perform its intended function under anticipated service
loads. The typical checks for serviceability are to ensure that deflections of members are within the
guidelines of a particular design code or a design basis document (section 7).
Stability – ensure the structure remains stable under service and ultimate loads. Typical stability
checks ensure that the structure will not overturn, slide or fail the foundation soil bearing capacity.
Natural frequency / resonance –ensuring the structure is designed so that its natural frequencies
avoid those of any driving vibrating forces. These vibrating forces can be generated by machine, by
man or by nature eg. Pumps, walking loads, wind vortex shedding etc. In general the resonant
frequencies of main structural members should be at least 2.5 times greater or smaller than the
driving frequency.
Fatigue – ensure that the fatigue cycles and stress range experienced by the structure do not exceed
those anticipated in design. Fatigue is generally a problem in structures, which are loaded cyclically
throughout their lifetime eg bridges, gantry cranes, frames supporting vibrating equipment. Special
detailing is required to avoid fatigue particularly in steel structures although concrete can also suffer
from fatigue problems.
The above limit states all depend on the structural system selected.

6.4 Braced Frames/ Trusses


A truss may take many forms whether two-dimensional or three-dimensional however they all typically
work on the same principals. When the overall structure resists a force it does so by generating axial
force within the internal members. Depending on the load case one member may experience tension /
compression, shear, bending or torsion although the forces are primarily axial. In a simple two-
dimensional truss under dead load the top chord will be in compression, the bottom in tension & the
struts and ties (web members) in either tension or compression.

Trusses are efficient in long span situations as the moment is resisted by an axial couple formed by the
chord members separated by the depth of the truss (i.e. level arm). Lighter steel members normally
make up the web members since they typically resist forces of smaller magnitude than the chords.
Trusses may be designed with either fixed or pinned connections. Pinned connections are typically used
to connect the web members to the chords although the use of rigid connections may assist in keeping
some redundancy within the structure if a connection or member were to fail. Chord members are
generally continuous to minimise the number of connections required.

6.5 Sway verses Braced Frames


Sway frames are typically formed with members of a high bending stiffness with rigid moment carrying
connections. The vertical and horizontal forces are transferred by bending and shear. Base connections
are generally pinned but not always. Lateral restraint is incorporated into the structure at critical points to
reduce the long effective lengths, which would otherwise render the members likely to fail by lateral
buckling in the minor axis. In a typical portal frame shed it would be normal to assume lateral restraint to
the top flanges at purlin and girt locations and lateral restraint to bottom flanges at fly brace locations.
Braced frames are generally stiffer than sway frames due to the more efficient placement of steel.
However sway frames are more cost effective in some circumstances due to the higher cost of
fabrication & erection of braced frames.

Used as a guide only, QA procedures must be followed 29


42/01012/07/44062 GHD Structural Design Manual
Revision 2
6.6 Diaphragms
In building structures it is economical to use the roof and floor systems as horizontal diaphragms. A
number of features of diaphragm design must be examined in detail:
The diaphragm span: if the vertical resisting elements of approximately equal stiffness are equally
spaced, the diaphragm connecting these elements need only be nominally stiff. If the diaphragm
must span from one end of the building to the other, consideration shall be given to horizontal steel
bracing.
Rigid diaphragm: use of a rigid diaphragm is almost mandatory to distribute the horizontal forces
and torsion in proportion to the relative rigidities of the vertical elements.
Deep Beams: diaphragms can be designed as “deep beams”, with the deck and slab carrying the
shear, and the flanges, such as edge beams at the edges, resisting bending moments. Penetrations
in the floor/diaphragms should be analysed similarly to a plate girder with web openings.
Connections: particular attention is required for the detail of connections when the length of contact
between the vertical resisting element and the diaphragm is limited.

6.7 Floor Structure Systems


The column grids of the structure directly impact the design of the floor structure. Table 10 summarises
the common floor structure systems in accordance with span ranges:

Table 10 Common floor structure systems

Span Material Common systems

5 m ~ 6 m (short span) Reinforced concrete One-way slabs


Two-way slabs
Bondek (Wdek) slabs

Precast concrete Ultrafloor


Floor planks
Transfloor

Steel Simply-supported beams

Timber Bearers and joists

6 m to 10 m (medium span) Reinforced concrete Flat slab


Flat plate
Banded slabs
Two-way slabs
Beam and slab
Waffle slab
Ribbed slab

Used as a guide only, QA procedures must be followed 30


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Span Material Common systems

Prestressed concrete Flat slab


Flat plate
Banded slabs
Two-way slabs
Beam and slab
Waffle slab
Ribbed slab

Precast concrete Ultrafloor


Floor planks
Transfloor

Steel Continuous beams


Composite beam and slab
Trusses

Timber Gluelam
Composite beam and slab
Trusses

More than 10 m (long span) Reinforced concrete Beam and slab


Waffle slab
Ribbed slab
Two-way banded slab

Prestressed concrete Banded slab


Two-way banded slab
Beam and slab
Waffle slab
Ribbed slab
Flat slab

Precast concrete Floor planks (hollow-cored)

Steel Composite beam and slab


Portal frames
Fabricated plate girders
Trusses
Deep castellated I beams

Timber Portal framesTrusses

Used as a guide only, QA procedures must be followed 31


42/01012/07/44062 GHD Structural Design Manual
Revision 2
6.8 Structure Robustness and Redundancy
For multi-storey buildings, there is a robustness requirement to prevent the structure from
disproportionate collapse in British and Australian Standards. It is recommended to ensure in the design
practice that all critical vertical and horizontal members are tied together in all directions.
For structures supported by several independent structures, a certain percent of redundancy based on
the set out of supporting structures needs to be allowed for in the design of supporting structures. This is
to prevent the disproportionate collapse of a structure due to the accidental failure of any one support.
Floor elements such as beams and slabs will need to be designed for non-flexural, large-displacement
action that prevents actual collapse.

6.9 Structural System Selection


Prior to commencement of design it is essential to determine a structural system with consideration to
the above guidelines. In many cases a structural system will chose itself due to its characteristics,
although in other cases options studies may be required to determine the optimum solution.

The selection of a structural system depends on those items mentioned above and also on:
Available space/geometry requirements
Architectural input
Client / builder requirements
Potential end user requirements;
Future use requirements
Environmental factors such as large temperature ranges, flooding risks, corrosion potential and
service life requirements
Ground conditions
Loading regimes such as high earthquake zones or cyclonic regions
Historically it is evident that the cost of labour is much greater than the cost of materials. This is apparent
when one examines old factory sheds with truss portals and riveted bridge structures. Current economic
conditions are driving labour costs higher meaning that using single deeper beam sections, although
heavier can be less expensive to fabricate, coat and erect. Ongoing maintenance costs may also favour
this approach since the surface area for coatings may also be reduced for a single member rather than
multiple smaller members. Off site prefabricated items such as precast concrete & pre assembled steel
frames can also reduce many of the additional costs associated with installation.
In the case of concrete, post tensioned/prestressed beams and slabs are the norm due to smaller steel
percentage requirements in the concrete and the advantages available in longer spans with fewer
supports. However reinforced concrete still has a major part to play in complex components, deep
beams, foundations and slabs on ground.

6.10 References and Further Reading

Used as a guide only, QA procedures must be followed 32


42/01012/07/44062 GHD Structural Design Manual
Revision 2
7. Serviceability

7.1 Introduction
AS 1170.0 Structural design actions Part 0: General principles Clause 1.4.16 defines serviceability as the
“Ability of a structure or structural element to perform adequately for normal use under all expected
actions”.
Serviceability is further described in AS 5100.1 Bridge design Part 1: Scope and general principles
(Clause 6.3.3) to include:
(a) Deformation of foundation material or a major load-carrying element of sufficient magnitude
that the structure has limitation on its use, or is a public concern.
(b) Permanent damage due to corrosion, cracking or fatigue, which significantly reduces the
structural strength or useful service life of the structure.
(c) Vibration leading to structural damage or justifiable public concern.

7.2 Deflections

7.2.1 Concrete structures


Always size for serviceability, reinforce for strength (not the other way around).
Calculations for deflections must consider the effective second moment of area of the member (I e) and
long term deflection due to creep. Section 8.5 of AS 3600 Concrete structures and AS 5100.5 Bridge
design Part 5: Concrete provides methods for calculating the effective second moment of area and creep
for beams. These methods can also be utilised for slabs. For thin structural elements such as slabs (and
other “thin” concrete members) with top and bottom reinforcement, it is incorrect to assume that one
layer of reinforcement (e.g. top layer in the case of simply supported slabs) will be the compressive
reinforcement (Asc) that can be used for determining the ksc factor for calculating long term deflection.
This is because the neutral axis could be above this layer of reinforcement; hence it is not compressive
reinforcement. Failure to account for this can cause major deflection problems. The Designer must
determine whether reinforcement located on the “compressive” face of the element provides effective
compression reinforcement (Asc).
Common floor deflection limits are summarised in Table 11. Use these limits and those specified in the
relevant codes unless otherwise specified by the Client or other disciplines.

Table 11 Common limits for calculated deflection of beam and slab

Type of member Dead Incremental Live Total Load

All non-cantilever 1/250


members

All cantilever 1/125


members

Used as a guide only, QA procedures must be followed 33


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Type of member Dead Incremental Live Total Load

All non-cantilever 1/500 when 1/250


members with provision is made
brittle masonry to minimise the
partitions (the movement effect
deflection which of brittle masonry
occurs after the partitions,
addition or otherwise 1/1000
attachment of the
partitions)

Cantilever 1/250 when 1/125


members with provision is made
brittle masonry to minimise the
partitions (the movement effect
deflection which of masonry
occurs after the partitions,
addition or otherwise 1/500
attachment of the
partitions)

All non cantilever 1/360 30 mm max or 1/500 1/300


members with
non-brittle
masonry
partitions

Compactus areas 1/750 1/500

Transfer beams 10 mm max

Joints between 15 mm max


new and old
structure with
flashings

Bridges: non- 1/800


cantilevered
members

Bridges: 1/400
cantilevered
members

Notes:

i. Compression reinforcement must be always provided in areas with masonry partitions

ii. For long span floors, consideration should be given to adopting an absolute limit for deflection, which will be more stringent
that the values quoted in table above.

In addition to the above, long term deflections for roof members shall be limited to prevent water ponding
and maintain falls to drainage outlets. A minimum fall of 1:100 is suggested.

Used as a guide only, QA procedures must be followed 34


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Limits for lateral drift caused by wind and earthquake force are (where hs = storey height):
Wind forces
Inter-storey drift by service wind hs / 500
Overall drift by service wind total height / 150
Earthquake forces
Inter-storey drift by seismic force 0.05hs k (k refer cl 2.10 of AS 1170.4)

7.2.2 Steel frame structures


The deflection of floor structures is similar to Table 11. Table 12 summarises deflection limit for steel roof
rafters.

Table 12 Common limits for calculated deflection of steel roof rafters

Building type Dead Live Wind Total Load

Industrial and 1/360 1/240 1/150 1/150


commercial
without ceiling,
roof pitch > 3o

Industrial and 1/500 1/240 1/150 1/150


commercial
without ceiling,
roof pitch < 3o

Commercial light 1/360 1/300 1/250 1/250


weight ceiling,
roof pitch > 3o

Commercial light 1/500 1/300 1/250 1/250


weight ceiling,
roof pitch < 3o

Commercial 1/500 1/600 1/600


plasterboard and
acoustic ceilings

Farm sheds 1/250 1/180 1/100


Pre-camber may be used to aid in achieving the above limits.

Roofs with low pitches should be checked to ensure that water ponding does not occur.
The limits for relative horizontal deflection between adjacent frames at the eaves level of industrial
buildings due to wind loads are summarised in Figure 13.

Used as a guide only, QA procedures must be followed 35


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 13 Common limits for calculated horizontal drift of steel frames

Building type Deflection limit Remarks

Building clad in metal sheeting Column height / 150 Relative deflection between
without gantry cranes, no adjacent frames
Frame spacing / 200
ceilings and without internal
partitions against external walls

Building clad in metal sheeting Column height / 250 Column height /300 should be
with gantry cranes, no ceilings used for heavy crane
Frame spacing / 250
and without internal partitions
against external walls

Building with masonry walls Column height / 250 Relative deflection between
supported by steelwork adjacent frames
Frame spacing / 200

Farm sheds Column height / 100 Relative deflection between


adjacent frames
Frame spacing / 100

7.2.3 Timber framed structures


Refer to Table 14 for deflection limits of timber structural elements.

Table 14 Common limits for calculated deflection of timber structure

Building type Dead Live Service wind DL + LL

Domestic,
commercial

Joists, bearers 1/260 1/300 1/300


and beams
8 mm max 8 mm max
supporting a floor
or roof deck
subject to foot
traffic

Bearers or joists 1/200 1/200 1/150


cantilevered to
5 mm max 5 mm max
support balconies

Vertical studs in 1/300 1/300 1/200


wall frames
8 mm max 8 mm max 12 mm max

Wall plates 1/200 1/200 1/100


5 mm max 5 mm max 5 mm max

Rafters and 1/300 1/300 1/150


beams supporting
18 mm main span 12 mm main span 25 mm main span
a roof deck not
subject to foot 12 mm eave
traffic

Used as a guide only, QA procedures must be followed 36


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Building type Dead Live Service wind DL + LL

Ceiling joist and 1/300 1/300


hanging beams
18 mm max 12 mm max

Lintels over 1/300 1/200


windows
8 mm max 8 mm max

Industrial

Roof beams 1/200 1/200 1/150


25 mm max 25 mm max 40 mm max

Floor beams 1/200 1/200 1/300


18 mm max 18 mm max

7.3 Durability

7.3.1 Corrosion

7.3.2 Crack Control for Concrete Structures


Crack control in concrete structures needs to be addressed with respect to public concerns and
durability.

In most non-aggressive environments, a maximum crack width of 0.3mm is generally deemed to be


acceptable. In more aggressive environments, a maximum crack width of 0.1mm to 0.2mm may need to
be considered to ensure the durability of the member is maintained.

AS 3600 Concrete structures is generally applicable to building type structures. Compliance with the
reinforcement stress limits and/or the reinforcement spacing of Section 8.6 (beams) and 9.4 (slabs) will
provide the requisite control over cracking commensurate with this type of structures (i.e. maximum crack
widths of approximately 0.3mm).
For bridge structures, similar provisions are provided in AS 5100.5 Bridge design Part 5: Concrete
Section 8.6 (beams) and 9.4 (slabs). As bridge structures are designed with a 100 year design life, these
provisions should be used with care and due consideration is to be given to the macro and micro-
environment.
The commentary to AS 3735 Concrete structures for retaining liquids specifies the mean crack widths for
structures that are required to retain (or exclude) liquid. Mean crack width limits of 0.1mm and 0.15mm in
tension and flexure respectively are specified where the member is continuously submerged; 0.1mm in
both tension and flexure where the member is subject to intermittent wetting and drying.
Maximum crack widths may be assumed to be 1.5 times the mean crack width.
Methods of calculating the crack widths are available in BS 8007 Design of Concrete Structures for
Retaining Aqueous Liquids, Eurocode BS EN 1992 Eurocode 2 Design of Concrete Structures and ACI
Manual of Concrete Practice 224R (to be finalised). The method proposed in BS 8110 is recommended.

Used as a guide only, QA procedures must be followed 37


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Further guidance can also be found in the CEB-fib Structural Concrete Textbook on Behaviour, Design
and Performance published by the federation Internationale du beton (fib).

7.3.3 Fatigue
Fatigue is to be considered for structural members subject to significant repetitive loading. This is
especially the case for railway bridge structures.
Fatigue requirements for steel structures are contained in Section 11 of AS 4100 Steel structures and
Section 13 of AS 5100.6 Bridge design Part 6: Steel and composite construction.
For fatigue requirements for concrete structures, refer to Section 2.5 of AS 5100.5 Bridge design Part 5:
Concrete.

7.4 Vibrations
Structures with large spans or heights may be considered for the dynamic effect caused by vibrations.
Dynamic analysis may be necessary when:
Structural steel floor spans of more than 8m
Steel portal frame spans of more than 50m
Reinforced concrete flat plate spans of more than 12m
Other types of structure spans of more than 25m
Cantilevers and masts of a height more than 10m
Core Aspect Ratio of a building (height to width) is more than 8: 1
Composite steel and post-tensioned concrete floors are extremely sensitive to vibration, particularly with
mechanical equipment loads. Structural engineers should advise the mechanical consultant to provide
vibration isolation accordingly.
Dynamic response is also influenced by structural form. The trend for lighter, stiffer floors, which follow
from consideration of static load, will bring a trend to higher natural frequencies. This should have the
effect of making the floor less likely to respond to loads moving over it, but perhaps more likely to
respond to disturbance from installed machinery.

Footfalls (pedestrian bridges, walkways, long span structures) with people loading are generally in the
range of 1.5 to 2 Hz. Therefore avoid a natural frequency of 1 to 3 Hz. AS 5100.2 Bridge Design Part 2:
Design Loads Section 12.4 says that for pedestrian bridges with resonant frequencies for vertical
vibration inside the range of 1.5Hz to 3.5Hz, the vibration of the superstructure shall be investigated as a
serviceability limit state.

7.4.1 Response of structure to vibration


(Notes for Car Parks from The Institution of Structural Engineers, June 2002, Design recommendations
for multi-storey and underground car parks, THIRD EDITION)
Empty car-park structures lack the damping normally provided in buildings by partitions and finishings.
However, the dynamics of most car park structures are generally found to be satisfactory when the
design complies with normal section sizes that give natural frequencies above 5Hz. Spans of 16.5m can
have natural frequencies in the range 2–4Hz, which will require special consideration of user comfort.

Used as a guide only, QA procedures must be followed 38


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Unusual or long-span structures may be more sensitive to dynamic loading and their responses should
be checked. In service, dynamic response usually becomes less of a problem as parked vehicles
increase the mass.
Additional guidelines for vibrations are available from the following publications:
King, Andrew B, BRANZ Study Report No 57 (1999) Serviceability Limit State Criteria for New
Zealand Buildings

7.5 References and Further Reading


AS 1170.0 Structural design actions Part 0: General principles
AS 3600 Concrete structures
AS 3735 Concrete structures for retaining liquids
AS 4100 Steel structures

AS 5100.1 Bridge design Part 1: Scope and general principles


AS 5100.5 Bridge design Part 5: Concrete
AS 5100.6 Bridge design Part 6: Steel and composite construction

BS 8007 Design of Concrete Structures for Retaining Aqueous Liquids


BS EN 1992 Eurocode 2 Design of Concrete Structures

ACI Manual of Concrete Practice


Federation Internationale du beton (1999) CEB-fib Structural Concrete Textbook on Behaviour, Design
and Performance.
King, Andrew (1999) BRANZ Study Report No 57 (1999) Serviceability Limit State Criteria for New
Zealand Buildings.

Used as a guide only, QA procedures must be followed 39


42/01012/07/44062 GHD Structural Design Manual
Revision 2
8. Design Loadings

8.1 Introduction
This section specifies general procedures and criteria for structural design. It covers design actions and
combinations of actions. This section is applicable to the structural design of whole buildings or
structures and their elements.
The design of a building or structure shall allow for, as a minimum, the following loads/actions:
Permanent / Dead Loads
Imposed / Live Loads
Wind Loads
Snow and ice Loads
Earthquake Loads
Construction Loads
Liquid pressure
Ground water
Earth pressure
Temperature
Refer to design criteria for basic loading and check for any special requirements by the Client. Provide a
loading summary sheet at the beginning of design calculations (Section 3.7).
A building shall be designed for the loadings required by the Building Code of Australia (BCA).

8.2 Design Standards


The following Standards and Guidelines apply for designs within Australia:

Table 15 Structural Design Standards and Guidelines

BCA Building Code of Australia

AS/NZS 1170 part 0 General Principals

AS/NZS 1170 part 1 Permanent, imposed and other actions

AS/NZS 1170 part 2 Wind actions

AS/NZS 1170 part 3 Snow and ice actions

AS/NZS 1170 part 4 Earthquake actions

Design Guidelines for Australian Public Cyclone Shelters

Queensland Government Department of Public Works

Used as a guide only, QA procedures must be followed 40


42/01012/07/44062 GHD Structural Design Manual
Revision 2
8.3 General Principals
Refer to AS/NZS 1170.0 for general principals relating to design loads - Parts 1 to 4.
Design shall be carried out by adopting the importance level and the design life for the building or
structure and the associated annual probabilities of exceedance for wind, snow and earthquake as
follows:
For Australia, as given in the Building Code of Australia or Appendix F of AS/NZS 1170 part 0;
For New Zealand, as given in Section 3 of AS/NZS 1170 part 0.
Loading combinations shall be as per section 4 of AS/NZS 1170.0.

8.4 Permanent / Dead Loads


Refer to AS/NZS 1170.1 in general and relevant standards for densities of specific materials.

In the absence of specific details, use the following:


Floor finishes (screed 50 mm) 1.2 kPa
Ceiling boards 0.4 kPa
False ceiling 0.25 kPa
Normal services 0.25 kPa
HVAC services 0.4 kPa
Demountable lightweight partitions 0.5 kPa
Use Table 16 for typical dead load data unless otherwise specified:

Table 16 Typical Dead Loads for Structures

Type Composition Dead load (kPa)

Floors Domestics timber floor (20 mm boards, 0.50


framing and 13 mm thick plasterboard
ceiling

Normal commercial floor (carpet + 50 mm 2.5


screed + 13 mm thick plasterboard ceiling
and 0.4 kPa for HVAC services) excluding
structure selfweight

Toilet floor (tile + 50 mm thick screed + 13 2.65


mm thick plasterboard ceiling and 0.1 kPa
for services) excluding structure
selfweight)

Walls Autoclaved Aerated Concrete, 100 mm 0.62~0.84


thick (CSR Hebel)

Glass blocks, 10/100 mm thick 1.0~1.5

Lightweight partition wall (steel studs & 2 0.20


layers of 13 mm plasterboard)

Used as a guide only, QA procedures must be followed 41


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Type Composition Dead load (kPa)

190 mm thick core-filled concrete 4.8


blockwork

300 mm thick core-filled concrete 6.2


blockwork

200 mm thick concrete walls plastered 5.0


both sides (12 mm thick)

Curtain wall (glazing + spandrel) 1.0

Demountable partitions 0.5~1.0 on plan

Roofs Domestic tiled roof (terracotta tiles, 1.0


framing & 13 mm plasticboard ceiling)

Metal deck roof (Trimdek sheeting, steel 0.15~0.25


purlins and 0.1 kPa services)

Metal deck roof with ceiling (Trimdek 0.25~0.35


sheeting, steel purlins 13 mm plasterboard
ceiling and 0.1 kPa services)

Metal deck roof with 1-hour suspended fire 0.40~0.50


rated ceiling (Trimdek sheeting, steel
purlins, 1 layers 16mm Fyrchek and 0.1
kPa services)

Sprinklers (low bound values are 0.07~0.10


assuming no services)

Bitumen roofing (3 layers) 0.3

Asphalt (19 mm, 25 mm) 0.41, 0.58

Tiles (clay layed to 100 mm gauge) 0.62~0.70

Concrete tiles interlocking 0.48~0.55

8.4.1 Earth and Hydrostatic Pressures


Earth and hydrostatic pressures shall be considered if these are applicable.

8.5 Imposed / Live Loads


Refer to AS/NZS 1170.1 section 3.0 for imposed / live loads.
It is recommended estimation of loads is conservative at concept and preliminary design stages, this
allows for change of use and flexibility of building.
Live load reduction should be considered in accordance with code requirements, although it should only
be considered to produce the most adverse effect during Concept and Preliminary Design stages. If not
specifically instructed otherwise, live load reduction should be applied as per code.

Used as a guide only, QA procedures must be followed 42


42/01012/07/44062 GHD Structural Design Manual
Revision 2
8.5.1 Imposed Loads on Barriers in buildings
Refer to AS/NZS 1170.1 Section 3.6 for Imposed Loads on Barriers.
The horizontal imposed action on barriers required to withstand the accidental impact from vehicles
during parking shall be taken as follows:
For light traffic areas (applied at 0.5 m above floor level):
– Barriers: 30 kN
– Barriers at the end of straight ramps (>20m long) for downward travel: 240 kN
For medium traffic areas (applied at 1.0 m above floor level): 40 kN

8.5.2 Construction Loads


Where the structure or element will be subjected to construction loads, the structure or element shall be
designed for these loads. The designer shall state on the drawings the construction loads assumed in the
design of the structure or element.
Examples of construction loads are as follows:
Steel beam supporting wet concrete during placement. Note this is applicable to composite design;
Material staking loads i.e. pallets of concrete block or bricks;

8.6 Wind Actions


Refer to AS/NZS 1170.2 for Wind Actions. This Standard sets out procedures for determining wind
speeds and resulting wind actions to be used in the structural design of structures subjected to wind
actions.

The Standard covers structures within the following criteria:


Buildings less than 200 m high.
Structures with roof span less than 100 m.
Structures other than offshore structures, bridges and transmission towers.
Refer AS 5100.2 for wind Actions on Bridges.
Refer AS 3995 for design wind speeds and wind loads for steel lattice towers and masts.
Generally wind loads on buildings and structures shall be determined in accordance with the
requirements of AS/NZS 1170.2.
For regions C & D designers shall refer to relevant state authority for further guidance on designing
structures in cyclonic regions. The Queensland Government has produced design guidelines for
Australian public cyclone shelters which incorporate recommendations of the Queensland Tropical
Cyclone Coordination Committee (QTCCC). The document provides guidance on shelter location,
structure loads, human factors and shelter management.

8.6.1 Internal Pressure Co-efficient


Internal pressure coefficients shall be determined in accordance with AS/NZS 1170.2 section 5.3.

Used as a guide only, QA procedures must be followed 43


42/01012/07/44062 GHD Structural Design Manual
Revision 2
In regions C and D, internal pressure resulting from the dominant opening shall be applied, unless the
building envelope (windows, doors and cladding) can be shown to be capable of resisting impact loading
equivalent to a 4 kg piece of timber of 100 mm × 50 mm cross section, projected at 15 m/s at any angle.

8.7 Earthquake Load


Refer to AS/NZS 1170.4 for Earthquake loadings.
The purpose of designing structures for earthquake loads is to:
Minimise the risk of loss of life from structure collapse or damage in the event of an earthquake;
Improve the expected performance of structures;
Retrofit existing structures to improve their performance to earthquake; and
Improve the capacity of structures that are essential to post-earthquake recovery to function during
and after an earthquake, and to minimise the risk of damage to hazardous facilities.
The standard AS 1170.4 and AS 5100.2 contain provisions for earthquake design and based on the
design category & structure regularity, earthquake loads shall be determined by either:
Static analysis; or
Dynamic analysis.

8.7.1 Design to Australian Codes


All structures and building shall be designed to resist the earthquake design loads specified in AS
1170.4.
Design load combinations for Limit State and Serviceability design shall comply with AS/NZS 1170.0.

The design earthquake loads shall be computed from AS 1170.4 for the appropriate structure Importance
Level.
Building and structures shall be classified as Importance Level 1,2,3 or 4. Annual Probability of
Exceedance, based on Importance Level, shall comply with Table B1.2b of the Building Code of Australia
(BCA).
Annual Probability of Exceedance for Earthquakes are tabulated below incorporating the Annual
Probability of Exceedance design events proposed by the BCA (2008).

Table 17 Annual Probability of Exceedance - Earthquake

Importance Level Annual Probability of Exceedance BCA 2008 (May 2008)

1 1:250

2 1:500

3 1:1000

4 1:1500

The annual Probability of Exceedance relates to the annual probability that the design event will be
exceeded in 50 years in 1:500 (10% in 50 years), 1:1000 (5% in 50 years) and 1:1500 (3.3% in 50
years).
Used as a guide only, QA procedures must be followed 44
42/01012/07/44062 GHD Structural Design Manual
Revision 2
Earthquake Loads computed from AS 1170.4 are Limit State Loads.

8.7.2 Earthquake Design Categories


Buildings and Structures shall be categorised into one of three Earthquake Design Categories listed in
AS 1170.4 in EDC1, EDC11 or EDC111.
The Earthquake Design Categories require earthquake assessments as follows:
EDC1 Simple Static check to AS 1170.4
EDC11 Static Analysis to AS 1170.4
EDC111 Dynamic Analysis to AS 1170.4
The minimum Earthquake Design Category for all structures and buildings shall be EDC11 regardless of
whether an EDC1 can be justified in accordance with AS 1170.4, hence as a minimum a Static Analysis
in accordance with Section 6 of AS 1170.4 shall be undertaken for all structures. Torsion effects, as
applicable, shall be applied to all structures.

Detailing of structures and foundations shall be to AS 3600 and AS 1170.4.


All foundations shall be tied together as per Section 5.2.2 AS 1170.4 regardless of the structure
importance level.

8.7.3 Earthquake Design Factors


Earthquake Design Factors shall be provided by the Geotechnical Investigation Report and AS 1170.4.

Hazard Factor ( )

Hazard Factor Maps are produced by Geoscience Australia (GA) are reproduced in AS 1170.4. The
Hazard Factor ( ) is equivalent to an acceleration co-efficient with an annual probability of exceedance of
1:500 (i.e. a 10% probability of Exceedance in 50 years).

The site hazard factor may be obtained from AS 1170.4 Fig 3.2 (B) or as detailed in the Geotechnical
Investigation Report.
Probability Factor (Kp)

The Probability Factor (Kp) shall be taken from AS 1170.4 Table 3/1 appropriate to the structures
Importance Level and associated annual probability of Exceedance.
Site Sub-Soil Class

The site sub-soil class shall be determined from the site geology as defined in the Geotechnical
Investigation Report.

8.7.4 Design to New Zealand Codes


All structures and building shall be designed to resist the earthquake design loads specified in NZS
1170.5.

Design load combinations for Limit State ad Serviceability design shall comply with AS/NZS 1170.0.

Used as a guide only, QA procedures must be followed 45


42/01012/07/44062 GHD Structural Design Manual
Revision 2
The standard NZS 1170.5 and NZS 3106 contain provisions for earthquake design and based on design
category and structural regularity, earthquake loads shall be determined by either:
Equivalent static method;
Model response spectrum method; or
Numerical integration time history method.
The method of analysis depends on:
The height of the structure;
The structural configuration and regularity; and
The largest translational period.
The earthquake loads on the structure depend on:
Site sub-soil classification;
Geographic location of the site and the shortest distance to the nearest major fault;
Fundamental period of translation;
Structural ductility; and
Importance level of the structure and its design working life.
Other considerations in the analysis and design for earthquake loads:
P-Delta effects;
Capacity design; and
Design lateral deflection of the structure and the design inter-story deflection.

8.8 Loading Calculations


Table 18 provides useful methods and software for calculating loads:

Table 18 Types of loading calculation

Calculation type Calculation method Software

Gravity load Tributary area method Load chase down module of


TEDDS
Validated spreadsheet

Wind load Based on AS 1170.2 Wind load calculation module


of TEDDS
Validated spreadsheet

Earthquake load Based on AS 1170.4 Validated spreadsheet


SpaceGass
ETABS

Used as a guide only, QA procedures must be followed 46


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Calculation type Calculation method Software

Geotechnical load Based on interaction between Wallap


soil and retaining structure
Plaxis
Validated spreadsheet

8.9 References and Further Reading


Holmes, J. & King, A. 2005, A Guide to AS/NZS1170.2:2002 – Wind Actions, Warren Publishing,
Mentone, Australia.

Used as a guide only, QA procedures must be followed 47


42/01012/07/44062 GHD Structural Design Manual
Revision 2
9. Standard Technical Specifications

9.1 Introduction
TBA

9.2 Technical Specifications


Standard technical specifications for different trades with various project scales are available in GHD.
The JM needs to:
Decide if the specification will be on drawings or separate documentation
Choose appropriate specifications
Make necessary amendments
Incorporate specific technical information
The list of specifications is shown on Table 19 and can be found in server (file path: TBA).

Table 19 List of standard specifications

Document reference Document title Revision number

Concrete

Concrete (for small projects)

Steelwork

Steelwork (for small projects)

Masonry

Masonry (for small projects)

Timber

Timber (for small projects)

Aluminium

9.3 Performance Specifications


Standard technical brief on geotechnical investigation for large-scale projects
Standard technical brief on geotechnical investigation for small-scale projects
Standard structural performance specification for large-scale projects
Standard structural performance specification for small-scale projects

Used as a guide only, QA procedures must be followed 48


42/01012/07/44062 GHD Structural Design Manual
Revision 2
10. Foundations

10.1 Introduction
The investigation and design of foundations is a broad subject. The guidance provided in this section is
intended to provide a basic understanding of the broad principles/design methods for relatively straight-
forward structures. It is not intended to replace Standards. However, care should also be taken in rigidly
applying Standards based methods, or presumed design values for soils/rocks.
Soil/structure interaction is complex and simplified methods may provide poor representation of actual
behaviour. GHD geotechnical engineers will be able to advise on appropriate design methods.

10.2 General
A foundation transmits structure loads to the underlying ground. Foundations must provide a factor of
safety against shear failure of the ground and be designed to keep within tolerable settlements/angular
distortion to avoid unacceptable damage to the structure (i.e. remain serviceable).
Foundations are either classified as shallow or deep foundations, and common types are:

Isolated pad footings – a small slab that supports an individual column


Strip footings – a long thin slab e.g. supporting a load-bearing wall
Combined strip/pad footing – supporting individual columns and load-bearing walls
Raft – relatively large slab supporting the whole structure
Stiffened raft – a raft stiffened with beams
Piers – shallow isolated bored or driven
Piles – driven, bored, jacked, screw types that transfer load to depths in the ground
Piled raft – a proportion of load is taken up by the slab and piles, with the piles used for settlement
reduction. Designed very differently to individually piled slabs, where the piles transfer all load.
Caissons – box or cylinders sunk or jacked into the ground during excavation of the ground internally.
Formed by incrementally casting insitu or by precast segmental liners
Shallow foundations are usually more economical. However, piles are commonly adopted if the soil near
the surface is incapable of supporting the structure without excessive settlement or shearing, or if highly
reactive soil conditions exist.

Where a certain foundation type is not thought feasible (e.g. raft), it may be possible to improve or treat
the ground more cost effectively than adopting other more expensive foundation types (e.g. piles).

10.3 Ground Investigation - Geotechnical Investigation Report


The designer should not proceed with detailed design without adequate interpreted ground investigation
data and advice from an appropriately experienced geotechnical engineer.

Used as a guide only, QA procedures must be followed 49


42/01012/07/44062 GHD Structural Design Manual
Revision 2
The geotechnical investigation and interpretation is critical to the understanding of the site, the
identification of potential risks and uncertainties, and as input for the design of any structure. Substantial
construction cost overruns, design implications and lost opportunities can result from poor geotechnical
investigation and interpretation. Where necessary, the designer should take advice on the contents and
recommendations given in reports.
Generally, there are two types of geotechnical reports produced:
A factual report, containing a description of geology, work methods, fieldwork findings, and laboratory
test results, and
An interpretative report containing geological ground model, ground properties, and
recommendations for design and construction.
In Australia, the two reports are quite commonly combined. Apart from providing information for design,
the reports will provide a basis for “foreseeable” contractual conditions and identification of potential
risks.
Preferably, all investigation reports and foundation designs should be at least reviewed by a GHD
geotechnical engineer.

10.4 General Foundation Design Approach


The design of a foundation system is generally as follows:
Calculation of the foundation loads and bearing pressures from the superstructure
Determination performance requirements (superstructures’ tolerance to total settlement, differential
settlement (angular distortion), heave, and lateral movement),
Selection of potential foundation systems
Sizing of the foundation based on an estimated allowable bearing capacity to maintain an adequate
factor of safety on shear failure of the ground for each footing type and dimensions,
Estimation of settlements and verification of tolerable limits,
Constructability,
Consideration of the potential risks and uncertainties, and
Cost.

10.5 Foundation Bearing Pressure and Net Bearing Pressure

10.5.1 General
Foundation bearing pressure (qb) is the pressure exerted at the base of the foundation on the ground by
the superstructure. It is not related to bearing capacity, which is the resistance provided by the ground to
counteract the foundation bearing pressure exerted. Net foundation pressure (qn) is the net increase in
pressure at the base of the foundation, after taking into account the weight of ground permanently
removed below existing ground level. It is expressed as:
qn=qb- D

Used as a guide only, QA procedures must be followed 50


42/01012/07/44062 GHD Structural Design Manual
Revision 2
where,
is the unit weight of the soil removed, and
D is the depth to the underside of foundation below existing ground level.

10.5.2 Calculation of Bearing Pressure for Vertical Loading Only


Where vertical loads only are acting on the foundation, the bearing pressure is the vertical load divided
by the footing area (L B).
qb=Rv/(L B)

10.5.3 Calculation of Bearing Pressure for Combined Vertical and Inclined/ Horizontal Loading
Inclined and/or horizontal loading applied to a foundation in combination with vertical loading causes the
force resultant (i.e. summation of force and force direction) on the foundation base to act inclined rather
than vertical. This has an overturning effect on the foundation. The calculation of bearing pressure is
modified to take this into account by reducing the foundation width, B, (in the direction of loading) to the
“effective width”, B’:

B’=B-2e
where,
e is the eccentricity of the resultant from the centre of the base

qb=Rv/(L*B’)
This is commonly referred to as the Meyerhof bearing pressure and is represented as a rectangular
stress block across the effective width rather than triangular stress distribution over the whole footing
width.

10.6 Bearing Capacity Estimates

10.6.1 Bearing Capacity – Ultimate and Allowable


“Ultimate bearing capacity”, qu, is the lowest pressure applied at foundation level to cause shear failure of
the ground. “Allowable bearing capacity”, qa, is the maximum pressure that may be applied to the ground
to provide an adequate factor of safety against shear failure of the ground and reduce settlements. There
is no one value of ultimate or allowable bearing capacity for each soil or rock type and strength. Bearing
capacity is a function of the following variables:
Ground conditions, including material type, density and strength
Groundwater location
Depth of embedment of the foundation
Dimensions of the foundation
Orientation and inclination of loading
Topography of the site (sloping)
Static or dynamic loading

Used as a guide only, QA procedures must be followed 51


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Earthquake and potential for ground strength loss
Generally, for non-cohesive (sand, gravel) soils the allowable bearing capacity would quite often be
dictated by allowable settlement, unless the ultimate bearing capacity was affected by sloping ground.
For cohesive soils, both the ultimate bearing capacity (i.e. shear failure) and settlements are important.

10.6.2 Concept Design with Presumed Bearing Capacity


Soils

At concept design only, preliminary sizing of pad footings with vertical loads could be based on the
presumptive bearing capacity values given in Table 20. Care should be taken in adopting presumed
bearing capacity values from international standards as the soil classification system may be different to
Australia (e.g. very soft clay in the UK has a shear strength less than 20 kPa, whereas it is less than 12
kPa in Australia).
It should be noted that the presumptive bearing capacity values only consider a factor of safety against
shear failure. They do not consider settlement. It would be advisable to seek confirmation from a GHD
geotechnical engineer. In general, foundations on soft cohesive soils and those with larger dimensions
(e.g. rafts), should have further consideration of settlements at concept stage.

Table 20 Presumptive Bearing Capacities to AS 1726 Soil Descriptions

Bearing soil stratum (assuming consistent soil conditions for Presumed bearing
minimum 3*footing width below footing base) capacity (kPa)

Strip footing Circular or


(B/L=0) Square
footing
(B/L=1)

Very soft clay (min 0.75 m below ground level), cu<12 kPa Seek Seek
advice, <35 advice, <40

Soft clay (min. 0.75 m below ground level), cu=12-25 kPa 35 – 55 40 – 65

Firm clay (min. 0.75 m below ground level), cu=25 – 50 kPa 55 – 100 65 - 115

Stiff clay (min. 0.75 m below ground level), cu=50 – 100 kPa 100 – 185 115 – 220

Very stiff clay (min. 0.75 m below ground level), cu=100-200 kPa 185 – 350 220 - 425

Hard clay (min. 0.75 m below ground level), cu>200 kPa >350 >425

Used as a guide only, QA procedures must be followed 52


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Bearing soil stratum (assuming consistent soil conditions for Presumed bearing
minimum 3*footing width below footing base) capacity (kPa)

Basis of presumed bearing capacity values for cohesive soils:


Assumed flat ground
Factor of Safety = 3.0
Applicable for vertical static loading only
Bearing capacity factor, Nc, of 5.14 for strip footing (B/L=0) and 6.2 for circular or square footing (B/L=1)
Footing taken as minimum 0.75m below ground level, with unit weight of soil above footing level=18
kN/m 3. Values should be reduced by 13 kPa if footing is at ground level
Long-term consolidation and creep settlement should be verified regardless of bearing capacity

Silty/clayey sands (SM/SC)

**Very loose clean sands and gravels (N<4) = 28-30 Seek Advice

**Loose sands and gravels (N = 4-10) (min. 0.75 m below ground level) 100
= 30

Medium-dense sands and gravels (N = 10-30) (min. 0.75 m below ground 180
level) = 34

Dense sands and gravels (N = 30-50) (min. 0.75 m below ground level) 300
= 38

Very dense sands and gravels (N > 50) (min. 0.75 m below ground level) >400
= 41

Basis of presumed bearing capacity values for non-cohesive soils:


Assumed flat ground
Groundwater table at footing level
Factor of Safety = 3.0
Applicable for vertical static loading only
1m minimum foundation width
3
Unit weight of soil above footing level=17 kN/m
** Vibration Sensitive

It is strongly advised that a GHD geotechnical engineer be consulted for appropriate values for mixed
soils such as silty or clayey sands. See below for discussion.
It is further noted that the capacity of foundations in coarse-grained soils is heavily dependent on
groundwater levels e.g. for a surface footing in sand, the dry bearing capacity is about double the
saturated bearing capacity.

Used as a guide only, QA procedures must be followed 53


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Mixed Soils (Silty/Clayey Sands) – Comment on Australian Soil Classification System (AS 1726)

It is strongly advised that a GHD geotechnical engineer be consulted for appropriate values for mixed
soils such as silty or clayey sands, or sandy silts. Many structural designers are uncertain on design
properties for mixed soils. Few mainstream geotechnical textbooks provide detailed guidance on
properties, but typically present a wide variety of soil properties and behaviour of sand and clay.
Wherever possible, grading analysis should be carried out on such mixed soils to determine the
proportion of silt/clay fines rather than relying on visual descriptions on borehole logs.
In Australia, the soils classification system (AS 1726) for silty/clayey sands (SM/SC) is very broad, and
the sand may have anywhere between 12 and 49% silt/clay fines. This is because the Australian
Standard for soil classification (AS 1726) delineates the dominant soil type (shown on borehole logs in
CAPITALS) on the basis of proportion of soil material passing the 0.075 mm sieve size. This sieve
represents the boundary between sand and silt sized soil particles. The Standard states this should be
done on a 50:50 basis. For instance, a soil with 49% clay and silt, and 51% sand, would be a clayey/silty
SAND, but a soil with 51% clay and silt, and 49% sand would be a sandy CLAY/SILT. For that example
there would be little difference in soil properties and behaviour for the two soils, which would both behave
as a fine-grained soil like clay/silt. However, the classification of a silty/clayey SAND also extends to a
soil with only 12% clay/silt and 88% sand, which would behave more as a poorly draining sand rather
than silt/clay. The challenge to designers is when the proportion of clay/silt in the sand starts to dominate
behaviour to the extent that the material basically has clay/silt properties. Adopting the DOMINANT soil
type only as the design basis may prove highly inaccurate.
The New Zealand and UK Standards both delineate dominant soil type on the basis of engineering
behaviour rather than a 50:50 proportion basis, with the difference being set at 35% clay/silt:65%
sand/gravel. Even in the 15 to 35% fines range, soil properties can be substantially different from that of
a clean sand or gravel.
Rock

The bearing capacity for rocks is significantly affected by the orientation, frequency and condition of
joints in the rock, which affect rock mass strength (ultimate bearing capacity) and stiffness (foundation
settlements). A wide variety of published presumed allowable bearing capacity values are available for
rocks, but considerable care should be taken in selecting an appropriate value, even for concept design.
A GHD geotechnical engineer should be asked for confirmation of appropriate detailed design values,
commensurate with settlement performance requirements. Some guidance is provided in Table 21, and
Figure 7.

Table 21 BS 8002 Rock Groups for Use with

Weathered and fractured rock masses:

Group 1 Pure limestones and dolomites Refer Figure 6a


Carbonate sandstones of low porosity

Used as a guide only, QA procedures must be followed 54


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Weathered and fractured rock masses:

Group 2 Igneous Refer Figure 6b


Oolitic and marly limestones
Well cemented sandstones
Indurated carbonate mudstones
Metamorphic rocks, including slates and schists (flat
cleavage/foliation)

Group 3 Very marly limestones Refer Figure 6c


Poorly cemented sandstones
Cemented mudstones and shales
Slates and schists (steep cleavage/foliation)

Group 4 Uncemented mudstones and shales Refer Figure 6d

Used as a guide only, QA procedures must be followed 55


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 6 Rock allowable bearing capacity for square pad foundations with settlement not
exceeding 0.5% of foundation width (taken from BS 8004)

Used as a guide only, QA procedures must be followed 56


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 7 Design Values for Foundations on Sandstone and Shale in Sydney Region (after
Pells, 1978)

10.6.3 Settlement Estimates


Estimates of foundation settlement should preferably be carried out by a suitably qualified geotechnical
engineer, taking into account the variation of pressure with depth. Consideration should be given to the
following:
Immediate settlement on application of load
Used as a guide only, QA procedures must be followed 57
42/01012/07/44062 GHD Structural Design Manual
Revision 2
Consolidation
Secondary consolidation or creep over the nominated structure design life
Differential settlements/ angular distortion
Settlement estimates in silts and clays should be carried out based on oedometer test results, and take
account of preconsolidation pressure (i.e. stress history) of the soil. Estimates of settlement in sands and
gravels should ideally be made using the Burland and Burbridge method, although a wide variety of
approaches are available in text books.
Various guidelines are available on tolerable settlement and angular distortion limits of particular
structures, but each structure should be considered uniquely. If necessary, prescribed displacements can
be introduced into structural models to check the impact of differential settlement on the structure.

10.7 Foundation design

10.7.1 Foundations for Residential Developments, AS 2870


This Standard (AS 2870) sets out the requirements for the classification of a site and the design and
construction of a footing system for a single dwelling house, townhouse or the like which is classified as
Class 1 and 10a under the Building Code of Australia.
The standard may also apply to other forms of construction including some light industrial, commercial
and institutional buildings if they are similar to houses in size, loading and superstructure flexibility. It is
also quite commonly used for classification of other sites in terms of reactivity of the ground. Although the
foundation recommendations in AS 2870 may not be applicable to other types of structures, the meaning
of the reactivity classifications is generally understood by designers.

10.7.2 Foundations in Reactive Soils


Considerable cost is incurred throughout Australia due to damage from reactive soils. These are soils
that shrink and swell with seasonal cycles of wetting and drying. Ground surface movements can amount
to variations over a dry/wet cycle of in excess of 100mm in extremely reactive soils, generally well
outside the tolerable range of movements that most structures are capable of accommodating unless
specifically designed. Therefore, it is critical that sound advice is sought from a geotechnical engineer if
reactive soil conditions are thought present.
Various solutions are available, as follows:
Design the structure to accommodate the predicted movements
Suspended slab with void formers on deep ground beams or piles to accommodate seasonal
movements without undue swelling pressure on the underside of slab
Excavate the zone, or part of the zone, of soil affected by seasonal moisture change and replace with
non or low-reactive soils
Treat the ground to reduce reactivity of the soil, e.g. lime stabilisation

Used as a guide only, QA procedures must be followed 58


42/01012/07/44062 GHD Structural Design Manual
Revision 2
10.7.3 Design of Pad Footings
In Australia, the design of pad footings can be carried out to AS 5100.3-2004 – Bridge Design, Part 3:
Foundations and soil-supporting structures. However, advice should be taken from a geotechnical
engineer, where possible, on the applicability of this Standard to the ground conditions, settlement and
movement tolerances, and intended purpose for the structure. This Standard may not be suitable in all
circumstances.
The Pad Design Module in TEDDS can be of assistance in the design of pad footings.
Pad Footings on Other than Rock

For foundation material up to 1000 kPa allowable bearing capacity, pads should be designed as flexural
members in accordance with AS 3600.
Pad Footings on Rock

For rock over 1000 kPa either refer below or design pads to transfer load in strut/tie or shear with a
greater depth. For widely jointed rock masses with an allowable bearing capacity greater than 1000 kPa,
pads may be designed assuming a strut-tie model with nominal reinforcement in the top of the pad, as
shown in Figure 8. This relies on containment of the lateral thrust component by the rock foundation
sides. Consideration of the strain compatibility between the concrete footing and the rock should be
made.

Figure 8 Pad on rocks

Used as a guide only, QA procedures must be followed 59


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Lateral Loads on Pad Footings

Aside from the reduction in bearing capacity of pad footings caused by lateral and/or inclined loading,
footings should be designed against sliding based on the frictional resistance on the base and/or passive
resistance of the ground. For the latter, consideration of the following needs to be taken into account:
Potential shrinkage of the soil away from the footing
Necessary movement required to mobilise passive resistance, and the need for a suitable limiting
factor of safety
Impact on footing design if excavation was to occur adjacent to the footing, i.e. loss of passive
resistance
BS 8004 provides further guidance on parameters and calculation of resistance to sliding.

10.7.4 Design of Raft Foundations


A raft foundation would be preferable when:
The ground has relatively low bearing capacity; and
Variable ground conditions and potential for excessive differential settlement between isolated pad
footings
The structural design of rafts can be complex. A raft should not be designed as an inverted floor slab on
unyielding supports. To design the slab on the assumption that its whole area is loaded to the allowable
bearing capacity can be highly inaccurate. Deflections induced bending moments and shear should be
checked. Programs such as SLABS and Strand 7 provide some indication of the distribution of
foundation bearing pressure on the underside of a foundation and bending moments throughout the slab,
assuming elastic behaviour of the ground and foundation. However, the springs approach may not be
representative of the ground response. More rigorous finite element methods can be used where non-
linear behaviour of the ground is expected.

10.7.5 Design of Piled Foundations


The design of piles should be carried out to AS 2159 and AS 5100.3, as appropriate. Further guidance is
also provided in BS 8004. It is strongly recommended that a GHD geotechnical engineer be consulted to
assist with the estimation of geotechnical pile capacity.
Piles are relatively long and slender members used to transmit foundation loads deeper into the ground.
They are also used to resist high uplift forces or for the resistance of horizontal loads. Piles are a
convenient method of foundation construction for works over water.
All piles should be designed for an out of position of 75mm as per AS 2159 to allow for drilling/driving
tolerances. This may require piles/pile groups to be tied together with footing beams to rectify the
eccentricities. This value of 75mm may be increased for long piles as the verticality of the pile may not
fall within this value. Similarly design columns for 75mm eccentricity when supported on piles or piers.

Used as a guide only, QA procedures must be followed 60


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Buckling Check

Piles should be checked for buckling, particularly long slender piles through soft soils. Where part of the
pile extends above finished ground level, that length should be designed as a column in accordance with
the relevant AS. The effective length is dependent on the lateral loading, and degree of fixity provided by
the ground and by the structure that the pile supports. The depth to the point of contra flexure varies with
soil type and strength/density.
Pile capacity

Pile capacity shall be calculated in accordance with AS 2159, unless using specialist design calculation
methods (e.g. design of rock socketed piles based on settlement criteria). Shaft friction and end bearing
contribute the geotechnical capacity. A GHD geotechnical engineer should be consulted for appropriate
design values.
Lateral Forces
Piles subject to lateral forces should be checked for moment, shear and deflection, taking into account
the stiffness of the pile and response of the ground. The following methods are available:
Hand calculations (e.g. Broms method)
Beam and spring models, using software such as Space Gass or Microstran
P-Y methods, using software such as LPILE
Finite element methods (e.g. Plaxis, Phase2)
It is strongly recommended that advice be sought from a GHD geotechnical engineer on an appropriate
soil/structure design method for each structure. Representation by linear elastic beam and spring models
can be highly inaccurate in some circumstances. The P-Y method is a slight improvement in that non-
linearity can be introduced into the springs, but the springs still work independently of each other. Finite
element methods are generally preferred but may not be warranted in all cases.
Deflection calculations should give consideration to the cracked section properties of the pile.
Pile groups subject to lateral forces should be referred to a geotechnical engineer, as the lateral capacity
of individual piles can be severely reduced if piles are less than about 6B apart (the passive support
zones start to overlap).

Negative Skin Friction


Down drag will be caused on piles installed through settling ground. A GHD geotechnical engineer
should make estimates of this negative skin friction. It should be used to check the structural strength of
the pile.

10.7.6 Pile Groups


For friction piles the minimum centre-to-centre spacing should be at least greater than the perimeter of
the pile or 2.5 to 3 times the pile diameter. End bearing should not be closer than one pile width between
the shafts of the piles. Behaviour of the group and individual piles should be considered. The stress bulb
of an individual pile is relatively shallow, but the individual bulbs merge in a group to form a large bulb
related to the width of the group. A GHD geotechnical engineer should check the settlement and bearing
capacity of the overall group.

Used as a guide only, QA procedures must be followed 61


42/01012/07/44062 GHD Structural Design Manual
Revision 2
10.7.7 Piled Rafts
The design of piled rafts, where the raft and the piles transfer a proportion of load, should be carried out
in conjunction with a GHD geotechnical engineer. Piled rafts are used to reduce settlements to tolerable
levels. They generally suit stiff clays and dense sand, so that a reasonable proportion of load can be
taken by the raft on ground.

10.7.8 Caissons
A caisson is a structure that is sunk through the ground and becomes part of the permanent works.
Some guidance is given in BS 8004. Geotechnical design of caissons needs to consider the following:
Method of sinking the caisson
Estimates of the skin friction to overcome when sinking the caisson
Ground loss and associated movements, and the impact on adjacent structures, services and other
infrastructure
Base heave and bearing capacity
Control of groundwater

10.8 Other Issues


Careful consideration should be given to the load combination where the foundation is subject to
uplift, lateral force and change in water table, etc.
Advice should be sought from a GHD geotechnical engineer if dewatering is required
Bridging structures should be provided when services are running undercross the building
For aggressive soil, refer to special requirement from TN57 of Cement & Concrete Association (acid
soil) and AS 2159 (sulphates)

10.9 References and Further Reading


Australian Standards:

AS 1726 Geotechnical site investigations


AS 2159 Piling – Design and installation
AS 2870 Residential slabs and footings
AS 5100.3-2004 – Bridge Design, Part 3: Foundations and soil-supporting structures
Other Standards:

BS 8004 Code of Practice for Foundations


Apart from textbooks and published data, other useful sources of information are:
CIRIA
BRE
Hong Kong Geoguides
USACE

Used as a guide only, QA procedures must be followed 62


42/01012/07/44062 GHD Structural Design Manual
Revision 2
11. Retaining Walls

11.1 General
A retaining wall is a structure that holds back earth or rock. Retaining walls stabilise soil from downslope
movement or erosion and provide support for vertical or near-vertical grade changes. Retaining walls are
generally made of masonry, stone, concrete, steel or timber.
The most important consideration in design and installation of retaining walls is that the retained material
is attempting to move forward and downslope due to gravity and surcharge. This creates an earth
pressure behind the wall (depending on the angle of internal friction and the cohesive strength c of the
material, see Table 24). This pressure will push the wall forward or overturn it if not properly designed.
Additional horizontal hydraulic pressure on the wall, and beneath the base, shall also be considered if
any groundwater behind the wall is not dissipated by a drainage system. Proper drainage behind the wall
is critical to the performance of retaining walls. Drainage materials will reduce or eliminate the hydraulic
pressure and increase the stability of the fill material behind the wall.

11.2 Retaining Wall Types

11.2.1 General
Common types of retaining structures are:
Gravity wall
Cantilever wall
Sheet piling or shoring
Bored pile wall
King-post pile wall
Basement wall
Reinforced soil wall
Cut slope stabilised by ground anchors or soil nails

11.2.2 Gravity walls


Gravity walls are made from a large mass of stone, concrete or composite material. Gravity walls depend
on the size and weight of the wall mass to resist pressures from behind. Gravity walls will often have a
slight setback, or batter, to improve wall stability by leaning back into the retained soil. For short,
landscaping walls, gravity walls made from dry-stacked (mortarless) stone or segmental concrete units
(masonry units) are commonly used.
Taller retaining walls are increasingly built as composite gravity walls such as: geosynthetic or steel-
reinforced backfill soil with precast facing; gabions (stacked steel wire baskets filled with rocks), crib
walls (cells built up log cabin style from precast concrete or timber and filled with soil) or soil-nailed walls
(soil reinforced in place with steel and concrete rods).

Used as a guide only, QA procedures must be followed 63


42/01012/07/44062 GHD Structural Design Manual
Revision 2
11.2.3 Cantilevered walls
Cantilevered walls are the most common type of taller retaining wall. Cantilevered walls are made from a
relatively thin stem of steel-reinforced, cast-in-place concrete or mortared masonry (often in the shape of
an inverted T). These walls cantilever loads (like a beam) to a large, structural footing; converting
horizontal pressures from behind the wall to vertical pressures on the ground below. Cantilevered walls
are sometimes butressed on the front, or include a counterfort on the back, to improve their stability
against high loads. Buttresses are short wing walls at right angles to the main trend of the wall. These
walls require rigid concrete footings below seasonal frost depth. This type of wall uses much less
material than a traditional gravity wall.

11.2.4 Sheet piling


Sheet pile walls are often used in soft soils and tight spaces. Sheet pile walls are made out of steel, vinyl,
fiberglass or plastic sheet piles or wood planks driven into the ground. Structural design methods for this
type of wall exist but these methods are more complex than for a gravity wall. As a rule of thumb; 1/3
third above ground, 2/3 below ground. Higher sheet pile walls usually require a tie-back anchor placed in
the soil some distance behind the wall face that is tied to the wall face, usually by a cable or a rod.
Allowance for articulation and settlement of the ground surrounding the tie-rods should be made to
ensure ongoing structural integrity. Anchors must be placed sufficiently far back from the wall to ensure
the active zone behind the wall does not overlap with the passive zone from the anchor. Alternatively
raking piles could be placed at the front side of the anchor.

11.2.5 Bored Pile Wall


Bored pile walls are generally used to support excavations, and can be constructed either with a small
gap between piles (contiguous) or slightly overlapping (secant).

11.3 Presumed Geotechnical Design Parameters for Concept Design


In the absence of reliable test data, presumed design parameters for concept design are given in Table
22 and Table 23).

Table 22 Typical Presumed Unit weights of granular soil (after AS 4678 and BS 8002)

3 3
Material m: moist bulk weight (kN/m ) s: saturated bulk weight (kN/m )

Loose Dense Loose Dense

Gravel 16.0 18.0 20.0 21.0

Well-graded sand 19.0 21.0 21.5 23.0


and gravel

Coarse or 16.5 18.5 20.0 21.5


medium sand

Well-graded sand 18.0 21.0 20.5 22.5

Fine or silty sand 17.0 19.0 20.0 21.5

Rock fill 15.0 17.5 19.5 21.0

Used as a guide only, QA procedures must be followed 64


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Material m: moist bulk weight (kN/m3) s: saturated bulk weight (kN/m3)

Brick hardcore 13.0 17.5 16.5 19.0

Slag fill 12.0 15.0 18.0 20.0

Ash fill 6.5 10.0 13.0 15.0

Table 23 Typical Presumed Unit weights of cohesive soil

Material m: moist bulk weight (kN/m3) s: saturated bulk weight


(kN/m3)

Peat 12.0 12.0

Organic clay 15.0 15.0

Soft clay 17.0 17.0

Firm clay 18.0 18.0

Stiff clay 19.0 19.0

Hard clay 20.0 20.0

Stiff or hard glacial clay 21.0 21.0

In the absence of reliable data for concept design, effective shear strength parameters may be assumed
from AS 4678 (reproduced in Table 24) or BS 8002. However, the values quoted in AS 4678 should be
treated with some caution. A GHD geotechnical engineer should be asked for suitable design values.

Table 24 Presumed Effective Strength Parameters (after AS 4678)

Soil group Typical soil in group c’ (kPa) ’ (degrees)

Poor Soft and firm clay of medium to 0-5 17 - 25


high plasticity, silty clays, loose
variable clayey fill, loose sandy
silts

Average Stiff sandy clays, gravely clays, 0 - 10 26 - 32


compact clayey sands and sandy
silts, compacted clay fill

Good Gravelly sands, compacted sands, 0-5 32 - 37


controlled crushed sandstone and
gravel fills, dense well-graded
sands

Very good Weak weathered rock, controlled 0 - 25 36 - 43


fills of road base, gravel and
recycled concrete

Used as a guide only, QA procedures must be followed 65


42/01012/07/44062 GHD Structural Design Manual
Revision 2
11.4 Surcharge Loads

Designers should refer to Standards, but typical surcharge allowances are as given in Table 25. A guide
to their application is provided in Figure 9.

Table 25 Common surcharges for retaining walls

Load condition Unfactored surcharge (kPa)

Functional roads classes 1, 2, 3, 6 or 7 20

Other functional class roads 10

Other imposed ground actions Selfweight of ground finished + live load


recommended in AS 1180.1

Foundations
See
Figure 9

Used as a guide only, QA procedures must be followed 66


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 9 Surcharge calculations

Used as a guide only, QA procedures must be followed 67


42/01012/07/44062 GHD Structural Design Manual
Revision 2
11.5 Design

11.5.1 Design Considerations


Retaining wall design involves:
Selection of appropriate wall types
Groundwater levels and potential seasonal changes
Earth pressure caused by soil, surcharge and ground water
Selection of earth pressure coefficients
Calculation of earth and groundwater pressures on the wall
Check of sliding on the base, overturning and base bearing pressures
Strain compatibility (can resistance be mobilised within tolerable movements?)
Consideration of drainage requirements
Constructability
Structural design and detailing

11.5.2 Horizontal Earth Pressures


Three common earth pressures referred to are:
At-rest, Ko
Active, Ka
Passive, Kp
The actual earth pressure imposed on the structure is dependent on the amount of movement or rotation
mobilised in the ground.
At-Rest Earth Pressure
The ground prior to any disturbance (excavation/filling, etc) has locked in stress, known as the at-rest
pressure. It is related to the vertical stress using the coefficient of at-rest pressure, Ko, as follows:

Lateral pressure at-rest, po = Ko v

Where,
Ko = 1-sin normally consolidated soils only)
Ko=(1-sin OCRsin (overconsolidated soils)

v is the vertical pressure at a given elevation


is the internal friction angle of soil
OCR is the overconsolidation ratio of the soil
A GHD geotechnical engineer should provide the OCR and appropriate Ko values.

Used as a guide only, QA procedures must be followed 68


42/01012/07/44062 GHD Structural Design Manual
Revision 2
At-rest pressure is important for retaining wall design as it represents the starting stress state in the
ground. Whether the at-rest pressure is appropriate to base earth pressure calculations or for retaining
wall design depends on the type of wall, the ground conditions and the amount of movement or rotation
expected. It is incorrect to base all retaining wall designs on the active pressure. The at-rest pressure is
particularly important for the design of embedded pile walls; rigid basement walls; tied back or propped
walls; flexurally stiff walls like thick concrete gravity walls founded on rock (where horizontal sliding will
not tend to occur); or any type of wall/ground conditions where movement is restrained.
At-rest pressure may be due to the weight of soil at present time (e.g. normally consolidated soil) or a
function of some earlier much larger weight of soil (since eroded), ice or even compaction of backfill (e.g.
overconsolidated soil). In the latter case, the locked in horizontal stress is a function of the highest stress
previously felt by the ground not the present vertical stress. At-rest pressures for overconsolidated soils
can be significantly higher than the present vertical stress.
Active and Passive Earth Pressures

Retaining walls represent a classic example of soil-structure interaction where the degree of movement
of walls by translation or rotation under the pressure of the retained soil can modify considerably the
magnitude of the pressure acting behind the walls or the resistance of the soil to movement in front of the
walls. These opposing forces are known as the active pressure (pa) and passive pressure (pp)
respectively.
Active pressures, pa, is smaller than po and only small wall outward movement (of the order 0.1-1.0% of
retaining wall height, depending on ground conditions and wall types) is necessary to mobilise pa.
Passive pressures, pp, is larger than the at-rest pressure, po, and large wall inward movement (of the
order 3-10% of wall height) is necessary to mobilise full pp.

Active pressure, pa = Ka v’

Passive pressure pp = Kp v’

Where,
Ka is the active coefficient of earth pressure
Kp is the passive coefficient of earth pressure

v’ is the vertical effective stress at a given elevation


For a cantilever wall, Ka and Kp values can be worked out based on Rankine’s Theory or Coulomb’s
Theory. For sheet pile, bored pile or anchored basement walls geotechnical design software such as
Wallap or Plaxis would be preferable, as they apply earth pressure relative to the stiffness of the wall,
propping or anchoring systems and as a result of mobilised movements.

11.5.3 Software
Wallap - beam and spring model, or finite element
Plaxis - finite element modelling.
TEDDS - RC retaining wall design & gravity retaining wall design
Capacity tables shown in CMAA references for masonry cantilever retaining walls.

Used as a guide only, QA procedures must be followed 69


42/01012/07/44062 GHD Structural Design Manual
Revision 2
11.6 References and Further Reading
Standards Australia has now released a code (AS 4678) that deals with most of the relevant areas of
design for earth retaining structures including retaining walls. This code is a reasonable starting point for
the design of retaining walls, but designers should be aware that certain authorities (e.g. Queensland
Department of Main Roads) are requesting retaining walls be designed to BS 8002. BS 8002 provides a
more rigorous guide to retaining wall design.
Australian Standards:
AS 1726 Geotechnical site investigations
AS 2159 Piling – Design and installation
AS 4678 Earth-retaining structures
AS 5100.3-2004 – Bridge Design, Part 3: Foundations and soil-supporting structures
Other Standards:
BS 8002 Code of practice for earth retaining structures
BS 8004 Code of practice for foundations
BS 8006 Code of practice for strengthened/ reinforced soils and other fills
Apart from textbooks and published data, other useful sources of information are:
CIRIA
BRE
Hong Kong Geoguides
USACE
FHWA

Used as a guide only, QA procedures must be followed 70


42/01012/07/44062 GHD Structural Design Manual
Revision 2
12. Dynamic Foundations

12.1 Introduction
This section is a guide for the design of foundations subjected to dynamic loads from vibrating
machinery.

12.1.1 Principles of Design


The natural frequency of a body, f n is directly proportional to the square root of its stiffness divided by its
mass as given by the following equation:

k
fn
m
It follows that the greater the stiffness, the higher the natural frequency for the same mass. Common
ways of increasing natural frequency (without an undue increase in mass) include:
Increasing the depth, D, of the member – here stiffness increases by D3, but mass only increases by
a multiple of D
Providing additional bracing
Increasing the bearing contact area of the foundation
Stiffening the subsoil.
When designing foundations, the mode frequency should be ±50% of the speed of the machinery.

12.1.2 General Guidelines for Design


There are several general guidelines that provide a good starting point for design, which will generally
give an acceptable answer. However as the structure is also dependant on other factors such as the soil
condition an analysis of the structure needs to be undertaken.
The following, including Figure 10, are the guidelines for the design of foundations for vibrating
machines:

The common centre of gravity of the system (i.e. of the foundation and machine) is located as near
as possible to the same vertical line as the centroid of the foundation area in contact with the soil. In
any case the eccentricity in the distribution of masses should not exceed 5% of the length of the side
of the contact area.
The height difference of the centroid of the machine and foundation should be kept to a minimum but
with due consideration to operational and maintenance requirements. Where possible the combined
centre of gravity of the machine and foundation in elevation should be within the height of the
foundation. Generally foundation thickness will be a minimum of 600 mm.
To ensure reasonable stability, the width of the foundation (measured at right angles to the
crankshaft) should at least be equal to the distance from the centre of the shaft to the bottom of the
foundation.
The proportions of the foundation block should be such as to ensure stability against rocking.

Used as a guide only, QA procedures must be followed 71


42/01012/07/44062 GHD Structural Design Manual
Revision 2
The foundation should be isolated from adjacent foundations and structures.
Continuous homogeneous concrete construction should be ensured without pour breaks or other
zones of weaknesses.
As a guide the foundation mass should be not less than 3 times the mass for rotating machinery or 5
times the mass for reciprocating machinery however, some references suggest 2 times and 3 times
respectively.
Bearing pressure under static loads should be limited to approximately 50% of the allowable soil
bearing pressure.
Reciprocating machines shall be supported directly on rigid foundations.
For rotating heavy mass machinery (such as mills and crushers) a rigid footing should be provided
with ground contact extending +/- 30 degrees either side of the centre of mass for the machine’s
longitudinal axis.
The Frequency Ratio should be greater than 1.5. The Frequency Ratio is defined as frequency of the
disturbing force divided by the natural frequency of vibrations of foundation block and machinery
system.

12.2 Analysis of Dynamic Foundations


Analysis of dynamic foundations can be carried out using space frame programs such as Spacegass,
however results are limited as mass participation and amplitude of vibration are not given. It is therefore
suggested to use finite analysis packages such as Strand7.
Prior to any dynamic analysis, consultation with a geotechnical engineer is strongly recommended as a
means of identifying key soil parameters to be used. It should be noted that:
The dynamic shear modulus and actual shear modulus are different parameters
The choice of the Poisson’s Ratio value can make a considerable difference to results.
In the analysis, the dynamic effect on the immediate supports and on the main structure shall be
evaluated using un-factored actual masses. Additionally, it is imperative that absolutely all masses are
accounted for in the analysis.

12.2.1 Equipment Supplier / Vendor Input


Prior to any analysis, all input data should be collated. The equipment supplier should, as a minimum,
provide:
Certified vendor drawings complete with design parameters and operating frequencies for each piece
of equipment.
Equivalent static loads, which already take into account appropriate dynamic factors including any
out of balance forces that may occur.
Values for static loads, impact forces in the horizontal and vertical directions for start-up and during
normal operation and machine operating speeds.
In the absence of any of the above, estimates of dynamic and static loads for similar equipment may be
used for preliminary design. However the final design shall be checked using certified values provided by
the equipment vendors.

Used as a guide only, QA procedures must be followed 72


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 10 Dynamic Foundations – General Guidelines

Isolated from other

structures/foundations

Machinery L

Foundation

Plan

Difference between

Machinery centroids < ±5% ‘L’

Machinery

Height of 600 min


CCG < D Foundation
thickness, D

Continuous, homogenous Qstatic 50% Qallowable

concrete foundation
Elevation Longitudinal Axis of

Machinery
Notes:

- Denotes Centre of Gravity (CG)

- Denotes Combined Centre of Gravity (CCG)

For Rotating Machinery: M Foundation 3 * M Machinery 30° 30°

For Reciprocating Machinery: M Foundation 5 * M Machinery

(Only applicable to heavy mass machinery)

12.2.2 Calculation of the Natural Frequency


For calculation of the natural frequencies or the eigen values the first choice the user needs to make is
how many modes should be considered for the analysis?

Unfortunately, there is no universally correct answer for how many modes are significant for the analysis
as it varies from case to case and also depends on what is being evaluated, such as checking the
resonance, checking the transient response or checking the response against earthquake.

Used as a guide only, QA procedures must be followed 73


42/01012/07/44062 GHD Structural Design Manual
Revision 2
The most rational basis for choosing the number of modes to analyse is based on the modal mass
participation factor which should always be the basis of arriving at the number of significant modes to be
considered for dynamic analysis when carrying out a resonance check.
When choosing the number of modes to consider for a dynamic analysis or resonance check, a method
based on the modal mass participation factor should be used. This method is explained below:
– As a first step, start with five or six significant modes and check the frequency against the
operating speed of the machine as well as the modal mass participation factor for these modes. If
the mass participation is of the order of 50 to 60% more modes should be considered. The
number of modes that excite at least 95 to 99% of the mass should be the basis of the number of
significant modes to be considered in the analysis.
The reason for 95-99% is explained below:
– Suppose for the first six modes it is found that the natural frequency of the system is below the
operating speed of the machine by 20% but has only excited 60% of the mass, while higher
modes which are in the vicinity of the operating frequency, have excited 89% of the mass (say
the 9th or the 10th mode). It is clear that these modes will excite the structure much more – an
observation that would not have been made had the analysis been restricted to a preconceived
six-mode analysis only.
Additionally, although there will always be an eigen value that matches or falls closely to the operating
frequency of the machine, as nearly 100% of the mass has already participated in the vibration in the
earlier modes, there will be no effect on the response of the structure despite the frequency falling in the
vicinity of the resonance range. It should be noted that this can only be predicted confidently if it is known
exactly how much mass has already participated in the vibration.

12.3 Design

12.3.1 Design Criteria


All machine foundations should satisfy two fundamental criteria:
Resonance does not occur between the frequencies of the pulsating loads and the natural
frequencies of the foundation/soil system.
Amplitude of any vibration does not exceed safe limits.
Foundations shall be designed such that:
The machine can function efficiently without excessive wear
The foundation suffers no damage or settlement sufficient to cause the machine to run inefficiently
The operation of the machine shall not cause discomfort or excessive vibration to personnel or
adjacent structures or plant
The foundation adopted will be the most economical which will meet all the above requirements.
The foundation is isolated from the main structural frame.

Used as a guide only, QA procedures must be followed 74


42/01012/07/44062 GHD Structural Design Manual
Revision 2
12.3.2 Minimum Reinforcement in Block Foundations:
Reinforcement in block foundations shall satisfy the following:
Requirements of codes regarding minimum reinforcement should apply in all reinforced elements of
concrete foundations.
Reinforcement should be used at all faces.
If the foundation is over 1 m thick, shrinkage reinforcement should be provided spaced at maximum
600 mm in three directions (cube reinforcing with minimum bar diameter of 16 mm).
The minimum reinforcement in a concrete block should consist of bars at least 16 mm diameter
spaced at maximum 300 mm centre to centre, extending both vertically and horizontally near all the
faces of the foundation block.

12.4 Foundations for Reciprocating Machinery


The foundation shall be designed using the methodology of British Code of Practices CP 2012 Part 1 as
a guide. The analysis shall consider rigid body motions of all six degrees of freedom and where close
coupling of natural frequencies exist then coupled motions shall be considered.

Unless vendor data requires more restrictive vibration levels, the limiting vibration levels calculated for
selected points of the machine casing shall be as given for a Group D machine as stated in ISO 10816.
These limits of this class of machine are as per Table 26:

Table 26 Machine Vibration Level Classes

Vibration Frequency Machine Operating Machine At Stand-still


Component

< 10 Hz s 100 µm s 23 µm

10 Hz v 4.5 mm/s v 1.0 mm/s

s = Single amplitude displacement


s
v = RMS velocity, where vrms (2 f )
2
Note that in Table 26, allowable values are given for vibration of the machine even when that machine is
not operating. These vibration levels are stated in order to limit the level of vibration transmitted to the
machine from other adjacent rotating machinery. Transmission from one machine to another can be a
problem, especially with adjacent, large reciprocating compressors.
The engineer’s attention is drawn to the fact that with large reciprocating compressors special attention
shall be paid to the design of pipe supports for piping close to the compressor units.

Used as a guide only, QA procedures must be followed 75


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Frequency Ratio

The engineer’s attention is drawn to clause 3.4.2 of CP 2012-1, which stipulates desirable frequency
ratios between the forcing frequencies and the natural frequencies of the foundation/machine system.
These ratios shall be targeted where possible. However, if the damped forced response analysis
indicates that calculated levels of vibration are acceptable, then the frequency ratios shall be taken as for
machines of a `lesser importance', as indicated in clause 3.4.2 of CP2012-1.

12.5 References and Further Reading


The following section has been compiled based on the following texts:
CP2012-1 (British Code for Foundations of Machinery)
Construction Industry Research and Information Association (CIRIA)
FOUNDATION GUIDELINES A. J. Smalley and P. J. Pantermuehl, Mechanical and Fluids
Engineering Division, Southwest Research Institute, January 1997
Design of structures and foundations for vibrating machinery – Suresh C Arya, Michael W. O’Neill,
George Pincus, 1979.

Used as a guide only, QA procedures must be followed 76


42/01012/07/44062 GHD Structural Design Manual
Revision 2
13. Concrete Columns and Walls

13.1 Introduction
Columns shall be designed to resist axial forces and bending moments caused by design loads and
additional bending moments produced by slenderness effects.
Columns are defined as either short or slender, and braced or unbraced.

13.2 Preliminary Design


Column dimensions are usually sized so they are short. Short columns are designed with zero additional
bending moments due to slenderness. Columns are commonly braced by core/shear walls.
The minimum dimension in the direction being considered for braced rectangular columns to be deemed
short is:

D > Le/7.5 (AS 3600 Clause 10.3.1)


Compensation for the moment effects caused by eccentricities, can be approximated by multiplying the
axial load from the floor immediately above the column being considered by the following:
Internal columns – 1.25
Corner columns – 2.0
Edge columns –1.5
Columns should be made the same size for the two top storeys, as top storey columns tend to have large
moments in relation to axial loads (IstructE/ICE ‘Green Book’).

Ultimate compressive strength


Nu = 0.6x(0.85xfc’xAc +fyxAs); fc’ = 40 MPa; f sy = 500 MPa (AS 3600 Clause 10.6)

Table 27 Ultimate Compressive Strength

Column Size Area of Section p=1% p=2% p=3% p=4%


mm x mm mm2 kN kN kN kN

300x300 90000 2106 2376 2646 2916

350x350 122500 2867 3234 3602 3969

400x400 160000 3744 4224 4704 5184

450x450 202500 4739 5346 5954 6561

500x500 250000 5850 6600 7350 8100

600x600 360000 8424 9504 10584 11664

Note: p = is total steel percentage As/Ag

For short braced columns with small bending moments the design axial strength may be taken as
0.75xNu and the bending moments disregarded (AS 3600 Clause 10.3.3). Refer also to Warner Rangan
Hall Faulkes section 22.5.3.
Used as a guide only, QA procedures must be followed 77
42/01012/07/44062 GHD Structural Design Manual
Revision 2
Slender columns are designed with additional bending moments taken into account by multiplying the
largest design bending moment by the moment magnifier; depending on the method of analysis for
determining the axial forces and bending moments.
From a constructability and cost viewpoint it is usual to size columns based on a reinforcement ratio of 1
– 1.5%. If architectural requirements require minimal section sizes it maybe better to increase concrete
strengths.
Square columns are usually the most efficient solution in terms of load carrying capability followed by
circular columns. Blade columns may have architectural planning advantages e.g. car parks.
For detailing purposes it is recommended that the least number of larger diameter bars is used to reduce
congestion of reinforcement splices and junctions. Four sided columns should ideally have multiples of
four bars and circular columns a minimum of six bars.

13.3 Detailed Design


Column interaction diagrams are commonly used to show column strength in combined bending and
compression. These diagrams are illustrated in the Concrete Cement Aggregates Association (CCAA),
Concrete Design Handbook for common column sections or can be generated using software such as
RAPT or Excel spreadsheets.
Examples of reinforcement detailing of columns refer to Concrete Institute of Australia (CIA)
Reinforcement Detailing Handbook and GHD Standard Structural Blocks
(N:\AU\Brisbane\Projects\14\0156106\Project Tools\Standard Structural blocks).
Below are some specific comments that may be useful in detailed design:

Consideration should be given to axial shortening of columns and differential movement of


supporting members between different materials and non-load bearing elements.
Where columns change size or shape and their centroids do not align, the column should be
designed for the resulting bending moment or, where appropriate, the eccentricity resolved through
the floor system.
The bending moments and lateral deflections induced by eccentricities can be significant. Elastic
lateral deflections due to dead loads will increase three fold (x3) due to creep.
Where floor concrete is a lower strength than the column concrete, AS 3600, Clause 10.8 needs to
be considered.
Where columns change size or orientation, for example a square column over a blade column, the
bearing stress at the overlapping area of the columns needs to be checked.
Ensure footings are made deep enough such that the column starter bars have their required
development lengths.
The cranking offset of vertical column bars at lapped splices results in lateral thrusts at the bend
points in the bars. Normally the bottom of the crank is located at the soffit level of the floor and is
considered to be restrained by the adjacent beam or slab. In situations where this is not possible
additional restraining ties are required. Refer to AS 3600, Clause 10.7.4.5.

Used as a guide only, QA procedures must be followed 78


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Downpipes in columns are to be discouraged. The column design must allow for the reduced cross-
section area for axial loads and the column be made of sufficient size to fit the down pipe in.
Reinforcement will need to be displaced to allow the pipe to enter and exit the column and may
produce reinforcement congestion issues.
Ensure vertical bars and ties are detailed in accordance with earthquake requirements of AS 3600,
Appendix C where required.

13.4 Walls
Walls subject to in-plane vertical forces may be designed as columns or by simplified methods of AS
3600 Clause 11.5.
In-plane horizontal forces shall also be designed for in accordance with AS 3600 Clause 11.6.
Walls designed as columns commonly do not always require the restraint provisions of vertical
reinforcement, refer AS 3600 Clause 11.7.4.

Walls subject principally to horizontal forces perpendicular to the wall may be designed as slabs.

13.5 Standard details


GHD Structural Standard details are available in AutoCad and PDF format and can be inserted into
drawings (see Appendix B).

Designers can obtain a copy of GHD Standard Details at: N:\AU\Brisbane\Projects\14\0156106\Project


Tools\Standard Structural blocks.

13.6 References and Further Reading


AS 3600 – 2009 Section 10 & 11
CIA, Recommended Practice, Reinforcement Detailing Handbook for Reinforced and Prestressed
Concrete.
Charles E. Reynolds and James C. Steedman, Reinforced Concrete Designer’s Handbook.
CCAA, Concrete Design Handbook.
Warner, R. F., Rangan, B. V., Hall, A. S., Faulkes, K.A. 1998. Concrete Structures. Longman,
Melbourne, 1998.
StructE/ICE Manual for the design of reinforced concrete building structures (‘Green Book’) (1985)

Used as a guide only, QA procedures must be followed 79


42/01012/07/44062 GHD Structural Design Manual
Revision 2
14. Reinforced Concrete In-situ - Elements

14.1 Introduction
The purpose of this section is to enable preliminary sizing for insitu reinforced concrete members to be
obtained with little effort but reasonable accuracy. It should not take the place of detailed design.

14.2 General
Reinforced concrete (RC) in-situ elements could be adopted considering the following:
High strength for compression and bending;
Good fire rating;
High construction quality;
Durability;
Easy to adapt to odd shapes; &
Generally well skilled tradespersons and world class equipment throughout Australia.
However, adoption of RC in-situ maybe not so advisable considering:
High costs;
Local unavailability;
Difficulty in concrete placement;
High self weight;
Time delay before load-bearing;
Tight schedule; &
Lack of skilled labour.

14.3 Preliminary Sizing


Member design should always start with member sizing, if for no other reason than to establish the self-
weight. Always size for serviceability and reinforce for strength (not the other way around).
Preliminary sizing to satisfy deflection limitations may be based on the following simple methods as a first
step towards the detailed design. Once the minimum span/depth ratio has been established, add the
concrete cover and select the overall depth.
For slabs, selecting a suitable slab thickness normally eliminates the need for shear reinforcement, but in
the case of beams, shear may be a governing condition. Beam width is an important consideration in
sizing the beam.
In all of the following cases (to Figure 17), ‘d’ is the effective depth to the tensile reinforcement, and
concrete strength at 28 days (fc’) is 32 MPa. For beams, the total self-weight is 6 kPa. Any load above 6
kPa should be added to the live load.

Used as a guide only, QA procedures must be followed 80


42/01012/07/44062 GHD Structural Design Manual
Revision 2
14.3.1 Slabs (per 1 metre width)
One-way slabs, span/d ratio,.
Two-way slabs, two adjacent edges discontinuous, span/d ratio,
Two way slabs, three edges discontinuous (one long edge continuous), span/d ratio,
Two way slab, four edges discontinuous, span/d ratio,
Flat plate, span/d ratio, Figure 15

14.3.2 Rectangular beams – Span/Depth


Simply-supported beams - Span/d ratio,
Continuous Beams - Span/d ratio,
Cantilever Beams - L/d < 7

14.3.3 Rectangular beams – Deemed to Comply


The deemed-to-comply span-to-depth ratio formula shown in AS 3600 Clause 8.5.4 could be simplified
to:

Lef/d = [T x bef / (2G + Q)]1/3


Where
Lef = the effective span of a member

T = deflection constant; taken as


330,000 for simply-supported beams,
690,000 for continuous beams and
34,000 for cantilever beams.
G = linear dead load (kN/m)
Q = linear live load (kN/m)
bef = width of the beam (m), 1m for one-way slab

Used as a guide only, QA procedures must be followed 81


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 11 Span/d ratio for one-way slabs

Used as a guide only, QA procedures must be followed 82


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 12 Span/d ratio for two-way slabs (two adjacent edges discontinuous)

Used as a guide only, QA procedures must be followed 83


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 13 Span/d ratio for two-way slabs (three edges discontinuous, one long edge
continuous)

Used as a guide only, QA procedures must be followed 84


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 14 Span/d ratio for two-way slabs (four edges discontinuous)

Used as a guide only, QA procedures must be followed 85


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 15 Span/d ratio flat plate

Used as a guide only, QA procedures must be followed 86


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 16 Span/d ratio for simply supported beams

Used as a guide only, QA procedures must be followed 87


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 17 Span/d ratio for Continuous beams

14.4 Preliminary Reinforcement


Preliminary sizing shall be further justified by estimating the required reinforcement.

Used as a guide only, QA procedures must be followed 88


42/01012/07/44062 GHD Structural Design Manual
Revision 2
14.4.1 Preliminary check
Case 1: Minimum Reinforcement
From AS 3600 clauses 3.1.1.3 & 8.1.6.1
Ast/bd > 0.22 (D/d)2 (0.6 fc’)/fsy
Case 2: Bending moment
Based on Clause 8.1 of AS 3600, the following approximate formula can be established:
Mu / d = 0.8 Astf sy x [1 – (0.6 x Ast / bd) x (fsy/fc’)], assume fsy = 500 MPa and fc’ = 32 MPa
As a first order check this can be approximated to Ast = Mu/(0.75 d fsy)
– When Ast/bd = 0.002, Ast = 2550 Mu / d (Mu in kNm, d in mm)
– When Ast/bd = 0.005, Ast = 2625 Mu / d (Mu in kNm, d in mm)
– When Ast/bd = 0.010, Ast = 2760 Mu / d (Mu in kNm, d in mm)
– When Ast/bd = 0.0125, Ast = 2835 Mu / d (Mu in kNm, d in mm)
Case 3: Shear stress

Based on clause 8.2 of AS 3600,


Vuc = 1 2 3bvdo x (Astfc’/ bvdo)1/3, assume fc’ = 32 MPa, 1= 1.1, 2= 3= 1.0
– When Ast/ bvdo = 0.002, Vuc = 0.44 bvdo
– When Ast/ bvdo = 0.005, Vuc = 0.60 bvdo
– When Ast/ bvdo = 0.010, Vuc = 0.75 bvdo
– When Ast/ bvdo = 0.0125, Vuc = 0.81 bvdo
Minimum shear reinforcement (beams only) = 0.35 bv s / fsy.f

Maximum bending shear stress = 0.10 fc

Maximum punching shear stress = 0.34 fc

14.5 Detailed design (Strength – moment and shear; Deflection – short-term &
long-term; Crack – flexural & shrinkage)

14.5.1 Software
GHD Validated Spreadsheets (N:\Global\Apps Resources\GHD Spreadsheets)
Slab 3.2 for slab design (strength, deflection & crack)
TEDDS
Microstran for beam & column design (strength & deflection)
Space Gass for beam and column design (strength & deflection)
Sam for section subject to biaxial and uniaxial bending (strength & deflection)
Rapt (prestressed & RC members)

Used as a guide only, QA procedures must be followed 89


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 18 Cross-sectional area for bars

Figure 19 Cross-sectional area per metre width for bar diameters

Used as a guide only, QA procedures must be followed 90


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 20 Mesh specification

14.6 Standard Details


GHD Structural Standard details are available in AutoCad and PDF format and can be inserted into
drawings (see Appendix B):
Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks) for reference. There is a reference number for every standard detail;
Designers should quote the number when it is adopted.

Used as a guide only, QA procedures must be followed 91


42/01012/07/44062 GHD Structural Design Manual
Revision 2
14.7 References and Further Reading

AS 3600 – 2009

CIA, Recommended Practice, Reinforcement Detailing Handbook for Reinforced and Prestressed
Concrete.

Charles E. Reynolds and James C. Steedman, Reinforced Concrete Designer’s Handbook.

CCAA, Concrete Design Handbook.

Warner, R. F., Rangan, B. V., Hall, A. S., Faulkes, K.A. 1998. Concrete Structures. Longman,
Melbourne, 1998.

Used as a guide only, QA procedures must be followed 92


42/01012/07/44062 GHD Structural Design Manual
Revision 2
15. Prestressed and Post Tensioned Concrete

15.1 Introduction
The purpose of this section is to enable preliminary sizing for prestressed concrete members. It should
not take the place of detailed design.

15.2 General
Advantages of using prestressed concrete include:
Common throughout Australia;
Could be designed in-house or by subcontractor;
Increase clear span;
Thinner slabs;
Lighter structures;
Reduced cracking and deflections;
Reduce storey height;
Watertightness;
Effect of restraint to floor shortening
Floors must be able to shorten to enable the prestress to be applied to the floor.
Due to the stiffness of cores, prestressing of slabs between stiff core walls maybe ineffective
because the slab cannot compress and the prestress forces are held within the cores and not the
slab.

15.3 Materials

Table 28 Specification for Prestressed Strand

Nominal Nominal Nominal Minimum Minimum Min.elong Relaxation Modulus


Diameter Steel Mass Breaking Proof to After of
(mm) Area (kg/m) Load (kN) Load fracture 1000hrs at elasticity
2
(mm ) (0.2% in 600mm 0.7 (MPa)
Offset) (%) breaking
(kN) load (%)

12.7 100.1 0.786 184 156.4 3.5 2.5 180-


205x103

15.2 143.3 1.125 250 212.5 3.5 2.5 180-


205x103

Specify low relaxation, super grade strand.

Used as a guide only, QA procedures must be followed 93


42/01012/07/44062 GHD Structural Design Manual
Revision 2
All post-tensioning companies manufacture slab anchorages of 4 and 5 strand capacity. The most
economical solution is a 5 strand anchorage as it reduces the number of castings to be used on a
project. Smaller size anchorages are available by some contractors, for example VSL has a 2 strand and
single strand anchorage.
Multistrand anchorages (3 strand to 55 strand anchorages) may be required for beams and bands. Try to
use slab anchorages over multistrand anchorages as they are easier to handle and are more
economical.
Ducts
For standard duct sizes and strand eccentricities refer to Figure 21.
Ducts are manufactured from galvanized sheet metal. VSL manufacture a plastic duct system, which
enhances durability (corrosion prevention) and reduces friction losses.

Figure 21 Standard Post-tensioning Duct Sizes

Used as a guide only, QA procedures must be followed 94


42/01012/07/44062 GHD Structural Design Manual
Revision 2
15.4 Tendon Profile
The following tendon profiles are most commonly adopted. Refer Figure 22.
Parabolic – this profile is most commonly adopted.
Harped – this profile is used for large concentrated loads, e.g. transfer columns.
Parabolic-straight – this type of profile is used for slab tendons in banded slabs and for column strip
tendons in flat slabs (middle strip tendons to be parabolic).
Concentric – this type of profile is to provide a uniform compressive force in the structure, e.g.
between band beams, slabs on ground. Ensure the tendon follows the centroid of the section, e.g.
roof. Slabs may vary in thickness due to falls; the tendon must follow the centroid to ensure
secondary effects are not induced.

Figure 22 Tendon Profiles

Parabolic

Used as a guide only, QA procedures must be followed 95


42/01012/07/44062 GHD Structural Design Manual
Revision 2
15.5 Preliminary Sizing
Typical span/total depth ratios for a variety of section types of multi-span prestressed floors are shown in
Table 29.

Table 29 Typical Span/depth for multi-span prestressed floors

Type of floor Total imposed Span / depth ratio Min slab thickness
loading (kPa) (mm)

Solid flat slab 2.5 40 250 for 4 panels

5.0 35 200 for 8 panels

10.0 30

Solid slab with drop 2.5 44


panel
5.0 40

10.0 34

One-way slab with 2.5 Slab: 45; Beam: 25


band beam (width
equal to span/5) 5.0 Slab: 40; Beam: 22

10.0 Slab: 35; Beam: 18

One-way slab with 2.5 Slab: 42; Beam: 18


narrow beam (width
equal to span/15) 5.0 Slab: 38; Beam: 16

10 Slab: 34; Beam: 13

Note: Use the above values as a guide only.

15.6 Standard Details


GHD Structural Standard details are available in AutoCAD and PDF format and can be inserted into
drawings (see Appendix B):

STD-To be advised
Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks) for reference.

Used as a guide only, QA procedures must be followed 96


42/01012/07/44062 GHD Structural Design Manual
Revision 2
16. Structural Steel

16.1 Introduction
The purpose of this section is to enable preliminary sizing for steel members and provide preliminary
information on connection types to be obtained with little effort but reasonable accuracy. It should not
take the place of detailed design.

16.2 General
Structural steel is adopted mainly because of its high strength and labour saving factor. A minimum mass
solution does not necessarily result in a minimum cost solution. Connection detailing and fabrication of
components should be simplified as much as possible. Generally shop welding and field bolting are the
most economical form of connections.
Fire protection in the form of fire spray or fire-proof box-up is used where a fire resistance is a
requirement.

16.3 Steel Grades and Sizes


Currently all hot rolled and welded Onesteel Sections are grade 300 although second hand steel
(frequently used in domestic construction) may be 250. Grade 300 is the nominal grade and the actual
strength grade to be used for design may be less (or higher). Actual yield strength is given in the
Onesteel guide online. Plate is generally 250 although ‘merchant bar’ with strength of 300 may be used
for small connections. It is recommended that for most situations grade 250 be assumed for connection
design. Other grades are available to special order, with long lead times. These should be only
nominated when approved by the Project Director. International grades are different again and if these
are to be used advice should be obtained from the suppliers as to the grade to be provided.
For historic sections considerable information is available in the BCSA publication “Historic Structural
Steelwork Handbook”, available from the Australian Institute of Steel Library.

Tubes are available in a variety of grades. Circular (CHS), square (SHS) and rectangular hollow sections
(RHS) to, AS 1163 are available with a yield stress of 250, 350MPa or 450MPa. For larger tubes,
specialist advice should be obtained as to the grade, suitability for structural work, and welding
requirements.

16.4 Frames
Typical steel frames used are:
Portal Frames (single storey)
Sway Frames (multi storey)
Braced Frames
Space Frames

Used as a guide only, QA procedures must be followed 97


42/01012/07/44062 GHD Structural Design Manual
Revision 2
16.4.1 Frame Analysis
Elastic

Generally an elastic analysis is sufficient to be carried out on a frame to find design moments and forces.
A second order analysis incorporating P- and non-linear analysis effects is available in most analysis
software. This is particularly useful for analysis of sway frames and also enables direct calculation of
member forces and moments in braced frames without the need to apply the amplification factors noted
in AS 4100 clause 4.4.2.
Plastic Analysis

Where serviceability and sway deflection requirements are not critical, a plastic analysis and design can
provide slightly lighter frames and members. Although there is no in-house software dealing with plastic
analysis, relevant literature can be consulted for methods to be used e.g. Limit state design of portal
frame buildings – Australian Institute of Steel Construction (AISC).
It should be noted that AS 4100 has some special requirements for plastic analysis and design, see
Clause 4.5.

16.4.2 Roof and Wall Bracing


This may be in the form of double diagonal tension bracing in
which one of each pair of diagonals is assumed to act in tension
or single diagonal compression bracing. Remember when using
this type of bracing, the shear load is transferred through the
joint. If high shear is a problem for braced bays then consider
using a K brace (shown right), this will halve the shear load.
Tension bracing may be rods or flats for smaller forces, or angles or tubes for larger capacities. Struts
are most efficient as tubes. Struts should be designed for self weight in addition to strut loads (not
necessary for small angles and rods as tension members).
All members assumed to be part of a bracing system should be designed for the relevant loads. This
may include purlins and/or eave beams. Eccentricities caused by out-of plane effects or non-concentric
members should be designed for.
Where bracing is horizontal such as in the roof, hangers may be required to prevent excessive sag in
small tension members. As an alternative, some form of pre-tensioning device may be used e.g. rods
with turnbuckles. For compression bracing, the reduction in axial load capacity due to self-weight
deflections should be taken into consideration (Woolcock, Kitipornchai & Bradford)

16.5 Interaction of Shear and Bending


For a simply supported steel beam subject to uniform load, the shear capacity of the cross-section is
generally calculated under pure shear as maximum shear and bending is assumed not to occur at the
same location.

For a continuous beam or cantilever beam, it is possible that maximum shear and bending occur at the
same location. The shear capacity in this case shall be checked considering the presence of bending
moment.

Used as a guide only, QA procedures must be followed 98


42/01012/07/44062 GHD Structural Design Manual
Revision 2
The calculation of nominal shear capacity for the interaction of shear and bending can be carried out by:
Proportioning method: bending moment is assumed to be resisted by the flanges (refer clause 5.12.2
of AS 4100)
Shear and bending interaction method: bending moment is assumed to be resisted by the whole of
the cross-section (refer clause 5.12.3 of AS 4100)

16.6 Connections
Connections should be designed for the loads and moments induced by the connecting members. Care
should be taken to ensure that:
The design reflects the assumptions made in the analysis e.g. full moment connections, pinned
connections etc
The size of member is adequate for the connection configuration
A variety of connections may be used, some of which are listed in Table 30:

Table 30 Typical Connection Types

Connection type Example

Flexible Connections Cleat plate


Flexible end plate
Angle seat
Bearing pad
Angle cleat

Rigid connections Fully welded


Bolted moment end plate
Bolted moment cap plate

Splices Welded splices


Bolted splices

Column bases Rigid bases


Pinned bases

Splices in long members should be made to keep lengths as near to standard rolled lengths as possible
to reduce wastage (refer to Onesteel online for availability and structural lengths). They should also be
positioned at points of minimum moment. Splices should be designed and checked for the forces and
moments at the position the splice occurs. Dimensions to splices should be clearly marked on the
drawings. It is possible to avoid bulky moment splices by the inclusion of a hinge at the splice position
and then designing the splice for shear only. (Check deflections if this procedure is adopted).

Used as a guide only, QA procedures must be followed 99


42/01012/07/44062 GHD Structural Design Manual
Revision 2
16.6.1 Erection of Building Steel Works
Normally it is the Designer’s responsibility for consideration of the safety in erection of steel structures.
To design constructible details, the Designer shall consult a senior engineer who is competent in the
design and erection of steel structures.
Close collaboration with the fabricator of the steel and the principal contractor/erector is essential, both at
the planning stage and throughout the erection process to ensure that a safe and efficient system is
planned for and achieved.

16.7 Protective Coatings


The Designer can choose from a selection of protective coating systems based on the expected service
life to first maintenance for various environments.

Many of the system options will provide equal performance in the field. The system chosen will depend
on many factors including location, material supply, available equipment, labour and cost. Typical
systems include:

Metallic spray coatings


Hot-dip galvanized coatings
Paint coatings
General atmospheric classifications, as indicated in AS 2312 “Guide to the Protection of Iron & Steel
against Exterior Atmospheric Corrosion” are as follows:
Category B (mild): Internal or External. Further than 50km from coastline and no industrial pollutants
Category C (moderate): Urban and light industrial areas 1 to 50km from coastline
Category D (Coastal): Less than 1km from coastline (not splash zone)
For special environmental conditions, e.g.: severe industrial, swimming pools, splash zone areas, etc.,
seek specialist’s advice.
Corrosion protection of structural steelwork should be planned by the job manager (JM) in the preliminary
design stages to ensure that the best performance is achieved from the system chosen. Designs should
be simple and excessive complexity should be avoided. Details to be avoided in the detailed design
include the following:
Pocket-shaped structures
Narrow crevices
Ledges
Undrained flat surfaces
Flat surfaces in loose contact
Sharp edges and corners
Intermittent welding.

Used as a guide only, QA procedures must be followed 100


42/01012/07/44062 GHD Structural Design Manual
Revision 2
16.8 Sizing

16.8.1 For steel floor or roof beams, with a depth of d and a span of L;
Case 1: Non composite UB L/d = 15~18
Case 2: Composite, UB L/d = 18~21
Case 3: All for UC L/d = 21~24
Case 4: Plate girder, WB L/d = 10~12
Case 5: Cantilever beam L/d = 7

16.8.2 For truss roof structures, with an effective depth of d (from the centre of top chord to
centre of bottom chord) and a span of L;
Case 1: Pitch roof truss, Pratt, Howe or Find L/d = 4~7 (L = 6~12 m)
Case 2: Mansard truss L/d = 7~8 (L = 15~30 m)
Case 3: Parallel chord truss (Pratt or Howe Lattice Girder) L/d = 15~25 (L = 6~50 m)
Case 4: Portal Frame (with haunches) L/d = 55 (rafter)
L/d = 50 (column)
Haunch Length = L/10
Haunch Depth = 2 rafter depth
Case 5: Space truss L/d = 20~24 (L 50 m)

16.8.3 For braced columns, calculate the column load N* and check the sectional size based on
following formula:
N* x 0.5 x fy x Ag
Where,
= 0.9
Ag = section area

16.8.4 For flooring:


Case 1: Chequer plate (thickness t, span L) L/t = 120
Case 2: Steel Grating to Webforge’s Catalogue

16.8.5 Purlins and Girts


Purlins and Girts should be sized using the manufacturers design tables or manufacturers design
programs.
Typically a maximum spacing of 1200 mm is adopted for purlins supporting ceilings. Lapped Z purlins are
the most efficient form and should be used where possible. A minimum of one row of bridging should be
provided for purlin spans greater than 3m.

Used as a guide only, QA procedures must be followed 101


42/01012/07/44062 GHD Structural Design Manual
Revision 2
For other than flat roofs consider bending about both axes particularly for tiled roofs. Use additional rows
of bridging and Z purlins for non-continuous spans to prevent twisting. Design ridge and eave members
for reactions from bridging. For curved roofs specify adjustable bridging to the manufacturer’s
specification. Roof pitches less than 3o are more suitable for small rigid structures and should not be
specified unless agreed by the client.

16.8.6 Suggested Maximum Slenderness ratio Le / ry:


Top / Bottom compression chord = 150;
Top / Bottom tension chord = 300;
Main column = 100;
Minor column = 250;
Main brace / Diagonal chord = 200;
Minor brace / Diagonal chord = 300

16.9 Capacity Check

16.9.1 Preliminary check:


Design Capacity Tables from AISC
Design Guide Suite
Portal Frame Design Charts (Figure 23, Figure 24 and Figure 25)
Slenderness ratio limits (refer 16.8.6)

16.9.2 Detailed design (Strength – moment and shear; Service – Deflection):

16.9.2.1 Software
Microstran for beam & column design (strength & deflection)
Space Gass (strength, deflection & connection)
LimCon (connection design)
Standard calculation (TEDDS)
– Steel beam analysis and design (strength, deflection, no angles)
– Steel column design (strength)
– Comprehensive pad design (fixed and pinned base)
GHD verified Excel spreadsheet (N:\Global\Apps Resources\GHD Spreadsheets)

Used as a guide only, QA procedures must be followed 102


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 23 Portal design charts (cross wind knee moment)

Used as a guide only, QA procedures must be followed 103


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 24 Portal design charts (longitudinal wind knee moment: 5 m frame spacing)

Used as a guide only, QA procedures must be followed 104


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 25 Portal design charts (longitudinal wind knee moment: 9 m frame spacing)

Used as a guide only, QA procedures must be followed 105


42/01012/07/44062 GHD Structural Design Manual
Revision 2
16.10 Standard Details
GHD Structural Standard details are available in AutoCAD and PDF format and can be inserted into
drawings (see Appendix B).
Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks) for reference. There is a reference number for every standard detail.

16.11 References and Further Reading


Australian Steel Institute Structural Steel Design Suite. (www.steel.org.au)
Woolcock, S. T., Kitipornchai, S and Bradford, M.A. Limit State Design of Portal Frame Buildings. AISC
1999.
Davison, B & Owen, G. W. (2003). The Steel Construction Institute – Steel Designers’ Manual, 6th
edition. Blackwell Publishing, UK.
Branko, E., Gorenc, A., Tinyou, R. (2005). Steel Designers’ Handbook, 7th edition. UNSW Press, Sydney.

Used as a guide only, QA procedures must be followed 106


42/01012/07/44062 GHD Structural Design Manual
Revision 2
17. Concrete Masonry

17.1 Introduction
The purpose of this section is to enable preliminary sizing for masonry members. It should not take the
place of detailed design.

17.2 General
Concrete masonry is a generic term covering many building systems that incorporate blocks of many
different shapes, sizes and strengths. It includes:
Decorative face masonry such as coloured, polished, textured or split blocks;
Plain blocks used as the load-bearing leaf in cavity construction and non load-bearing partitions;
Reinforced corefilled blockwork in large building panels and retaining walls; and
Mixed construction consisting of unreinforced, corefilled and reinforced blockwork.
Concrete masonry is adopted for the following advantages:
Non-repetitive construction;
Reasonable load bearing capacity for low-rise commercial and industrial buildings;
Effective resistance to horizontal loads; and
Less expensive than reinforced concrete

17.3 Preliminary Sizing

17.3.1 A wall of thickness t, and height, h;


Case 1: unreinforced, free standing: h/t= 6
Case 2: unreinforced, top lateral support by light (steel / timber) roof: h/t=27
Case 3: unreinforced, top lateral support by concrete slab: h/t=36
Case 4: unreinforced, isolated pier, top lateral support by light roof: h/t=13
Case 5: reinforced, free standing, reinforcement continue into support h/t=12
Case 6: reinforced, top lateral support by light (steel / timber) roof: h/t=36
Case 7: reinforced, top lateral support by concrete slab: h/t=48
Case 8: reinforced, isolated pier, top lateral support by light roof: h/t=30

17.3.2 A wall of thickness t, and width w;


Case 1: unreinforced, one edge laterally supported w/t=12
Case 2: unreinforced, two edges laterally supported w/t=36
Case 3: reinforced, one edge laterally supported, reo pass support w/t=24
Case 4: reinforced, two edges laterally supported w/t=48

Used as a guide only, QA procedures must be followed 107


42/01012/07/44062 GHD Structural Design Manual
Revision 2
17.4 Capacity Check

Preliminary check:
Check wall thickness based on sizing limits (refer Section 17.3)
Design charts in Design of Concrete Masonry Buildings by CMAA
Design charts in Reinforced Concrete Masonry Cantilever Retaining Walls by CMAA
Capacity tables, Figure 26, Figure 27 and Figure 28).

Detailed design
AS 3700- Masonry Structures
TEDDS
GHD verified structural spreadsheets (N:\Global\Apps Resources\GHD Spreadsheets)

Used as a guide only, QA procedures must be followed 108


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 26 Wall properties and wall compressive load capacity

Used as a guide only, QA procedures must be followed 109


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 27 Distributed-load Capacities for Continuous-Span Lintels

Used as a guide only, QA procedures must be followed 110


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 28 Distributed-load Capacities for Continuous-Span Lintels

17.5 Standard Details


GHD Structural Standard details are available in AutoCAD and PDF format and can be inserted into
drawings (see Appendix B).

Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project


Tools\Standard Structural blocks) for reference when they carry out design tasks. There is a reference
number for every standard detail; Designers should quote the number when it is adopted.

17.6 References and Further Reading

Used as a guide only, QA procedures must be followed 111


42/01012/07/44062 GHD Structural Design Manual
Revision 2
18. Timber Structure

18.1 Introduction
The purpose of this section is to facilitate preliminary sizing for timber members and provide preliminary
information on connection types. It should not take the place of detailed design

18.2 General

18.2.1 Type of Timber


Timber has various hardwood and softwood species that are divided into a number of stress grades.
These grades range from F2 through to F34. (Refer to Table H2.1, H2.2 and H2.4 of the Timber
Structures Code AS 1720.1.) The stress grade given to a timber element defines the basic working
stresses, the modulus of elasticity (E) and modulus of rigidity (G) (see Table H2.1 in AS 1720.1).
Softwoods vary from F4 through to F11. A common grade of softwood is unseasoned Oregon with a
stress grade of F7. Hardwoods tend to have a higher stress grade, a reasonable stress grade for an
unidentified hardwood is F11.
Timber can be seasoned or unseasoned however be careful specifying large or long unseasoned timber
members that will have finishes attached to them as shrinkage could be significant and cause problems.
The types of timber used also affects the structure’s durability characteristics so check the exposure the
members will be subject to. Termite resistant timber may also be a requirement (eg. treated pine,
cypress pine, some hardwoods).
The class of timber must be specified, i.e. Class 1 to 4. Classes 1 and 2 should be specified for exterior
timbers where Class 3 and 4 have less tolerance to weathering.

Specifying a hardwood does not necessarily mean it is a durable outdoor timber; the class of timber will
ensure durability. Remember certain hardwoods have very high shrinkage levels that can result in
warping, twisting and joint separation.

18.2.2 Timber design


Timber elements should be designed in accordance with AS 1720.1 and can be sized using the tables in
AS 1684 or the National Timber Framing Code of NSW Timber Framing Manual (for domestic
construction). Check deflection, especially with unseasoned members. Avoid bouncy timber floors. For
preliminary sizing refer to section 18.3 or Figure 29, Figure 30 and Figure 31
For small domestic construction of standard layout, it may not be necessary to design the timber
members covered by the relevant Framing Manuals as the builder can size the elements using the
standard tables. The job manager should always check in the scope of works document that detailed
timber layouts are excluded by GHD.
Check hold down requirements for light roofs. In timber-framed construction, stability is usually provided
by braced walls (see Figure 32 and Figure 33). Ensure sufficient bracing is allowed for and the studs are
adequately held down.

18.2.3 Timber connection

Used as a guide only, QA procedures must be followed 112


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Care should be taken with the design of connections involving timber. The edge distances and bolt
spacing are much greater than steel.
Timber width, number of bolts used, type of timber and the direction of load relative to the grain, will all
affect bolt capacity. Shear connectors and split rings may be used to increase bolt capacity – be careful
to check required edge distances.
For critical bolted connections, eg truss joints; it is common to specify retightening of the bolts, say 6
months after installation, especially if unseasoned timber has been used to allow for shrinkage. Checks
should be carried out for the additional deflections caused by rotation at the connection.
When using screw and coach screw joints for critical connections, consideration should be given to pre-
drilling the hole to a 40~50% diameter of the fastener. This will reduce the tendency for splitting.

18.3 Sizing

18.3.1 For bearers and joists, with a joist spacing of 450 mm, depth of d and a span of L;
Case 1: Simply-supported bearers L/d = 10~15
Case 2: Continuous bearers L/d = 12~17
Case 3: Simply-supported joists L/d = 20~25
Case 4: Continuous joists L/d = 25~30

18.4 Capacity Check

18.4.1 Preliminary check:


Design Capacity Tables from manufacturers (Figure 29, Figure 30 and Figure 31) or check for the
latest information at following web pages:
– Hyne and Son: http://www.hyne.com.au
– Laminated Timber Supplies: http://www.lamtim.com.au
– Laminated Veneer Lumber: www.chhfuturebuild.com
– Pryda Australia: http://www.pryda.com.au

18.4.2 Detailed design (Strength – moment and shear; Deflection):


Software
– Microstran for beam & column design (strength & deflection)
– DesignInHyne
– TEDDS

Used as a guide only, QA procedures must be followed 113


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 29 Timber post supporting roof or floor loads

Figure 30 Bearers

Used as a guide only, QA procedures must be followed 114


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 31 Joists

Used as a guide only, QA procedures must be followed 115


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 32 Strap brace – Speedbrace Type A Unit

Used as a guide only, QA procedures must be followed 116


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 33 Strap brace – Speedbrace Type B Unit

18.5 Standard Details


GHD Structural Standard details are available in AutoCAD and PDF format and can be inserted into
drawings (see Appendix B):

Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project


Tools\Standard Structural blocks) for reference. There is a reference number for every standard detail;
Designers should quote the number when it is adopted.

18.6 References and Further Reading

Used as a guide only, QA procedures must be followed 117


42/01012/07/44062 GHD Structural Design Manual
Revision 2
19. Composite Construction

19.1 Introduction
The combination of reinforced concrete and steel systems forms the basis of composite construction.
Composite slab construction generally consists of a cold-formed steel profiled floor deck, which acts as
the permanent formwork system for the concrete slab during construction. The decking spans in one
direction across supports and can be propped or unpropped depending on the span and depth of wet
concrete supported.
In its final state, the profiled steel decking acts as bottom reinforcement with additional top steel in the
form of mesh and bars provided to allow for continuity and to control cracking.
Where steel beams are used to support the decking, they may also be designed compositely to provide a
more economical solution. Apart from larger fully composite steel framed buildings, profiled decking is
also commonly used as “lost formwork” in positions where conventional timber forms cannot be removed
following casting of the slab eg. suspended ground floor slabs on walls or beams.
Note that for smaller jobs the option of non-composite construction should also be considered as studs
are expensive and can make erection more difficult.

19.2 Structural Arrangement


The main advantage of composite construction using profiled metal sheeting (eg. Bondek) is the
omission of formwork, and wherever possible, propping. This can lead to an extremely efficient and quick
construction procedure.

It is therefore important to select a structural arrangement that takes full advantage of the permanent
formwork system. Commonly for a rectangular grid this takes the form of secondary beams spaced at
regular intervals that allow the metal sheeting to span without props.
Primary beams span the full distance between columns in the opposite direction. In unequal rectangular
grids, the secondary beams should span the longer distance.
Be aware of the preliminary mechanical services design as large penetrations in steel beams can cause
problems and should be avoided eg. avoid putting a deep, heavily loaded beam across the path of an AC
run if possible.

19.3 Sizing

19.3.1 For flooring


Case 1: Bondek (refer to www.lysaght.com)
Case 2: W-dek (refer to www.lysaght.com)

19.3.2 For beam


Case 1: Onesteel’s Span Tables for Simply Supported Composite Beams (see Figure 34, Figure 35,
Figure 36 and Figure 37).

Used as a guide only, QA procedures must be followed 118


42/01012/07/44062 GHD Structural Design Manual
Revision 2
19.4 Capacity Check

19.4.1 Metal decking design


Lysaght have produced full design tables for their profiled decking ‘Bondek’ & ‘W-dek’ and these are
usually used for the basis of decking design indicated below:
Strength and serviceability under construction stage
Strength and serviceability under combined stage
Fire resistance period as specified in BCA requirements.

19.4.2 Steel beam design


Steel beams may be designed as propped or unpropped depending on spans and the method of
construction. Generally in steel framed construction unpropped construction is preferable. The steel
beam needs to be checked for –
Construction loads (non-composite action)
Ultimate loads (composite action)
Edge beams need to be carefully checked, particularly for their ductility requirements. Preliminary sizing
of steel beam can be undertaken from Figure 34, Figure 35, Figure 36 and Figure 37.
Unpropped Construction

The beams are designed elasticity for construction loads, wet concrete and self-weight; continuity may
be used where applicable. Deflection must be assessed and added to the working deflection in the final
composite state.

In the final state the beams are checked as simply supported composite elements under the working
loads. Stresses induced in the steel beam during construction remain and must be taken into account
when the final check is made. Steel flange stress often governs for an unpropped beam design and to
help this Grade 350 steel can be specified.
To reduce dead load deflections use pre-cambered beams.
Propped Construction

For continuously propped arrangements the beam is assumed to be unstressed during construction and
therefore the system is checked in the final state to withstand all the loads by composite action. Again
the elements are designed as simply supported.
If an alternative propping arrangement is used eg single prop at midspan, it may be possible to use the
stresses induced in the structure during the construction phase to advantage for the structure in its final
composite unpropped state.
Shear Connectors

Shear connectors facilitate composite action between the steel beam and concrete slab. These are fixed
to the beam either by:
Profiled sheeting on site, or
In the fabrication shop.

Used as a guide only, QA procedures must be followed 119


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Design of shear connectors should be in accordance with the AS 2327.1: Composite Structures, AS
5100.6 Steel & Composite Construction and the AISC Composite Beam Design Manual and Safe load
tables. Note that for profiled sheeting transverse to the steel beam, shear connectors or studs must fall
between the longitudinal stiffeners in the decking profile.
Additional transverse reinforcement may also be required in the concrete slab to transfer longitudinal
shear.

19.4.3 Software
OneSteel
Lysaght W-dek Design
Lysaght Bondek Design

Used as a guide only, QA procedures must be followed 120


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 34 Span tables for simply supported composite beams (office floors)

Used as a guide only, QA procedures must be followed 121


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 35 Span tables for simply supported composite beams (retail floors)

Used as a guide only, QA procedures must be followed 122


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 36 Span tables for simply supported composite beams (plant rooms)

Used as a guide only, QA procedures must be followed 123


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 37 Span tables for simply supported composite beams (car parks)

Used as a guide only, QA procedures must be followed 124


42/01012/07/44062 GHD Structural Design Manual
Revision 2
19.5 Standard Details
GHD Structural Standard details are available in AutoCAD and PDF format and can be inserted into
drawings (see Appendix B).
Designers can obtain a copy of GHD Standard Details N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks for reference. There is a reference number for every standard detail;
Designers should quote the number when it is adopted.

19.6 References and Further Reading


Patrick, M. Composite Beam Shear Connection Design & Detailing Practices for Australian Steel Decks.
University of Western Sydney, 2004.

Used as a guide only, QA procedures must be followed 125


42/01012/07/44062 GHD Structural Design Manual
Revision 2
20. Pavement Design

20.1 Introduction
The section describes the design of flexible pavements subjected to the heavy loads.

20.2 Failure Mechanism


To successfully design a slab it is necessary to know the possible failure mechanisms. Listed below is an
explanation of the types of failures, which can occur:
Structural failure – due to heavy load for the concrete strength / concrete thickness / subgrade.
Failure occurs on the hidden face of the slab and may take some time to become apparent.
Joint Failure – Shear failure at the joints. This may be exacerbated by water ingress through the joint
resulting in fines being pumped out from under the joint.
Fretting of joints – the exposed edges of the joints break off under the action of hard wheels.
Abrasion – Due to a large number of hard wheels traversing the slab the surface becomes crazed
and degrades to a gravel like appearance especially in heavy trafficked aisles particularly at corners.
Shrinkage – “Shrinkage” cracks are probably more an early thermal effect as they are usually
apparent a day or so after pouring, particularly in the middle of slab bays, or where slab corners are
restrained. These cracks are difficult to eliminate but good detailing greatly reduces their occurrence.

20.3 Preliminary Thickness Design


Concrete floor and pavement thicknesses are determined by taking into account the nature and
frequency of imposed loads, the strength of the foundation (subgrade / sub-base) support, and the
strength of the concrete.
In general, the objective is to keep the flexural tensile stress induced in the concrete by the imposed
loads within safe limits.
Other criteria to be considered: post load base plate to resist punching shear and settlement due to
excessive soil pressure under heavy distributed loads

20.3.1 Case 1 – Light loads


For lightly loaded commercial and industrial floors, normal thickness based on previous satisfactory
performance may be selected based on Table 31 below:

Table 31 Suggested Thickness Slab for Typical Loading Conditions

Typical Loading Classification of Subgrade Recommended Thickness of


Slab (mm)

Offices, shops, garages mainly Weak 150


for private cars, light industrial
premises with unit loading up to Normal 125
5 kPa

Garages mainly for commercial Weak 200


vehicles, industrial premises,
Used as a guide only, QA procedures must be followed 126
42/01012/07/44062 GHD Structural Design Manual
Revision 2
Typical Loading Classification of Subgrade Recommended Thickness of
Slab (mm)

Normal 175

All other loading and design See requirements of specific design


situation

20.3.2 Case 2 – Dynamic loads


For dynamic loads applied by wheeled vehicles such as trucks, or off-road vehicles i.e. forklifts and
straddle carriers. Thicknesses are determined by:
1. Calculating flexural static wheel loads
2. Limiting the stress according to the estimated number of load repetitions (Table 32)
3. Determining load stresses from Westergaard equations; or
4. Choose slab thickness from Table 33 for forklift trucks.

Table 32 Fatigue Strength of Concrete in Flexure

Stress Allowable Stress Allowable Stress Allowable Stress Allowable


Ratio Repetitions Ratio Repetitions Ratio Repetitions Ratio Repetitions

0.50 Unlimited 0.60 32, 000 0.70 2, 000 0.80 120

0.51 400,000 0.61 24, 000 0.71 1, 500 0.81 90

0.52 300, 000 0.62 18, 000 0.72 1, 100 0.82 70

0.53 240, 000 0.63 14, 000 0.73 850 0.83 50

0.54 180, 000 0.64 11, 000 0.74 650 0.84 40

0.55 130, 000 0.65 8, 000 0.75 490 0.85 30

0.56 100, 000 0.66 6, 000 0.76 360

0.57 75, 000 0.67 4, 500 0.77 270

0.58 57, 000 0.68 3, 500 0.78 210

0.59 42, 000 0.69 2, 500 0.79 160

Note: Stress Ratio = Load Stress / Flexural Strength

Used as a guide only, QA procedures must be followed 127


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 33 Recommended Thickness Slab (mm) for Typical Loadings from Forklift Trucks

8-hour working day 24-hour working day


Maximum axle

Movement per hour


Forklift truck
rating (kg)

4 10 20 4 10 20
load (kg)

Design life (years)


10 20 30 10 20 30 10 20 30 10 20 30 10 20 30 10 20 30

2000 5500 125 150 125 150

3000 7000 125 150 175

3750 8500 150 175 150 175 200

4500 10000 150 175 200 175 200 225

7000 15500 175 200 225 200 225 250 275

Notes:

i. The Table has been based on average axle loads, from typical tow-axle forklift trucks

ii. Axel loads are average, maximum fully laden axle loads

iii. Table is for “normal” subgrades. Add 25 mm for weak subgrades.

20.3.3 Case 3 – Static loads


Point loads or distributed loads could be very heavy in some industrial buildings and warehouses:

For point load, a load factor is used and correction has to be made if the loaded area is small
compared with slab thickness.
For distributed loads, one design objective is to limit tensile stresses in the top of slab in aisle ways
so that cracks will not occur.

20.3.4 Case 4 – Reinforcement


The basic objective of joint and reinforcement design is to:
1. Control the cracking that occurs in floors and pavements as a result of volume changes in the
concrete due to drying shrinkage.
2. Control the magnitude of warping stresses induced by differential temperature and / or moisture
conditions between the top and bottom of slab.
3. Divide the floor into sections for convenience during construction.
Types of pavement based on reinforcement are:
Jointed Unreinforced Concrete Pavements (JUCP): necessary reinforcement for stress condensation
appears at irregular slab pattern and contraction joint not exceeding 5 m (see Table 34)
Jointed Reinforced Concrete Pavements (RCP): light reinforced concrete with joint distance normally
in the range of 10~15 m (see Table 35)

Used as a guide only, QA procedures must be followed 128


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Continuously Reinforced Concrete Pavements (CRCP): no joints with sufficient reinforcement (see
Table 36)

Table 34 Maximum Recommended Joint Spacings for JUCP

Floor / Pavement Thickness (mm) Maximum Recommended Joint Spacing (m)

< 150 3.0

151-200 3.5

201-250 4.0

> 250 4.5

Table 35 Deformed Bar Reinforcement for CRCP

Slab thickness (mm) Reinforcement (mm2/m) Bar spacing using 16 bars


(mm)

150 900 220

175 1050 185

200 1200 165

225 1350 145

250 1500 130

Table 36 Reinforcement for RCP up to 13m Long

Slab thickness (mm) Mesh required Dowels required (diameter,


mm)

100 SL62 24

125 SL62 24

150 SL72 24

175 SL82 28

200 SL82 28

225 SL92 32

250 SL92 32

275 SL92 36

300 SL92 36

Used as a guide only, QA procedures must be followed 129


42/01012/07/44062 GHD Structural Design Manual
Revision 2
20.4 Detailed Design
Software
– Slabs
– Concrete Industrial Floors & Pavement Design V1.3

20.5 Standard Details


GHD Structural Standard details are available in AutoCad and PDF format and can be inserted into
drawings (see Appendix B):
Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks) for reference. There is a reference number for every standard detail;
Designers should quote the number when it is adopted.

20.6 References and Further Reading


Cement Concrete Aggregates Association (CCAA). (1999) Industrial Floors & Pavements – Guidelines
for design, construction & specification, 2nd edition.

Used as a guide only, QA procedures must be followed 130


42/01012/07/44062 GHD Structural Design Manual
Revision 2
21. Precast Concrete Structure

21.1 Introduction
There are two major categories of precast concrete:
Structural elements
– Slab systems (hollowcore slab, plank)
– Beams (single T beams, double T beams)
– Walls (tilt-up walls, precast wall)
– Cores
– Columns
Non-structural elements
– Facade
Consider the following key items when doing detailed design:
Material: The correct materials must be specified to achieve the quality of finish and colour required.
Size, colour and quality of the aggregate may need to be specified to obtain the final appearance.
Design for the strength and serviceability requirements of AS3600 and the Precast Concrete Handbook
(NPCAA & CIA) or use manufacturers’ design charts (Figure 38, Figure 39, , Figure 40, Figure 41, Figure
42 , Figure 43, Figure 44 , Figure 45, Figure 46, Figure 47 and Figure 48).

21.2 Hollowcore Slab (Plank)

21.2.1 Sizing
It is normal for the manufacturer to participate in the design process with the design team as well as
providing advice on costing. The design of hollowcore floors is usually undertaken in two stages:
Preliminary design – the general layout, the overall dimensions of the planks and the typical details
are selected to suit the building requirements; and
Final design – the details of the planks such as strand pattern connections, embedded items are
decided and the shop drawing produced.
For the purpose of conceptual design, size of hollowcore plank could be chosen from:
Live load capacity table (uniform area load, refer Figure 38, Figure 39, Figure 40, Figure 41, Figure
42, Figure 43); and
Equivalent load distribution method chart (FTP method, refer Figure 44, Figure 45, Figure 46 and
Figure 47).

Used as a guide only, QA procedures must be followed 131


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 38 Load capacity table for 150 mm thick hollowcore plank

Figure 39 Load capacity table for 200 mm thick hollowcore plank

Used as a guide only, QA procedures must be followed 132


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 40 Load capacity table for 250 mm thick hollowcore plank

Figure 41 Load capacity table for 300 mm thick hollowcore plank

Used as a guide only, QA procedures must be followed 133


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 42 Load capacity table for 350 mm thick hollowcore plank

Figure 43 Load capacity table for 400 mm thick hollowcore plank

Used as a guide only, QA procedures must be followed 134


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 44 Linear load distribution of point load (PCI)

Figure 45 Load distribution coefficients – centrally-placed line load (FIP)

Used as a guide only, QA procedures must be followed 135


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 46 Load distribution coefficients – point load (FIP)

Used as a guide only, QA procedures must be followed 136


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 47 Load distribution coefficients – point load at edge (FIP)

21.2.2 Self weight of planks

Table 37 Plank weight based on manufacturer’s design data

Section Name Manufacturer Dead load (kN/m2)

S150 Recrete Industries 2.42

S200 Recrete Industries 3.08

S250 Recrete Industries 3.64

S300 Recrete Industries 4.05

F205.6 Hollow Core Concrete Pty Ltd 2.55

F220.5 Hollow Core Concrete Pty Ltd 3.30

F300.4 Hollow Core Concrete Pty Ltd 3.73

F400.4 Hollow Core Concrete Pty Ltd 4.25

Used as a guide only, QA procedures must be followed 137


42/01012/07/44062 GHD Structural Design Manual
Revision 2
21.3 Precast Concrete Wall

21.3.1 Tilt-up construction


Tilt-up construction is a 2 step process where the section of Wall is cast horizontally on the ground then
tilted by crane to its final vertical position.
Similar details can be used for precast walls cast in a factory.
For recommendations on design and detailing of tilt-up panels refer The Tilt-up design and construction
manual and the Precast Concrete Handbook. Note the requirements for the stability of wall panels in fire.
The capacity of fixings must be strong enough to ensure the wall panel attached to steel structure is as
required by clause C1.11 in the Building Code of Australia (BCA).
Loss of the roof diaphragm can lead to collapse.
All design to be carried out in accordance with AS 3850 – 1990. It is common for the design of tilt-up
panel erection to be the responsibility of the contractor. This must be clearly defined in the specification.

Note that propping requirements may impose loads onto floor slabs and need to be reviewed.

21.3.2 Sizing (where t is the thickness of wall)


Overall length – 60 t
Length between fixings – 50 t
Cantilever length – 8 t
Slenderness ratio limit for fire resisting (refer Figure 48)

Figure 48 Structural adequacy, fire-rated wall

Used as a guide only, QA procedures must be followed 138


42/01012/07/44062 GHD Structural Design Manual
Revision 2
21.4 Standard Details
GHD Structural Standard details are available in AutoCad and PDF format and can be inserted into
drawings (see Appendix B).
Designers can obtain a copy of GHD Standard Details (N:\AU\Brisbane\Projects\14\0156106\Project
Tools\Standard Structural blocks) for reference. There is a reference number for every standard detail;
Designers should quote the number when it is adopted.

21.5 References and Further Reading


National Precast Concrete Association Australia (NPCAA) & Concrete Institute of Australia (CIA), (2002).
Precast Concrete Handbook (CIAZ48). Sydney.
Brooks, H. Tilt Up Concrete Association - The Tilt-up design and construction manual. USA, 2000.

Used as a guide only, QA procedures must be followed 139


42/01012/07/44062 GHD Structural Design Manual
Revision 2
22. Water Retaining Structures

22.1 Introduction
This section is intended to be further developed to include all aspects of design outlined in Key Issues,
below. Initially, it addressed only structure serviceability and joints, issues considered most likely to lead
to unsatisfactory performance in concrete water retaining structures.
The other issues raised, while important to the outcome of design and construction of a water retaining
structure do not, in GHD’s opinion, have the same immediate (negative) impact on a structure, and on its
fitness for purpose. That is not to say the other issues are not important, and their impacts on a structure
need to be considered.
There have now been additions made to this guide to include many of the other design aspects, such as
concrete quality and mix design, minimisation of cracking, estimation of peak temperatures, design for
earthquake, etc.

The types of water retaining structures that are required to be designed are (generally) those at sewage
treatment plants and those at water treatment plants
The principal difference is that water treatment plants are essentially above ground and sewage
treatment plants are below ground but there are some common requirements for these structures.

22.2 Key Issues

22.2.1 Fundamental Performance Requirements


The following list represents the most common issues associated with concrete water retaining
structures.
Ability to retain liquid

Control of cracking
Satisfactory joint performance
Quality of concrete and concrete placement (porosity, voids, boney areas)
Life Cycle Performance
Ease and cost of construction
Durability
Resistance to contained liquid over the long term, both in terms of leakage, and degradation of the
structure.
Ability of owner/operator to maintain structure

22.2.2 Fundamental Design Requirements


Capacity and structure performance requirements
Loads and load combinations
Design for serviceability and strength

Used as a guide only, QA procedures must be followed 140


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Design for durability
Material and construction requirements
Joints, waterstops, joint fillers and sealants
Testing of completed structures

22.3 Information Required Before We Start Structural Design


It is best to get involved with the process designers as early as possible and after getting a feel for the
site conditions, discuss founding levels and the type of structures, and then co-ordinate with the
geotechnical engineer before the geotechnical investigation is carried out.
The intention is to have a (preliminary) site plan showing the location and size of the structures, in order
to obtain a geotechnical report which provides the following information
Site classification
Bearing capacity recommendations and estimated settlements.
Borehole logs
Recommendations on ease of excavation and batter slopes
The natural ground water level.
Comments on dewatering and temporary works.
Finally, the 1 in a 100 year flood level needs to be confirmed. This is especially important for structures
that will be located below ground level.

22.4 Basic design


The following broad provisions can be considered to be basic solutions that form a baseline against
which detailed design results can be assessed.
They can be considered minimum solutions, and MUST be checked against detailed design provisions.
In preparing any design for walls over 3 metres high, it is recommend designers use a minimum wall
thickness of 300 mm with N16 at 150 centres horizontally as a base line case to control cracking.
If the structure is relatively small in area, designers should use as a minimum a 300 mm thick slab with
no joints other than construction joints.
For below ground structures, it is recommended that the water table is assumed to be at ground level.
Designers should take great care if electing to use solutions that are below these minimum.

22.5 Concrete Quality and Mix Design


Prior to AS 3735, AS 1480 was used for guidance on water cement ratio. The maximum water cement
ratio for water retaining structures was 0.5 and this was used for water treatment plants and 0.45 for
sewage plants.
The ability to place the concrete is very important in producing a structure which is durable and does not
leak. The aim is to have a slump of 80 mm and with simple water reducers this means about 180 litres of
water is required in the mix.

Used as a guide only, QA procedures must be followed 141


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Therefore the minimum cementitious content recommended is 360kg/m 3 for water retaining structures
and 400 kg/m3 for sewage structures.
Super plasticizers are more regularly proposed by the contractors today, and greater slumps are
achieved; this is a good thing particularly if reinforcement is congested.
For the reasons of durability, placeability and control of set rates and temperature gains, the use of fly
ash in the concrete mix is recommended, and it is suggested that 30% (by weight of cementitious
material) minimum should be specified.
One of the advantages of fly ash is the reduced heat of hydration generated in the setting process, which
results in a reduced amount of reinforcement to control restrained shrinkage or alternatively less
cracking. It also results in greater long term durability when cured correctly.

22.6 Estimation of Peak Temperatures


An early publication that assisted in the determination of the rise of temperature was a commentary on
BS5337.
It is simplistic, but is still and effective way to estimate temperature rise in a placed concrete section.
Key provisions of the BS5337 Commentary are:
For walls the basic value of T1 was 30
For slabs the basic value of T1 was 25
These assumed a 400 element thickness
It assumed an ambient placement temperature of 15 degrees.
It assumed a concrete temperature of 20 degrees.
It assumed cement content of 340kg/m3.
It assumed plywood formwork left in place.
It assumed normal British Portland cement.
If the thickness increased by 100 mm then 5 degrees was to be added on.
If the ambient temperature increased by 5 degrees then 5 degrees was to be added on.
If the cement content increased by 50kg/m3 then 6 degrees was to be added on.
Further research since has presented temperature rises for varying thickness of walls and slabs, for
various types of formwork, both ply and steel, and various stripping times. These can be considered an
improvement on the initial guide.
Over the years, other publications have appeared which have considered the use of fly ash and slag
mixes, and tables predicting temperature rises have been produced for elements up to 3 metres thick
Designers who are serious in preventing cracking in restrained concrete should be aware of these
publications and make use of them.
Two of the key recent publications are:
Early –Age thermal crack control in concrete 19:92 Ciria Report 91 T A Harrison
Early –Age thermal crack control in concrete 20:07 Ciria Report C660 P Bamford

Used as a guide only, QA procedures must be followed 142


42/01012/07/44062 GHD Structural Design Manual
Revision 2
22.7 Design for shrinkage and swelling
Design for shrinking and swelling can be done with coefficients for average temperature changes.
The shrinkage recommended in the Code of 60 microstrain is equivalent to an average change in
temperature of -5 degrees.
The swelling recommended in the Code of 110 microstrain is equivalent to an average change in
temperature of +10 degrees.
These are minor movements in comparison to the temperature induced strains and this explains why
they have been generally ignored.

22.8 Intent of AS 3735


AS 3735 states the following, which clearly sets out its intent and focus
“Although design for serviceability will generally dominate, design for strength shall be considered to
ensure that the load capacity and slenderness ratios for members or the structure are within acceptable
limits”

The Code Commentary further elaborates on the code philosophy, and provides background to the
rationale adopted.
Note that while the Code does cover all aspects of serviceability, it really only gives coverage to flexural
strength. This means that a working stress approach is mandated for serviceability and flexural strength,
but designers must use current limit state provisions from AS 3600 for shear strength.

22.9 Design for Strength


AS 3735 is the key design aid in assessing water retaining concrete structures. Use the provisions of the
Code in conjunction with Code Commentary. The Commentary identifies many useful design references
and details
The Code is NOT limit state, and has not (to date) been hijacked by academics. It still forms a sound
designer’s guide for water retaining concrete structures.
It does have limitations in that most of its design aids relate to cylindrical concrete water storage tanks. If
a structure is anything else (rectangular, designed as a transient storage, irregular in plan, etc.), further
references must be used to enable the serviceability provisions of the Code to be applied.
Key areas where attention must be paid to guarantee satisfactory performance are:
A sound understanding of the serviceability provisions of the Code.
Correct selection of joint type, locations, materials and capablities.
AS 3735 is one Code where a thorough reading of the commentary adds significantly to application of
the Code provisions.

22.9.1 Reduction Factors for Moment and Tension


The Commentary to AS 3735 has a table showing the reduction factor to be used for moment and
tension based on the percentage of reinforcement in the section for thickness up to 250 mm.

Used as a guide only, QA procedures must be followed 143


42/01012/07/44062 GHD Structural Design Manual
Revision 2
This factor is based on the amount of cracking that takes place and the reduction in stiffness of the
concrete.
A text book published in the 1980s referred to this as plastic redistribution, and suggested a value of
50%. (See also 22.10.3)
It is possible to calculate the reduction factor, for any given thickness and percentage of reinforcement in
accordance with the Commentary.
A designer who is serious in designing in accordance with the code should do this.
For example:
For a 300 wall with 50 cover and N16 at 150 the reduction factor is 0.33
For a 300 wall with 50 cover and N20 at 150 the reduction factor is 0.44
For a 400 wall with 50 cover and N24 at 150 the reduction factor is 0.5

22.10 Design for Serviceability

22.10.1 Overview
An overview of serviceability issues for concrete water retaining structures is listed below. More detailed
discussion is in the following sections.
AS3735 Section 3.1, serviceability dominates, check for strength
Relate Section 3 provisions to Commentary
Minimum reinforcement ratios set on the basis that cracks must be controlled in both width and
distribution
Key provision is to limit crack widths by limiting reinforcement stresses
Reinforcement must be considered in conjunction with a defined jointing rationale
Avoid joints wherever possible
Practical maximum regular structure size without expansion/contraction joints 50m x 50m, 50m
diameter
Quality foundation required
Design of rectangular tanks or sumps not explicitly covered in Code
Provide effective underdrains to control hydrostatic loads and assist in leak detection

22.10.2 Serviceability - Cracking


Placement temperature of concrete and air temperature at placement are important in cracking in
restrained concrete.
In design to the British code BS 8007, the crack width calculations and percentage of reinforcement
required the evaluation of two temperatures T1 and T2 which were added together and used in a crack
width formula.
T1 was the rise in temperature from ambient to the peak hydration temperature of the concrete.

Used as a guide only, QA procedures must be followed 144


42/01012/07/44062 GHD Structural Design Manual
Revision 2
T2 was the drop in seasonal change. For summer concreting this was taken as 20. It was clearly
recognised that concreting in winter did not require as much reinforcement as concreting in summer.
This fact is not emphasised in our Code, and seems to be forgotten by designers and supervisors.
Ambient was defined as the temperature at 9am on the day of placement, which represents the
temperature of the slab that the wall is cast against.
The British code specified a maximum crack width of 0.2 mm for submerged exposure. The maximum
crack width is defined as the crack which is greater in width than 95% of cracks.
This is believed to be based on a supposition that restrained shrinkage cracks are likely to autogenously
heal if they are less than 0.2 mm in width.
Our Code AS 3735 aims for a mean crack width of 0.15 mm.
Based on the paper by Zeman and Maltby, a mean crack width of 0.15 mm corresponds to a maximum
crack width of 0.27 mm, which is 30% greater than in the British code.
There are, however, logical reasons to believe that the average crack width is ¾ the maximum crack
width, and that the maximum crack width in the Australian code is actually 0.2 mm, as per the British
Code.
The reinforcement percentages in AS 3735 Table 3.1 for a given bar diameter are the same
percentages as a T1 and T2 of 40 degrees, and a coefficient of thermal expansion of 12 microstrain, with
a maximum crack width of 0.2 as per the British Code BS 80007.
The AS 3735 Commentary suggests that our Table 3.1 takes into account Australian conditions. Given
that Australian conditions are vastly different between Darwin and Hobart the value should be used with
caution.
The code does advise if sections are thick or cement contents are high, additional reinforcement is
needed and refers to the British Code for guidance.
If there is a 20 degree summer morning which generates a temperature rise of 45 degrees, and the
temperature of the structure in winter is 5 degrees, then we have a combined T1+T2 of 60 degrees. To
comply with AS 3735 you would then need to increase the reinforcement required in Table 3.1 by the
ratio of 60/40 at least.
An even distribution of cracks of mean width 0.15mm is sought, and the minimum ratios in AS3735 3.2.2
A) and B) are intended to achieve this where no other specific actions dominate.
Note that it is not expected that there will be no cracks. The Code simply seeks to control what cracks
do occur to the extent that they do not cause serious leaks, are able to autogenously heal, and cause no
long term damage to the structure durability.

22.10.3 Serviceability – Design Considerations


Given that joints are frequently the source of serious leakage, it is often desirable to ignore the provisions
for reinforcement area reduction by 25% for more closely spaced joints in 3.3.2 B), and complete the
entire design as fully restrained, with a minimum amount of (or no) jointing.

Used as a guide only, QA procedures must be followed 145


42/01012/07/44062 GHD Structural Design Manual
Revision 2
In limiting steel stresses, the Code implies that smaller diameter bars are more effective than larger
diameter bars. By limiting steel stresses, the Code effectively pre-determines the maximum crack
widths, and saves the designer having to calculate crack widths. Of course, higher steel stresses can be
used, but the designer will need to carefully assess crack widths to confirm that serviceability limits can
be met.
The Code makes no comment on preferred bar spacing.
This is because the effect of bar spacing is not uniform.

For large steel stresses, the effect of tension stiffening is small, and a reduction in spacing for a given bar
size and stress leads to a REDUCTION in crack widths.
Conversely, for small steel stresses, the effect of tension stiffening is large, and a reduction in bar
spacing for a given bar size and steel stress leads to an INCREASE in crack widths (because at small
steel stresses, better control of stress distribution leads to less cracks, but those that do occur are
larger).

In the case of temperature or shrinkage strains, increasing strain subjects the concrete to increasing
stress until the cracking strength of the section is reached. Further increase in stress is accompanied by
a decrease in section rigidity as the crack propagates. A point is eventually reached where the crack
propagation stops because the section reaches a rigidity capable of relieving the stress without further
deformation.
The section rigidity reduces to a value where stress cannot increase further, as there is no restraint to
strains which generate the stress.
The basic import of all of this is to carefully consider restraint on a structure, and reinforce the structure
accordingly. There is a tendency to take the reinforcement reduction factor of 25% in the Code as a
potential cost saving measure, and to justify using it by inserting joints.
In GHD’s experience, it has often been preferable to delete joints, treat the structure as fully restrained,
and reinforce accordingly. Often, better quality control can be maintained on reinforcement and concrete
placement than joint installation.
This also relates to temperature stresses that a structure will experience in service. In Melbourne and
Southern regions where daily and annual temperature variation can be substantial, no reductions should
be made. In Brisbane and tropical climates, the daily and annual temperature variation is much smaller,
and reinforcement reductions are possible.

22.10.4 Minimum Reinforcement Ratio for Restrained Concrete


GHDs experience is that the reinforcement specified in Table 3.1 in AS 3735 is often insufficient to
control cracking.
It is recommended that to control cracking and leakage in restrained walls for summer concreting
conditions, it is necessary to increase the percentage of reinforcement by 50 % for 16 diameter bars, and
even this will not ensure that some of the cracks will not leak.
A reasonable minimum reinforcement content is not less than N16 at 150 centres each face in a 300
thick wall.

Used as a guide only, QA procedures must be followed 146


42/01012/07/44062 GHD Structural Design Manual
Revision 2
22.10.5 Dealing with cracking
It is suggested that designers consider adding the following notes to their drawings, as cracks in fully
restrained structures are almost guaranteed.
1. Some restrained shrinking cracking is to be expected in the slabs and walls
2. The cracks may not heal autogenously during water testing
3. The contractor is required to repair cracks and stop leakage before the structure is backfilled (for
below ground structures)
4. Steps the contractor can take to minimise restrained cracking in concrete are:
– Chose placement temperatures that correspond to winter weather conditions
– If concreting in hot weather is unavoidable the concrete placement temperature should be
minimised by using ice for mixing water
– Strip formwork early to minimise heat build up in the walls
– Use a mix design which minimises the rise in temperature due to the cement hydration process
– Use placement techniques that minimise cracking

22.11 Waterstops
It has been common practice for 10 to 15 years to use hydrophilic waterstops which sit on the upper
surface of the base slab. Wall kickers on the base slab are no longer commonly used.
Provided the contractor seals the bottom of the form to prevent the grout leaking from the concrete, the
use of hydrophilic waterstops is now the preferred construction method.
In slabs, it is recommended that rearguard water stops be placed on sill beams. Sometimes, if the
designer specifies a no-fines concrete underdrainage layer, a fibre cement sheet can be embedded into
the surface of the no fines, and the rearguard can be placed on that.
This means there is a transition from a PVC rearguard waterstop system beneath the slab to a
hydrophilic system above the slab. The connection between the two can be done with either a vertical
mitred PVC waterstop or a hydrophilic waterstop over the depth of the slab.
If there are expansion joints, then a PVC system must be used in slab and walls.
Examples of typical details used for waterstops and joints are illustrated in Appendix H.

22.12 Joints
Given that movement joints in water retaining structures are often the source of major leaks, these joints
should be avoided wherever possible.
In most cases, this requires the designer to consider designing and detailing reinforcement for a fully
restrained structure, with consequent significant quantities of reinforcement to carry thermal and
shrink/swell loads.
When you add the serviceability (strain) load cases within the structure itself to externally applied actions
in larger structures, stresses in sections can quite quickly become extreme.

Used as a guide only, QA procedures must be followed 147


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Poor foundations are often a cause of unnecessary flexural actions within structures, and it is well worth
obtaining good geotechnical data to allow sound footing design and foundation treatments.
A structure at the larger end of the range mentioned above is similar in footprint to a high rise office
tower, and no-one would contemplate designing an office tower on the strength of one borehole or trial
pit, possibly not even below the location of the structure. As a minimum, boreholes and trial pits should
allow for good working knowledge of the foundation conditions under the structure walls and floor slabs.
Further, the geotechnical investigation should consider issues such as;
Variability in bearing capacity
Effect of water exposure on the founding soils
Likely differential settlement performance
Likely total settlement performance
Water table location and potential maximum and minimum levels
Lateral soil loads on walls
Most effective forms of ground improvement where founding soils are less than ideal (piling, lime
stabilisation, cement stabilisation, compaction grouting, excavation and replacement, etc.)
Aim to reduce additional loads and deflection effects on the structure by providing a uniform foundation.

A functioning and effective network of underdrains linked to pits outside the main structure can assist in
reducing or preventing external hydrostatic loads. For constantly full structures, this may not be an
issue, but nearly all structures need to be emptied for maintenance, and substantial damage can occur
when structures are emptied (buoyancy loads).
Perhaps just as importantly, if the structure is leaking, flowing underdrains can reveal the leak and even
quantify the extent of leakage.

22.13 Design for earthquake


For below ground tanks, design for earthquake need not be considered for the external walls.
For internal walls, such as baffle walls a useful guide is the New Zealand Code NZS 3106 1986.
For above ground tanks where the effects of earthquake apply to both the external and internal walls, it is
suggested that designers use the New Zealand Code NZS 3106 1986.
In general, designers can expect that a combined load case including earthquake will not be critical in
Australia.
Although the moments and tensions may be increased, the additional 25% increase in allowable stresses
generally makes the earthquake load case non critical in Australia.
This may not be the case for locations where larger seismic loads must be applied. For the design of a
WTP at Hamilton in New Zealand, GHD found the design to be dependent on the load case including
earthquake.

Used as a guide only, QA procedures must be followed 148


42/01012/07/44062 GHD Structural Design Manual
Revision 2
22.14 References and Further Reading
In using AS3735, it should be noted that nearly all its guiding provisions on temperature effects and
shrink/swell are intended for all structures, but that guidance in deriving the strains and strain loads are
only provided for cylindrical tanks, both normally reinforced and pre-stressed. (Refer to the Code
appendices)
In order to design rectangular tanks or sumps, further guidance is required. The Code Commentary
references Portland Cement Association’s “Rectangular Concrete Tanks” Skokie, 1981. This is a useful
design guide, and should be obtained by anyone wishing to design rectangular tank structures.
A second guide is Monograph No.27 by the U.S. Bureau of Reclamation. Although the two guides
referenced are now somewhat dated, they are still entirely adequate to assist in designing functional
structures. At the least, they offer a useful cross reference or check on results from a strip, grid, frame or
finite element analysis.

Used as a guide only, QA procedures must be followed 149


42/01012/07/44062 GHD Structural Design Manual
Revision 2
23. Bin Design

23.1 Introduction
This section provides guidance in selecting bulk solids container geometry and components, references
in determination of loads and procedures in structural design.

23.2 General
Bulk solid container structures include bins, silos, bunkers and dump hoppers. They are typically used for
the storage of bulk materials in mining, agricultural, food and chemical industries.

23.3 Steps in the design


Typical steps used in the design are as follows.
Obtain geometric detail usually from client, process engineer or mechanical engineer. Factors that
affect geometry of the bin can include designer preference, location, outlet dimension, storage
volume and flow of material.
Determine properties of material to be stored. To ensure adequate storage, minimum density of
material is used in the calculation of storage volume.
Determine loads and load combinations to be used in structural design of the container.
Design the structure in accordance with the relevant standards.
Appendix A in AS 3774 provides a useful data sheet to record the bulk solid container specification.

23.4 Selection Criteria


It is desirable for the bin to be clearly of funnel or mass flow type, as flow in the unstable zone may
experience undesirable vibration during the emptying process. The flow type for material in the bin can
be checked using Figure 49. For particular types of materials, when reliable flow is desirable, mass flow
is often chosen. In this case, the standard suggests selecting the hopper half-angle using the lower
bound curve.
For material with free flowing characteristics, funnel flow may be selected in favour of mass flow to
minimise design force in wall-hopper transition. The hopper half angle for funnel flow is typically larger
providing the advantage of more storage volume for the same height of container.
Eccentric fill or discharge of a bin should be avoided as it produces complex loads. Especially for circular
bins, bending moments in the walls will be introduced in addition to the typical hoop stresses.

23.5 Material Properties


Parameters derived from material properties are used in the load calculations. AS 3774 contains
characteristic values of physical properties of commonly stored materials (bulk solids), as shown in Table
3.1 of the standard. Important material properties include unit weight, angle of wall friction and effective
angle of internal friction. These values are normally presented within a range between specified upper
and lower values. Note that different applications require appropriate set of upper and lower bound value
for the properties to provide the critical load effect as shown in Table 6.1 of the standard.

Used as a guide only, QA procedures must be followed 150


42/01012/07/44062 GHD Structural Design Manual
Revision 2
For a non-standard material, it is appropriate to determine properties of bulk solids by the testing
methods described in Appendix C of the standard. In Australia, specialised organisations such as
TUNRA at the University of Newcastle or ITC at the University of Wollongong, normally carry out the
tests. When available, test results are preferable to general parameters from the standard as properties
vary when bulk solids are from different sources.
Figure 49 The Boundaries between Mass Flow and Funnel Flow
(Ref. Figure 3.4, AS 3774 – 1996)

23.6 Design Loadings and Loads Combinations


Loads are classified into groups which are permanent loads (A), normal service loads (B), environmental
loads (C) and accidental loads (D) as shown in Figure 49 which also provides load combinations.
According to AS 3774, load combinations are to produce the most adverse effects in the particular
structural member being designed. Figure 50 illustrates suggested load factors for each load type. It is
noted that load combinations that are likely to occur, are used in the structural design. In particular, the
maximum gravity load from stored bulk solids is used to design support structure. Wall pressure and
traction from bulk material mainly control the design of container wall and stiffeners. Whereas wind and
earthquake loads are important in stability checks. The followings give individual load types and
determination of these loads based on AS 3774.

23.6.1 Permanent loads (A)


Permanent loads include self weight, equipment, platform, roof, gate and other structures supported by
the container.

Used as a guide only, QA procedures must be followed 151


42/01012/07/44062 GHD Structural Design Manual
Revision 2
23.6.2 Normal services loads (B)

Symmetrically Filled and Discharged


For symmetric filled and discharged container in general, the following steps are required in determining
normal services loads.
Calculate the initial normal pressure on vertical walls as per Clause 6.2.1. This Clause also provides
variation to initial normal wall pressure due to factors that can increase or decrease wall pressure
including calculation of impact load on the vertical wall by hard lumpy material. Initial normal pressure
on vertical wall typically increases with depth as an exponential curve.
Figure 50 Classification and Combination of Loads
(Ref. Table 4.1, AS 3774-1996)

Used as a guide only, QA procedures must be followed 152


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 51 Load Factors for Ultimate Strength Design
(Ref. Table 4.2, AS 3774-1996)

Calculate initial frictional forces on vertical walls as per Clause 6.2.2 of the code. These will increase
linearly with depth.
Determine initial normal pressures on hopper walls using Clause 6.2.3.3. of the code.
Where applicable, also calculate initial forces on container closures. These forces are initial pressure
on the flat-bottomed bin and horizontal shearing tractions acting outwards from the centre of the bin.
Calculate frictional tractions on hopper walls from initial material load from Clause 6.2.3.4. of the
code.
Hard lumpy material can increase hopper wall loads when dumped into the bin. Where insufficient
material is left in the hopper, local effect due to impact load on hopper wall needs to be investigated
using Clause 6.2.3.5. of the code.
The procedures are then repeated from the beginning to determine the loads induced by flow during
symmetrical discharge. Methods of determining these flow loads are in Clause 6.3.
Figure 51 shows pressures and tractions on the container wall and hopper from the bulk solids during
both filling and discharging. Abrupt increase of load is observed at the transition from container wall
to hopper.
6.2.2 Eccentrically Filled and Discharged
For special cases of eccentrically filled and discharged bins, initial and flow loads on bin walls are
calculated in accordance with Clauses 6.4 and 6.5 respectively.
6.2.3 Other Normal Service Loads
Other normal service loads acting on a bin and its support are covered in Clauses 6.6 to 6.12 in AS
3774. These loads include the following:

Loads on gates and feeders from stored material. Consideration should be taken for the increased
loads due to large drop height and rapid fill rates.

Used as a guide only, QA procedures must be followed 153


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Impact load on the walls from dumping of solids especially in empty bin.
Loads from feeders, breakers, sizers etc. For example, shear load on outlet enclosure due to
unattached feeders.
Live loads on platforms and roofs as per AS 1170.1 and AS 1657.
Load due to differential gas pressure.
Forces from lateral restraints.
Loads on internal structural members within the stored material.
Calculation of loads transmitted to supports is carried out using the upper bound density of the
material to be stored.
Spillage of material onto attached structures such as roofs and platforms.

Used as a guide only, QA procedures must be followed 154


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Figure 52 Examples of Loads from Bulk Solids during Symmetric Filling and Discharging
Conditions.

Loads from Bulk Solids

Pressure or Traction (kPa)


0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0
0.0

0.5

1.0

1.5
Effective Depth (m)

2.0

2.5

3.0

3.5

4.0

4.5
Norm al wall pressure - Filling Wall traction - Filling
Norm al wall pressure - Discharging Wall traction - Discharging

23.6.3 Environmental Loads (C)


Determination of the environment loads are covered in Section 7 of the Standard.

Used as a guide only, QA procedures must be followed 155


42/01012/07/44062 GHD Structural Design Manual
Revision 2
23.6.4 6.3.1 Wind Loads
AS 1170.2 provides methods to calculate external and internal pressures of rectangular and cylindrical
bins. These include wind pressure coefficients used to find overall drag force on the container wall, as
well as wind pressure on the roof and for the underside of an elevated container. Methods for
determining the external non-uniform wind pressure around the cylindrical bin is also provided. Negative
internal air pressure from the wind on un-roofed containers is considered where appropriate. Careful
consideration should be given in calculating wind pressure when a group of containers or silos are built in
the vicinity of each other.

23.6.5 6.3.3 Loads from Differential Settlement of Foundation


When designing foundation for bulk solid container, differential settlement should be assessed. If
differential settlement is allowed to occur under a single column, axial force in the affected column may
become zero thus increasing the reactions in the adjacent columns and leading to large stresses in the
bracing.

23.6.6 6.3.2 Earthquake Loads


Seismic loads are assessed using AS 1170.4. However, AS 3774 provided methods to calculate
earthquake loads specific for bulk solid containers. This method takes precedence over AS 1170.4.
Although not often considered in the design of bulk solid containers, methods in finding wall pressure and
frictional traction from material due to earthquake for ground-supported and elevated containers, both
rectangular and cylindrical, are available in AS 3774. The standard also provides the magnitude of total
horizontal force and point of application on the structure.

23.6.7 6.3.4 Other Loads


Other loads may need to be considered include effect from differential temperature and loads due to
swelling of stored material from a rise in moisture content. The wall may be designed to include extra
pressure from swelling of materials.

23.6.8 Accidental Loads (D)


Where applicable, accidental loads are to apply on the container and associated structure. They are
normally determined using rational methods.

Vehicle impact loads.


Lateral loads on ROM bin wall from surcharge material being pushed by dozer or loader when the bin
is full.
Pressure cause by internal explosion.
Forces due to accumulated water inside the bin.

23.7 Structural Design


Standards for the design of steel structures (AS 4100) and concrete structures (AS 3600) are typically
applied depending on type of material used in the construction of the bulk solid container.

Used as a guide only, QA procedures must be followed 156


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Once the bin geometry is defined, calculations to determine structural components are undertaken. They
include ring beams, stiffeners, wall plate, roof and support structure. Useful reference for the design of
steel components of rectangular and cylindrical bins can be found in AWRA Technical Note 14 which
contains empirical formulae. For cylindrical bins, a check of the wall capacity should involve a maximum
combined hoop and bending stress, and wall vertical buckling. The ring beam between the vertical wall
and the hopper is designed for vertical and lateral distributed loads. Torsion is critical around support
columns for cylindrical bin due to the offset of the load from column line.
For design of a steel rectangular bin, wall thickness can be checked using Roark’s formulae for plate
under pressure. The size of a wall stiffener for a rectangular container can be conservatively estimated
by simple beam design under uniform distributed load from wall pressure.
Critical load scenarios used in design include, but are not limited to, the followings.
Initial state of fully loaded bin for design of vertical wall, hopper wall and support structure,
Emptying (discharge) condition for wall-hopper ring beam and vertical wall (buckling capacity),
Opened top bin during construction/maintenance subjected to external pressure and internal suction
from wind to check vertical wall and ring stiffeners,
Empty bin subjected to storm wind or earthquake for stability of structure.
The design of the bin and the support structure is then optimised using Finite Element Analysis (FEA).
Plate or brick elements are used as appropriate to model the container structure. For a typical steel bin
with stiffeners, the FE model is created as follows.
Wall plates, stiffeners and ring beams are typically modelled with plate elements. However, wall
stiffeners sometimes are modelled using beam elements, and offset from the wall plates, to reduce
modelling time. This type of modelling technique ignores the full restraining effect from the
continuous weld between the wall and the stiffeners.
Support structure can be modelled with beam elements. Consideration is required when modelling to
avoid high stress concentration at connection between the bin and its support.
Ideally mechanical items should be supported and founded on their own support structure. Where this is
not possible, a dynamic analysis is required to ensure the structures natural frequency is not aligned with
that of the machine operating frequency.

23.8 Material of Construction


The mechanical properties are to be considered when selecting material for wall subjected to impact load
and wear. For steel container walls, material used in the construction can be carbon steels or quenched
and tempered steel of various grades. The minimum yield strengths range from 250 MPa to 700 MPa.
Aluminium and stainless steel may be used in some applications. When the plate thickness is over 25
mm, high yield strength steel such as Bisalloy-80 may be used instead to avoid excessive amount of
weld metal, bearing in mind cost implications for adopting higher strength material.

Used as a guide only, QA procedures must be followed 157


42/01012/07/44062 GHD Structural Design Manual
Revision 2
23.9 Corrosion

23.9.1 Internal Surfaces


Bulk solids can cause wear of a metal wall. Highly abrasive bulk solids include iron ore, alumina and wet
coal. Wall liners can prevent heavy wear of container wall plate. Where severe wear is possible, hoppers
can be lined with quenched and tempered low alloy steel plates. Low friction lining plate of stainless steel
or alumina tiles may also be used on the inside surfaces of hoppers to accommodate mass flow.
It is noted that corrosion and wear can act simultaneously. The standard states that an allowance is to be
made for the loss of thickness in the steel walls, hoppers, etc. due to wear and corrosion, where this is
likely to be significant.

23.9.2 External Surfaces


Depending on the severity of the environmental condition, an appropriate surface protection system can
be selected from many metallic and paint coatings available. The choice is governed by the severity of
exposure to aggressive agents, cost, expected life of the coating, ease of access, possible damage to
coating and appearance. Guidance to surface treatments on structural steel can be obtained from
AS/NZS 2312.
Sometimes, no corrosion protection is required, such as in a very dry environment, surfaces may be only
painted for aesthetic reason.

23.10 Maintenance
To avoid problems with bulk solid containers in service, regular inspection should be carried out
according to a preventative maintenance plan. Factors in determining the frequency of inspection include
the type of container and bulk solids, nature of current defects, type and frequency of overloading, future
use, degree of difficulty of repairs, environment, and consequence of failure.

23.11 References
AS/NZS 1170.2-2002 “Structural design actions – Wind actions”
AS 1170.4-2007 “Structural design actions – Earthquake actions in Australia”

AS/NZS 2312-2002 “Guide to protection of structural steel against atmospheric corrosion by the use of
protective coatings”
AS 3600-2009 “Concrete Structures”

AS 3774-1996 “Loads on bulk solids containers”


AS 4100-1998 “Steel structures”
AWRA Technical Note 14, “Design and Construction of Welded Steel Bins”, Australian Welding
Research Association (AWRA), December 1984.
Symposium – “Steel Bins for Bulk Solids”, Australian Institute of Steel Construction (AISC) and Australian
Welding Research Association (AWRA), 1980.

Used as a guide only, QA procedures must be followed 158


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24. Durability in Design, Construction and Maintenance

24.1 Introduction

24.1.1 Why is Durability Planning Important?


Major capital works, whether government or privately owned assets, must achieve the intended service
life for acceptable return on capital employed. A durability philosophy throughout the project delivery will
provide capital investment optimisation by appropriate design, construction and maintenance measures
to achieve the asset owner’s intended service life.

24.1.2 Objective of Durability Planning


In engineering terms, durability planning is cost-effective selection of materials to achieve the intended
service life. The technical analysis determines the nature and rate of materials deterioration for given
environmental conditions.

24.1.3 Durability and Design Life


Design and construction to Standards in a particular country, or international Standards, may not achieve
the asset owner required design life in a specific corrosive environment. Significant premature
maintenance could be required. Asset owners may require a design life of 20, 50, 70, 100, 150, 300
years whilst Standards may state 40 to 60, 50, 100 or no comment on design life. Durability assessment
evaluates, explains and provides solutions to all parties.

24.1.4 Durability Tasks during Asset Life

24.1.4.1 Preliminary Design


Preliminary design durability tasks include the below:

Characterise macro and microenvironments for the structure


Identify deterioration mechanisms for construction materials in the environments
Model deterioration mechanisms to determine future service life of materials
Select materials to achieve the required service life to lowest whole of- life cost
Identify potential materials related construction difficulties
Prepare materials specifications
Develop cost-effective inspection and testing programs

24.1.4.2 Detailed Design


Detailed design durability tasks include the below:
Durability audit of final design options
Evaluate specific material options to stated performance criteria

Used as a guide only, QA procedures must be followed 159


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Review detailed material specifications for project elements with design and construction teams
QA/QC requirements for durability sensitive critical elements of materials and workmanship
Critical element programs to inspect, monitor and maintain

24.1.4.3 Construction
Construction durability tasks include the below:
Develop optimum materials technical performance
QA/QC focusing on durability critical issues
Site trials relevant for all materials
Contractor’s method statements
Materials evaluation
Critical durability construction processes
Cathodic protection and prevention systems
Corrosion monitoring systems
Troubleshooting construction difficulties
Resolution of technical noncompliance and any disputes
Review to confirm acceptable durability has been achieved prior to issue of works practical
completion certificate

24.1.4.4 Maintenance during Operation


Maintenance durability tasks include the below:
Asset management inspection audits and maintenance recommendations throughout the service life
to achieve lowest whole-of-life cost
Maintenance Plan/Asset Management Plan based on materials deterioration expectations and costs
to achieve acceptable asset owner performance predicted for 5, 10, 20, 50 years

24.1.5 MTG Durability Consultant & Durability Assessment Reports


GHD Materials Technology Group (MTG), including, has been engaged as a Durability Consultant and
completed Durability Assessment Reports for civil and building structures since the early 1990’s,
including Taywood Engineering Limited (TEL) experience prior to October 2001 when TEL Australia
became part of GHD.
Past reports are available via GHD Library, Structures and Materials Technical Forum and GHD Search
2.0 ranging from brief to construction completed on the below (contact MTG SLL or SLC for support).
Energy & Resources
Power Station seawater structures
Property & Buildings

Buildings, including exterior façade and below water level basement levels

Used as a guide only, QA procedures must be followed 160


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Transportation

Bridges
Elevated Viaducts
Tunnels
Wharf
Water
Water Tanks
Water Treatment Plants
Wastewater Treatment Plants
Pump Stations
Desalination Plants
Dams

24.1.6 Structural versus Durability Design

24.1.6.1 Structural Focus is Accepted by All and Formally Designed


Structural design and formal design by Structural Engineers are accepted by all construction parties.
Structural inadequacy or failure of civil and building structures is typically not a concern. Structural failure
without additional loading from earthquake, weather conditions, impact, etc. is a rare occurrence and
recall of such failures in recent times are the Singapore New World Hotel in 1986 that killed 30+ people,
a South Korean bridge in 1990’s with cars plunging into the river, a South Korean building interior
walkway in 1990’s that killed shoppers and a United States of America bridge that collapsed in 2008.
Investigations of these and similar failures typically identify construction defects, lack of quality control
during construction or lack of appropriate maintenance as the primary cause.
Structural Engineers are accepted as critical to provide structural adequacy using Standards, proprietary
software, technical society publications and other guides.
Standards stated design life is normally readily achieved for structural adequacy, and typically well
beyond.

24.1.6.2 Durability Focus is Expected by All and Not Formally Designed


Durability design is expected by all construction parties but formal design by Durability Engineers is
rarely a specified requirement. The common informal expectation is someone completes the durability
design within the design process and in the absence of a named person the Structural Engineer must
have completed the task. This is not a reasonable obligation for the Structural Engineer who does not
have durability training and /or experience. An alternative view is the Australian Standards take full
account of durability such that structural design being acceptable equates to acceptable durability
design. However, Australian Standards provide requirements for separate project durability assessment
in localised wording. Furthermore, reliance on durability provided by a Standard is not an acceptable
legal defence for premature durability damage to a structure where a reasonable Engineer is expected to
have awareness of other Technical Society publications that require additional durability provisions.

Used as a guide only, QA procedures must be followed 161


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Is premature deterioration or unacceptable maintenance a present day problem? Yes in some corrosively
aggressive environments, where materials selection is inadequate or building techniques are inadequate.
Therefore acceptable durability is not always being achieved to the level expected.
A common present day practice is durability design delivered without Durability Engineer input. Durability
Engineers are accepted and required by selective asset owners (e.g. Australian Water Authorities
particularly for desalination plants and Road Authorities). The formal durability design is delivered
primarily using in house knowledge and software, technical society publications, Standards and other
guides. Note that Australian Standards are not the primary design tool.
Australian Standards stated design life will not always achieve durability adequacy, with some possible
premature maintenance and repair. How will asset owners react to the risk of premature durability
defects?

24.1.7 Durability Input Guidance for Asset Owners


A formal durability process would provide input during the below listed design, construction and design
life stages for a structure. Guidance for asset owners is given in the remainder of this section.
Asset Owner Brief
Tender
Design & Construction
Concept/Preliminary Design (say 30%)
Detailed Design (say 85% and 100%)
Construction (during the works and practical completion)
Asset Management Materials Durability Focus

24.1.7.1 Asset Owner Brief


The asset owner should understand the impact of:
Durability of construction materials and impact on serviceability and maintenance, complementary to
structural adequacy.
Service life performance, not just at construction.
Structures rarely collapse whilst corrosive environments can initiate premature deterioration.
Materials selection and building techniques impact durability.
Unexpected maintenance can result with possible serviceability failure.
The asset owner Brief should state:
Design Life.
Durability requirements and by whom.
Operational conditions/constraints that impact design life are identified, or the need to identify
understood.

Used as a guide only, QA procedures must be followed 162


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.1.7.2 Tender
The durability requirements approach will be similar (some minor alternative words) for the contractual
nature of Designer, Contractor, Design and Construct, Alliance, Build Operate and Transfer, etc.
Durability input should be:
Identified
Understood
Stated in Tender Documents
Cost allocated for the works
Tender submission clearly conveys Durability Engineer input so the asset owner is aware of the
durability input offered.

24.1.7.3 Design and Construction


Durability Assessment Reports (DARs) are deliverables by a Durability Engineer that outline durability
requirements and assess durability compliance during design and construction processes of:
DAR: Preliminary Design (DAR:PD)
DAR: Detailed Design (DAR:DD)
DAR: Verification Report (DAR:VR)
The Brief and Tender submissions must effectively define what will be delivered during design and
construction. The asset owner should implement durability from the starting Brief.

24.1.7.4 Asset Management Materials Durability Focus


Asset Operation and Maintenance Management should be developed with a materials durability focus to:
Define specific requirements with the asset owner.
Provide compatibility with asset owner design life and maintenance strategy.
Conduct appropriate condition audits and analysis.
Complete appropriate maintenance works when any defects are identified.
Provide technical support on request.

24.2 Project Durability Requirements and Asset Design Service Life

24.2.1 Project Durability Requirements


Project Scope of Works and Technical Criteria (SWTC) durability requirements are to be identified for
compliance. Search the project documents for “durability”.

Used as a guide only, QA procedures must be followed 163


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.2.2 Project Asset Design Life
A project review of design for compliance will be completed including the below:
a) Design life for all structure components must be identified and agreed with the Client, asset owner
and/or project team (will vary with the contractual arrangement).
b) SWTC design life stated should be identified and checked.
c) Relevant Standards design life becomes the default where the project brief does not state a design
life.
d) All assets will require maintenance (including replacement where appropriate) during the design life
appropriate for the specific asset using access available in operation. The intended maintenance can
be stated in the durability assessment to clarify the design intent.

24.2.3 Design Life Definition


Many definitions can be found in Standards (including Australian and International), guideline
documents, technical society publications, technical papers, etc.
Australian Standards AS 4997:2005 for Maritime Structures and AS 5100:2004 for Bridges are the most
recent where Australian Engineer’s have technically debated words deemed appropriate, reached
agreement and the definition has been published for Australian use. Extracts are given below and MTG
prefers the AS 5100 definition however, other Engineer’s can have a different opinion.

24.2.3.1 AS 5100 Definition of Design Life Words (check present day Standard)
Serviceability Limit States include:
– Permanent damage due to corrosion, cracking or fatigue, which significantly reduces the
structural strength or useful service life of the structure.
End of Service Life is Defined to occur when:
– Deterioration progresses to a level that makes the structure unsafe or unserviceable.
– The level of maintenance necessary to maintain the functionality of the structure becomes
uneconomical.

24.2.3.2 AS 4997: Definition of Design Life Words (check present day Standard)
The period for which a structure or structural element remains fit for use for its intended purpose with
appropriate maintenance.
Condition at end of design life:
– Adequate strength to resist ultimate loads
– Be serviceable
– Further deterioration Inadequate structural capacity
Concrete structures need individual assessment to be durable.

Used as a guide only, QA procedures must be followed 164


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.2.4 Standards Design Life
Australian Standards (and all International Standards) have variable design life periods as listed in Table
38

Table 38 Design Life Period in Australian Standards

Australian Standards Design Subject Nominal Design Life (years)

AS 3600-2001 General concrete structures 40-60

AS 3600-2009 General concrete structures 50 +/- 20%

AS 3735-2001 Concrete structures for retaining 40-60


liquids

AS 2159-1995 Piling 40-60

AS 4997:2005 Maritime Structures Temporary works 5


Small craft 25
Normal commercial 50
Special/residential 100

AS 5100:2004 Concrete elements for bridges. 100

’92 AUSTROADS, not an Reinforced concrete elements of 100


Australian Standard but a bridges
relevant bridge reference

24.3 Environmental Exposure Classifications

24.3.1 Exposure Classifications


The aggressivity of local environments will be assessed with respect to durability and an exposure
classification designated. Australian Standards used for assessment may include the below depending
on the structure and environment. All these Codes and Guidelines provide exclusions or words of
warning that project specific exposure must be considered in durability assessment and not rely on the
Code/Guideline alone.
AS 2159 Piling
AS 2312 Guide to the Protection of Structural Steel Against Atmospheric Corrosion by the use of
Protective Coatings
AS 3600 Concrete Structures
AS 3735 Concrete Structures for Retaining Liquids
AS 4997 Maritime Structures
AS 5100 Bridge Design

Used as a guide only, QA procedures must be followed 165


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Project specific and asset owner documents will be considered (e.g. Water Authority or Road
Authority Standards or Guidance documents)
Exposure classifications determined for all structure components are summarised in table format to focus
on the key issues and reduce word content.

24.3.2 Environment Designation and Description


Various types of environments can be identified and described with an example in Table 39. These
environments are used in the summary and detailed durability assessment of material components.

Table 39 Environment Descriptions Example

Environment Designation Description

Environment 1 Soils below Soils immediately below a permanent body of water


permanent water including oceans, rivers, lakes, channels etc.

Environment 2 Within permanent Permanent bodies of water including oceans, rivers, drains,
water estuaries, lakes etc.

Environment 3 Natural ground The local soils below the peak of the water table. A region
below water table of these soils will be subject to season wetting and drying
with the rise and fall of the water table.

Environment 4 Natural ground The local soils above the peak of the water table. However,
above water table still subject to contamination and wetting due to flood
waters, irrigation and precipitation.

Environment 5 Specified fill above Fill with no PASS or significant residual salts (e.g. no
water table chlorides, no sulphates) and neutral pH.
Potential for contamination by saline ground/flood waters.

Environment 6 Intertidal and splash The region surrounding the mean saline water level subject
saline zone to changes in the water level. Includes tidal movement,
waves, seasonal variations, frequent flooding and splashing
from water traffic.

Environment 7 Spray saline zone Region immediately above the intertidal and splash saline
zone. Only considered in locations where wave action
generates spray.

Environment 8 Upper spray zone, Locations where airborne salinity is likely.


saline exposed

Environment 9 Atmospheric Encompasses the atmospherically exposed areas that are


exposed not protected from rainfall and other precipitation but is
remote from sources of airborne saline water.

Environment 10 Atmospheric Atmospherically exposed areas that are protected from


protected rainfall and precipitation.

Environment 11 Atmosphere interior Interior environments including inside buildings, internal


cavities, tunnels etc.

Used as a guide only, QA procedures must be followed 166


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.3.3 Reinforced Concrete Exposure Classifications
Standards used to determine exposure classifications of the various environments are listed below:
a) AS3600-2009 Concrete Structures for 50 +/- 20% (40-60) years design life.
b) AS5100.5-2004 Bridge Design Table 4.3 for 100 years design life. AS3600 (Concrete Structures)
and AS2159 (Piling) give alternative criteria. However, these standards only give guidance for
achieving a design life of 40-60 years, while AS5100 accounts for a design life of 100 years.
c) AS 4997 Maritime Structures for less than 5 to 100 years design life for a range of structure types.
Classifications are considered similar to AS3600 (Concrete Structures) with AS 3600 preferred as a
Code rather than a “Guideline”.

24.3.4 Steel Atmospheric Exposure Classifications


Standards used to determine exposure classifications of the various environments are listed below:
a) AS2312:2002, Guide to the protection of structural steel against atmospheric corrosion by the use of
protective coatings.

24.3.5 Steel Pile Exposure Classifications


Standards and assessment criteria used to determine exposure classifications of the various
environments are listed below:

a) AS2312:2002, Guide to the protection of structural steel against atmospheric corrosion by the use of
protective coatings.
b) AS2159 Supp1-1996, Steel piling.
c) German Gas and Water Works Engineers Association Soil Corrosivity Assessment Technique.
d) Langelier Saturation Index, and Ryznar Stability Index: Scaling tendency.

24.3.6 Acid Sulphate Soils Assessment


Durability relevant testing should be recommended for the geotechnical investigation as materials
durability required tests for Acid Sulphate Soil (ASS) are different to normal geotechnical engineering
tests.
Geotechnical investigation testing for ASS will be reviewed. Boreholes that indicate presence of
PASS/ASS will typically require additional geotechnical investigation.
Mitigation measures for ASS/PASS soils are not specifically addressed in current Australian Standards,
some guidance can be obtained from consideration of sulphate and pH levels in relevant classification
tables. Additional, more specific guidance can be obtained from:
rd
BRE Special Digest 1:2005 (3 Ed) “Concrete in aggressive ground”.
“Acid Sulfate Soils: Concrete Structures – Advice for Design and Construction”, Roads & Traffic
Authority of NSW, 1997.
GHD MTG and Environmental staff

Used as a guide only, QA procedures must be followed 167


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Consideration is given to both the current soil status and potential future condition, and to effects of
construction and other activities in the area. Mitigation measures typically include:
a) Design to avoid disturbance of or intrusion into the relevant soil horizons.
b) Design and construction practice to minimise disturbance of relevant soil horizons.
c) Construction material selection.
d) Use of sheet or liquid applied acid resistant barriers for isolation of elements from the ground.
e) Sacrificial concrete or steel thickness.
f) Excavation and imported fill.

24.3.7 Ground Water and Water Courses Corrosion Aggressivity


Information is obtained and used for evaluation of each structure component covering the below:
a) Site investigations and the Government Department of Water are sources of specific water quality
data at various sites including ground water bores and water courses.

b) Salinity data from water courses are assessed.


c) Scaling tendency of the ground waters and waterways is assessed with respect to the Langelier
Saturation Index (LSI), Ryznar Stability Index (RSI), Leaching Corrosion Index (LCI), and Spalling
Corrosion Index (SCI).

24.3.8 Soil Investigation Assessment


Soil field investigation test data relevant to durability assessment is used for the evaluation of each
structure component. Durability relevant testing should be recommended for the geotechnical
investigation as materials durability required tests for soils deterioration impact on construction materials
are different to normal geotechnical engineering tests

24.3.9 River and Ground Water Levels


River and ground water levels are determined and used for durability assessment.

24.3.10 Location Saline Water


The location of nearby saline water is shown on plan views including sea, river, inlet, inland, process
spill, bore, etc.

24.3.11 Fill Materials


Fill materials are expected to comply with AS 3798.

24.4 Deterioration Mechanisms


Deterioration mechanisms for materials are summarised in Table 40. More detailed durability issues for
construction elements are included in the following sections.
These deterioration mechanisms are used as the basis for identifying the project specific Potential
Durability Issues.

Used as a guide only, QA procedures must be followed 168


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Table 40 Deterioration Mechanisms Summary

Construction Deterioration Mechanism Factors Controlling Rate of Deterioration


Elements

Reinforced Chloride induced Corrosion Environmental characteristics


Concrete
Carbonation induced corrosion Concrete mix quality
Macrocell corrosion Crack control
Localised corrosion at cracks and Reinforcement cover
joints
As placed concrete quality, particularly cover,
Thermal/Restraint and Shrinkage compaction and curing
Cracking
Joint preparation
Coatings

Concrete Acid Sulphate Soils Environmental characteristics


Alkali-Aggregate Reactivity Aggregate properties
Sulphate attack Crack control
Acid attack from contaminated Binder type
groundwater and soil
Concrete quality
Delayed Ettringite Formation
Surface preparation
Thermal/Restraint and Shrinkage
Construction methods
Cracking
Process Chemical attack

Exposed Steel Corrosion Environmental characteristics


and Other
Coating/protection failure Material selection
Metallic objects
Cathodic protection interference Dissimilar metals
Stray currents Detailing
Fatigue Electrical insulation
Wear Coating selection/maintenance
Acid Sulphate Soils Cathodic Protection System performance
Process Chemical attack

Movement Excessive movement leading to poor Capacity for replacement


Joints running surface
Design detail
Concrete cracking, delamination and
Workmanship
spalling
Prevention of water penetration
Wear
Reinforcement detailing
Corrosion
Traffic load and volume
Process Chemical attack

Used as a guide only, QA procedures must be followed 169


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Construction Deterioration Mechanism Factors Controlling Rate of Deterioration
Elements

Coating Vandalism deterioration Design detail


Coating failure by cracking, flaking, Workmanship
peeling, spalling
Material quality and barrier resistance
Reinforced concrete corrosion properties
induced defects by lack of coating
environmental barrier resistance
(where appropriately required)
Process Chemical attack

Waterproof Construction loading Design detail


Membranes
Environmental degradation Workmanship
Prevention of water penetration
Material quality

Fibre Ultra violet (UV) light Design detail


Reinforced
Fatigue Material quality and barrier resistance
Plastic (FRP)
properties
Wear/Abrasion
Workmanship
Process Chemical attack

Roadway/ Deflection & settlement Foundation design


Pavement Poor running surface Environmental characteristics
Joint failure – Cracking
Wear
Overloading

Bridge Bearings Construction loading Capacity for replacement


Environmental degradation Positioning
Wear Protection
Design
Traffic load and volume

Public Artwork Vandalism Design


Material specific degradations such as Confinement
corrosion
Materials selection
Protective coatings

Used as a guide only, QA procedures must be followed 170


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.5 Reinforced Concrete Typical Potential Durability Issues

24.5.1 General
This section outlines typical potential durability issues for reinforced concrete elements.
A general overview of reinforced concrete durability issues is separately available from MTG. The
general overview is reviewed and those project specific potential durability issues for reinforced concrete
elements are taken into account at detailed design.

24.5.2 Concrete Degradation


a) Alkali Aggregate Reaction (AAR): AAR deterioration of concrete can occur with certain aggregate
types, which react and decompose in the presence of moist alkaline conditions (within concrete).
Using non-reactive aggregate and/or low alkali cements alleviates the potential for this form of
deterioration. Evaluation of the concrete supplier materials will be completed.
b) Backfill & Soil Chemical Induced Deterioration: The concentration of sulfate, as well as highly
alkaline or highly acidic conditions of the proposed back fill materials may be harmful to concrete.
Additionally, microbiological activity within the soil or the concrete is known to cause damage to
concrete. These problems are addressed by chemical analysis of the proposed fill materials and the
local soil conditions. Review is completed once the geotechnical investigation has been completed.
c) Delayed Ettringite Formation (DEF): DEF is a potential degradation mechanism that may occur in
steam cured precast reinforced concrete elements. It is generally accepted that to effectively prevent
concerns relating to DEF, the temperature of the concrete during steam curing has to be monitored
rather, than the steam temperature, and that for concrete temperatures of 65 C or less, DEF is not
likely with normally available cements. Precast concrete review will be completed.
d) Acid Sulfate Soils (ASS): ASS are naturally occurring soils containing pyrites, or chemical
precursors of pyrite, which have begun to oxidise through exposure to oxygen. When water passes
through ASS, sulphuric acid is leached out. Potential Acid Sulfate Soils (PASS) are similar to ASS in
nature but are in an unoxidised state. Engineering operations on ASS and PASS, such as
excavation, dredging and draining accelerate the exposure of pyritic material to air and speeds up
the production of acidic waters. These problems are addressed by chemical analysis of the proposed
fill materials and the local soil conditions. Review is completed once the geotechnical investigation
has been completed.

24.5.3 Reinforcement Corrosion


The deterioration of steel inside concrete is a corrosion process, which can be initiated by one or both of
two mechanisms.
a) Chloride contamination: the presence of a critical concentration of chloride ions at the
reinforcement surface will cause local breakdown of the passive layer, even at a high pH. In
structures exposed to a saline environment this is the primary corrosion risk for the reinforced
concrete.

Used as a guide only, QA procedures must be followed 171


42/01012/07/44062 GHD Structural Design Manual
Revision 2
b) Carbonation: the gradual penetration of atmospheric carbon dioxide in unsaturated concrete will
neutralise the protective alkaline environment surrounding embedded steel. If moisture is present,
the steel will corrode.
This deterioration mechanism is heightened by cracking, low cover depth, inappropriate material
selection or workmanship.

24.5.4 Concrete Crack Widths Durability Impact


The design crack width should be reviewed for durability in the given environment for the required design
life to achieve acceptable operation and avoid premature structure damage.

24.5.5 Concrete Crack Effect on Water Penetration


Liquid retaining and liquid excluding structures require special consideration of concrete crack impact on
water penetration. Overview considerations are below.
a) Not all cracks in concrete will necessarily leak. A number of cracks that appear wide at the surface
may not penetrate the full depth of the section. Others may follow an irregular path that limits the flow
of moisture or may be blocked by contaminants.

b) Cracks as thin as 0.1 mm can allow the penetration of moisture under piezometric head. The rate of
flow will depend on the piezometric head, thickness of section and tortuosity of the crack.
c) Where cracks form at an early age some level of sealing will occur due to a process known as
autogenous self-healing. This process will occur at early age and after water filling a Tank. The
extent of such self-healing is difficult to predict being dependent on a number of factors including the
crack width, cement binder type, rate of flow, water composition, time of water flow through the
crack, etc. However, self-healing can be confirmed by core sampling and examination of the
retrieved concrete core, and/or observation of water leakage through cracks over time (i.e. whether
self-healing seals the crack).

d) Where flow through cracks exceeds acceptable limits cracks, the crack may be sealed by a variety of
methods.

24.5.6 Thermal/Restraint and Shrinkage Crack Risk Assessment

24.5.6.1 General
The thermal/restraint and shrinkage modelling of concrete early age behaviour completed can result in
additional reinforcement recommended above reinforcement ratios compliant to design by Australian
Standards to limit predicted crack widths less than specific values (e.g. say less than 0.3 mm width for
general structural concrete and 0.1 mm width for water retaining structure). The thermal/restraint
modelling is completed to approaches given in Concrete Society Technical Reports, Construction
Industry Research and Information Association publications and GHD’s in house experience. Risk
management decisions are considered when relevant, such as suggestions to include additional
reinforcement based on the reduced likelihood of crack widths greater than specific values.

Used as a guide only, QA procedures must be followed 172


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.5.6.2 Assessment Methods
The following approach is adopted by MTG:
a) GHD reinforcement for crack control is based on GHD’s predictive model. The fundamental
knowledge was developed by GHD Materials Technology Group when formerly Taywood
Engineering Limited Australia (prior to October 2001 and included CIRIA Report 91 Early Age
Thermal Crack Control in Concrete second edition), and has been continually enhanced by new
knowledge with feedback from actual concrete project performance. The GHD approach is in general
accordance with CIRIA C660 Early Age Thermal Crack Control in Concrete.
b) Drying shrinkage assessment uses the CEB-FIP Model Code 1990 to determine long-term
shrinkage. The model in this case assumes concrete that achieves a 56-day shrinkage result of 600
microstrain, to the Concrete Specification requirement of 600 microstrain, when tested in accordance
with the Australian Standard AS 1012.13. Creep relief of the shrinkage strains has been determined
by applying the approach adopted in CEB-FIP Model Code 1990.
c) CIRIA C660 alone is not recommended for GHD analysis. Key matters include:
i) Guideline is for United Kingdom (primary intent) and European use, and is not world wide use
intent.
ii) Author Phil Bamforth is former Taywood Engineering Limited UK person who initiated the initial
GHD model together with GHD staff, when all worked for Taywood. MTG staff are familiar with
the majority technical content of CIRIA C660, which contains significant Taywood knowledge
held by GHD from the Taywood Australia operations transfer in October 2001.
iii) MS Excel model allows anyone to easily complete a concrete thermal/restraint and shrinkage
crack analysis without any background technical experience, which is a significant risk for
accuracy in the results.

24.5.6.3 Concrete Thermal/Restraint and Shrinkage Crack Formation Highest Risk Positions
The hydration reaction that takes place when cement is mixed with water is exothermic and the volume
of concrete will expand and contract as it heats up and then cools back down to the ambient
temperature. If the concrete is restrained (unable to move freely) in certain locations, this expansion and
contraction may result in the concrete cracking. Commonly the highest risk locations for crack formation
are:
a) Wall vertical cracks from restraint by prior cast floor, primarily forming from the floor base up the wall
and between prior cast wall segments.
b) Floor slab cracks from restraint by prior cast floor segments, forming from the prior cast floor
perpendicular inwards and between prior cast parallel floor segments
c) Roof cracks from restraint by prior cast walls, forming from the walls perpendicular inwards and
between parallel walls.
d) Mass foundation cracks from restraint by the base, or prior cast base and adjacent vertical elements.
The widths, location and occurrence of these cracks can be predicted using concrete Thermal/Restraint
and Shrinkage Modelling and can be controlled by a range of measures (e.g. suitable placement of
reinforcement).

Used as a guide only, QA procedures must be followed 173


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.5.6.4 Crack Risk Management for Thermal/Restraint and Shrinkage Induced Concrete Cracks
The general approach to minimise the risk of thermal/restraint and shrinkage induced concrete cracks,
particularly for liquid retaining structures, for consideration at detailed design will be:
a) Concrete pour sequence review to maximise pour size and take account of practical constraints on
concrete placement, including:
i) Wall pours to full height without horizontal construction joints or minimising horizontal joints
where full pour height cannot achieve acceptable concrete quality.
ii) Wall pour lengths to minimise vertical construction joints.
iii) Floor and roof slab sizes to minimise vertical construction joints.
iv) Floor slabs continuous reinforcement across plan area (i.e. reinforcement continue through
construction joints).
v) No expansion joints in floors and walls to reduce future inspection and maintenance
requirements.

b) Reinforcement can be used to control crack widths to less than the design limits, after the pour
sequence is finalised by the project team (i.e. input from designers, construction staff, durability
consultant, premix supplier on concrete mixes, etc.).

c) Concrete placement temperature kept as low as practically possible via interaction with the project
team and chosen premix supplier.
The concrete thermal/restraint behaviour and crack risk assessment completed at Detailed Design stage
is a guideline only. The concrete thermal/restraint behaviour and crack risk assessment must be re-
assessed after the final construction sequence is confirmed.
To resolve uncertainty from design assumptions to actual construction environmental conditions and
selected method of construction the Specification is recommended to include a requirement for the
Contractor to complete a Thermal/Restraint and Shrinkage Crack Risk Assessment. The Contractor
would need to take appropriate actions to achieve the target acceptable concrete crack width.

24.5.7 Construction Joints

24.5.7.1 General
Vertical and horizontal construction joints in reinforced concrete will create restraint for the concrete
element cast against prior cast concrete. The concrete pour sizes will be selected to practically minimise
the number of construction joints.

24.5.7.2 Past Project Drawing Details


The reason for construction joint detail design typically cannot be determined from past project drawings
alone. Design Reports and/or Durability Assessment Reports should give guidance on the joint selection
reason.

24.5.7.3 Selection Influences


Factors influencing construction joint selection include:
a) Structure dimensions

Used as a guide only, QA procedures must be followed 174


42/01012/07/44062 GHD Structural Design Manual
Revision 2
b) Method of construction including formwork selection for maximum wall height, practical maximum
concrete pour size, staff available for concrete placement and finishing, etc.
c) Structural movement impacts on the joint.
d) Predicted most likely crack risk formation positions.
e) Concrete batch plant supply capacity.
f) Construction joint materials, including practical placement in the concrete, past history of
successful use, performance history, expected service life and new products available

24.5.7.4 Construction Joints Restraint Induced Concrete Cracks


Construction joints have an impact on concrete cracks as outlined below:
a) Vertical and horizontal construction joints in reinforced concrete will create restraint for the concrete
element cast against prior cast concrete.
b) To minimise the crack risk, concrete pour sizes can be selected to practically minimise the number
of construction joints.

c) A sloped fall inwards for all below floor thickened sections (e.g. ring beam below external wall inner
face) instead of a vertical inner thickened face will provide reduced floor base restraint and reduce
the need for construction joints.

d) Provide no restraint joint details if the structural design will allow.

24.5.7.5 Construction Joint Selection


Recommended approach can include the below:
a) Durability assessment to complement structural design.
b) Constructability review meeting of all relevant parties at the earliest possible design stage for
improved design and construction productivity. Approach will vary with contract type (i.e. design,
design and construct, alliance, etc).
c) Review products with material manufacturers at the design stage, including past performance and
new products. Review should not start with the Contractor at time of construction materials.

24.5.8 Other Matters


The durability review will cover matters understood to be of most significance at the specific stage of the
project. MTG will not attempt to review all possible durability matters, which is commonly understood to
be beyond the normal scope.

24.6 Metal Elements Typical Potential Durability Issues

24.6.1 General
This section outlines typical potential durability issues for metal elements.

The MTG Current Practice Guideline on Steel Corrosion gives a general overview of durability issues
and is separately available from MTG. The general overview is reviewed and those project specific
potential durability issues for metal elements are taken into account at detailed design.

Used as a guide only, QA procedures must be followed 175


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.6.2 Structural Roof Beams, Purlins and Bridging
a) Structures roof frame includes all the beams, purlins and bridging. The elements of the roof frame
are typically constructed from galvanised steel and sometimes aluminium to gain sufficient durability
in highly aggressive environments. These elements are exposed to the internal atmospheric micro-
environment.
b) Main roof purlins should have down turned lip to the bottom flange for water shedding purposes and
positioned where possible to be self draining in high humidity environments (e.g. water tank roof).
The durability of these elements is determined by many factors including the material, the presence
and type of coating systems and the corrosivity of the atmosphere.
c) Localised corrosion attack where individual elements are connected or lay on one another is one of
the major issues. This problem can be minimised by the use of electrically insulating materials or
specific coatings where elements are in contact with one another. Additionally, localised attack may
initiate at the location of fixing points or damage to protective coatings/galvanising. The risk of
corrosion at these locations can be minimised by the selections of appropriate fixing materials, visual
identification of defects at the time of construction and localised coating repairs.

24.6.3 Roof Sheeting


a) Roof sheeting is exposed to both the external and internal atmospheric environments. The durability
of the roof sheeting material is determined by many factors including the base material, coating
systems and exposure conditions. The main long-term durability issue for this structural element is
corrosion. Assuming the use of appropriate clad steel sheeting, corrosion will typically initiate at the
location of defects or flaws in the protective coatings, or at fasteners and points of contact with other
metals. Prompt repair of localised damage and the use of suitable insulation (e.g. between fasteners
and sheeting) can minimise corrosion initiation at these locations.

b) In general, proprietary brand products have an acceptable record of service for specific intended
service life, so long as the correct grade of material is selected for the exposure conditions.

24.6.4 Fasteners
The category of fasteners includes fasteners for roofing/cladding, holding down bolts, structural steelwork
and mechanical/electrical works. Fasteners are susceptible to the general atmospheric and aqueous
corrosion. However, there are three other important corrosion issues related to their application that
should be addressed specifically.
a) Galvanic corrosion can occur when dissimilar metals are in physical contact in the same electrolyte.
To avoid this problem the design should provide for appropriate fixing materials and insulators
specified for the given element.
b) Crevice corrosion can occur along interfaces where two elements are joined. For example, where
beams are butted against one another, or between the head of a bolt and the surface of a beam
web. These areas are susceptible to increased times of moisture residency and the formation of
concentration cells where depleted oxygen levels can promote localised corrosion. This problem can
be alleviated by the careful design of joints, material selection and the use of insulation between
joints.

Used as a guide only, QA procedures must be followed 176


42/01012/07/44062 GHD Structural Design Manual
Revision 2
c) Localised damage of coating systems can occur at points where mechanical fasteners mesh with
other elements. These sites will be likely sites for corrosion initiation and should be locally repaired
using a suitable method prior to commissioning.

24.6.5 Mild Steel Pipeline

24.6.5.1 Below Ground


The type of pipe typically used for water carrying installation is Mild Steel Cement Lined MSCL, with the
application of a fusion bonded medium density external coating such as Sintakote (by Tyco Water or
similar approved). The main durability issue from the external environment is corrosion. The external
coating is typically considered to give the pipe steel adequate protection against significant degradation
due to corrosion. Corrosion will generally initiate at the locations of flaws in the external coating. Flaws
will predominately occur at location of joints or due to miss handling. It is important that proper coating
patch repair, wrapping, placement and handling procedures are followed and all pipe coatings are
inspected prior to being backfilled.

Corrosion in soils can be accelerated by:


a) Low resistivity.
b) High sulphate concentrations.
c) High chloride concentrations.
d) Low pH.
e) Stray currents.
f) Fluctuating water table.
g) Microbiological influenced corrosion.
h) Organic content of soils.
i) Redox potentials.
Where soil conditions are determined to be ‘overly aggressive’, cathodic protection may be required to
minimise external corrosion.

24.6.5.2 Above Ground


The factors affecting the above ground durability of the pipes are considered equivalent to those of the
structural steel exposed to the atmosphere. Inorganic Zinc coatings are typically used to protect MSCL
pipes above ground. In some cases aesthetic coatings may also be applied. The coating system should
be selected to provide adequate and economical protection for the steel against atmospheric corrosion
for the given environment especially at the transition from above to below ground. All above ground
coatings will require periodic maintenance. Sintakote (by Tyco Water or similar approved) coated pipes
typically used for pipelines is understood from the manufacturer to provide resistance to UV light and
should be considered for the transition zone from below ground to above ground.

24.6.5.3 Internal
The main durability issues for the internal surfaces of MSCL pipes are corrosion and wear. The following
can all cause premature failure of the internal coating system and pipe:
a) Cracks or debonded cement lining.

Used as a guide only, QA procedures must be followed 177


42/01012/07/44062 GHD Structural Design Manual
Revision 2
b) Localised coating failures.
c) Erosive flow conditions.
d) Aggressive water properties.
e) Microbiological influenced corrosion.
To address these issues:
Pipeline design should consider flow properties and water quality.
Internal inspection and patch repairs should be conducted prior to commissioning of the pipeline.
Scheduled internal inspections and maintenance should be conducted to minimise the risk of
premature failure occurring

24.6.6 Structural Metal


All structural metal elements will be evaluated for durability matters covering:
a) Grade.

b) Exposure conditions.
c) Galvanising grades as appropriate.
d) Coating systems.

e) Accessibility for future maintenance/replacement, especially of protective coating systems.


f) Detail isolation from dissimilar metals.
g) Details to prevent water ponding and accumulation of solid (particularly hygroscopic) contaminants
(e.g. Z purlins with down turned lip).
h) Need for Cathodic Protection
i) Review Water Authority available water structure manuals.

24.6.7 Stainless Steel


All stainless steel elements will be evaluated for durability matters covering:
a) Steel grade.
b) Exposure conditions.
c) Issues that potentially affect corrosion susceptibility, including, but not limited to: welding, cleaning,
finishing and fabrication procedures; mounting and fixing details; and gasket/insulator composition.
d) Detail isolation from dissimilar metals.

24.6.8 Pipelines (inclusive of all supports and protective measures)


All pipework elements including associated supports and protective measures will be evaluated for
durability matters covering:
a) Exposure conditions (above ground, below ground e.g. salty soil, aggressive soil, caustic in
groundwater, etc.).

Used as a guide only, QA procedures must be followed 178


42/01012/07/44062 GHD Structural Design Manual
Revision 2
b) Fluid properties (type, pH, temperature, pressure, flow, abrasions etc).
c) Pipe material selection.
d) Joins including cleaning, finishing and fabrication procedures and materials;
e) Pipe fitting material selection.
f) External protective coatings.
g) Internal linings and/or protective coatings inclusive of Cement Mortar Liners.
h) Jointing details, including gasket, seal or insulator composition.
i) Mounting and fixing details.
j) Detail isolation from dissimilar metals.
k) Voltage induction and mitigation.
l) Earthing.
m) Need for Cathodic Protection.
n) Review Water Authority available water pipeline manuals.

24.6.9 Other Metal Components


The durability review will cover matters understood to be of most significance at the specific stage of the
project. MTG will not attempt to review all possible durability matters, which is commonly understood to
be beyond the normal scope.

24.7 Rubbers/Elastomers, Protective Coatings, Sealants and Adhesives Typical


Potential Durability Issues

24.7.1 General
This section outlines potential durability issues for rubbers/elastomers, protective coatings, sealants and
adhesives.

24.7.2 Rubbers/Elastomers
Rubber materials are typically used as gaskets, o-rings and seals in the desalination plant. The types of
rubbers typically used are ethylene propylene diene monomer (EPDM), Nitrile (NBR) and Viton.

24.7.2.1 Loads
Rubber gaskets can be mechanically damage by over tightening of flanges. The use of torque wrenches
and controlled operational procedures will increase the durability of these materials.

24.7.2.2 UV Degradation
Rubber materials exposed to UV can degrade and thus must either contain UV stabilisers, pigments or
carbon black evenly dispersed in the compounds.

Used as a guide only, QA procedures must be followed 179


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.7.2.3 Chemical Degradation
The resistance of the rubbers to the chemicals used in structures (e.g. desalination plants) depends on
the rubber type and the compound used. The chemical concentration and temperature also affect the
rate and extent of chemical degradation. It should be remembered that whilst some rubbers may be
suitable for chemical service at low temperatures they may not be suitable for the same chemical service
at higher temperatures. A detailed assessment of the chemical resistance of the various rubbers will be
conduced however in general:
a) EPDM is resistant to most acid and alkaline solutions but can chemically degrade in chlorinated
service, fluorosilicic acid and 98% sulfuric acid.
b) Nitrile can chemically degrade in chlorinated service, fluorosilicic acid, sodium hydroxide and 98%
sulfuric acid.
c) Viton is resistant to most acid and alkaline solutions but can chemically degrade in fluorosilicic acid
and sodium hydroxide.

The rubber type and grade selected shall resist deterioration in the conditions of operation.

24.7.2.4 Temperature
Rubber materials selected must be suitable for the temperature of operation. Long term exposure of
rubber to elevated temperatures can cause degradation. Effects shall be evaluated prior to material
selection.

24.7.3 Protective Coatings


Protective coatings applied to any exposed steelwork or concrete/cementitious, composite or plastic
surface will be evaluated to determine that the following are achieved:
a) adhesion
b) durability

c) protective barrier properties


All protective coatings applied will have all required barrier properties confirmed on an Australian NATA
endorsed Certificate of Test, issued no more than two years prior to application of the paint. The
reasonably known protective coating standards will apply where relevant.
Protective coatings are expected to require reapplication during the service life of the structures.
Therefore consideration must be given to maintenance painting issues and satisfactory reapplication of
protective coatings, in particular issues of access and surface preparation where appropriate.

24.7.3.1 Standards
For potable water tanks, the internal coating must have AS/NZS 4020 approval and appropriate
conditioning procedures followed.
AS 2312 Guide to the Protection of Structural Steel against Exterior Atmospheric Corrosion shall be
followed where applicable. Specifications for each coating system will be required. APAS (Australian
Paint Approvals Scheme) can be used.
Surface preparation is essential to the durability of all coatings. For coating of metals, AS1627 should be
followed.

Used as a guide only, QA procedures must be followed 180


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.7.3.2 Inspection
Inspection during application of coatings must check for:
a) Specified surface preparation.
b) Specified atmospheric conditions appropriate for coating application.
c) Specified coating thickness and recoat times are achieved.
Coating inspectors should be certified with a current Coating Inspectors qualification provided by the
Australasian Corrosion Association or NACE.

24.7.4 Sealants and Adhesives


Movement of the joint and likelihood of moisture and contaminant penetration must be assessed.
Sealants are expected to be replaced during the service life of the structures and therefore consideration
must be given to how and when all sealants will be replaced or injected.
Particular issues to be considered include, but are not limited to, access and joint preparation
requirements for satisfactory sealing and resealing.

24.7.5 Other Components


The durability review will cover matters understood to be of most significance at the specific stage of the
project. MTG will not attempt to review all possible durability matters, which is commonly understood to
be beyond the normal scope.

24.8 Waterproofing Reinforced Concrete Typical Potential Durability Issues

24.8.1 General
This section outlines typical potential durability issues for waterproofing of reinforced concrete structures.
Waterproofing of cracks in reinforced concrete structure due to concrete thermal/restraint and shrinkage
is covered in the Reinforced Concrete section which includes sealing of cracks with proprietary products.
This section looks at potential durability issues of waterproofing of joints in construction.

24.8.2 Waterstops
Waterstops are typically PVC or hydrophilic material embedded in concrete joints and some waterstops
allow movement of the joints. Multiple waterstops are often utilised to minimise the risk of water leakage.
Waterstops are unlikely replaced during the service life of the structures therefore appropriate material
must be assessed for likelihood of moisture and contaminant penetration through the joint causing
degradation of the waterstops.
Particular issues to be considered include, but are not limited to, access and joint preparation
requirements for satisfactory installation including water ponding test.

Used as a guide only, QA procedures must be followed 181


42/01012/07/44062 GHD Structural Design Manual
Revision 2
24.8.3 Reinjectable Hose Sealing Systems
These systems are installed within a construction joint, generally near the centreline with at least 100 mm
cover, and provide the advantage of being able to seal non-visible internal surface irregularities after the
appearance of leakage. Proprietary systems are based on the use of either microcement or polymeric
grouts, allowing some ability to select grouting materials to suit joint design and potential exposure.
The waterstops are contained in the concrete and protected by the concrete from physical and ultraviolet
light damage. Provided the waterstop is correctly installed in accordance to manufacturer’s requirements
and not damaged the joint is expected to be water tight. Degradation of the waterstop materials due to
potable water contact is not expected and the material shall be manufactured to meet AS 4020.

24.8.4 Sealants
Sealants are generally used as waterproofing barrier in additional to waterstops to minimise the risk of
water leakage. Suitable sealants must be selected to take up joint movements and likelihood of moisture
and contaminant penetration through the joint causing degradation of the sealant and loss of
waterproofing barrier. Sealants are expected replaced during the service life of the structures and
therefore consideration must be given to how and when all sealants will be replaced or injected.
Particular issues to be considered include, but are not limited to, access and joint preparation
requirements for satisfactory resealing including water ponding test, and specific requirements for
exposure to chemical spills in affected areas.

24.8.5 Waterproof Coatings


Waterproof coatings may be used as a water barrier to avoid water penetration. Waterproof coatings
must be assessed for the likelihood of degradation due to mechanical effects, moisture or chemical
contaminant. Waterproof coatings are expected replaced during the service life of the structures and
therefore consideration must be given to how and when all waterproof coatings will be replaced or
reapplied.
Particular issues to be considered include, but are not limited to, access and surface preparation
requirements for satisfactory application including water ponding test, and specific requirements for
exposure to chemical spills in affected areas.

24.8.6 Other Waterproofing Methods


The durability review will cover matters understood to be of most significance at the specific stage of the
project. MTG will not attempt to review all possible durability matters, which is commonly understood to
be beyond the normal scope.

24.9 Composites and Plastics Typical Potential Durability Issues


Section will be added in the future. Contact MTG staff for present support details.

Used as a guide only, QA procedures must be followed 182


42/01012/07/44062 GHD Structural Design Manual
Revision 2
Appendix A
Standard Calculation Summary Sheet

42/01012/07/44062 GHD Structural Design Manual


Revision 2
42/01012/07/44062 GHD Structural Design Manual
Revision 2
Appendix B
Standard Details in AutoCad

42/01012/07/44062 GHD Structural Design Manual


Revision 2
42/01012/07/44062 GHD Structural Design Manual
Revision 2
Appendix C
Template for Design Information

42/01012/07/44062 GHD Structural Design Manual


Revision 2
Design information
.1. Structural System
[brief description of structure to be design]
[superstructure – gravity system: describe the force transfer route of vertical loads]
[superstructure – lateral load resisting system: describe the force transfer route of lateral loads]
[substructure: describe the structure in contact with soil]
[foundation: describe the foundation system]
The design life of new structure is [??] years.
The design fire resistance period of new structure is [??, check with architect] minutes
Expected service life to first maintenance is [??, check with architect] years (for steelwork structure only)
2. Loading Criteria
2 .1 Imposed Dead loads

[refer to Table 16 of Design Manual for typical imposed dead loads]


2 .2 Imposed Live loads

[refer to Design Manual for typical imposed live loads]


2 .3 Wind load

Level of existing structure above ground [??] m

Wind category [?]


Terrain category [2 or 2.5 generally]
Building classification [check with architect]
2 .4 Earthquake loads

[check if the new structure is classified as “post-earthquake recovery structure” before list out all the
parameters]
Structure classification [?]
Acceleration coefficient [?]
Site factor [?]
Importance level [1.25 or 1.00?]
2 .5 Geotechnical loads
[lateral earth pressure: as recommended by GIR]

[buoyancy: as recommended by GIR]


Surcharge [10?] kPa

42/01012/07/44062 GHD Structural Design Manual


Revision 2
2 .6 Crane (monorail, jib crane) load
[contact crane manufacturer for the details of wheel loads]
Maximum wheel load = [??] kN
Minimum wheel load = [??] kN
Skewing force = [??] kN
Baffle reaction = [??] kN
Lateral loads due to longitudinal travel drives = [??] kN
3. Limit states
3 .1 Ultimate design limits

For all the structure design, the individual load cases will be combined in accordance with relevant load
combination specified in AS 1170. Following load cases shall apply [delete inapplicable ones]:

Load Imposed Dead G Q Feq Wu


case

Case 1 1.2 1.2 1.5

Case 2 1.2 1.2 0.7 1

Case 3 0.9 0.9 1

Case 4 1 1 0.7 1

Case 5 0.8 0.8 0.56 1

3 .2 Serviceability limits

Following deflection limits are suggested for the serviceability limit states:

Element Type of deflection Limit

[Slab] [refer Table 11 for typical deflection limits] [1/??]

[Beam] [refer Table 11 for typical deflection limits] [1/??]

[Column] [refer Table 11 for typical deflection limits] [1/??]

Following load case should apply for serviceability:

Load G Q Ws Imposed
case Dead

Case 1 1.0 0.7 1.0 1.0

Case 2 ?? ?? ?? ??

42/01012/07/44062 GHD Structural Design Manual


Revision 2
4. Material specification
The following structural materials to be used during construction are subject to the material standards
cited in the nominated codes. Typical values for the properties of these materials are included here and
may be adjusted where appropriate.
4 .1 Concrete
[concrete grade 1] N32
[concrete grade 2] N40
4 .2 Steel Reinforcement
Deformed bar; fsy= 500 MPa, normal ductility (N Class)
Plain bar; fsy= 250 MPa, normal ductility
4.3 Structural Steel

UB, UC, PFC grade 300

RHS, CHS grade 350


Flat bar, plate grade 250

5. Geotechnical information
Refer to Geotechnical Investigation Report prepared by [??? on ???].
5 .1 Subsoil condition

[describe strata information based on GIR]


5 .2 Foundation system

[foundation system 1 and its supporting members]


[foundation system 2 and its supporting members]

6. Design reference and software


6 .1 Software [add in more if necessary]

Name Version Purpose Producer

Microtran V8.10f General structural analysis Engineering system Pty. Ltd.

LimCom V3.0 Connection design Engineering system Pty. Ltd.

SpaceGass V10.22 General structural analysis Integrated Technical Software (ITS)

TEDDS V9.0 Load Chase down CSC international

6 .2 Design standards [add in more if necessary]

S / No. Criteria Codes

42/01012/07/44062 GHD Structural Design Manual


Revision 2
01. Dead and imposed load AS/NZS 1170.0 – 2002 and AS/NZS 1170.1 – 2002 for building and
shelter design

02. Wind loading AS/NZS 1170.2 – 2002, AS 5100.2 – 2004

03. Steel design AS 4100 – 1998,

04 Concrete design AS 3600 – 2001

05 Foundation AS 2159 - 1995

42/01012/07/44062 GHD Structural Design Manual


Revision 2
Appendix D
Analysis Formulae

42/01012/07/44062 GHD Structural Design Manual


Revision 2
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1077

CA NTIL EV&S

'iw
c b—4 a
L I' L -

Wx2
Wa
Mmax 7
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

Frjp Izçjg
cwrved—raight I—
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

° W(&+/8a2b+/2ab3b3)
24E1
dmaxfJ(' +Ja) _______________________
2W
___ I/A J/Q
a b— c— ____ ____
L. L

N
i aj
______
RAW
4.— curved .4strai'ht f_-

dnaxjx c/C /5(1


5b
umax. ,(/#
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1077

CA NTIL EV&S

'iw
c b—4 a
L I' L -

Wx2
Wa
Mmax 7
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

Frjp Izçjg
cwrved—raight I—
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

° W(&+/8a2b+/2ab3b3)
24E1
dmaxfJ(' +Ja) _______________________
2W
___ I/A J/Q
a b— c— ____ ____
L. L

N
i aj
______
RAW
4.— curved .4strai'ht f_-

dnaxjx c/C /5(1


5b
umax. ,(/#
Steel Designers' Manual - 6th Edition (2003)

1078 Bending moment, shear and deflection

CA NT/LEVERS

9 w I 2W

________L________ b
-I ________L 'I

M _T()_2] Mmax = w( 4)
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

i-.--— curved —H straight k—

1 IIi
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

llWp—j
°'C 60E1 W(2#SOo2b,4Oab2-H/b)
/SbI dmox 60E1
dmax Jf(i+ i7)

IA a b— a
L L s.j

Mx IX Mx = MC
M,,,x— P a

No shears
A[ I

1 N. B. For ant/—clockwise moments


the deflect/on is upwards.
—curved k—
fCU/hLk
SEX C 2E1
Sb'
d=E2'(,
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1079

SIMPLY SUPPORTED BEAMS

w/z
IA B
T L L
RB

// —
T
N
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

R8
W
RA=RB T 94 R5

= S WL' Wa fSL2— c2j


To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

"max 384 El dà,cx.= 96(1

_ 012
2°I'ZL/
2/ when x1—
ien .;—

RAfl\ 'RB
_______ RB

When xa
djmt2n(2n)d#n2(242m]
When xo
dmax j.j (8?_4Lb2#b3) d4where mx/L and naa/L
Steel Designers' Manual - 6th Edition (2003)

1080 Bending moment, shear and deflection

L 84
p8

M Mx = Wx( 2)
A4nqx. —O•/28WL — WL/6
when x,=0•5774L
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

II
= WIS
.x;. I R8
RA=RB =
RB 2W/S
4,dmcx Jdmax
0.0/304 wi! WL5
dinax 60(1
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

51
when x =05/PJL

AA'F
T.—a b
BA
-F a
L
"A B

I' 4 ('- 3J Ij x 2X2


"mar. Mx L"YLJ
Mmax' Wi/i2

A5
RR8 wft
j dc,s i,dmax.

d, ur=ëi(85,7aLL4c2L4I) —3iffk1
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1081

SIMPLY SUPPORT(O BEAMS


-
N yzW/a
,l. ti.
Lb

£k"\
M- '"cx. 6 Mmcx _(_m# 'j)
N iW,en x—
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

NRa
'A=B= W/2

Wm
R8 —

lOab# 5b2)
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

_____ RB

z_
RA[ when x= a/i—

— R8—w/2 _______IA
4/max.
B .7

dmxJi(/5a2#2Oabi'.5b2)
Steel Designers' Manual - 6th Edition (2003)

1082 Bending moment, shear and deflection

SIMPLY SUPPORTED BEAMS


P p
1

L
pa

V A4nc P
M,,,— T
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

I I I I

I IB
'4—R8— . RA-R8-P
J/nax. J/nax.
°;nax.
PL3 -
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

a>c

"C ,,Ppib*Zc)
L
Pc(b#2a)
MD— L

I I
— Pb/L Po/L
P(b *)
LH
L
always occurs within Pot central deflection
00774 L of tfie centre of the beo,n odd the values for each P
When ba, derived from the formula
d PL3rsa Ia i7
centre 48E1L L (LII in the adjacent diagram.
This value is a/ways with/n
S % of the maximum value.
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

RAI

MC=ME
I
Steel Designers' Manual - 6th Edition (2003)

I
=P
Mmax r

dmax = 304(1
PL

/9 P1!
dmczx. 23PL3
648(1
dp
JB

A*C
MC Mf —
PL
MD

— — 2P

768(1
4/PL3
din ax. — SJPL
dinax.
Bending moment, shear and deflection

SPL
1083
Steel Designers' Manual - 6th Edition (2003)

1084 Bending moment, shear and deflection

SiMPLY SUPPOATED BEAMS


pp p p pPppPP P
A b F AB LA (n—I) forces B

'9

When n is odd.
MDME!fL Mmax —
(nL /) p1
When n is even.

MCfl Mmcx. — n. PL/8


This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

R8
'4 B("')'%
A4 —A8 "2P
;,2'max.

6JPL
dmax. 1000E%
When n is odd
PL3 r i7r it /
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

When n /5 even

dmaxiuu,p4,. . nft_ :# )J
TOTAL LOAD

When n >10, consider the load uniformly distributed


The reaction at the supports = W/2, but the maximum SE
at the ends of the beam — W(n;/)11,AW
The value of the maximum bending moment — C. WL
The value of the deflection at the centre of the span — k.

Value otn A C k
2 0 250O 0.1250 O• 0/05
S
4
0.3333
03750
0/I/I
0.1250
00118
00/24
5
6
Q.4
04/67
01200
O•IZSO
0.0/26
OO/27
7 O4286 0/224 0•0/28
8 0•4375 0•1250 00/28
9 04444 0/236 00l29
/0 O•4500 0.1250 0•0/29
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1085

SIMPLY SUPPORTED BEAMS

L '.1

A CI BA MA(,
a
L
+ b—I ___________'M8
1 MA
®M4>M5
MA
Al8

_ Al. GIL MC8 — Al . bIL Al4 1jt'—M8


(Al8 antic/ockw,M8
_________________ Shear diagram when MA "M8
PAl Va R4( Vs
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

Al4 — MB
A4 A5 M/L
As shown

M.cb 'a b'


dc-31 (zz) When M4M8.
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

ML2
For anti-clockwise moments dmcxaii —y

KiI
the deections are reversed ____________________________
2nd degree_parabola. W

L
'°1
Complement of parabola.

Mx -i (m4-2m#m) Mx (mJm2#4m1_2m4)
— Mmq .LfL

RAflJ R ______
/6

A4—R5—W/2 ,IA_R8_W/z
dmqx

a'maz — 61WL3 dmax —


28W1
Steel Designers' Manual - 6th Edition (2003)

1086 Bending moment, shear and deflection

SIMPLY SUPPORTED BEAMS

w unit /oao'
/W.unFt /oaoç
C A 0 B CADBfl
L NH A5

- MA=MB=-

RD
L,
AAI
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

—4..',,
rTA
_

d4f(
RA=RBWN

dc=o_4(jn3,Ln2_ i) fC
(s-)
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

wL2 N
/6(1
Where
boo w— unit
/W.unit
CiA
-HNI
BAD
'1QH- L
0 BE
—j O•5774L

IQH
L
_4J4 A
Max. upward deflection is at 0.

NIA
___ W(L#N)(LN)

m.x/L ,,1.N/L
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1087

8(J/LT—/N BEAMS

1p1/W wI2 W12

H
-- b-
L -4

WL
MA=MB
EN
WL
MC
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

4B WI? = = w/2
—.10g/L .—O s8L—.12/L p—
____ I /dflfl

= WI! -J Wa / I

384(1 vinax=4f% (L—OJ


To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

'-jvl
V4—a-4..—b- "+ c—4 'Ak—- L— -
in

MA

MA,b [e(4L-Je)- c 31'4L —Jc)] MA=-.in (Jnr"—em+o)


B,bLY4t--7°" (4L —saj M8_ jrn2 (4-3m) ,'Mmax
RA 'L,,2( 3si,)
/2
k—x- When x=2(n,3—2m42)
2
When r is the support reaction
M-M
sinp/e
M-M
A4 L _r3# L RA = W(m-2) 8 = Wm(2-m)
2 2ri,
dna.
u
When a=4/2 and x1=O-445L
When a = c. WI!
.Ey(L3#2Lba #4Lc?—8a) 3JJ(%
WI!
c/C =
384(1
Steel Designers' Manual - 6th Edition (2003)

1088 Bending moment, shear and deflection

B/JILT-/N BEAMS

LA L
8 A C ___
M
MA ___________M8 MA V NMB
- WL 7/Qx 9x ) M- SWL

#Mmax WL/233 N5Qn x=OS5L MWL//6


MA = —WI//S M8 =— WL/iO
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

-
RB
R8
R=O.SW R8=O7W RA=RB=W/2
—4J•22LF—O S6L —Ø22L

WI! — 1.4W!!
max. - 384E1
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

when x, = 0• 525L

W/2 2W/L HI'a

La- b—+_aJ
L—4 ____C
F I.

MA M8 (SL44aL_402) MA = M8 -WL//6
M=WL/48

R8 W/2 A = R3 W/Z
JRB
. L/2 —H ¼

O-WL3
364(1
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1089

B(//LT-/N BEAMS
2W/a
fwTw/a
L ____ L
b

MAV NMa
MA MA =-J'2(Jo/obL)
M3
mAC.
Mx =P.X+M3- 2W(x.-bP
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

InCB.Mx RB.X#M5
AB W/2
A = (/OL—sLa42a)
____ I /d____
____ , prO,,
2
4 —

dmax = Wa /
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

-Y HJ
W/2

—a I
W/a

b
L
-f
W/2

I.
W

L
b
.1

1NMB MA
MB
MA Maz_R(4L_3a) MA — _!10L_15#502)
M3 _lOL2(5L4t

= = W/2
= (/oL-/sLa'#8a)
R = '/5L —ec)
dmax = (/SL—Ma)
Steel Designers' Manual - 6th Edition (2003)

1090 Bending moment, shear and deflection

BUILT—/N BEAMS
parabolic total /oao W con,c*ment
parabo/p

JM8 MAV NMB


A -M5=-WL//O Aq rM8i.._WL/2O

JRa
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

'!4 W/2 RA R8 W/2

F4axJ d,,, 384E!


1.3 WL3 O•4WL3
dma 384 (2'
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

r
Any symmetrical load W

A ______
symmetrical diagram
rA

2br
a rb
2 -2br----—-
NMB
aab r&_
MA = M8 —A,JL
where A5 is the area of the 'free'
bending moment diagram
CM. f(Ja—L)M8c -M(3b-L)
When
A
RB '4I
A =R8 = W/2 = P8 = slope of moment diagram
M#M M#M
Eab
- - I The fic.re shown
A, is ha/f the bending
* + moment diagram
fr- X, - C 1* and +7
H-X--- LareC.GI When '2/L = in,
A5x —AIX1 M. L2m2(/_m)2(1_2m)
lmaxatC ZET 2E%
Where A' is the area of the fixing For ant/clockwise moments
moment diagram reverse the deflect/ons
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1091

BUILT—/N BEAMS
p p

L/2 L/2 I-.


F. L
b 8JJ

MA L—' Al5 MA LZ EEE3 M8


- MA = - M5 = M = PL/8 MA=- M5 = —
2Pa2b
_______
__________ Mc 1!
A1_____ _____ A__________
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

= = /,,2 =P (iT) (I#2 *)


1Q8 = z

4mcx p9
dinax = /92
I.— X
dC—
— __
Po3b3
El
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

d _ZPa2b3 WñQnX=
"3EI(3L2a)2 JL-Za
p p p p

MA—MB—
Pa MA/ \MB
- __Pa(L-q) MA=MB=_JPLh6
L
MC MD = P02,'L MC =MD =
RAfl
LJ AI I

PA=P'P
II
dmax PL
= /30(a)7
Steel Designers' Manual - 6th Edition (2003)

1092 Bending moment, shear and deflection

BUILT—/N BEAMS
p
______
fA C
P

D
p p
Irg
'.—L/y —1--L/j —.—L/j. l,L/3 P.-L/3L2
MA MB MA LF- 5N M8
MA =M8 = — 2PL/9 MA =M = —/9PL/72
MC =MD = PL/p MD =1/

AL I
PArJ__LJA
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

RA = JP/2

S PL3 4/PL3
dmax 548?1
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

p pp PPpp
II
A C D +
£ +
BL 34C D E 1I
PB
L/4--L/4 L/4 L/4 L/L/4 +L/4 +L/4 4L14

MAVNMB MAL/L NEiNMB


MA=MB =r—SPL/,6 MA =M3=—I/PL/j2
= JPL/15 MD = Mf = SPL/32

2P

— PL3
dmax —
96(1 max. = 96(1
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1093

BUILT—/N BEAMS
p pp p ppppppp
A CVt
L/5 +Lg +L,/s+L4
P 4 (n-i) forces
I I
B
sôcev,naLJ,

MAMALNMB
M52PL/5
MA P1,'/)
MA=MB=—
MD wM — PL/5

AArL
HJ
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

'A B=2'°
-
Y1F,ennisoø6 ,I / /
dmax.=1000
13PL3
El
Cimox 114-4c (,ji
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

Whennis even,

EfIiiiitiIII
•:.:i••
COLUMN
LOAD PER SA4NW
O...o.&. ,.
n CONTINUOUS BEAM
}d L/n#-+-L/n4L/n+-L/-#L1-+-44
L
4;.' •O4•4•.

-
L. - When n >10, consider the load .nitorrrdy oYstiibuted
The load on the outside stringers is carried c'/rect/y by the supports
The continuous beam Is assumed to be horizontal at each support
The reaction at the supports for each s,oan = W/2. but the maximum
shear force in any span of tM continuous beam = V/J.IAW
The value of the fixing moment at each support = — B. WL
The value of the maximum positive moment for each span = C. W4
The value of the maximum deflection for each span —0'0O26
Value ofn A B C
2 02500 00625 0•0 625
3 0•3333 0074/ 00370
4 03750 0078/ 00469
S 04000 0-0800 0•0400
6 04/67 Qc// 0•0439
7 04286 0O8/5 00408
B Q•4375 00820 00 430
9 04444 00823 004/3
10 04500 00825 00425
Steel Designers' Manual - 6th Edition (2003)

1094 Bending moment, shear and deflection

PROPPED CAN T/L(VERS

/w W

A C B ___ a C
-f-b
MA

3L/8 - MA =— f (2n)2wñere a/Ln


= —WL
--- C—
— 9W!.
#Mmax
RA -n (4-n)]
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

—--_J P8 1

P3 =f RAi4J[8n2(4n)]
x/Lm P3 Hn2( fl)
LA'max.

d=j(m —Jm3#2m) dC(/_I2n#7n2_n)


To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

dmax.— WL3
- ____

/W
A C B BA
L
b .f- c—.J
Web L°'
MA
'P MA7tS
MA=— MA =— d2-c2)(29c2_c12)
AI
J2 (o - n 2)
-____
L R8=r5—
= f (n'-n-pi.e) Where and r3 are the simple
support reactions for the beam
(MA being considered positive)
a, d=J42/n3(3n?-6b5
When x a.
dj[2p4—p3n (n-6n,'.8)+
_________ pn2(jn2_8n #6)]
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1095

PROPPED CANTILEVERS

W/2 W/2
1'

CU BA
—f— a -H
L
MA

if = % then between B and U.


MAZ_ —H X/L = m
1027SLf x
Al,. = ftx4xa(4 -Sm+Zm9]
+M,n0x. when x=I'4Jm#2rn2, Mx(2Qm'—27m#7)
A 7WL '''max.—
—20)
=0 67L]
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

'9:4

RB
Vx i(9_rn2)
= (2i."#soi. 42) A PB =

p8 = (2L2_30L#40')

- OO0/WL3
El
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

When x=0•598L

W/? W/2

-L
RB
MA

3WL
MA
#MmaxO0454 WL
[When x 0 283L]

=4W
T RB = * p_L
A- /3W
B - 32

- 0•0047WL'
dmax (I —ii
When x=O•447L When x=0404L
Steel Designers' Manual - 6th Edition (2003)

1096 Bending moment, shear and deflection

PROPPED CANT/LEVERS

a
_____
L
C BA
A a _____
iR8 L A5
___S___
I
0•577b
0/28 Wa
MA
MA
_________________ JL
BQt we en
CandA, Mx=R5.x-j!('x-b) M5=R5.x -
=_ j9(si.2_j,2)
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

(JaL/5aL#2OL2)
#Mmox when xb# fI/-

___ RAt NJR5


A8 =('5L —a) = 9 (si!—,/)
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

W P5 A8 = (t+SaL2)
2W
2W W
__________
C
____________ -
____ a " b—' 4—a______
.. L b____I
L R3
H0577aH 0/28 Wa
f—x —
Wab
H042Jb—
MA

When ,n=a/L 2
3m#2 I
MC=AB.b MA=_f/0L2-3b2)
Between AandC RA I

=RAW52/a2
_______________
}.—x--H \ IRS
Between C and B \
N
I

Vx=RA_Wrc24Z

A8 = '('/SL —4a) A3 = [L(//L -/S,i#(5L- a)]


RA=W-RB RA=W-RB
Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1097

PROPPED CANTILEVERS

W.iw4,
fAA 51 CD
1. L

I Wa2
M5=—2M— F

c-'3'L
8TN
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

.jwap

d0 WL4f2(ep#4?q#dp3(p#i)J
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

- Omax. —
54(1

IJA
'1. L
8+ CD
1.
M
'1
84 'C 0

—-Pa M3 -2M4 =—M


=

4I
q
Tfl
- 8 =-

— dmax I0'O
Ido
= j4p2,pq#Jp #Jq) a#b)#a2(2# VJ
d,n0x '-dmax"
27(1
Steel Designers' Manual - 6th Edition (2003)

1098 Bending moment, shear and deflection

PROPPED CANT/LEVERS

W/2

+
AB na/L *7/q
Wa

MA _—4(2L_a) MA=— f-(4L—3a)


This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

W'?Qn X< a,
tl =(9,r?x —/2nx# /2x— 4xg2)
+Mmax occurs whQn

A r—\
NB
A -(4L?#ZaL— az) (4L2 #4aL—JoT
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

R8 W—
W
I w

a —+—b —+-—a —4
L kO.4/SLtq
—L——-——--R8

MA7Z
I—

MA

#MmaxOO948 WL MArn (5L2+40L—4G2)

A4fl R, RAI\ I '?&

32 A Bj A477(2/1!#4aL 4a2)
/ A8

Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1099

PROPPED CANT/LEVERS

!F051
coIr4oleraQnt of parabola

SWL — SWL

M .('/Om-ZOm#7m) Mx =(-4om+8om-om4I7m)
#My,OO888WL,when X #Mmax*O0399WL,whQn x —
0 •3965L 02343L
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

A6
7W
B—
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

MA

MAW-' lz' ,nax.MA-O./93PL


wMn baO577L
M=('Z-# p)maxA1-0./74PL
1•• —
A5

F9 7PL
7(%
dmax'OOOPJ2
ft3
A3=f3b#2L) RAPR5
Steel Designers' Manual - 6th Edition (2003)

1100 Bending moment, shear and deflection

PROPPED CANT/LEVERS

Pp P

'4
ACDE
¼+V4+L/4 +-/4
MA MA

MA = — ISPL

D "E7
/7PL JJPL
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

PAL
4P 2P P AJ3 J2

Lj
dx 00152
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

dmax00209 zr

P P P P PP P
AC D EB* .4C DEFBA
L/L/3 Lb
MA MA

/9PL
MA=— 48
M0 -.96
M ....SJPL
£ 288

dxOQ/69 0nczx. — O.OZdS


Steel Designers' Manual - 6th Edition (2003)

Bending moment, shear and deflection 1101

PROPPED CANILEVEPS
pppp P PPPP Pp
D E F19 forces B
L/4tL/4tL/44L/
MA

JSPL
MAE7ZTJ
PL(,,?l)
MA=— -— MA= 8n
Ad /S7PL
rT(— 5/2
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

P8
RAEL (5n2-4n-i)
RB = (34z_4fl#/)
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

whennis
GO•OZZl !or9e, dmax. /85(1
Any symmetrical load W

JA ____
I.
AreaR L AreoS'B
a=L
MA
VAr000 X
M-
If AArea of free B.MDiagram ® _r—
a>O423L

cO423L
a< O423L
A8
M1 9(2_6n#sn')
P82
=!—
L MCA (2— 6n# Pn?3ni
-Sn
It'B
dmax occurs at point corresponding L
to Xon M diagram, the area A In Case!, R — SM/ZL
being equal to the area 0 CaseS, A= M/L
Vmax = Area SXx
Steel Designers' Manual - 6th Edition (2003)

1102 Bending moment and reaction

EQUAL SPAN CONTINUOUS BEAMS


UNIFORMLY DISTRIBUTED LOADS
Moment = coefficient x W x L
Reaction coefficient x W
where W is the UD.L. on one span only and L is one span

W\ —OO6J
0•096
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007

W, —0.100 VA —0100 VA
ooeo O•OZS o 0080 0
o — — o
W 0.0s0 W
0•101 0101
o 0 0 0
o.o5o W\ 0.o50
0075
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/

—0•117 W\
0073 0054
o "II 0

0072 0.06/ 0•098


ol
—0036 W. —Q./07
______________ 034
0054 0056

00P4
VA _______________
— 0049 W1 —0054
0
6
#00/4
0074
0 01 0
Steel Designers' Manual - 6th Edition (2003)

Bending moment and reaction 1103

EQUAL SPAN CONTINUOUS BEAMS


CENTRAL POINT LOADS
Moment coefficient x Wx L
Reaction = coefficient x W
where W is the Load on one span only and L is one span

1W
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/
Steel Designers' Manual - 6th Edition (2003)

1104 Bending moment and reaction

EQUAL SPAN CONTINUOUS BEAMS


POINT LOADS AT THIRD POINTS OP SPANS
Moment = coefficient A' Wx L
React/on = coef/cIenf x W
where Wis the total boo' on one span only a L is one span
'Yz Wiz Wk "lz
This material is copyright - all rights reserved. Reproduced under licence from The Steel Construction Institute on 12/2/2007
To buy a hardcopy version of this document call 01344 872775 or go to http://shop.steelbiz.org/
Appendix E
Useful Design Data

42/01012/07/44062 GHD Structural Design Manual


Revision 2
Appendix F
Soil and Rock Description

42/01012/07/44062 GHD Structural Design Manual


Revision 2
Appendix G
Prestress Design Notes

42/01012/07/44062 GHD Structural Design Manual


Revision 2
Contents

1. INTRODUCTION
1.1 Introduction
1.2 History of Prestressed Concrete
1.3 Definitions

2. REVIEW OF PRESTRESSING PRINCIPLES


2.1 Objectives of Prestressing
2.2 The Basic Principles of Prestressing
2.3 Applications of Prestressing
2.4 Prestressing Methods

3. PRESTRESSING COMPONENTS
3.1 Prestressing Steel
3.2 Prestressing Ducts
3.3 Live End Anchorages
3.4 Dead End Anchorages
3.5 Coupling Anchorages
3.6 Flat Jacks
3.7 Prestressing Companies

4. GENERAL DESIGN THEORY


4.1 Design for Flexure
4.2 Design for Shear
4.3 Deflection of Prestressed Members
4.4 Prestress Losses
4.5 Anchor Block Design
4.6 Continuous Beams
5. PRESTRESSED STRUCTURES
5.1 Post-tensioned Slab Systems
5.2 Pre-tensioned Floor Systems
5.3 Industrial Floors
5.4 Tanks

6. DESIGN DETAILING
6.1 Introduction
6.2 Post-tensioned Slabs
6.3 Multi-strand Post-tensioning

7. CONSTRUCTION
7.1 Construction Operations
7.2 Safety Precautions
1. INTRODUCTION

1.1 Introduction

Many people incorrectly consider it important to view prestressed concrete as a distinct


technique, and not purely as an extension of reinforced concrete to overcome certain
technical difficulties, thereby widening the field of the economic application of concrete.
There is no doubt that the growth in prestressed concrete that has existed over the past will
continue. It is therefore important for designers to become more familiar with the design
principles of prestressed concrete and to realise that the concept is not as complex as is often
believed.
This course is aimed at those engineers who are engaged in reinforced concrete design and
have a good knowledge and understanding of the behaviour of reinforced concrete elements.
Upon completion of the course, one should feel equipped to design a post-tensioned
structure, and be confident that the design is correct and economical.

1.2 History of Prestressed Concrete

The first person clearly to define the purpose of prestressing, as inducing compressive forces
in the tensile zone of a beam, was the Austrian engineer J. Mandl in 1896. His aim was to
reduce the tensile stresses in the concrete under load, and thus utilise the strength of the
concrete as much as possible. The idea was further developed by the German engineer
Koenen.
In these proposals, the loss of prestress due to elastic shortening was taken into account, but
the effects of shrinkage and creep were ignored, and mild steel was tensioned to the stresses
permitted for reinforced concrete. However, this initial tensioning stress was to low and
shrinkage and creep soon reduced the prestressing force to zero.
Freyssinet claims 1904 as the year when prestressing was born. He resolved to substitute
“permanently acting forces for the elastic forces developed in the reinforcement”. In 1912
Freyssinet encountered difficulties due to creep when he built the bridge at Veurdre; but he
was unable to follow up the problem before 1926. From this investigation he concluded that
high-strength steel should be used for tensioning; this in turn led to effective prestressing by
the introduction of “permanently acting forces”.
In the 1920’s, RH Dill in the USA recognised that high strength wire could be used to produce
a satisfactory prestressed member. The first engineer to use the properties of high-strength
strands initially in construction, before Freyssinet, was Torroja in 1925. Another pioneer, the
French engineer Coyne successfully employed tensioned high-strength steel cables for
anchoring dams in 1928. However, only Freyssinet persevered with his idea of introducing
prestressing.
Freyssinet carried out the first successful practical designs in prestressed concrete in Europe
in the 1930’s, when the time-dependent creep and shrinkage behaviour of concrete was
better understood. After World War 2, Freyssinet designed a number of successful and highly
acclaimed bridges in France, which led to wide acceptance of prestressed concrete.
1.3 Definitions

By definition, a prestressing system is one that will pre-compress the member in such a way
that when it is subsequently subjected to load, the tensile stresses induced by the applied
loads are either eliminated or counteracted to a desired degree.
There are two basic forms of prestressing, namely ‘pre-tensioned’ and ‘post-tensioned’. In
both cases, a prestressing force is usually applied by means of tendons, which comprise
high-strength wire, strand or bar. The term ‘post-tensioned’ is frequently misapplied in the
same manner as the term ‘pre-tensioned’. In the correct sense, these terms should be applied
only to the steel, and not to the member or the concrete.
q In pre-tensioning, the tendons pass through the mould, or moulds for a number of
similar members arranged end to end, and are tensioned between external anchor
blocks, by which the tension is maintained while the concrete is poured. When the
concrete has hardened sufficiently the ends of the tendons are gradually released
from the anchor blocks. During this operation, known as transfer, the force in the
tendons is transferred to the concrete by bond stress. The length required at each
end of a member to transmit the full tendon force to the concrete is termed the
‘transmission length’.

q In post-tensioning, the concrete member is cast incorporating ducts for the tendons.
When the concrete has hardened sufficiently, the tendons are tensioned by jacking
against one or both ends of the member, and are then anchored by means of
anchorages that bear against the member or are embedded in the concrete.

The level of prestress adopted in the design can have a significant influence on the cost.
Frequently the most economical design is achieved by using limited prestressing, which does
not prevent cracking under the full service design load, but which is sufficient to control
deflections and crack widths and hence achieve good service load behaviour. Hence,
depending on the level of the applied prestress force, the member can be classified as being
either ‘fully prestressed’, or ‘partially prestressed’.
q In a fully prestressed member, there is sufficient prestress to ensure that there is no
cracking of the concrete under the applied service load. In the British Standard, this is
classified as either a ‘Class 1’ type structure if there is zero tensile stress, or ‘Class 2’
if the tensile stress is limited so that there is no visible cracking of the structure.
q In a partially prestressed member, tensile cracking is allowed to occur, with crack
control being provided by bonded tendons and often supplementary reinforcement. In
the British Standard, this is classified as a ‘Class 3’ type structure.

‘External post-tensioning’ refers to the practice where the prestressing tendons are placed
on the outside of the physical cross section of the structure. The prestressing forces are
transferred to the structure at the anchorages and at ‘deflectors’. The use of external post-
tensioning has major applications in the design of long-span bridges, and in the strengthening
of existing structures.
2. REVIEW OF PRESTRESSING PRINCIPLES

2.1 Objectives of Prestressing


The primary objective of prestressing is to offset or counteract the stress, which the
subsequent loading will produce, thereby extending the range of stress to which the member
may safely be subjected.
Prestressing offers a number of advantages; the main ones are:
i. Unlike reinforced concrete, prestressed concrete can be kept free of cracks, thereby
improving its durability. This is particularly advantageous in corrosive atmospheres,
aggressive ground conditions and marine structures; and its watertightness in liquid
storage tanks and containers.
ii. Less steel and concrete are used, reducing the load on columns and foundations,
and leading to longer spans and fewer supports.

iii. Post-tensioned floors can save construction time by permitting early removal of
formwork and reduced back propping.
There are virtually no disadvantages, though there are certain restrictions. Prestressing
requires high quality throughout – of design, detailing, materials and workmanship.

2.2 The Basic Principals of Prestressing

In the design of reinforced concrete beams, it is assumed that the tensile strength of the
concrete is negligible, and the tensile forces created by the bending moments are resisted by
reinforcement. Cracking and deflections are essentially irrecoverable. The reinforcement
exerts no forces on the member on its own account.

In prestressed concrete, the primary purpose of the prestressing is to apply a force to the
concrete either by bond or by means of special anchoring devices. Hence, the whole of the
concrete can be made to act structurally. The prestressing steel is used actively to preload
the member and cracking and deflections are recoverable.
However, it should be noted that under overload conditions, as soon as the flexural tensile
strength of the concrete has been overcome, prestressed concrete behaves in a similar
manner to reinforced concrete, and at the ultimate load, the tensile and compressive
resistances required to withstand these conditions are the same for both the reinforced and
prestressed members.

2.2.1 Equivalent Load Concept


The effect of the prestress on a concrete element can best be visualised using a
procedure called the ’equivalent load method’. Using the concept, all the forces
exerted on the concrete member by the prestressing tendon can be determined. The
forces occur at the anchorages and wherever the cable changes direction.
Most post-tensioned beams have tendons with curved profiles. The figure below
shows a free body diagram of a small segment of a curved cable.
Over the length ∆x, the change in slope of the cable is ∆θ. It can be shown that, for
equilibrium of the segment, the force per unit length is equal to:
wp = Pκ
where κ is the curvature of the cable at the point considered, and wp is the force
exerted by the cable on the concrete. This force acts radially, but for small angles can
be considered to act vertically with negligible error.
For a parabolic cable profile, the curvature is constant along the length of the tendon,
and therefore the cable exerts an equivalent uniformly distributed load on the
concrete given by:

where: h is the mid-span sag

At the anchorages, the prestressing cable exerts a concentrated axial load in the
direction of the tendon. If the anchorage is located eccentrically relative to the
centroid of the beam, then the horizontal component of the prestressing force will also
result in an applied moment.
Consider the case of a simply supported beam with a parabolic cable profile, with
eccentricity e1 at midspan, and anchorage end eccentricities eo above the centroid of
the beam. The resulting equivalent loads for the beam are indicated below:
2.2.2 Benefits of Prestressing
The benefits of prestressing can best be demonstrated by way of an example.
Consider two rectangular beams of identical dimensions. The first beam is a
reinforced concrete section with 3-N28 bars. The second beam is fully prestressed
with a single tendon, comprising 4-15.2 diameter strands. The area of the
reinforcement and prestressing has been selected to give the same ultimate bending
moment capacity for each member.
The key results for each of the beams, subject to an applied service state bending
moment of 200 kN-m, are illustrated below:

Ast = 1860 mm2 Asp = 572 mm2


dna = 140 mm dna = 387 mm
∆ fsp = 6 N/ mm2
2
fst = 251 N/mm
Ιeff = 1574 x 106 mm4 Ιeff = 4167 x 106 mm4
φ Mu = 304 kN-m φ Mu = 304 kN-m

From the above results, it is evident that the prestressed concrete section requires
less than 1/3 the amount of steel compared with the equivalent reinforced concrete
section. It must, however, be recognized that the difference in initial cost is not
proportional to the difference in weights of the steel. High-strength steel is required
for the prestressing steel, and thus the unit cost is higher than for the reinforcement in
the concrete beam.
Another key benefit of prestressing is evident - the effective moment of inertia of the
partly cracked prestressed section is almost 3 times that of the reinforced section, as
the majority of the concrete section is effective for the prestressed element. This
results in a substantially stiffer element, with a corresponding reduction in deflections
and potential cracking.
2.3 Applications of Prestressing
Bridges and floor slabs account for a large percentage of prestressed structures, but there are
many other general and specialised applications that are sometimes overlooked.
In bridge construction, the advantages are obvious. For large spans, low rises and small
structural depths which would be impractical in reinforced concrete can be achieved, whilst for
shorter spans, construction depths are less so that there is improved appearance, lighter
foundations and greater durability.
The prestressing of building frames and floor slabs achieves considerable reductions in dead
loading with savings in foundation costs; reduces the number of columns, giving more
flexibility in planning, reduces cracking; and reduces the deflection of beams and slabs.
The prestressing of marine structures gives a longer life, due to the higher durability of
prestressed concrete. Railway sleepers are now almost entirely of prestressed concrete; and
tanks and other liquid containers, eliminates cracking, seepage and leaks. Pipes, poles,
masts, towers and other tall structures are often axially prestressed in order to resist bending
in any direction.

Specialist applications include the stressing of industrial ground slabs, roads and runways to
prevent cracking. In nuclear power station construction, the reactor pressure vessels are
typically prestressed because of increased safety characteristics.

By controlling the amount of prestress, a structure can be made either rigid or flexible without
affecting its ultimate strength. A flexible structure is obviously more resilient and can absorb
considerable energy before failure due to impact and thus show improved performance under
seismic and dynamic conditions. Examples of such structures include fender-piles of wharfs
or jetties. Alternatively, a very rigid structure is more suitable to resist heavy vibrations.

2.4 Prestressing Methods


The permanent stresses imposed on the concrete can be produced in several ways. The
most appropriate methods for concrete structures, subjected to bending or tension, are those
in which tensioned steel wires, strands, cables, or bars are used.
As noted in Section 1.3, there are two basic forms of prestressing, namely pre-tensioned and
post-tensioned.

2.4.1 Pre-tensioned Concrete


In the pre-tensioning process the tendons are stressed by direct forces, either:

a) Against abutments (ie. anchor blocks cast in the ground);


b) Against struts extending over the length of the casting bed underneath the member
to be cast, or
c) Against the form itself (ie. the mould within which the concrete is to be cast).

The procedure is best shown diagrammatically:


Pre-tensioned concrete is most appropriate to factory manufacture, in which a permanent
casting bed is used for repetitive applications. Factory batched concrete can be handled
directly to the forms resulting in tighter control of concrete quality. In addition, permanent
cranage for the removal of completed members and formwork is possible.

Casting beds are commonly set up employing the abutment method, where the
abutments are placed a substantial distance apart, thus permitting a number of similar
units to be cast end-to-end around continuous strands tensioned against the anchorages
at each end of the casting bed. This is known as the “long line” process.
The prestressing wires or strands pass through the ends of the casting bed formwork and
end abutments. The strands are then stressed and the forms subsequently filled with
concrete. After the concrete has attained the specified strength, the strands are gradually
released at the abutment positions and the strands are severed at the ends of the
member. The prestress is transferred into the member by the elastic shortening of the
concrete.
During this operation, known as transfer, the force in the strands is transferred to the
concrete by bond stress. The length required at each end of a member to transmit the full
tendon force to the concrete is termed the ‘transmission length’.
In the manufacturing process, the frictional weight of the concrete member on the casting
bed may prevent longitudinal shortening, until the member is lifted. During this period,
shrinkage or thermal cracking may occur unresisted by the prestress since the member
has not shortened and is effectively not prestressed. Similarly, in lifting the member from
the casting bed, the initial lifting force may cause cracking. Special care must therefore be
taken to ensure that the bond between the concrete and the casting bed is readily broken.
Pre-tensioned strands, by nature of their being stressed before concreting, must be
straight between points of support. Where a profiled strand is required, it is possible to
deflect the strands using ‘hold-downs”. These can be attached to the soffit, or thrust down
from above. In order to avoid high friction forces developing at the change in direction at
the hold-downs, rollers or shoes can be used.
It should be noted that the use of deflected strands is not commonly adopted in Australia.

2.4.2 Post-tensioned Concrete


Post-tensioning can be utilised for either precast or cast-in-situ construction. In either
case, the stressing of the tendons is carried out after the concrete has been cast and
attained the required strength. The prestressing force is transferred to the concrete by
means of end anchorages.
There are many systems for post-tensioning, a number of which are described in Section
3. The prestressed strands cannot be bonded to the concrete before tensioning or the
attainment of the required elongation would not be possible. The tensioned steel may be
individual wires or strands, cables composed of separate wires or strands, or alloy bars,
and may be placed in ducts cast into the concrete, or outside the concrete altogether.
For post-tensioned work, complete in-situ form and pour construction is generally
adopted. However, there may be advantages worth considering from the following
alternative methods of construction:

q Precasting. Repetitive beams or planks, especially when required for constricted


or remote construction sites, are well suited to this form of construction.
Transportation must be considered, including lengths, widths and weights.
Commonly, special arrangements with traffic authorities are required, and close
attention to clearances and wheel loading along possible transport routes. Where
these conditions can be satisfied, precasting can be a viable option.

q Composite Construction. A row of parallel precast prestressed beams or planks


are erected and designed to act as a permanent formwork, after which an in-situ
deck is applied. This is a frequently used system in buildings, and is quite
competitive with the alternative in-situ floor systems. This form of construction is
particularly appropriate to a uniform, repetitive floor layout.

q Segmental Construction. Precast lengths of beam are lifted into place and
supported on temporary falsework, after which joints are poured in-situ. Joints
may be either spaced and grouted, using special grout, or adjacent surfaces may
be match-cast, and coated with an epoxy glue immediately prior to offering the
surfaces into contact. In either case, the consequences of imperfect matching
must be allowed for, and the effects of possible loss of prestress must be
considered.

q Incremental Launching. Where intermediate supports are costly or impractical,


this may be attractive. Lengths of section are cast adjacent to, and in line with
proposed span, and then jacked forward out of the forms, allowing the next
section to be cast. Sections are progressively prestressed to achieve structural
integrity. Stressing patterns must allow for the stress reversals which occur at the
various stages of launching, especially the strong cantilever action which occurs
at the early and subsequent stages of construction. In the early stages of such
work, the overall stability of the advancing structure must be ensured.
3. PRESTRESSING COMPONENTS
There are now available a wide range of prestressing systems. Three main factors govern the
choice of the system: the magnitude of the required prestressing force, the geometry of the
section (space available for the tendons) and the cost.
The tendon is the basic element of a post-tensioning system. It comprises one or more
strands, constrained at both ends by an anchorage, and encapsulated within a duct.

3.1 Prestressing Steel


3.1.1 Strand
In Australia, the two most common types of strand are both made of 7-high tensile wires
tightly twisted together in a similar manner to a steel rope. They have a nominal overall
diameter of 12.7 or 15.2 mm.

The wire is typically low relaxation, stress relieved, super grade. The term low relaxation
refers to how much the strain relaxes in the strands under constant load. The term super
grade refers to the strength of the wire itself.

The vast majority of strands used are 12.7mm diameter. The use of 15.2mm strand is
normally used for large beams, or slabs with a span greater than about nine or ten
metres. Although 12.7mm diameter strand has a higher strength per unit weight when
compared to 15.2mm strand, the latter is more economical where it leads to a reduction in
the number of tendons and a subsequent reduction in anchorage costs. In many
situations this may not be possible, particularly for shorter span slabs, where the
maximum tendon spacing is typically limited to 8 times the slab thickness.

The most common form of tendons for prestressed structures is:

Material type Nominal Area Minimum Characteristic


and standard diameter (mm2) breaking load Strength
(mm) (kN) (MPa)
5 19.6 33.3 1700
Wire – AS 1310
7 38.5 65.5 1700
9.3 54.7 102 1860
7-wire super
12.7 100 184 1840
strand – AS
15.2 143 250 1750
1311
23 415 450 1080
Bars – AS 1313
29 660 710 1080
(super grade)
32 804 870 1080
38 1140 1230 1080
CHARACTERISTIC STRENGTH OF COMMONLY USED WIRE STRAND AND BAR
3.1.2 Stress Bars
Stressbar is most suited to lengths less than 10 metres. It is a rigid and self-supporting
system making it suited to vertical applications. The rolled thread ends eliminate draw-in,
which is most important for short tendons.

A range of diameters is available to give a wide selection of tendon forces.


The prestressing force is anchored at the end of the bar by a rolled thread, nut, washer
and bearing plate. Where necessary bars cam be joined with threaded couplers, and
clevis fittings may be used where pin connections are required.

3.2 Prestressing Ducts

Post-tensioned tendons are encapsulated within the concrete in a duct that is usually
manufactured in corrugated steel with a wall thickness of between 0.3 and 0.5mm. The ducts
are normally supplied in 4-6 m lengths, and are then coupled on site.

For enhanced corrosion protection and fatigue resistance of the tendons, the use of
corrugated high strength polyethylene ducts is recommended.

The ducting can be installed with or without the tendons in place prior to concreting. If the
tendons are not installed prior to concreting, it is recommended that temporary stiffeners,
such as flexible polythene tubing, be installed to ensure any accidental damage or slurry
penetration will not restrict later installation of the tendons.
For a post-tensioned member, the ducts are grouted with a cement-based grout after the
stressing operations are completed. Grouting has two major objectives:
i. To protect the steel against corrosion; and
ii. To effect the bond between the prestressing steel and the concrete.
Care must be taken to ensure that grout tubes are provided, and venting allowed for ensuring
that all the air within the ducts will be removed during the grouting process.

3.3 Live End Anchorages

In post-tensioned systems, the prestressing force is transferred to the concrete through


bearing type anchorages. A specialist prestressing contractor usually supplies the anchorage
device as an integrated proprietary item.
The following details illustrate the “Structural Systems” post-tensioning system, and are
typical of most proprietary systems.

3.3.1 Mono Strand Slab Systems


The post-tensioned slab system is used in shallow structures. It comprises up to 6-
strands placed in a flat duct with corresponding flat anchorages. All the mono strand
systems utilise a barrel and wedge to transfer the strand force to the casting and in turn to
the concrete.

3.3.2 Multi Strand Systems


The “Structural Systems” multi strand system consists of up to 61-12.7mm or 55-15.2mm
strands to form a single tendon. The individual strands are anchored in a common anchor
head with a wedge grip system and the strands are simultaneously stressed using a
hydraulic jack.

Details of the Structural Systems ‘Type M1’ anchorage are illustrated below:

As the number of strands increases, then so too must the size of the jack.
Several other ‘Live end’ anchorage types are also available. The type of stressing
anchorage and tendon configuration selected will vary depending on the application.
Note must be taken of the physical dimensions of the anchor. These dimensions are
dictated by the load capacity of the tendons and are not restricted to a proprietary system.
Care must also be taken in preparing any design to ensure that the particular jack
required for the system chosen can actually be inserted in the space provided.
Refer to specialist prestressing supplier brochures for details of the full range of available
anchorages.

3.4 Dead End Anchorages

Dead end anchorages are embedded in the structure and so take advantage of the bond
between the strand and concrete. Most systems comprise either the ‘bulb-type’’ (onions) or
the ‘swage-type’ anchorage.
The ‘bulb-type’ anchorage is an economical system suitable for any number of strands. It
transfers the force from the tendon by a combination of the bond of the strand and the
deformed bulb.

3.4.1 Mono Strand Slab Systems


The following details illustrate the “Structural Systems” mono strand ‘dead end’
anchorage system, and are typical of most proprietary systems.

The 'swage' type anchor is suited to cases where the tendon force must transfer to the
concrete right at the swage plate. The strand must be debonded between the swage plate
and the duct.
The ‘bulb’ type anchor provides a more gradual transfer of the stressing force into the
concrete. It is also a ductile anchor.
3.4.2 Multi Strand Systems
The “Structural Systems” multi-strand ‘dead end’ anchorages are designed to the same
principles as the mono-strand anchorages, varying only in size and number of strands as
illustrated below:

3.5 Coupling Anchorages

Couplers are used to give continuity to the tendons, which due to their length or the
construction methodology cannot be installed or tensioned as one unit. The first-stage of the
tendon is stressed and anchored in the normal way and the dead end of the second-stage
tendon is then assembled around it.

Couplers are expensive but their cost can be offset against savings in formwork,
reinforcement and sliding bearings required with an expansion joint. Construction joints in
prestressed slabs would normally be no further apart than about 40 metres.

Standard coupler components exist for all mono-strand slab tendons, and a full range for all
multi-strand systems.

MONO STRAND COUPLER MULTI STRAND COUPLER


The location of the coupler is typically selected at approximately 0.2 to 0.25 of the span
length. The coupler itself is located at the centroid of the cross-section to avoid any bending
moments from eccentricity of the anchorage.

3.6 Flat Jacks

The compressive force required to prestress concrete need not necessarily be produced by a
steel tendon, but can be obtained by the application of an external force. One way of applying
such a force is the “flat jack”, devised originally by Freyssinet, which is relatively cheap and
simple. Large numbers can therefore be employed where required, even if they are not
recoverable. The amount of movement obtainable, however, is limited.
In principle, the flat jack consists of a flattened metal canister that is inserted into a joint in the
structure and inflated as required by a fluid under high pressure. The shape may be elliptical,
although circular jacks are more usual. Rectangular jacks with rounded corners are also
available.

3.7 Prestressing Companies

There are several proprietary prestressing systems in use in Australia. Details of the major
prestressing companies are indicated below:
1. Structural Systems
2. VSL Australia Pty. Ltd.
3. Austress Freyssinet
4. GENERAL DESIGN THEORY

4.1 Design for Flexure

All prestressed concrete members must be designed to ensure an adequate degree of safety
and serviceability at all relevant limit states. The usual approach is to design based on the
most critical limit state and then to check that the remaining limit states will not be reached.

It is important to ensure that the design is not, however, totally controlled by one limit state
alone, as this may suggest an uneconomical design. The design should attempt to optimise
all aspects of the structure. This can only be achieved if the designer has a total
understanding of the member’s behaviour, and is able to progressively refine the design to
ensure an optimal solution.

The two main limit states are as follows:

i. Ultimate Limit State :

The member must have sufficient strength to support, without collapse an “ultimate
load”, which should represent the greatest load that might occur under the most
unfavourable conditions.

ii. Serviceability Limit State :


a) The deflection under serviceability limit states should not adversely affect the
appearance or efficiency of the structure.

b) The cracking of the concrete should not adversely affect the appearance or
durability of the member.

In assessing the likely behaviour of a prestressed concrete member, flexural cracking must be
limited. The requirements of AS3600 Clause 8.6.2 are noted below:
4.1.1 Behaviour of Uncracked Section
In the uncracked range of behaviour, concrete stresses remain relatively small, so that the
short-term response of the concrete to stress changes is nearly linear and elastic. Under the
effect of permanent stresses due to prestress and dead load, considerable creep occurs in
the concrete and significant additional long-term deformations occur.
Consider a post-tensioned beam with a curved cable profile.

At midspan it has an applied force ‘P’ at an eccentricity from the centroid of the cross-section
of ‘e’. The extreme concrete fibre stresses can be calculated using simple linear elastic
theory. For the combined effect of the prestress and an applied moment ‘M’, the resulting
stresses are given by:

For the uncracked section, the section properties are normally based on the gross concrete
sections. The effects of reinforcing steel can be included by determining the section
properties based on a transformed section, but this is generally not necessary.
For statically determinate members, the longitudinal stresses due to prestress depend only on
the magnitude and the eccentricity of the prestressing tendon. This is not necessarily true for
statically indeterminate beams, due to potential hyperstatic reactions. This is discussed in
Section 4.6.
In the uncracked section there is negligible change in magnitude of the prestressing force as
the external load is applied. This property is beneficial when fatigue resistance is a design
consideration. A fully prestressed design is thus particularly appropriate for structures
subjected to large repeated live loading.
4.1.2 Behaviour of Cracked Section
As the applied moment increases above the cracking load in a bonded prestressed member,
cracks develop and extend progressively into the section. This behaviour is different to that of
a typical reinforced concrete beam without prestress. In the latter case, the cracks form
suddenly and immediately extend well into the section. As the applied moment increases, the
lever arm remains almost constant, with the increase in moment capacity resulting from an
increase in stress in both the concrete and reinforcement. For prestressed concrete
members, significant increase in moment capacity occurs through the progressive increase in
the lever arm, as opposed to an increase in force in the tendon.
The general flexural theory for a cracked prestressed section is illustrated below:

Under the action of prestress alone, prior to the application of any external moment, the strain
in the concrete at the tendon level is εce (compressive), and the strain in the tendon is εpe
(tensile). The applied moment induces a tensile strain εcp in the concrete at the level of the
tendon. Assuming perfect bond, the same change in strain must occur in the tendon. The
resulting strain in the tendon is therefore:

Similarly, the strain in the reinforcement can be expressed in terms of the top fibre concrete
strain and the neutral axis depth as follows:

For normal design applications in the working stress range, the concrete stresses are
primarily in the lower linear region of the stress-strain curve (stresses less than 0.45 f’c). For
the above section, the compressive force in the concrete is given by:

These expressions allow the internal forces to be determined. By equating the total
compressive and tensile forces within the section, a relationship between εo and dn is
obtained. For an assumed value of dn, the corresponding internal moment of resistance can
be determined and compared against the applied moment. An iterative trial and error
approach can then be used to determine the correct neutral axis depth.
4.1.3 Ultimate Limit State
As the applied moments increase, the concrete’s performance becomes non-linear. The
stress in the tendon, which increases steadily after cracking, now increases more rapidly, the
depth of the neutral axis reduces, and the extreme concrete fibre stress increases. At this
stage, a prestressed member behaves similar to a reinforced concrete section.
The ultimate capacity is reached when the tendon is stressed to its ultimate strength, or when
the compressive strain capacity of the concrete is reached. For under-reinforced beams,
failure is initiated by yielding of the tendon. The associated large tensile strains result in
increased crack widths and the neutral axis depth reduces. The increased concrete stresses
acting on the reduced compressive area ultimately result in a compression failure of the
concrete. The large steel strains produce visible cracking and considerable deflection of the
member before the failure load is reached.
In an over-reinforced member, failure occurs when the compressive strain limit of the
concrete is reached. This type of failure occurs suddenly with little warning.
The general flexural theory for determining the ultimate capacity of a prestressed concrete
member is similar to that described in Section 4.1.2 for a cracked section.

For normal strength concrete up to 50 MPa, AS3600 adopts the value εu = 0.003. It is
possible to construct the concrete compressive stress block by using a linear variation in
strain above the neutral axis and assuming a suitable stress-strain relationship. It is found that
the calculated moment capacity is not very sensitive to the parameters chosen to define the
shape of the concrete stress block.
In AS3600, use is made of the equivalent rectangular stress block, in which the nonlinear
stress distribution is replaced by a uniform stress of magnitude 0.85 f’c which acts over an
area bounded by the edges of the section and a line parallel to the neutral axis located at a
distance γ dn below the extreme compressive fibre.

In assessing the stress in bonded tendons at ultimate, AS3600 Clause 8.1.5 permits the use
of a conservative, empirical expression to evaluate the stress σpu:

where:
While the above equation may lead to a simpler calculation, in some circumstances a more
accurate analysis may be necessary, requiring an iterative trial and error approach similar to
that described in Section 4.1.2.
In most practical situations the actual stress-strain curve for the prestressing steel can be
better approximated by an idealised bi-linear, elastic-plastic distribution in which the ‘yield’
stress for the tendon is taken as the 0.2% offset stress.

Where greater accuracy is required, the following equation provides a smooth transition from
the elastic curve to the strain-hardening curve, and closely represents the actual stress-strain
curve.

Where the exponent ‘n‘ can be chosen to properly simulate the actual transition from the
elastic part of the curve to the strain-hardening part after yielding.

The ‘trial and error’ approach for determining the ultimate flexural capacity of a prestressed
concrete beam has been adopted in the GHD ‘Excel’ computer program “Prestressed Beam”.
This program uses the latter more accurate stress-strain curve for the prestressing tendons,
as well as a non-linear stress-strain relationship for the concrete.

4.1.4 Effect of Un-bonded Tendons


The above analysis assumes that the tendons are fully bonded to the concrete. For ungrouted
post-tensioned tendons, slip occurs between the strand and the concrete, and consequently
the strains in the strand tend to equalise over significant lengths of the beam. This reduces
the peak stress levels that would otherwise occur at cracks. The local increase in strain in the
tendon at peak moment regions will therefore be considerably less than the value (εce + εcp)
which applies to a bonded tendon.
The value of the reduced strain is dependent on numerous factors, such as the cable profile
along the beam, and the amount of friction between the tendon and the duct. Even where
testing has been carried out, the results cannot confidently be extrapolated to similar
structural systems. Various conservative approximations are therefore introduced to allow the
tendon stress to be estimated at the ultimate section capacity.
In assessing the stress in un-bonded tendons at ultimate, AS3600 Clause 8.1.5 permits the
use of a conservative, empirical expression to evaluate the stress σpu.

Other than in the design of major bridges, the use of un-bonded tendons is not generally
recommended or adopted in Australia.
4.2 Design for Shear

Shear failure is an ultimate load condition. Its influence on the behaviour of a prestressed
member at service load is not usually significant. The types of cracking that occur in a
prestressed concrete beam at ultimate load are illustrated in the figure below.

In regions of negligible shear, flexural cracks occur at the extreme tensile face when the
tensile strength of the concrete is exceeded. These cracks, located in region ‘1’ in the figure
above, extend vertically into the section.

In the adjacent region ‘2’, both the shear force and the bending moment are large, and the
cracks initiate as vertical cracks in the tensile face. As they extend upwards with increasing
load, they become inclined under the influence of the shear force. These cracks are referred
to as ‘flexure shear’ cracks.
In region ’3’, where the shear force is high but the bending moment is relatively small, inclined
cracks initiate in the web of the beam if the principal tensile stress reaches the tensile
strength of the concrete. These cracks are referred to as ‘web-shear’ cracks and are most
likely to develop in beams with thin webs and high prestress.

4.2.1 Web-Shear Cracking


As noted above, web-shear cracking occurs in regions previously uncracked in flexure. The
typical distribution of compressive stress and shear stress in a thin-web member is illustrated
below:

The direct stress due to the prestressing force and applied moment varies with depth. The
value of the tensile stress is determined as follows:
As the section is uncracked, distribution of shear stress across the section can be calculated
using the classical equation:

The resulting principal tensile stress can be determined using the Mohr circle, and is given by:

The shear force required to produce shear cracking can then be determined by equating the
maximum principal tensile stress to the tensile strength of the concrete. In AS3600 the
principal tensile strength of the concrete is taken as:

The corresponding shear force Vt is found by setting the principal tensile stress (determined
using the Mohr circle equation) equal to the concrete tensile strength.
The maximum principal tensile stress does not necessarily occur at the centroidal axis, where
the maximum vertical shear stress occurs. For Ι-sections, the junction of the flange and web
is frequently the critical location. It is therefore necessary to determine the stresses at
numerous depths within the section.

Within the section, the direct stress ‘σ ‘ depends on the value of the applied moment, which in
turn is related to the shear force. A trial and error calculation is therefore required to
determine the value of Vt. This involves an iterative procedure in which the applied loads on
the beam are proportionally increased (or decreased), until the corresponding maximum
principal tensile stress within the section, determined using the above equations, reaches the
principal tensile strength.
The above process will be demonstrated by way of any example during the two-day course,
using the ‘excel’ spreadsheet program “Shear Capacity of Prestressed Beams”.
The ultimate shear strength at web-shear cracking is given in AS3600 as:

Vuc = Vt + Pv
where: Pv = the vertical component of the prestressing force at the section

The force Pv usually adds to the shear resistance provided by the concrete. As the slope of
the tendon is liable to be near its maximum value in regions prone to web-shear cracking, the
contribution of the vertical component of the tendon force can be significant. In unusual
cases, the vertical component of force may subtract from the shear resistance of the
concrete.
4.2.2 Flexure-Shear Cracking
In the zone of the beam subject to ‘flexure-shear’ cracking, inclined cracks develop as an
extension of the vertical cracks initiated by flexure. In AS3600, the resistance to ‘flexure-
shear’ fracture is assumed to comprise of three components, which add together to give an
estimate of the total shear resistance:
§ the vertical component of the prestressing force in the tendon, Pv, as described in
Section 4.2.1.
§ a crack initiation force, which is that value of shear force which corresponds to
the bending moment that would cause a crack to open at the section, it being
assumed that the crack already exists. The bending moment adopted for this
case is taken as the de-compression moment, Mo, and the shear force
corresponding to this moment, Vo. For the case of a simply supported beam, the
value of V can be taken as Vo = [Mo / M * ] V *.

§ A crack propagation force, which is the additional shear force after cracking until
failure of the section occurs. This is taken to be equal to the shear resistance
determined for an equivalent reinforced concrete beam, allowance being made
for both prestressed and non-stressed steel.
The ultimate shear strength at ‘flexure-shear’ cracking is given in AS3600 Clause 8.2.7 as:

4.2.3 Ultimate Shear Strength


The ultimate shear strength of the prestressed beam, excluding the contribution of the shear
reinforcement, is taken to be not greater than the lesser of the values determined previously
for ‘web-shear’ cracking and ‘flexure-shear’ cracking, unless the section is cracked in flexure,
in which case only the latter ‘flexure-shear’ capacity applies.
Where the resistance Vuc is inadequate, or where the depth of the section exceeds 750mm,
shear reinforcement is introduced to augment the total resistance of the section. The sizing
and spacing of stirrups, and the location and anchorage of longitudinal steel, follow the same
procedures as for reinforced concrete beams. However, for prestressed slabs, or for wide
shallow beams, shear reinforcement may only be required when shear forces are high.
The contribution to the ultimate shear strength of the beam by the shear reinforcement is
given in AS3600 Clause 8.2.10:
§ For perpendicular shear reinforcement:
§ For inclined shear reinforcement:

The overall design strength of the beam is taken as:

φ Vu = φ (Vuc + Vus)

The above analysis assumes that the shear reinforcement yields before crushing of the
concrete in the web occurs. An upper limit Vu.max is therefore set on the value of Vu, as given
in Clause 8.2.6:

4.3 Deflection of Prestressed Members

[Not Yet Written]


4.4 Prestress Losses

The initial prestressing force applied to the live anchorage is transmitted along the tendon, but
decreases as a consequence of instantaneous and long-term losses.
The losses can be classified as follows:
q Instantaneous Losses
These losses are due to:
a) Friction along the duct;
b) Draw-in of the anchorage wedges
c) Elastic shortening of the concrete

q Long Term Relaxation Losses

These losses are due to:


a) Shrinkage of the concrete
b) Creep of the concrete
c) Relaxation of the steel

4.4.1 Friction Losses in the Duct


During tensioning operations, friction between the tendon and the duct results in a loss of
prestress along the length of the tendon, increasing with distance from the jack. These
losses only apply to post-tensioned systems.
Losses due to friction are made up of two components:
i. Wobble friction – due to unintentional variations in the duct. This applies to both
straight and profiled tendons;
ii. Friction due to curvature of the duct – this friction loss is dependent on the angle
turned through and the coefficient of friction.
The prestressing force Px at any distance ‘x’ from the jack is given by:

– µ (α tot + Β Lx)
Px = Po e

where: µ = coefficient of angular friction (radians -1)


α tot = sum of the absolute values in radians of successive
angular deviations of the tendon over a distance Lx
Β = wobble coefficient in radians per metre.
Lx = the length of the tendon from the jacking point [m]

The friction coefficients depend on various factors such as the condition of the duct inner
surface, the condition of the strand external surface and the tendon layout.
4.4.2 Draw-in of the Anchorage Wedges
A loss of prestress occurs when the load is transferred from the stressing jack to the
anchorage of the tendon. When the strands are ‘locked off”, there is a small slip (or draw-
in) of the strands through the jack before the anchorage wedges “bite” into the strand and
hold it firm. The slip is normally about 6mm (dependant upon the jacking system used).
The movement in the strand is in the opposite direction to the tensioning process and is
opposed by the same friction mechanism as opposed the initial tensioning, (ie. µ and Β
are the same). Hence the friction loss profile is effectively modified by a “mirror image”
profile near the jack (ie. the slope of the ‘draw-in’ line is the same as the friction loss line,
but in the reverse direction).
The energy dissipated in the strand is given by:

KE = δ Ep A p

The point where the friction overcomes the ‘draw-in’ can be determined by equating the
energy dissipated in the strand to the area enclosed by the two friction loss lines.

4.4.3 Elastic Shortening of the Concrete


When the prestressing force is transferred to the concrete, the stress produced in the
concrete is accompanied by a corresponding elastic strain. The reduction in stress in the
tendon associated with this strain of the concrete is the loss of prestress due to elastic
shortening.

For pre-tensioned members the loss of force in the tendon at transfer is given by:

∆P = Ap σcp . Ep / Ec

where: σc = stress in the concrete at the level of the tendon

For post-tensioned members, all the tendons are not stressed simultaneously. Assuming
that all tendons experience a uniform shortening and are stressed one after the other in a
unique operation, the loss of force in the tendons can be determined using the following
expression, where ‘n’ is the number of tendons:

∆P = n – 1 . Ap σcp . Ep / Ec
2n

For a single tendon, all the elastic shortening occurs while the tendon is still attached to
the jack. No loss to prestress due to elastic shortening therefore occurs. For several
tendons, the loss tends to half the product of the modular ratio and the stress in the
adjacent concrete.

In most cases the loss due to elastic shortening is relatively minor compared with losses
from other sources.
4.4.4 Losses due to Concrete Shrinkage
As the concrete shrinks with time so the prestressing steel undergoes a corresponding
shortening which results in a loss of the prestressing force. With pretensioned members
the loss is based on the total ultimate shrinkage strain, while for post-tensioned members
only the shrinkage strain that occurs after stressing is taken into account.

The loss in prestress in the tendons due to shrinkage of the concrete is given by:

∆P = Ep . εcs . Ap

The design shrinkage strain ( εcs ) can be estimated from shrinkage data provided in
AS3600. The shrinkage is dependent upon the environment in which the member is
drying and the exposed surface area of the member from which moisture is lost (as
measured by the hypothetical thickness, th). For normal strength concrete, AS3600
suggests that the design shrinkage at any time after the commencement of drying can be
estimated from:

εcs = k1 εcs.b

where εcs.b is a basic shrinkage strain which in the absence of measurement, can be
taken to be 850x10-6; k1 describes the development of shrinkage with time and is
determined by interpolation from Figure 6.1.7.2 in AS3600.

If a prestressed member has reinforcement distributed throughout the cross-section,


Clause 6.4.3.2 allows the design shrinkage strain, εcs, to be reduced by dividing by
(1+15As/Ag ) for the purpose of estimating the loss of prestress.

4.4.5 Losses due to Concrete Creep


Concrete under stress continues to undergo increasing strain with time. This shortening
strain results in a corresponding loss of prestressing. The magnitude of this loss varies
with the age and strength of the concrete.

In the absence of more detailed calculations and provided that the sustained stress in the
concrete at the level of the tendons at no time exceeds 0.5 f’c, the loss in prestress in the
tendons due to creep of the concrete is given by:

∆P = Ep . εcc . Ap

The design creep strain ( εcc ) can be estimated from :

εcc= φcc ( σci / Ec )


where: φcc = the design creep factor
σci = the sustained stress in the concrete at the level of the centroid of the
tendons, calculated using the initial prestressing force prior to any time-
dependent losses and the sustained portions of all the service loads.
In accordance with AS3600, the design creep factor can be estimated from:

φcc = k2 k3 φcc.b

In the absence of tests, the basic creep factor (φcc.b ) may be taken as the values given in
Table 6.1.8.1. The factor k2 depends on the hypothetical thickness, th, the environment
and the time after loading and is given in Figure 6.1.8.2(A). The factor k3 is a function of
the age at first loading and may be determined from Figure 6.1.8.2(B) in the Standard.

4.4.6 Losses due to Tendon Relaxation


A limited amount of strain relaxation occurs in the tendons under the continued high level,
long-term stress in the strands.

In the absence of detailed test data, AS3600 Clause 6.3.4 allows the basic relaxation (Rb)
of a tendon, for Australian manufactured material, to be taken as 1% for low relaxation
wire, 2% for low relaxation strand, and 3% for alloy steel bars. The basic relaxation is the
assumed relaxation 1000 hours after stressing, for a prestress level of 0.7 f’c, and an
average temperature of 20oC.

The design relaxation of the tendon is determined from:

R = k4 k5 k6 Rb

where, k4, k5 and k6 are coefficients to account for the duration of the prestressing force,
the level of the stress in the tendon, and the average annual temperature.

As the force in the tendon will gradually reduce due to the loss in prestress from the long-
term effects of creep and shrinkage in the concrete, the long term relaxation of the
strands is modified as a percentage of theses losses.

The loss in prestress force due to tendon relaxation is given by:

∆P = Ptransfer . k4 k5 k6 Rb (1 – [∆Pshrinkage + ∆Pcreep] / Ptransfer )

where: Ptransfer = average force in the tendon immediately after transfer


(i.e. after all immediate prestress losses)

4.5 Anchor Block Design

In pre-tensioned members, the tendon forces are transmitted by bond between the tendon
and the concrete. In a post-tensioned member, the forces are transmitted through bearing
type anchorages. These mechanisms are entirely different, and are considered separately. In
both systems, highly concentrated forces occur, and these must be safely dispersed into the
structure.
The most critical time in the life of an anchor is during stressing. The concrete strength at
transfer is usually specified as less than the full design compressive strength for the member,
and the highest stresses occur during prestressing, which is before any deferred loss of
prestress occurs.

4.5.1 Post-tensioned Anchorages


In post-tensioned members, the force in the tendon is transferred to the concrete by steel
bearing plates or proprietary anchorages. At these points there are regions of high
compressive stresses immediately behind the anchorage plate, which in turn result in high
tensile stresses as the force distributes itself into the member.
The stresses are influenced by numerous factors such as the size and shape of the
anchorage, the relation between the area over which the load is applied and that of the
anchorage, and the proximity of other anchorages. The stress distribution in the end block
is complex, but can be demonstrated as indicated below:

The possible modes of failure in the anchorage zone are:


i. ‘Bearing failure’ of the concrete beneath the anchor plate.
ii. ‘Bursting failure’ (or splitting) where the tensile stress along the axis of the
member is not sufficiently restrained.
iii. ‘Spalling failure’ of the surrounding concrete, at edges and surfaces, as a result of
strain deformations and high local tensile stresses in the concrete.
In order to accommodate the tensile stresses, adequate reinforcement must be provided
in both the vertical and horizontal directions.
A considerable amount of research has been carried out on the stress distribution in the
end blocks of post-tensioned members, and this has led to the derivation of empirical
design procedures as well as numerous analytical procedures.
Rules for the design of end zones are given in Section 12.2 of AS3600. They are
restricted to cases having a maximum of two anchorages either vertically or horizontally.

q Stresses Directly Behind the Anchorage

The average stress behind the anchor plate, allowing for a deduction for the
trumpet size, is calculated at the time of transfer. The allowable value for this stress
is given in AS3600 Clause 12.3 as follows:

where: A1 = bearing area of the anchorage


A2 = the largest area of the bearing surface
geometrically similar and concentric to A1

Cast-in proprietary anchors will generally transfer higher forces than external plates
of similar area due to the transmission of forces along the anchor. Proprietary
anchorages are generally designed on the basis of full-scale tests in accordance
with AS1314. The tests often indicate that the code limit is conservative.

q Bursting Stresses
In the end anchorage zone of a post-tensioned member, the concentrated
anchorage force is dispersed until a linear stress distribution is achieved across the
overall cross-section. The length of the dispersion zone is approximately equal to
the width or depth of the member. As the longitudinal compressive stresses fan out
in curved trajectories away from the anchorage, a bursting tensile force is produced
perpendicular to the tendon axis.
Maximum bursting usually occurs directly behind the anchorage, and for a
prestressed member with a single anchorage, can be represented as follows:
The transverse stresses occurring along the line of the anchor force vary with the
distance from the anchor head. These stresses are both compressive and tensile
as shown, forming a moment couple.
The tensile bursting force can be calculated using the empirical formula given in
AS3600 Clause 12.2.4, which has been shown to give reliable results:

T = 0.25P (1 - kr)

where: P = maximum force at the anchorage at transfer


kr = the ratio of the depth, or breadth, of the anchorage
plate to the corresponding depth, or breadth, of the
symmetrical prism.

The variation of T / P with kr is shown below:

The formula varies slightly from experimentally derived results in regions of low kr,
but this is not significant in most practical cases.
For multiple anchorages, the same principle used for a single anchorage can be
adopted by dividing the member into a series of symmetrically loaded equivalent
prisms. Each prism is then treated individually using the empirical formula noted
previously.

The design of the anchorage zone should also consider the possibility of the
tendons being stressed separately. During the stressing operation higher splitting
forces can occur than in the final condition with all anchors stressed. The design
must allow for the corresponding anchorage / symmetrical prism geometry at the
time of stressing. It may be necessary to specify a sequence of stressing in order to
avoid placing additional reinforcement in the anchorage zone.

q Spalling Stresses
For anchorage zones where the tendon is highly eccentric to the member
centreline, or for multiple anchorages spaced widely apart, tensile ‘spalling’ forces
may occur at the loaded concrete surface. This surface tensile cracking between
anchorages is additional to the bursting stresses that occur directly behind the
anchorage.
The magnitude of the tensile spalling force can be calculated using a concrete
beam analogy that considers a free-body of the anchorage zone. This approach is
described in more detail below.

q ‘Transverse Moment Method’ of Analysis


In this design approach, the anchorage zone is represented as a free-body. The
force in the tendons is distributed over the depth of the anchorage plate, and is
resisted by a uniform or trapezoidal stress distribution remote from the anchorage
zone. The stresses remote from the anchorage zone are linearly distributed over
the depth of the section, and are determined using:

The integration of the stress distribution is equal to the total force in all the tendons.
Using the free-body, one can determine the bending moment diagram through the
depth of the member. The bursting moments are taken as positive, whereas the
spalling bending moments are negative. The bending moment is divided by a
suitable estimate of the ‘lever-arm’ depth to obtain a corresponding tensile force.

The lever-arm dimension, based on experimental evidence, is calculated as follows:


a) For Bursting Moments:
§ For a single anchorage located symmetrically in the end zone, the
lever-arm is taken to be 0.5 times the overall depth of the member.

§ For multiple anchorages or eccentrically located single anchorages,


the lever arm is taken to be 0.5 times the depth of the symmetrical
prism. For a rectangular end block with multiple anchorages, the
lever-arm can be simplified and taken as D/2n, where D is the overall
depth of the section and n is the number of anchors.

b) For Spalling Moments:


§ For a single eccentric anchorage, the lever-arm is taken to be 0.5
times the overall depth of the member.
§ Between pairs of widely spaced anchorages, the lever-arm is taken
as 0.6 times the spacing of the anchorages.

The foregoing method provides an easily understood design approach, and is


easily adapted to unusual anchorage layouts and non-rectangular cross-sections.
q Detailing of Anchorage Reinforcement
Reinforcement is provided to resist the entire calculated transverse tensile forces.
AS3600 limits the reinforcement stress to 150 MPa in order to control crack widths.
It is recommended that only deformed bars, not exceeding 20mm diameter, be
used in the end zone.
The ‘bursting’ moment reinforcement should theoretically occupy the zone of
maximum tensile stress, but is generally distributed uniformly from 0.2De to 1.0De
from the loaded face in order to ease congestion. The bar size and spacing
determined for this zone should be continued to the loaded face.

In multi-strand systems, it is common to weld closed links together to ease


congestion caused by tension laps.
Anchorage reinforcement requirements for slabs have been established through
testing. The usual recommendation by the prestressing companies is to have a
continuous cage of reinforcement running around the perimeter of the slab. The
cage consists of 4-N12 bars, one in each corner, and 2 or 3-N10 wire ties,
depending on the number of strands, to either side of the anchor as shown in the
figure below. The N12 bars have the added benefit in that they provide the spalling
moment resistance at the edge of the slab between anchorages.

q Surface Tensile Stresses


The figure below shows that tensile stresses occur near the surface of an end block
and at the corners. Although the maximum value of these tensile stresses can be
relatively high, the resulting tensile force is generally not significant, as they exist
over a limited area. It is recommended that the surface tensile force be taken as
0.04P, where P is the sum of the prestressing forces applied to the anchorages.
Suitable nominal reinforcement should be provided to resist the above forces, and
to prevent corners from cracking off. The reinforcement should continue around
corners to the top, bottom, and side faces.

q Spalling Forces in T-Beams


For T-beams and box girders transverse tensile stresses must be investigated in
both the webs and slab. Generally all the post-tensioning is applied to the web. The
portion of the prestressing force carried by the slab is transferred from the webs to
the slab through shear stresses. The anchorage zone stress distribution is relatively
complex, but can be simplified using the ‘Transverse Moment Method’ of analysis
as illustrated below:
4.5.2 Pre-tensioned Anchorages
In pre-tensioned elements, anchorage is achieved by mechanical interlock, and direct
frictional contact between the strand and the concrete. At transfer, the tendons shorten;
and the prestress force is transferred into the concrete within a length Lp called the
‘transmission length’ or ‘development length’.
In the absence of substantiated test data, the transmission length is given in Clause
13.3.2 of AS 3600:

Some de-bonding occurs with time over a short length from the end of the strands / wires,
and it is therefore assumed that a completely unstressed zone of length 0.1 Lp develops
at the end of the tendon.

Where the strand or wire is untensioned, the development length is taken as not less than
1.5 times the value given in Table 13.3.2. AS 3600 does not permit the use of plain wire
for pre-tensioned work (Clause 19.3.1.3).

4.6 Continuous Beams


5. PRESTRESSED STRUCTURES

5.1 Post-tensioned Slabs

Prestressing offers larger spans with reduced structural depth, resulting in larger column-free
areas. Positive deflection and crack control and, if necessary, crack-free watertight slabs offer
the designer the opportunity to overcome some of the limitations of reinforced concrete and
structural steel.

5.1.1 Column Layout


The column layout can have a major influence on the structural efficiency of the floor
system. Although the layout may often be dictated by other than structural considerations,
it is important to be aware of the best layout for a post-tensioned slab.

A column grid spacing of between 8m and 10m usually results in the most economical
structure while maintaining architectural flexibility.
It is generally uneconomical to vary the prestress level in every span. If the following
points are given consideration then a constant amount of prestress may be used
efficiently throughout the length of the structure with resultant cost savings:
§ End spans approximately 80% the length of interior spans;
§ Cantilevers approximately 30% the length of interior spans;
§ Expansion joints, where necessary, located at ¼ points within a span;
§ Tendon length should be limited to a maximum of 60 metres;
§ The use of tendons shorter than 8 metres is generally not economical;

The above points are not always possible, but will generally lead to greater economy
where used.

5.1.2 Slab Profiles


The three most common floor systems used for building structures such as offices,
shopping centres and carparks are the flat plate, flat slab and band beam.

q Flat Plate
The recent increase in formwork price has encouraged the use of flat plate structures.
They are best suited to square grids with a span of 7m to 9m. The increase in
concrete, reinforcement and prestress must be offset by the formwork cost savings.
The depth of a flat plate is generally dictated by punching shear requirements.
Thinner slabs or longer spans can be achieved by providing shear heads.
Loadings Span/Depth Prestress Reinforcement
Carpark – 2.5 kPa 38 – 42 1.6 – 2.0 MPa 3 – 4 kg/m2
Retail – 4.0 kPa 35 – 38 1.6 – 2.0 MPa 4 – 5 kg/m2
q Flat Slab
These are more efficient than flat plates for spans greater than 8m. Drop panels are
typically 30-35% of column grid. This is a widely used system as it maintains a flat
soffit, simple formwork, as well as flexibility for services reticulation.
The drop panels increase the flexural stiffness of the floor as well as improving its
punching shear capacity. The system provides the thinnest floors and can lead to
overall floor height reductions. The economic span range is up to 12m.
Loadings Span/Depth Prestress Reinforcement
Carpark – 2.5 kPa 42 – 47 1.8 – 2.2 MPa 3 – 4 kg/m2
Retail – 4.0 kPa 38 – 42 1.8 – 2.4 MPa 4 – 5 kg/m2

q Band Beam and Slab


This is the most common form of construction representing approximately 70% of all
post-tensioned floor solutions. It is well suited to buildings with variable spans and
rectangular grids. In most instances, the width of the band beam is chosen to suit
standard formwork sizes. The sides of the band beam can be either square, or
tapered.
The widths of band beams should be standardised and importantly, the width of the
slab between the bands should be selected as a multiple of 1200 / 1800mm to suit
the standard formwork sheet widths.
The band beam has a relatively wide, shallow cross section that reduces the overall
depth of the floor while permitting longer spans. The post-tensioned tendons should
not be interwoven, leading to simpler installation.
The band beam depth allows for simple expansion joint details.
Loadings Span/Depth Prestress Reinforcement
Beam Slab
Carpark – 2.5 kPa 28– 32 36- 38 1.4 – 1.8 MPa 4 – 5 kg/m2
Retail – 4.0 kPa 26– 28 35- 38 1.6 – 2.0 MPa 5 – 6 kg/m2

5.1.3 Load Balancing


Control of deflections is an important design requirement. With conventional tendon
layouts, the draped tendon profile induces deflections in the slab that are in the opposite
direction to the deflections resulting from the self-weight and applied loads. By varying the
amount of prestress and tendon profile, the magnitude of the prestress-induced
deflections can be selected to ‘balance-out’ any desired proportion of the applied loads.
The selection of the load to balance is an important factor in the overall cost. Selection of
too high a load to balance may incur high prestressing costs reducing the overall
economy of the solution. A combination of a lower level of ‘balanced load’ and the
addition of normal reinforcement at peak moment regions may be more economical.
Generally, the loads to balance indicated in the following table will usually result in an
economical solution:
Building Partition & Design Live Load to Balance
Occupancy Superimposed Dead Load
Loads

Office Buildings 0.5 – 1.0 kPa 3.0 kPa (0.8 – 1.0) x swt
Shopping Centres 0.5 – 2.0 kPa 4.0 kPa (0.9 – 1.1) x swt
Residential 2.0 – 3.0 kPa 1.5 kPa swt + 50% of
Buildings (brick partition walls) partition loads

Car Parks Nil 2.5 kPa (0.7 – 0.9) x swt

For structures designed for relatively large partition and superimposed dead loads, care is
required at the time of stressing due to the absence of the partition loads. The problem
can be avoided by stressing in stages as the superimposed load is applied

5.1.4 Level of Prestress


The level of average prestress serves as a useful check on the economics of the design
and the serviceability of the structure:
Average Comments
Compression
Below 1.4 MPa Amount of compression is generally inadequate to resist
cracking, and is too low for waterproofing. The tendons will
probably not have adequate ultimate capacity alone and will
require supplementary reinforcement at both positive and
negative regions. Generally not economical.
1.4 – 2.4 MPa Typical design range for post-tensioned slabs. For typical
conditions, slabs stressed in this range will perform well with
respect to cracking and shortening. Use 2.0 MPa minimum for
watertight slabs. Will usually result in adequate ultimate flexural
capacity sagging regions, but may require supplementary
reinforcement in negative regions.
Above 3.5 MPa Recommended maximum value. Slabs stressed in this range
will undergo excessive shortening due to axial creep and
elastic shortening. Heavily loaded transfer beams would be in
this range.

5.1.5 Waterproof Slabs


Slabs that are required to be waterproof often act as roof slabs and are subject to high
temperature differentials because of daily exposure to direct sunlight on their upper
surface.
For a slab to be waterproof, the formation of shrinkage and temperature cracks must be
prevented. This can be done by providing the slab with a minimum residual prestress
level, after all losses, of 2.0 MPa in each direction. The tendons should also be distributed
uniformly across the width of the slab.
Additional reinforcement should be placed at the upper face of the slab to help resist the
additional stresses generated by differential creep and shrinkage between the slab and its
restraints and temperature differentials. It will also help to control early shrinkage cracking
prior to the application of the prestress. Common practice is to provide SL82 mesh in the
upper face of all slabs with exposed surfaces.

5.1.6 Tendon Layout


In determining a suitable tendon distribution, consideration must be given to ensuring that
there is no interference of the tendons crossing at the low and high points of the tendon
drapes. This is best achieved by layering the tendons to avoid ‘clashes’. The designer will
need to determine which cables will occupy the ‘primary’ layer, and which the ‘secondary’
layer early in the design stage as this affects the effective depth of the tendons, and
hence the capacity of the respective sections.
The distribution of the tendons across the design panel strip should be determined in
accordance with the requirements of AS3600 Clause 7.5.5. For continuous spans, 75% to
100% of the tendons should be placed in the ‘column strip’, with the remainder placed in
the ‘middle strip’.
Also, Clause 9.1.2 stipulates that in two-way flat slabs, at least 25% of the total negative
moment in the design panel strip, is to be resisted by reinforcement or tendons located in
a cross-section of slab centred on the column. The width of the cross-section is to be
taken equal to twice the overall depth of the slab or drop panel plus the width of the
column.

Although AS3600 does not specify a maximum spacing of tendons, the recommendations
of AS1481 Clause 4.13.2.5, that the tendon spacing not exceed 8 times the slab
thickness, is considered appropriate and should be adhered to.

5.2 Pre-tensioned Floor Systems

Precast pre-tensioned floor systems have many advantages over conventional post-tensioned
insitu suspended slabs.

When precast is being considered for any project it is essential that all facets ranging from the
detailing, manufacturing, erection and finishes are taken into consideration. This will ensure
that the final solution utilised satisfies not only the architectural and structural requirements,
but also economics and buildability of the project.

There are numerous precast pre-tensioned floor systems available in Australia. The systems
most currently in use are:

5.2.1 Hollow Core Planks

5.2.2 Ultrafloor

5.2.3 Transfloor
5.2.4 Solid Slabs

5.2.5 Tee-Beams

5.2.6 Beam Shells

5.3 Industrial Floors

[Not Yet Written]

5.4 Tanks

[Not Yet Written]


6. DESIGN DETAILING

6.1 Introduction

The design details adopted for any structure can have significant effects on the speed and
quality of construction. Detailing should be standardised and as simple as possible.
Congested areas, in particular, should be carefully assessed and appropriately detailed.
The design stage is the best time to allow for the site requirements.

6.2 Post-tensioned Slab Systems

The design of post-tensioned slab systems is covered in some detail in Section 4.1. The
following points should be considered in arriving at the final project detailing:

6.2.1 Formwork
Dimensions of beams and slabs should be based on standard formwork sheet sizes to
minimise formwork cutting. In particular, the distance between beams should permit
standard form ply sheets to be used (i.e. 1200mm / 1800mm module).

Changes in the level of the soffit form should be minimised as much as possible. Where
the slab is required to be thickened to accommodate a set down in the top of the slab, the
soffit of the set down area should extend to between the bands to simplify the formwork.
For wide shallow band beams, with a beam / slab depth ratio of less than 2, the side
forms should be vertical rather than sloping as this simplifies construction of the
formwork. For deeper band beams, the side forms are normally splayed to minimise the
potential shrinkage differential between the slab and the beam, which could otherwise
result in transverse cracking of the slab between the band beams.

6.2.2 Reinforcement
The most common on-site problem with post-tensioned structures involves clashes
between the reinforcement and tendons, which are unfortunately often only picked up on
site, potentially leading to delays. The following general recommendations should be
followed:
ii. The reinforcement should be detailed so that the bottom bars can be placed
first followed by the prestressing tendons, and then the top bars.
iii. The reinforcement should be detailed using standard bar lengths as far as
practicable. Ideally, bent bars and stirrups should be kept to the same size and
shape where feasible.
iv. Open ligatures should be detailed. This allows the tendons to be laid in position
following the placement of the bottom layer of reinforcement, rather than having
to be threaded through closed ligs. If closed ligatures must be used, it is
preferable to use a ‘U’-shaped lig with cogged ends, and a separate closer bar.
v. The bottom layer of slab reinforcement should not extend too far into the band
beam as this could clash with the tendons.
vi. Edge beams should not be to narrow. Transverse slab tendons (with
associated bursting reinforcement) can provide an obstacle to the longitudinal
tendons in the edge beam. Slab anchorages are typically up to 300mm long,
and consequently a narrow edge beam can be very difficult to prestress. A
minimum edge beam width of approximately 1000mm is therefore generally
adopted. Also, check that the top reinforcement does not interfere with the
anchorages.

vii. The reinforcement required in the anchorage zones must be carefully detailed
to avoid congestion, and to ensure ease of placement, particularly where it
coincides with reinforcement in columns and other beam reinforcement.

6.2.3 Tendons
Where possible, use only mono-slab systems throughout the project. Avoid the use of
tendons larger than 6-12.7mm or 5-15.2mm strands, as the larger jacks are difficult to
handle on site. Obvious exceptions apply to large transfer beams, where an
uneconomically large number of tendons would be required.
For overall economy, the designer should always attempt to minimise the number of
tendons by maximising the tendon spacing - the maximum tendon spacing should be
limited to 8 times the slab thickness. For less highly stressed slabs, tendons comprising
12.7 mm-diameter strands are generally more cost effective, provided they are spaced at
the maximum allowed. For thicker, more highly stressed slabs, it may necessary to
substitute 15.2 mm-diameter strands to maintain the maximum tendon spacing.

The cost of prestressing depends primarily on the length of the tendons. Short tendons
are relatively expensive in comparison with long tendons as the fixed costs per tendon,
such as the number of anchorages required and the stressing operation, which is
independent of the length, are prorated over a lesser tonnage. A practical slab and beam
length for single end stressing is approximately 35m, and for tendons stressed from both
ends, 60m.
In flat slabs, transverse tendons should be located below the longitudinal primary tendons
(ie. a 20mm reduction in drape). This avoids weaving tendons at the columns.

Tendons should not be located close to columns as they may clash with the column
reinforcement. This problem is often exacerbated where the column reinforcement
terminates in the slab, and the cogs fan out into the slab or band beam. Preferably, the
tendons should be located a minimum of 600mm either side of the column. This will
ensure that the tendons achieve their maximum drape over the line of the supports.
It is recommended that tendons always be independently supported and kept separate
from reinforcing. Where slab tendons cross wide ban beams, ensure adequate support is
provided to the tendon.
Standard anchorage recess sizes and dimensions should be detailed wherever possible.
Beware of clashes of tendon high points with holding down bolts and baseplates, or
services blockouts.
A minimum cover of 25mm should be specified to the duct – this prevents the concrete
from popping off when high-pressure grout is pumped into the ducts.

6.2.4 Concrete Specification


Concrete strengths of 25 or 32 MPa are normal, with 32 MPa permitting an earlier transfer
of the prestressing. The concrete mix should be tested to ensure that the shrinkage value
falls within the specified limits. A maximum shrinkage strain of 600x10-6, as per AS1012
test, is a suitable requirement for most structures.

6.2.5 Crack Control


Often a slab may have minimal continuous reinforcement additional to the prestress. In
these cases, the application of an initial prestress is required to minimise the effects of
initial shrinkage and plastic cracking of the concrete. In such conditions, it is normal to
stress all tendons to 25% of the jacking force at 24 hours after completion of the pour.
The second and final stage prestress is applied when the concrete has reached the
strength nominated in the design.
In banded slabs, the differential shrinkage between the band beam and the slab can
cause cracking. Cracks develop due to the faster rate of concrete shrinkage in the thinner
slab relative to the deeper band beam. The probability of cracking increases as the beam
/ slab depth ratio exceeds a value of 2.

Slabs intended to be watertight require particular care to eliminate cracks. Prestress


levels of 2.0MPa minimum and continuous top mesh reinforcement are required. For
banded beams, additional top reinforcement, between and parallel to the band beams, of
not less than 0.16%, should be provided.

6.2.6 Building Restraints


During the initial planning stage, ensure that the prestressed elements are not locked
between stiff elements. Ideally, stair cores and lift shafts should be located close to the
centre of the pour. If this cannot be avoided, then pour strips or movement joints may be
required.
In very long elements, the total shortening of the slab will affect the outer columns. This is
most significant in the ground to first floor columns. Consideration should be given to
reducing the stiffness of the columns, or providing expansion joints in the first floor to
reduce the length (and hence the amount of creep and shrinkage) of the first floor spans.

6.2.7 Joints
The required construction joint layout will be determined to a large extent by the area of
concrete the builder can finish in a day. Commonly, the maximum area is about 1200 to
1600 square metres. Construction joints are commonly formed using standard coupling
anchorages.
[To be Expanded]

6.2.8 Slab Penetrations


Penetrations through existing post-tensioned slabs require careful construction.
Whenever possible, it is recommended that the tendon locations be permanently marked
on the underside or top of the slab.
Any coring of a post-tensioned slab should be preceded by a pilot hole drilled to verify
clearance from tendons.
Future penetrations that have been included in the design should be located clearly on
the slab soffit (eg. a 3mm x 5mm recess around the perimeter of the future opening).

6.3 Multi-Strand Post-tensioning

For building elements with multi-strand post-tensioning, the detailing requirements can be
more onerous due to the increased number and size of tendons. This will generally result in
greater congestion of tendons and reinforcement, and therefore requires greater attention to
detail to ensure that potential on-site problems are prevented.

6.3.1 Minimum Spacing of Tendons


A contributing factor to the quality of the concrete is good and thorough vibration. It is
therefore essential that the tendons be arranged in such a manner that a normal vibrator
head can penetrate easily between the tendons and the secondary reinforcement without
dislodging or damaging the ducting.
In members with high duct curvature, the order of stressing needs to be carefully
considered to prevent a tendon from bursting through the intervening concrete into an
adjacent unstressed duct during stressing.

Since the stresses applied to the concrete immediately behind the anchorage bearing
plate are considerable, the arrangement of the bearing plates must provide a minimum
edge distance. The minimum required distance of the bearing plates to the concrete
edges, and to adjacent anchorage bearing plates is dependent on:
§ the post-tensioning force to be transmitted
§ the concrete strength
§ the bearing plate dimensions
§ the reinforcing steel behind the bearing plate
Specific requirements are generally provided in the technical literature supplied by the
various prestressing companies.

6.3.2 Anchorage Layout


It is normal practice with tendons less than about 30m long to stress from one end only,
using a fixed anchorage at the far end. Where a considerable number of tendons is
required, as in bridge decks, it is desirable to stress adjacent cables from alternate ends
in order to develop a more uniform stress pattern throughout the structure.

For tendons of greater length, it is normal practice to stress from both ends.

6.3.3 Cable Profiles


To ensure a smooth transfer of the prestressing force to the anchorages and concrete,
and not to exceed the estimated friction losses, it is necessary to maintain a smooth
tendon curvature at all locations. The various prestressing companies generally specify a
minimum straight length of tendon required adjacent to the anchorage, and the minimum
radii to which tendons should be profiled.
In the absence of specific details, the following information can be used as a guide:

A reduction of the minimum radius of curvature should only be contemplated in extreme


situations, and then the local reactions to the structure and the additional stresses in the
tendon due to the curvature, must be allowed for.
6.3.4 Cable Supports
In order to place the tendons exactly according to the design, rigid supports to the
tendons are essential. To avoid local deflection between the supports, they should be
spaced between 1.0 to 1.5m apart.

When only the duct is placed prior to concreting, supports must resist buoyancy forces.

6.3.5 Anchorage Block-outs


Until the stressing operation is completed, the stressing anchorages must be accessible.
Normally, this is achieved by providing block-outs in the concrete. After stressing and
grouting, the block-outs are filled with concrete, thereby providing a corrosion protection
for the anchorage.
The bearing plate, sleeve, and spiral form one assembly that is attached to the formwork
with nails or bolts, depending on the size of the anchorage. Attention must be given to the
requirement that the tendon axis be exactly perpendicular to the bearing plate.

6.3.6 Space Requirements for Jacks


Sufficient space must be available behind the anchorage to allow access for placement of
the jack, and for its stroke when stressing. In many cases, availability of access will be
restricted due to physical constraints or due to construction planning, and the layout of the
dead and live end anchorages will be influenced by these considerations.

Details of required clearances should be sought from the prestressing company.


7. CONSTRUCTION

7.1 Construction Operations

Post-tensioned concrete relies for its structural adequacy and serviceability performance on
correct detailing and on-site operations. This involves the interaction of all the structural
elements (concrete, reinforcement and post-tensioned tendons), and the timely execution of
the on-site activities. The timing of the stressing operations is important to the achievement of
the required construction cycle.

7.1.1 Construction Planning


Design and Detailing of Prestressing Systems
Key to the success of any project is that the design be considered in terms of
buildability as well as structural adequacy. As far as possible the systems and
layout of tendons should be arranged to minimise potential hazards. If
necessary, stressing pockets should be located in the top of the slab to
ensure adequate access to live end anchorages. Stressing from the
underside of the slab is no longer an acceptable practice.
In the case of a project fully designed and documented by the consulting
engineer, the prestressing contractor will normally provide workshop
drawings for the post-tensioning system, showing detailed tendon profiles,
profile laying sequences and jacking forces.

The forces at the anchorages are very high. Consequently, the concrete
strength is critical in this zone. For the same reason, the reinforcement tends
to be congested in the vicinity of the anchorages, which makes compaction
more difficult. Good reinforcement detailing by the designer, and careful
consideration, by the reinforcement fixer, of the sequence in which the
reinforcement is to be placed is important.

Planning of the Construction Work


The installation of the post-tensioning is a relatively straightforward process,
the main issue being to co-ordinate the site activity with other trades,
particularly steel fixing, to minimise the time spent on critical path activities.
Tendon prefabrication is often undertaken off the deck for this reason.
The prestressing works should be integrated in the general construction
schedule in a way that can minimise the hazards, such as stressing during
periods when other works in the affected area are at a minimum.

Personnel, Qualifications

Personnel working with prestressing systems should be well informed about


the hazards involved and the precautions to be taken.
The work must be performed under proper supervision and the personnel are
to be experienced in this kind of work, and familiar with all equipment to be
used.

7.1.2 Installing Tendons and Concreting


Prestressing Tendons
Prestressing tendons must be protected from corrosion. Strands and coils of
wire should be covered, and stored clear of the ground. The ducts and
completed assemblies are to be kept clean and dry during the construction
period by sealing the ends and all openings.
Prestressing strands must be inspected for corrosion, kinks, nicks etc. as this
can initiate fracture.
Prestressing steel is particularly sensitive to heat. Welding, burning, and
flame cutting operations must not be permitted near the tendons. It is also
possible to get local damage by sparking if the return current from an electric
welder is allowed to stray through prestressing steel on its way back to the
welding set.

Accurate alignment of the anchorages normal to the tendon is essential.


Anchor plates should be fixed to the end forms to prevent any movement
during concreting.

All connections between anchorages, ducting and grout connection must be


securely fixed and sealed. Tape is most often used for this.

Pre-pour Inspections
The consulting engineer should carry out a pre-pour inspection of the post-
tensioning system. This entails a detailed inspection of the prestressing and
reinforcement, and importantly the anchorage reinforcement. This should
ideally only be carried out once the site supervisor has checked the
installation against the workshop drawings and is satisfied that the works
have been completed, and is ready for a final inspection.

Concreting
Care must be taken during concreting. Ducts must not be damaged or
displaced, either by poker vibrators, or by concrete being dropped from an
excessive height.
It is important that the concrete, particularly in the anchorage zones, is
properly placed and compacted. The majority of problems encountered
during the stressing process occur due to inadequate quality control at this
stage.
If the strands have been installed in the ducts before pouring the concrete,
there is a risk of grout leaking into the duct and the tendon becoming
immovable thereby preventing stressing. In order to avoid this problem, it is
recommended that the duct be flushed with water in order to wash out any of
the cement paste that might have leaked into the duct. In addition, the
strands should be moved within the duct, within about an hour of the
concreting operation being completed. This movement will break any bonding
that might have commenced between the tendon and the duct as a result of
the ingress of cement paste.

Concrete test cylinders should be cast at the same time as pouring the slab /
beams, so that the strength can be checked before tensioning takes place.
These cylinders are additional to the normal quality control testing
requirements. These should be site cured using the same procedures
adopted for the structure. These concrete cylinders will need to be tested
before stressing of the tendons is permitted.

7.1.3 Preparations before Stressing


Equipment for Prestressing
All equipment, jacks, pumps, hydraulic pipes etc. must be thoroughly cleaned
and checked before use. Pressure gauges and load cells need to be
calibrated by jacking against a load cell.

Tensioning Procedures
Prior to commencing tensioning, check that the following items have been
completed:
i. The strength of the concrete has attained the specified value. This
can only be checked by testing the prepared concrete cylinders. Do
not assume that elapsed time since casting is a sufficient assurance
of adequate strength.

ii. The structural drawings should specify the sequence of stressing.


iii. Check whether tendons are to be tensioned from one or both ends.

Working Area

The area around the stressing anchors and passive anchors should be
fenced off from the general working area.
If the area behind the anchors cannot be adequately secured in any other
way, the necessity for erecting rigid screens as safety barriers should be
considered. An alternative is to fix an anchor-ring to the jack or structure by a
chain.

7.1.4 Stressing and De-stressing


Personnel

Stressing must not be commenced unless competent, experienced personnel


are present at both cable ends.
All personnel must have proper protective equipment - gloves, safety boots,
hard-hat and safety clothes.

Stressing Equipment

During the entire stressing operation, the equipment should be checked for
possible malfunctions, leakages, pipe failures etc. The jacks should be
cleaned, threads oiled, and protective wrapping removed just prior to use.

Working Area
No personnel are to be allowed to access behind the anchors from the time
the stressing has commenced, until the full load has been transferred to the
anchorage.

Stressing Records
Stressing records must be maintained for all tendons. These records are to
be prepared by the prestressing subcontractor, and should include details of
the stressing force, and measured extensions of the strands. If the actual
extensions deviate from those calculated by more than +10% of the
theoretical value, the designer should be informed. Care should be taken
when restressing (if required) to ensure that the wedges do not clamp at the
same point on the strand.

The projecting tendon ends must not be cut after the jacking has been
completed, until the consulting engineer has reviewed the stressing records,
and given permission.

7.1.5 Grouting Operations


Grouting

Grouting of the tendons in bonded structures serves two major objectives:


i. To protect the steel against corrosion. The life of the structure
depends essentially on the quality of this operation.
ii. To effect the bond between the prestressing steel and the concrete.

The grout must be homogeneous and therefore mixed mechanically, and


have less than 3% bleeding. Industry practice is to use an anti-bleed, methyl
cellulose organic additive in the grout mix to ensure duct filling, limit bleeding
and improve grout fluidity.

Grout Venting Requirements

When the tendon profile exceeds 600mm, such as in continuous structures,


each high point must be vented to release trapped air and water. Under
certain conditions, drain tubes should be provided at low points of long
tendons and for tendons with a very high drape.
When grouting draped tendons with several high points, the first vent is
closed when the group consistency is comparable to the inlet grout, and
pumping is continued until a good mortar flows from the next vent, which is
closed in turn, and so on until the opposite end of the tendon is reached. If
the tendon is very long and the pressure high, grouting can be continued at
one of the vents, after closing off the original inlet point.

7.2 Safety Precautions

The operations involved in tensioning and de-tensioning prestressing tendons are not
dangerous provided that sufficient care is taken. Prestressing steel, when tensioned, does
however contain a considerable amount of stored energy, which in the event of any failure of
the prestressing steel, anchorage, or jack, may be released violently.

[To be expanded ]
Appendix H
Water Retaining Structures – Details

42/01012/07/44062 GHD Structural Design Manual


Revision 2
42/01012/07/44062 GHD Structural Design Manual
Revision 2
42/01012/07/44062 GHD Structural Design Manual
Revision 2
42/01012/07/44062 GHD Structural Design Manual
Revision 2
GHD

Level 1 42 Sturt Street Townsville QLD 4810


PO Box 930 Townsville QLD 4810 Australia
T: (07) 4720 0400 F: (07) 4772 6514 E: tsvmail@ghd.com.au

© GHD 2010
This document is and shall remain the property of GHD. The document may only be used for the purpose
for which it was commissioned and in accordance with the Terms of Engagement for the commission.
Unauthorised use of this document in any form whatsoever is prohibited.

Document Status

Author Reviewer Approved for Issue


Rev
No. Name Signature Name Signature Date
A Hong Su Alistair Moat

B Hong Su Leslie J Millist


Satyajit Datar

0 Hong Su Satyajit Datar

1 Hong Su Satyajit Datar


Clint Spencer
(Editor Various)
Alistair Moat

42/01012/07/44062 GHD Structural Design Manual


Revision 2
(Various)
Keith Hungerford
(Structural
Systems)
Russell Olsen
(Design Loadings)
Richard Wallis
(Reinforced
Concrete)
Richard Evans
(Foundations &
Retaining Walls)
Gerald Lovell
(Columns & Walls)
James Yip
(Serviceability)
Paul Clarke
(Timber)

2 L. Smit C. Spencer
(Software,
Prestressed
Concrete)
J. Pautasso
(Concrete)
I. Round
(Standard
Technical Spec)
J. Yip
(Water Retaining
Structures)
R. Chohan
(Water Retaining
Structures)
S. Elward
(Water Retaining
Structures)
G. Tomlinson
(Water Retaining
Structures)
R. Paull
(Durability)
L. Milst
(FEA Guidelines)
P. Vatin
(FEA Guidelines)
S. Konstanty
(Dynamic
Foundations/Bin
Design/Material
Flow)

Y. Kumai
(Design Loading)
T. Kershaw
(Bin
Design/Material
Flow)

42/01012/07/44062 GHD Structural Design Manual


Revision 2

You might also like