Sei sulla pagina 1di 216

Formwork Handbook

D
R
AF

by

Stephen A Ferguson and Douglas W Crawford


T
PART VB COPYRIGHT STATUTORY LICENSE ELECTRONIC WARNING NOTICE

COMMONWEALTH OF AUSTRALIA
Copyright Regulation 1969
WARNING

This material has been copied and communicated to you by or on


behalf of Curtin University of Technology pursuant to Part VB of
the Copyright Act 1968 (the Act)

The material in this communication may be subject to copyright


under the Act. Any further copying or communication of this
material by you may be the subject of copyright protection under
the Act.

Do not remove this notice


Formwork Handbook

D
R
Disclaimer
AF
While every effort has been made and all reasonable care taken to ensure the accuracy of the material
contained herein, the Author(s) shall not be held to be liable or responsible in any way whatsoever and
expressly disclaims any liability or responsibility for any loss or damage costs or expenses howsoever
incurred by an person including but without in any way limiting any loss or damage costs or expenses
incurred as a result of or in connection with the reliance whether whole or partial by any person upon
any part of the contents of this material.
Warning
This publication should not be used without the services of a competent professional person with
expert knowledge in the field, and under no circumstances should this Publication be relied upon to
T
replace any or all of the knowledge and expertise of such a person.

ii 2013 Stephen Ferguson


Preface

Preface
The motivation for creating this handbook comes from a strong belief that the health and safety of
construction workers should not be put at greater risk from structural failure than other workers.
Unfortunately; in Australia and elsewhere, this is not the case. In practice, the frequency of structural
failure and the general risk of death is much higher during construction than, later, during the service
life of the completed permanent structure.
In the past, a higher frequency of structural failure may have been tolerated because of an underlying
tacit attitude in the design and construction industry that temporary structures, such as formwork, are
less important than permanent structures and therefore greater risks are acceptable. However, this is
no longer the case.
Investigations into the causes of construction failure agree procedural inadequacies enable flaws in the
design and/or construction to go undetected. Examples of procedural inadequacies include:
communication difficulties and confusion of responsibilities amongst participants; inadequate briefing
of designers; the lack of design drawings; inadequate checking of designs (particularly those
containing novel features); unapproved modifications of the initial design; or failure to inspect work
prior to loading. Although this research was carried out 30 to 40 years ago, formwork collapse and
D
failure continues to occur all too often and the findings are just as applicable today as then.
Implicitly, the higher frequency of structural failure also casts doubt on the competence of those
involved in the design and construction of formwork and falsework. This premise is supported by
researchers who interviewed those responsible for the design and construction of the majority of
falsework in the UK. The researchers found that:
R
(a) At all levels of the industry there is a lack of understanding of the fundamentals and basic
principles involved in achieving the stability of falsework;
(b) The contracting sector does not appreciate the implications of the assumptions made by
suppliers who design the majority of falsework systems; and
AF
(c) There is a lack of checking and a worrying lack of design expertise.
In 2002, the Standing Committee on Structural Safety (SCOSS) in the UK warned that there is a need
to carefully consider the means by which falsework is currently procured, designed, constructed and
supervised. Furthermore, that judging from the evidence it is only a matter of time before a serious
event occurs. As recently as 2010, SCOSS reported that these concerns remain pertinent.
There is no evidence in the literature that the issues identified in the UK apply to Australia.
However, the frequency formwork and falsework collapse over recent years support this view.
To address the shortcomings identified SCOSS recommended compiling courses that include elements
on procurement, statutory responsibility, and managing the supply chain as well as the technical issues
T
associated with design.
To this end, this text may help by providing guidance on procedures, requirements and methods for the
design and construction of formwork and falsework to comply with Australian Standards and Work
Health and Safety Regulations.
Acknowledgements
I would like to acknowledge the contribution and thank: the staff at Syntect Consulting Engineers
(Simon Johnson, David Webster and Nicholas Ho) for all their help in putting this handbook together;
Harry Backes and John Woodside for their comments; and Ian Gilbert for his help with Chapters 9 and
10. I also wish to acknowledge the contribution and counsel of my co-author Douglas Crawford
(Chairman of Australian Standards Committee BD-043 responsible for AS 3610) and the work of all
BD-043 committee members.
Special thanks must go to Eur Ing Peter F. Pallet who kept me to update with changes happening to the
Euro and British Standards and freely shared his expert knowledge and the wealth of information he
has published.

2013 Stephen Ferguson iii


Formwork Handbook

Contents
Chapter 1 Introduction
1.1 Formwork and falsework .................................................................................................................. 1.1
1.2 Formwork importance ...................................................................................................................... 1.2
1.3 Formwork design requirements ........................................................................................................ 1.3
1.4 Scope ...................................................................................................................................... 1.3
1.5 Application ...................................................................................................................................... 1.5
1.6 Glossary ...................................................................................................................................... 1.5
1.7 Notation ...................................................................................................................................... 1.7
Chapter 2 Safe formwork design and construction
2.1 Introduction ...................................................................................................................................... 2.1
2.1.1 Participants .............................................................................................................................. 2.2
2.1.2 Scope........................................................................................................................................ 2.2
2.2 Project designer ................................................................................................................................ 2.3
2.2.1 Project documentation ............................................................................................................. 2.3
2.2.1.1 Location and magnitude of design service loads ............................................................... 2.4
2.2.1.2 Sequence and timing of concrete placement ..................................................................... 2.4
D
2.2.1.3 Limitations on the magnitude and locations of construction loads ................................... 2.4
2.2.1.4 Loads from the effects of pre-tensioning or post-tensioning ............................................. 2.4
2.2.1.5 Limitations on the use of the permanent or existing structure for formwork restraint ...... 2.4
2.2.1.6 Minimum stripping times and procedures ......................................................................... 2.5
2.2.1.7 Method of multistorey shoring .......................................................................................... 2.5
2.2.1.8 Requirements for composite construction or permanent form systems ............................. 2.5
R
2.2.1.9 Information about the foundation ...................................................................................... 2.5
2.3 Construction contractor .................................................................................................................... 2.5
2.3.1 Construction documentation .................................................................................................... 2.6
2.3.2 Formwork coordinator ............................................................................................................. 2.6
2.3.2.1 Formwork design brief ...................................................................................................... 2.7
2.4 Formwork supplier ........................................................................................................................... 2.9
AF
2.4.1 Work, Health and Safety .......................................................................................................... 2.9
2.4.2 AS 3610 1995 and AS 3610.1 2010 ........................................................................................ 2.10
2.5 Formwork designer ........................................................................................................................... 2.11
2.5.1 Hazard identification, elimination or control ........................................................................... 2.11
2.5.1.1 Risk control measures ....................................................................................................... 2.12
2.5.2 Practical requirements .............................................................................................................. 2.12
2.5.3 Formwork documentation ........................................................................................................ 2.13
2.5.3.1 General requirements ........................................................................................................ 2.13
2.5.3.2 Specific requirements of AS 3610.1 - 2010 ...................................................................... 2.13
2.5.3.3 Proprietary documentation ................................................................................................ 2.14
2.5.3.4 Safety report ...................................................................................................................... 2.14
T
2.5.3.5 Multistorey shoring ........................................................................................................... 2.14
2.6 Formwork checker ............................................................................................................................ 2.14
2.7 Formwork contractor ........................................................................................................................ 2.15
2.8 Formwork Supervisor ....................................................................................................................... 2.15
2.9 Formwork Inspector ......................................................................................................................... 2.16
2.9.1 Approval to load formwork ..................................................................................................... 2.16
2.9.2 Approval to remove formwork ................................................................................................ 2.16
2.10 Obligation and competence ............................................................................................................ 2.17
Chapter 3 General design requirements
3.1 Introduction ...................................................................................................................................... 3.1
3.2 Stability, strength and serviceability................................................................................................. 3.1
3.2.1 Limit states design ................................................................................................................... 3.1
3.2.2 Stability (limit states) ............................................................................................................... 3.3
3.2.3 Strength (limit states) ............................................................................................................... 3.3
3.2.4 Serviceability (limit states) ...................................................................................................... 3.4
3.2.4.1 Serviceability limits for formwork deformations .............................................................. 3.4
3.2.4.2 Serviceability limits for multiple-use equipment .............................................................. 3.7
3.2.5 Working load limit (WLL) ....................................................................................................... 3.8

iv 2013 Stephen Ferguson


Contents

3.3 Structural integrity ............................................................................................................................ 3.9


3.3.1 Robustness ............................................................................................................................... 3.9
3.3.2 Preventing progressive collapse ............................................................................................... 3.10
3.4 Design working life .......................................................................................................................... 3.10
3.5 Australian standards ......................................................................................................................... 3.10
3.5.1 Australian formwork standards: a brief history ....................................................................... 3.10
3.5.2 Other Australian standards relevant to formwork .................................................................... 3.11
3.6 Economy ...................................................................................................................................... 3.12
Chapter 4 Actions and action combinations
4.1 Introduction ...................................................................................................................................... 4.2
4.2 Actions ...................................................................................................................................... 4.2
4.2.1 Permanent actions (G).............................................................................................................. 4.2
4.2.1.1 Vertical actions from weight of formwork (Gf) ................................................................. 4.2
4.2.1.2 Vertical actions from weight of concrete (Gc) ................................................................... 4.3
4.2.2 Concrete pressure (P)............................................................................................................... 4.3
4.2.3 Vertical and horizontal variable actions (QvandQh) ................................................................. 4.3
4.2.3.1 Vertical actions from workmen, concrete mounding and equipment (Qw) ........................ 4.4
4.2.3.2 Vertical actions from stacked materials and equipment (Qm)............................................ 4.5
D
4.2.3.3 Horizontal actions from construction activity (Qah) .......................................................... 4.8
4.2.3.4 Vertical and horizontal actions on guardrails (QgvandQgh) ................................................ 4.9
4.2.3.5 Other vertical and horizontal actions (QxvandQxh) ............................................................. 4.9
4.2.4 Environment actions (wind, snow, water and earthquakes) ..................................................... 4.10
4.2.4.1 Wind (Ws and Wu) ............................................................................................................. 4.10
4.2.4.2 Snow (Ss and Su) ............................................................................................................... 4.11
R
4.2.4.3 Serviceability and ultimate water (Ss and Su) .................................................................... 4.11
4.2.4.4 Earthquake (Eu) ................................................................................................................. 4.11
4.2.5 Accidental actions (Av and Ah) ................................................................................................. 4.11
4.2.5.1 Vertical impact (Av) .......................................................................................................... 4.11
4.2.5.2 Horizontal impact (Ah) ...................................................................................................... 4.11
4.2.6 Notional actions (N1, N2 and N3) .............................................................................................. 4.12
4.2.6.1 Notional horizontal forces for initial out-of-plumb erection (N1)...................................... 4.13
AF
4.2.6.2 Notional forces for braces that reduce the effective length of compression members (N2)4.13
4.2.6.3 Notional forces to ensure a minimum level of structural integrity (N3) ............................ 4.14
4.3 Action combinations ......................................................................................................................... 4.15
4.3.1 Action combinations for serviceability limit states .................................................................. 4.15
4.3.1.1 Surface finish quality and dimensional control ................................................................. 4.15
4.3.1.2 Elastic behaviour in multiple-use formwork ..................................................................... 4.16
4.3.2 Action combinations for ultimate limit states strength and stability ..................................... 4.16
4.3.2.1 Action combinations for stability limit states .................................................................... 4.17
4.3.2.2 Action combinations for strength limit states .................................................................... 4.17
4.3.2.3 Strength load factor for "primary" members ..................................................................... 4.18
4.3.2.4 Duration of load factor for use with AS 1720.1 ................................................................ 4.19
T
Chapter 5 Concrete Pressure
5.1 Introduction ...................................................................................................................................... 5.1
5.2 CIRIA Report No 108 ...................................................................................................................... 5.1
5.2.1 Factors influencing concrete pressure ...................................................................................... 5.4
5.2.2 Plan shape and area of cast section .......................................................................................... 5.4
5.2.3 Concrete rate of rise ................................................................................................................. 5.6
5.2.4 Constituent concrete materials ................................................................................................. 5.6
5.2.5 Concrete temperature ............................................................................................................... 5.6
5.2.6 Vertical form height ................................................................................................................. 5.7
5.2.7 Other factors ............................................................................................................................ 5.8
5.2.7.1 Method of vibration........................................................................................................... 5.8
5.2.7.2 Formwork permeability ..................................................................................................... 5.8
5.2.7.3 Underwater concreting ...................................................................................................... 5.9
5.3 Rate of rise ...................................................................................................................................... 5.9
5.3.1 Minimum rate of rise for full depth hydrostatic pressure ......................................................... 5.9
5.3.2 Proposed method of concrete placement.................................................................................. 5.9
5.3.3 Proposed sequence of concrete placement ............................................................................... 5.10

2013 Stephen Ferguson v


Formwork Handbook

5.3.4 Duration of concrete placement ............................................................................................... 5.11


5.3.5 Formwork and reinforcement arrangement .............................................................................. 5.11
5.3.6 Economy .................................................................................................................................. 5.11
5.3.7 Limitations on formwork strength or serviceability ................................................................. 5.12
5.4 Statics of concrete pressure .............................................................................................................. 5.12
5.4.1 Vertical side formwork ............................................................................................................ 5.12
5.4.2 Inclined side or sloping soffit formwork .................................................................................. 5.13
5.4.2.1 Hydrostatic concrete pressure ........................................................................................... 5.13
5.4.2.2 Concrete pressure limited by setting ................................................................................. 5.15
Chapter 6 Side Formwork
6.1 Introduction ...................................................................................................................................... 6.1
6.2 Form ties ...................................................................................................................................... 6.4
6.2.1 Types of form ties .................................................................................................................... 6.4
6.2.2 Form tie capacity ..................................................................................................................... 6.5
6.2.2.1 Comment on the design rules for tension members resisting concrete pressure ............... 6.5
6.2.2.2 Recommendations on the design of tension members resisting concrete pressure............ 6.5
6.2.3 Serviceability limit states ......................................................................................................... 6.6
6.2.4 Form tie identification.............................................................................................................. 6.6
D
6.2.5 Precautions when using form ties ............................................................................................ 6.6
6.3 Double sided wall formwork ............................................................................................................ 6.6
6.3.1 Balanced concrete pressure ...................................................................................................... 6.6
6.3.2 Limit states design of wall form face and framing members ................................................... 6.7
6.3.3 Unbalanced concrete pressure .................................................................................................. 6.8
6.4 Single sided wall formwork.............................................................................................................. 6.10
R
6.5 Bracing ...................................................................................................................................... 6.10
6.5.1 Bracing for alignment .............................................................................................................. 6.10
6.5.2 Bracing for stability ................................................................................................................. 6.11
6.5.2.1 Robustness ........................................................................................................................ 6.11
6.5.2.2 Imposed actions ................................................................................................................. 6.11
6.5.2.3 Accidental actions ............................................................................................................. 6.11
6.5.3 Bracing anchors ....................................................................................................................... 6.12
AF
Chapter 7 Soffit Formwork
7.1 Introduction ...................................................................................................................................... 7.1
7.2 Load path for vertical loads through soffit formwork ...................................................................... 7.2
7.3 Loading patterns ............................................................................................................................... 7.2
7.3.1 Stage 1 Prior to concrete placement ...................................................................................... 7.2
7.3.2 Stage 2 During concrete placement....................................................................................... 7.3
7.3.2.1 Stability limit states ........................................................................................................... 7.4
7.3.2.2 Strength limit states ........................................................................................................... 7.4
7.3.3 Stage 3 After concrete placement ......................................................................................... 7.4
7.4 Analysis of soffit form members ...................................................................................................... 7.4
T
7.4.1 Point loads vs UDL .................................................................................................................. 7.4
7.4.2 Lateral buckling of beams ........................................................................................................ 7.5
7.4.3 Simply supported beams .......................................................................................................... 7.5
7.5 Sloping soffit formwork ................................................................................................................... 7.7
7.5.1 Vertical falsework .................................................................................................................... 7.7
7.5.2 Stability limit statesSliding .................................................................................................. 7.8
7.5.3 Stability limit statesOverturning .......................................................................................... 7.8
7.5.4 Out of vertical falsework ......................................................................................................... 7.9
7.6 Unbalanced concrete pressure Discontinuous soffit formwork ..................................................... 7.10
Chapter 8 Falsework
8.1 Introduction ...................................................................................................................................... 8.1
8.2 Falsework design actions .................................................................................................................. 8.2
8.2.1 Stage 1 - Prior to concrete placement ........................................................................................ 8.2
8.2.2 Stage 2 - During concrete placement ......................................................................................... 8.3
8.2.3 Stage 3 - After concrete placement ............................................................................................ 8.3
8.3 Factors Influencing falsework behaviour, stability and strength ...................................................... 8.3
8.3.1 Falsework restraint ................................................................................................................... 8.4

vi 2013 Stephen Ferguson


Contents

8.3.1.1 Free Standing .................................................................................................................. 8.5


8.3.1.2 Top restraint .................................................................................................................... 8.6
8.3.1.3 Intermediate restraint....................................................................................................... 8.7
8.3.1.4 Requirements for formwork to be considered top restrained .......................................... 8.7
8.3.2 Falsework - Sway or fully braced frames ................................................................................ 8.9
8.3.3 Falsework bracing .................................................................................................................... 8.12
8.3.3.1 Problems with existing design criteria for falsework bracing ......................................... 8.12
8.3.3.2 Concept of effective length ............................................................................................. 8.12
8.3.3.3 Minimum axial stiffness and forces in braces that reduce the effective length shores ... 8.14
8.3.3.4 Brace connection behaviour ............................................................................................ 8.17
8.3.3.5 Brace axial stiffness ........................................................................................................ 8.19
8.3.3.6 Plan bracing (diagonal bracing in the horizontal plane) .................................................. 8.21
8.3.4 Falsework base plates and screw jacks .................................................................................... 8.21
8.3.4.1 Eccentricity ..................................................................................................................... 8.21
8.3.4.2 Detrimental effect of eccentric loads or reactions ........................................................... 8.23
8.3.4.3 Minimum eccentricity ..................................................................................................... 8.24
8.3.4.4 Rotational stiffness .......................................................................................................... 8.27
8.3.5 Spigot connections ................................................................................................................... 8.27
8.3.5.1 Eccentricity ..................................................................................................................... 8.28
D
8.3.5.2 Angular change at joints .................................................................................................. 8.29
8.3.5.3 Structural model .............................................................................................................. 8.29
8.3.6 Out-of-straight compression members ..................................................................................... 8.30
8.3.7 Differential settlement and axial shortening ............................................................................ 8.31
8.3.7.1 Differential settlement ..................................................................................................... 8.31
8.3.7.2 Different axial shortening................................................................................................ 8.33
8.3.8 Knee buckling .......................................................................................................................... 8.33
R
8.4 Analysing falsework structures ........................................................................................................ 8.35
8.4.1 Structural models ..................................................................................................................... 8.35
8.4.2 Calculating falsework member capacity .................................................................................. 8.36
8.4.2.1 End eccentricity and member out-of-straightness ........................................................... 8.37
8.4.2.2 Example .......................................................................................................................... 8.38
8.4.3 Falsework member column effective length ............................................................................ 8.39
AF
8.4.3.1 Estimates of column effective length in sway frames ..................................................... 8.39
8.4.4 Falsework frame buckling........................................................................................................ 8.42
Chapter 9 Stripping Criteria
9.1 Introduction ...................................................................................................................................... 9.1
9.2 Methods for stripping horizontal forms ............................................................................................ 9.1
9.2.1 Single-stage stripping............................................................................................................... 9.1
9.2.2 Two-stage stripping ................................................................................................................. 9.2
9.3 Minimum Stripping times................................................................................................................. 9.2
9.3.1 Development of concrete strength with age ............................................................................. 9.3
9.3.2 Minimum stripping times for vertical forms ............................................................................ 9.4
T
9.3.3 Minimum stripping times for horizontal forms and removal of shores.................................... 9.4
9.4 Calculating the minimum early-age strength for stripping ............................................................... 9.7
9.4.1 Stripping the forms only .......................................................................................................... 9.7
9.4.2 Stripping formwork supports under reinforced concrete ......................................................... 9.9
9.4.3 Stripping formwork supports under prestressed concrete ........................................................ 9.11
9.5 Assessment of concrete strength at early age ................................................................................... 9.11
9.5.1 AS 3600 ................................................................................................................................... 9.11
9.5.2 Other methods of assessing early-age compressive strength ................................................... 9.12
Chapter 10 Multistorey Shoring
10.1 Introduction .................................................................................................................................... 10.1
10.1.1 Significance of multistorey shoring design ............................................................................ 10.2
10.1.2 Current design guidance in Australian Standards .................................................................. 10.2
10.1.2.1 AS 3600 2009 ............................................................................................................... 10.2
10.1.2.2 AS 3610 1995 ............................................................................................................... 10.2
10.1.3 Guidance provided in the literature ........................................................................................ 10.3
10.1.3.1 Problems with past practise ........................................................................................... 10.3
10.1.3.2 Recent guidance predicting load distribution ................................................................ 10.4

2013 Stephen Ferguson vii


Formwork Handbook

10.1.4 Guidance provided herein ...................................................................................................... 10.4


10.2 Factors influencing the design of multistorey shoring .................................................................... 10.5
10.2.1 Method and sequence of stripping and shoring ...................................................................... 10.5
10.2.1.1 Reshoring ...................................................................................................................... 10.5
10.2.1.2 Undisturbed shoring ...................................................................................................... 10.6
10.2.2 Flexural stiffness of each floor at time of loading ................................................................. 10.7
10.2.2.1 Effective second moment of area .................................................................................. 10.7
10.2.2.2 Concrete modulus of elasticity ...................................................................................... 10.8
10.2.2.3 Span and support conditions ......................................................................................... 10.8
10.2.3 Effective axial stiffness of shores .......................................................................................... 10.8
10.2.3.1 Single shore ................................................................................................................... 10.9
10.2.3.2 Multiple shores .............................................................................................................. 10.10
10.2.4 Preload in multistorey shores ................................................................................................. 10.13
10.2.5 Construction loads ................................................................................................................. 10.13
10.2.5.1 Imposed construction load ............................................................................................ 10.14
10.2.6 In-service design floor load ................................................................................................... 10.15
10.2.7 Floor construction cycle time ................................................................................................. 10.15
10.2.8 Other factors .......................................................................................................................... 10.15
10.2.8.1 Concrete creep ............................................................................................................... 10.15
D
10.2.8.2 Ambient temperature change......................................................................................... 10.15
10.3 Methods for calculation load distribution in multistorey shoring ................................................... 10.16
10.3.1 Relative stiffness method ....................................................................................................... 10.16
10.3.1.1 One level of multistorey shores ..................................................................................... 10.17
10.3.1.2 Two levels of multistorey shores................................................................................... 10.18
10.3.1.3 Three levels of multistorey shores................................................................................. 10.19
10.3.1.4 Four levels of multistorey shores .................................................................................. 10.20
R
10.3.2 Slab shore interaction method ................................................................................................ 10.21
10.3.3 Finite element analysis ........................................................................................................... 10.22
10.4 Analysis methods for reshores vs undisturbed shores .................................................................... 10.23
10.4.1 General ................................................................................................................................... 10.23
10.4.1.1 Floor numbering ............................................................................................................ 10.24
10.4.2 Reshoring ............................................................................................................................... 10.24
AF
10.4.3 Undisturbed shoring ............................................................................................................... 10.24
10.5 Special situations to consider ......................................................................................................... 10.25
10.5.1 Unloaded multistorey shores .................................................................................................. 10.25
10.5.2 Onset of cracking ................................................................................................................... 10.25
10.5.3 Foundations - settlement ........................................................................................................ 10.25
10.5.4 Props not directly over each other.......................................................................................... 10.26
10.5.5 Shores at the centre of the slab carry more load than those closer to the supports ................ 10.26
10.6 Acceptance criteria for early-age loading ....................................................................................... 10.26
10.6.1 Uncertainty............................................................................................................................. 10.26
10.6.2 Serviceability limit states ....................................................................................................... 10.26
10.6.3 Ultimate limit states ............................................................................................................... 10.27
T
10.6.4 Acceptable overload............................................................................................................... 10.27
Chapter 11 Concrete Finishes - Identification of defects
11.1 Introduction .................................................................................................................................... 11.1
11.2 Blowholes ...................................................................................................................................... 11.1
11.3 Face steps ...................................................................................................................................... 11.2
11.4 Concrete placing and compaction defects ...................................................................................... 11.4
11.5 Cleanliness of forms before concrete is placed .............................................................................. 11.5
11.6 Other common defects .................................................................................................................... 11.6
Appendix A Formwork importance
A.1 Level of Risk ................................................................................................................................... A.1
Appendix B Coefficients of static friction
B.1 Introduction..................................................................................................................................... B.1
Appendix C Recommended reading
C.1 Introduction..................................................................................................................................... C.1

viii 2013 Stephen Ferguson


List of Figures

List of Figures
Figure 1.1 Soffit formwork includes soffit forms and supporting falsework................................. 1.1
Figure 1.2 Wall formwork ............................................................................................................. 1.2
Figure 1.3 Falsework ..................................................................................................................... 1.2

Figure 3.1 Probability distributions for design action effects and design resistance ..................... 3.2
Figure 3.2 Cumulative deflections ................................................................................................. 3.4
Figure 3.3 Measuring surface undulations ..................................................................................... 3.6
Figure 3.4 Impact damage but not failure ...................................................................................... 3.9

Figure 4.1 Formwork and falsework design actions ...................................................................... 4.1


Figure 4.2 Typical construction activity during concrete placement (Stage 2) ............................. 4.4
Figure 4.3 Stacked materials (Stage 1) .......................................................................................... 4.6
Figure 4.4 Stacked materials and equipment (Stage 3).................................................................. 4.6
Figure 4.5 Horizontal actions from construction activity .............................................................. 4.9
Figure 4.6 Impact from moving crane load ................................................................................... 4.12
Figure 4.7 Impact from moving crane load ................................................................................... 4.12
Figure 4.8 Impact from moving vehicle ........................................................................................ 4.12
D
Figure 4.9 Notional loads required to take account of initial out-of-plumb erection ..................... 4.13
Figure 4.10 Notional forces to ensure braces have the minimum brace strength & stiffness .......... 4.14
Figure 4.11 Notional horizontal actions to ensure minimum levels of structural integrity.............. 4.15

Figure 5.1 Comparison between measured and calculated pressures ............................................ 5.1
Figure 5.2 Concrete pressure envelope .......................................................................................... 5.2
R
Figure 5.3 Influence of rate of rise on concrete pressure in walls ................................................. 5.5
Figure 5.4 Influence of rate of rise on concrete pressure in columns ............................................ 5.5
Figure 5.5 Influence of coefficient C2 on concrete pressure in walls ............................................ 5.6
Figure 5.6 Influence of concrete temperature on concrete pressure in walls ................................. 5.7
Figure 5.7 Measuring vertical form height of concrete discharge height ...................................... 5.7
Figure 5.8 Influence of vertical form height on concrete pressure in walls ................................... 5.8
AF
Figure 5.9 Concrete placed in layers ............................................................................................. 5.10
Figure 5.10 Vertical construction joints introduced to reduce the area of concrete to be placed .... 5.11
Figure 5.11 Concrete placed to full height over a shorter distance to avoid cold joints .................. 5.11
Figure 5.12 Lateral concrete pressure distribution on vertical formwork ........................................ 5.13
Figure 5.13 Hydrostatic concrete pressure distribution on an inclined soffit or lower surface........ 5.13
Figure 5.14 Vector components of hydrostatic concrete pressure distribution on an inclined soffit 5.14
Figure 5.15 Hydrostatic concrete pressure distribution on an inclined top or upper surface ........... 5.14
Figure 5.16 Vector components of hydrostatic concrete pressure distribution on an inclined top .. 5.15
Figure 5.17 Incorrect concrete pressure distribution (limited by setting) on an inclined soffit ....... 5.15
Figure 5.18 Component concrete pressure distribution limited by setting on an inclined soffit...... 5.16
Figure 5.19 Correct concrete pressure distribution limited by setting on an inclined soffit ............ 5.16
T
Figure 5.20 Concrete pressure distribution limited by setting on an inclined top form ................... 5.17
Figure 5.21 Component concrete pressure distribution limited by setting on an inclined top form 5.17

Figure 6.1 Wall formwork with secondary horizontal walers and primary vertical soldiers ......... 6.1
Figure 6.2 Wall formwork with secondary vertical studs and primary horizontal walers ............. 6.2
Figure 6.3 Load distribution through side formwork..................................................................... 6.2
Figure 6.4 Double sided wall formwork (bracing not shown) ....................................................... 6.3
Figure 6.5 Single sided wall formwork ......................................................................................... 6.3
Figure 6.6 Common types of form ties .......................................................................................... 6.4
Figure 6.7 Form ties balance concrete pressure on double sided formwork .................................. 6.7
Figure 6.8 Form ties balance concrete pressure on double sided inclined formwork .................... 6.7
Figure 6.9 Out of balance effects when opposing side forms are not parallel ............................... 6.8
Figure 6.10 Out-of-balance effects on inclined and tapered wall formwork ................................... 6.9
Figure 6.11 Action effects on single sided formwork ...................................................................... 6.10
Figure 6.12 Imposed and notional actions on side formwork .......................................................... 6.11
Figure 6.13 Bracing for accidental impact....................................................................................... 6.12

Figure 7.1 Simple suspended slab formwork................................................................................. 7.1


Figure 7.2 Load distribution through soffit formwork (when viewed from underneath) ............... 7.2

2013 Stephen Ferguson ix


Formwork Handbook

Figure 7.3 Line or point loads that arise from stacked materials during Stage I ........................... 7.2
Figure 7.4 Adverse partial loading of multiple span bearer ........................................................... 7.3
Figure 7.5 Importance of direction of pour.................................................................................... 7.3
Figure 7.6 Point loads from secondary beams (joists) acting on the primary beam (bearer) ......... 7.4
Figure 7.7 Narrow timber beams required lateral restraint ............................................................ 7.5
Figure 7.8 Coefficients for beam action effects ............................................................................. 7.6
Figure 7.9 Concrete and formwork held at rest on a sloping soffit by friction .............................. 7.7
Figure 7.10 Action effects on sloping soffit formwork ................................................................... 7.7
Figure 7.11 Destabilising and stabilising action effects on joists running perp. to the slope .......... 7.9
Figure 7.12 Concrete held at rest on a sloping soffit by friction, with out-of-vertical falsework .... 7.9
Figure 7.13 Examples of discontinuous soffit formwork ................................................................ 7.10
Figure 7.14 Examples of horizontal forces being transferred to falsework ..................................... 7.11

Figure 8.1 Unrestrained or freestanding falsework........................................................................ 8.5


Figure 8.2 Top restrained falsework .............................................................................................. 8.6
Figure 8.3 Intermediate restraint.................................................................................................... 8.7
Figure 8.4 Designation for the load paths required to provide full lateral and rotational restraint 8.8
Figure 8.5 Differing levels of top restraint provided by surrounding walls ................................... 8.8
Figure 8.6 Differing levels of top restraint provided by columns .................................................. 8.9
D
Figure 8.7 Freestanding falsework with sway and fully braced members ..................................... 8.10
Figure 8.8 Freestanding falsework with multiple column bracing ................................................ 8.10
Figure 8.9 Freestanding falsework with fully braced members ..................................................... 8.11
Figure 8.10 Top restrained and fully braced falsework ................................................................... 8.11
Figure 8.11 Effective length factors for members with idealised end restraints .............................. 8.13
Figure 8.12 Braced column with an initial out-of-straightness imperfection o .............................. 8.13
R
Figure 8.13 A series of parallel out-of-straight columns restrained by a line of bracing ................. 8.15
Figure 8.14 Brace stiffness multiplier for a series of parallel 48.3CHS4.0 columns ....................... 8.15
Figure 8.15 Single column braced at multiple points along its length ............................................. 8.16
Figure 8.16 Typical horizontal brace (ledger) to column (standard or shore) connection ............... 8.18
Figure 8.17 Hysteresis loops for horizontal brace to column connection ........................................ 8.18
Figure 8.18 Scaffold tube double coupler ........................................................................................ 8.19
Figure 8.19 Scaffold tube swivel coupler ........................................................................................ 8.20
AF
Figure 8.20 An example of eccentric end connection of a diagonal brace ...................................... 8.21
Figure 8.21 Examples of eccentric loading ..................................................................................... 8.22
Figure 8.22 Examples of irregular or variable stiffness bearing surfaces ........................................ 8.23
Figure 8.23 Example of eccentric reaction ...................................................................................... 8.23
Figure 8.24 Column strength curves for eccentrically loaded shore in "new" condition ................. 8.24
Figure 8.25 Illustrations of eccentricities of actions and reactions .................................................. 8.25
Figure 8.26 Effects of eccentric actions can be more severe in one direction ................................. 8.26
Figure 8.27 Falsework failure at a spigot joint during testing ......................................................... 8.27
Figure 8.28 Examples of good and bad practice in loading connections in compression members 8.28
Figure 8.29 Eccentricities arise at spigot joints ............................................................................... 8.28
Figure 8.30 Angular imperfections at joints .................................................................................... 8.29
T
Figure 8.31 Spigot structural model ................................................................................................ 8.30
Figure 8.32 Initially out-of-straight slender compression members ................................................ 8.30
Figure 8.33 Column strength curves for shores complying with the different out-of-straightness .. 8.31
Figure 8.34 Differential settlement due to the presence of concrete foundations ............................ 8.32
Figure 8.35 Load distribution due to beam flexural stiffness .......................................................... 8.32
Figure 8.36 Load redistribution due to differential axial shortening ............................................... 8.33
Figure 8.37 Euler buckling and knee buckling ................................................................................ 8.34
Figure 8.38 Knee buckling of formwork frames with extended screw jacks ................................... 8.34
Figure 8.39 Non-uniform deformation of timber loaded at right angles to the grain....................... 8.34
Figure 8.40 Models for Euler and knee buckling ............................................................................ 8.35
Figure 8.41 An eccentrically loaded pin-ended strut ....................................................................... 8.38
Figure 8.42 Effective length of members in top restrained frames with central pinned bracing ..... 8.40
Figure 8.43 Effective length of members in top restrained frames with pinned bracing at the base 8.40
Figure 8.44 Effective length of members in free standing frames with central pinned bracing ...... 8.41
Figure 8.45 Effective length of members in free standing frames with pinned bracing at the base 8.41
Figure 8.46 Effective length of members in free standing frames with pinned bracing .................. 8.41

x 2013 Stephen Ferguson


List of Figures

Figure 9.1 Single-stage stripping ................................................................................................... 9.1


Figure 9.2 Two-stage stripping leaving undisturbed shores ......................................................... 9.2
Figure 9.3 Two-stage stripping by back-propping ......................................................................... 9.2
Figure 9.4 Typical development of concrete strength with age ..................................................... 9.3
Figure 9.5 Typical compressive strength development of Portland cement .................................. 9.3

Figure 10.1 Example of multistorey shoring with one floor of formwork & 3 floors of shores .... 10.1
Figure 10.2 Indicative discrepancy in load distribution .................................................................. 10.3
Figure 10.3 Example of "thinning" multistorey shoring ................................................................. 10.4
Figure 10.4 Multistorey shoring with one floor of formwork and two floors of reshores .............. 10.5
Figure 10.5 Multistorey shoring with three floors of "undisturbed" formwork .............................. 10.6
Figure 10.6 Idealised model of one floor of formwork and two floors of multistorey shoring ..... 10.9
Figure 10.7 One level of formwork and three levels of reshores all on a 5 x 7 grid ....................... 10.11
Figure 10.8 As per Figure 10.7 except the lower two floors of reshores are on a 5 x 4 grid .......... 10.12
Figure 10.9 Different multistorey shoring load situations .............................................................. 10.14
Figure 10.10 One level of multistorey shoring ................................................................................. 10.17
Figure 10.11 Two levels of multistorey shoring ............................................................................... 10.18
Figure 10.12 Three levels of multistorey shoring ............................................................................. 10.19
D
Figure 10.13 Four levels of multistorey shoring ............................................................................... 10.20
Figure 10.14 The layout of formwork shores and multistorey shores differ from floor to floor ...... 10.22
Figure 10.15 Slab deflections under staged construction loads from multistorey shoring ............... 10.23
Figure 10.16 Idealised model with both the top and intermediate floor loadings ............................. 10.24
Figure 10.17 Idealised model of stripping the lowest level of "undisturbed" multistorey shoring ... 10.25

Figure 11.1 Acceptable Blowholes .................................................................................................. 11.1


R
Figure 11.2 Face step reduced by grinding ..................................................................................... 11.2
Figure 11.3 Acceptable quality repair of face steps ........................................................................ 11.2
Figure 11.4 Acceptable quality repair of face steps ........................................................................ 11.3
Figure 11.5 Measuring face steps ................................................................................................... 11.3
Figure 11.6 Honeycombing along the bottom edge of a concrete beam ......................................... 11.4
Figure 11.7 Poor compaction with board finish.............................................................................. 11.4
AF
Figure 11.8 Debris left when formwork not cleaned ...................................................................... 11.5
Figure 11.9 Rust stains left on the forms ........................................................................................ 11.6
Figure 11.10 Concrete surface damaged after removal of the forms ................................................ 11.6
Figure 11.11 Concrete surface contaminated after removal of the forms ......................................... 11.7
Figure 11.12 Concrete surface stains ................................................................................................ 11.7
Figure 11.13 Dirty faces of wall forms result on objectionable appearance ...................................... 11.8
Figure 11.14 Poor formwork sealing resulting in objectionable appearance ..................................... 11.8
T

2013 Stephen Ferguson xi


Formwork Handbook

List of Tables
Table 2.1 Key participants and their roles or responsibilities ...................................................... 2.2
Table 2.2 Appropriate levels of obligation and competence for different levels of risk .............. 2.17

Table 3.1 Acceptable form face deformations and surface undulations ....................................... 3.5
Table 3.2 Acceptable surface undulations expressed as span to deflection ratios ........................ 3.6
Table 3.3 Recommended serviceability limits for formwork member deflection ........................ 3.7
Table 3.4 Applicable Standards for various materials used in formwork .................................... 3.11

Table 4.1 Stages of formwork construction.................................................................................. 4.2


Table 4.2 Annual probabilities of exceedence for ultimate limit states events ............................ 4.10
Table 4.3 Ultimate limit states regional wind speeds, m/s ........................................................... 4.11
Table 4.4 Duration of load factor (k1) for strength ....................................................................... 4.19

Table 5.1 Values of coefficient C2 ............................................................................................... 5.3


Table 5.2 Factors affecting concrete pressure .............................................................................. 5.4

Table 8.1 Values for the factor in Equation 8.8 ....................................................................... 8.17
D
Table 9.1 Early-age mean strengths for normal-class concrete .................................................... 9.4
Table 9.2 Minimum compressive strength of concrete for stripping vertical forms..................... 9.4
Table 9.3 Minimum strength and curing requirements for concrete ............................................ 9.5
Table 9.4 Minimum times for stripping of forms between undisturbed shores ............................ 9.5
Table 9.5 Minimum times before removal of supports not supporting structures above ............ 9.6
R
Table 10.1 Loads in multistorey shoring taking account of shore axial stiffness and layout ......... 10.10
Table 10.2 Axial stiffness of each level of shores relative to the 28-day slab flexural stiffness .... 10.13

Table A.1 Level of risk for formwork in different situations ........................................................ A.1
AF
Table B.1 Nominal design coefficient of static friction for use in limit states design .................. B.2
T

xii 2013 Stephen Ferguson


Chapter 1 - Introduction

1
Introduction
1.1 Formwork and falsework
During construction, formwork supports and acts as a mould for wet concrete. Formwork is often
referred to as "temporary works" although, some formwork or part thereof, may remain part of the
permanent structure.
The term "formwork" describes both the forms directly in contact with concrete and a supporting
structure of braces and form ties, and where appropriate falsework. Part of the formwork may also act
as a temporary platform or scaffold; i.e. provide access, a working platform or an area of formwork
designated for loading or storing materials, plant and equipment.
D
Soffit forms for slabs and beams are supported by falsework, see Figure 1.1
R
AF
T

Figure 1.1: Soffit formwork includes soffit forms and supporting falsework
Side forms for concrete walls, columns, slabs, beams and foundations are supported by bracing and
often form ties in tension are used to balance concrete pressure, see Figure 1.2.

2013 Stephen Ferguson 1.1


Formwork Handbook

Figure 1.2: Wall formwork


D
The term "falsework" refers to temporary structures used to support not only formwork, but also parts
of the permanent structure until they become self-supporting, see Figure 1.3.
R
AF
T
Figure 1.3: Falsework
1.2 Formwork importance
Formwork is important because it has a major impact on the quality, cost and time to build concrete
structures. In addition, its sound design and construction is essential to ensure safety during
construction.
A high-quality off-form concrete surface finish, especially where colour control is specified, can only
be achieved with extreme care, careful planning and design, as well as a high level of workmanship.
The quality of formwork is also important in achieving durable concrete and nearly all formwork must
be dimensionally correct within relatively small tolerances.
Formwork is the major cost component of in-situ concrete and labour is the major cost component of
formwork. The cost of formwork labour alone will often exceed the sum of all other concrete cost

1.2 2013 Stephen Ferguson


Chapter 1 - Introduction

components; i.e. formwork material, reinforcement supply and fixing, as well as concrete supply and
placing.
Formwork is nearly always on the critical path of the schedule to construct any concrete structure.
Formwork is a heavily-loaded structure whose reliable performance is critical to safety.
The level of effort and rigour applied to formwork design and construction should reflect the
importance of formwork, specifically the risk and consequence of failure.
1.3 Formwork design requirements
In Australia, requirements for the design and construction of formwork are set out in AS 3610 1995
Formwork for concrete (SA 1995), including Amendment No 1 (January 2003), and AS 3610.1 2010
Formwork for concrete Part 1: Documentation and surface finish (SA 2010). AS 3610.1 2010
supersedes only part of AS 3610 1995, specifically: only Sections 2 and 3, Clause 4.7 and Section 5
of AS 3610 1995. Until AS 3610.2 is published (which appears unlikely to occur in the near future),
the requirements for formwork design and testing set out AS 3610 1995 Sections 4 (except for Clause
4.7) and Appendix A still apply.
Since they first appeared in AS 3610 1990, only minor amendments have been made to the
requirements for design and testing set out in AS 3610 1995. Some requirements are now out-of-date.
D
The purpose of this Handbook is to provide up-to-date guidance on formwork design that fulfils the
requirements of AS 3610.1 2010 and AS 3610 1995. The Handbook expands on the content of the
Standards that pertain to formwork design and documentation and also provides background
information.
R
The design methods set out in the Handbook comply with the requirements and methods, as well as,
where possible, the notation set out in the latest Australian material and design Standards. In addition,
the design guidance herein follows the limit states design philosophy and general principles set out in
ISO 2394:1998, General principles on reliability for structures (ISO 1998) and AS 1170.0 2002
Structural design actions Part 0: General principles (SA 2002).
Much of the information presented is sourced from the recommendations and information provided in
AF
the literature, authoritative references and other national Standards. Where necessary, the information
has been adapted to comply with Australian Standards and practice.
1.4 Scope
The primary focus of the Handbook is to provide guidance in areas of formwork design critical to
safety. Accordingly, the scope of the Handbook has been limited to addressing: design and
construction procedures; general design requirements; design actions, combinations and their
application to side and soffit formwork; falsework design; and stripping and multistorey shoring.
Information is also provided on identifying defects in the surface finish of formed concrete.
T
In addition to updating the content provided in Section 4 of AS 3610 1995, the Handbook introduces
concepts not covered in the Standards and information on the application, specifically:
Chapter 2 focuses on addressing shortcomings in procedural adequacies, which
researchers agree lead to failure and collapse. The Chapter introduces the concept of a
formwork coordinator who is responsible to manage and coordinate formwork design
and construction. Another important aspect covered is the preparation of a formwork brief
that sets out the requirements of the formwork design. The Chapter also sets out the roles
and responsibilities of, as well as the documentation required from, all those involved,
including: the project designer, construction contractor, formwork supplier, formwork
designer, formwork checker, formwork contractor, formwork supervisor, and formwork
inspector.
Chapter 3 sets out general formwork design requirements that must be satisfied. In
particular, Chapter 3 sets out the basic requirements to satisfy serviceability, stability and
strength limit states, as well as the need to provide minimum levels of structural integrity.
Complicated requirements in AS 3610.1 2010 pertaining to acceptable surface

2013 Stephen Ferguson 1.3


Formwork Handbook

deformations are translated into easy to use formwork member span-deflection ratios to
satisfy serviceability limit states.
Chapter 4 covers actions and action combinations. The Chapter includes practical
changes to the magnitude and application of loads for concrete mounding and stacked
materials. It also introduces the concept of notional loads, specifically: to address the
effects of permitted initial out-of-straightness; to take account of the forces and minimum
stiffness required in braces that reduce the effective length of compression members; and
to provide for a minimum level of structural integrity. Chapter 4 also introduces the
concept of different levels of risk for formwork and how these can be used to determine
the magnitude of environmental actions.

Chapter 4 also specifies action combinations for limit states design that are consistent
with AS 1170.0 and in a new format that lists vertical and horizontal action combinations
separately. This is intended to avoid confusion, assist in identifying critical combinations
and emphasise the importance of horizontal loads, which are too often underestimated or
neglected. The magnitude of the global load factor for primary members introduced in
Amendment No 1 to AS 3610 1995 (SA 2003) has been re-calibrated to the new action
combinations.
D
Chapter 5 covers concrete pressure calculated using the formula in CIRIA Report No 108,
which was adopted in AS 3610 1995. The values for the coefficient for the effect of
concrete cement and admixtures (C2) have been updated to include for silica fume and
self-compacting concrete. The Chapter includes an extensive discussion on the factors
that influence concrete pressure.
R
Chapter 5 also provides guidance on the statics of concrete pressure and methods of
simplifying concrete pressure distributions for inclined side or sloping soffit or top
formwork.
Chapter 6 provides a basic introduction to design of side formwork and bracing side
AF
formwork, for both single and double-sided forms. It also provides guidance on
identifying situations where the concrete pressure is out-of-balance and requires careful
attention to bracing details and load paths.
Chapter 7 provides a basic introduction to the design of soffit formwork, in particular:
loads, loading patterns, issues associated with sloping soffit formwork, unbalanced
concrete pressure and discontinuous soffit formwork.
Chapter 8 presents and discusses the fundamental aspects related to the design of
formwork falsework, namely: falsework design actions; falsework restraint; the
difference between sway and fully braced frames; requirements for falsework bracing;
T
and the influence of effective brace axial stiffness, connection behaviour, eccentricity,
member out-straightness, differential settlement and axial shortening.

Chapter 8 also provides guidance on analysing falsework, as well as on estimating


member effective lengths, calculating member capacity and frame buckling.
Chapter 9 provides guidance on the stripping criteria for vertical and horizontal
formwork. The Chapter also introduces different methods of stripping horizontal
formwork and presents methods for calculating the minimum early-age concrete strength
for stripping.
Chapter 10 discusses the guidance provided in the literature, past practice and factors
influencing the design of multistorey shoring. Three methods for calculating load
distribution in multistorey shoring are presented. Different methods for analysing
reshores and undisturbed shores are explained and the Chapter focuses attention on
special situations to consider. Guidance on acceptance criteria for early-age loading of
concrete slabs is also provided.

1.4 2013 Stephen Ferguson


Chapter 1 - Introduction

Chapter 11 presents a series of photographs of concrete surface finish defects commonly


encountered and provides commentary on the cause and possible repair.
Appendix A provides guidance on assessing formwork importance based on consequence
of failure.
Appendix B provides a table of coefficients of static friction for a range of commonly
encountered materials for using in limit states design.
Appendix C provides a list of recommended reading for those seeking to increase their
knowledge of formwork design.
The Handbook is not intended to be a stand-alone reference. It provides information fundamental to
formwork design and construction of commonly encountered concrete elements. To obtain a broad
understanding about formwork design and construction, the readers attention is drawn to the cited
literature referenced at the end of each Chapter and the recommended reading referenced in Appendix
C. For specialist information relating to particular proprietary equipment and products, readers should
consult the supplier proprietary documentation, method statements and safety reports. Readers should
also be familiar with the requirements of relevant Work, Health and Safety Regulations and Codes of
Practice that pertain to formwork design and construction.
D
1.5 Application
The Handbook is intended to be a useful reference for practicing and student engineers, project
designers, construction contractors, formwork contractors, formwork designers, formwork suppliers,
formwork checkers and formwork inspectors.
Despite focusing on formwork, many of the concepts presented herein equally apply, and could be
R
adapted, to the design of other temporary structures, including: falsework for other than formwork and
scaffolding.
1.6 Glossary
Action
AF
Load.
Action effect
Deformation, shear force, bending moment, axial force or torsion in a member, component or
connection under load.
Backprops (backpropping)
Shores installed under a suspended slab or beam, as part of the formwork stripping process, in a
manner that transfers the load in the formwork shores being removed to the backprops without any
additional stress in the suspended slab or beam.
T
Braces (bracing)
Horizontal and diagonal components required to align, support and stabilise formwork and falsework.
Construction documentation
Documents that set out details of the overall construction: method, schedule, equipment and logistics.
Falsework
A temporary structure used to support construction loads (e.g. soffit forms or parts of the permanent
structure until they are self-supporting).
Form (soffit, side, sloping or top)
Part of the formwork directly in contact with the concrete, which typically consists of a form face and
supporting framework of beams, which may be prefabricated as a single component (e.g. form panel)
or comprise an assembly of a form face sheet (e.g. plywood or steel), grillage of secondary and
primary beams.

2013 Stephen Ferguson 1.5


Formwork Handbook

Form tie
Tension member used to balance the concrete pressure on opposing forms.
Formwork
Formwork is a structure, usually temporary, erected to support and mould cast-in-situ concrete until it
becomes self-supporting. It consists of forms and, where appropriate form braces, form ties and
falsework (SA 2010).
Formwork design documentation
Documents (e.g. calculations, drawings, sketches, specifications, brochures, risk assessment, method
statements and instructions) that set out details of the formwork, including design details, components,
arrangements, safe work methods and safety hazards.
Gangform
A large prefabricated form or form assembly used and handled as a single form (i.e. not dismantled
after each use), usually built for forming walls.
Importance level
A structural category used to ensure the level of reliability is appropriate to the level of risk.
D
Level of risk
A framework for categorising risk based on the consequence of failure.
Multistorey shoring
R
Undisturbed shores or reshores that support floors, as well as permit the transfer and sharing of
construction loads between the floors they connect, during the construction of multistorey structures.
Permanent forms
Part of the permanent structure that acts as formwork.
Project documentation
AF
Documents (e.g. drawings, specifications and associated documents) that set out information required
to construct the project.
Proprietary documentation
Documents (e.g. brochures, catalogues, drawings, and specifications) that set out information required
for proprietary equipment.
Proprietary equipment
Multiple-use or mass produced construction components, equipment, and plant available for sale or
hire (e.g. prefabricated or manufactured formwork and falsework components, cranes, hoists, and
T
concrete pumps).
Reshores
Shores installed under a self-supporting suspended slab or beam; i.e. after removing all the formwork.
Screw jacks ('U' head and base)
Threaded falsework components that connect to the top and bottom of shores. They provide a bearing
surface and height adjustment and permit formwork stripping.
Shores (props, supports, or standards)
Falsework components, usually vertical but may be inclined, that act as columns (or struts).
Soffit formwork (soffit forms)
Formwork or forms for the underside of concrete elements; e.g. slabs, beams, and stairs.

1.6 2013 Stephen Ferguson


Chapter 1 - Introduction

Soleboards (Soleplates)
Falsework components used as "temporary footings" to spread the load from shores and reduce the
bearing pressure on the foundation material.
Tableform
A formwork assembly used and handled as a single form (i.e. not dismantled after each use), usually
built for forming suspended slabs and beams.
Undisturbed shores
Formwork shores left in place and untouched under a suspended slab or beam.
1.7 Notation
To avoid confusion, the notation adopted herein often varies from that used in the source document.
A = member cross-sectional area (see Section 8.3.3.5)
Ac = plan area of the concrete element to be cast (see Section 5.3)
Af = bearing area of the forms or other material (see Section 10.2.3.1)
D
Ah = accidental horizontal actions (see Section 4.2.5.2)
An = net area of the cross-section (see Sections 3.2.4.2 and 8.4.2.1)
As = cross-sectional area of the shore (see Section 10.2.3.1)
Ast = cross sectional area of longitudinal tensile steel reinforcement (see Section 10.2.2.1)
R
Av = accidental vertical actions (see Section 4.2.5.1)
a = surface undulation reading (see Section 3.2.4.1)
an = brace stiffness multiplier given in Figure 8.14 (see Section 8.3.3.3)
b = surface undulation reading (see Section 3.2.4.1)
AF
b = width of a joist (see Section 7.5.3)
C1 = coefficient for the effect of size and shape of formwork (see Section 5.2.1)
C2 = coefficient for the effect of concrete cement and admixtures (see Section 5.2.1)
C3 = coefficient for the effect of concrete temperature (see Section 5.2.1)
cm = factor for unequal end bending moments (see Section 8.4.2)
d1id = internal diameter of the outer member (see Section 8.3.5.2)
d2od = external diameter of the inner member (see Section 8.3.5.2)
T
d = depth of a joist (see Section 7.5.3)
dc = overall depth of the concrete section (see Section 9.3.2)
df = diameter of bolt or pin (see Section 3.2.4.2)
dl = thickness of the concrete layer (see Section 5.3.3)
E = Young's modulus of elasticity (see Section 8.3.3.2)
E = action effects (see Section 3.2.1)
Ecj = mean modulus of elasticity of the concrete at the relevant age (see Section 10.2.2)
Ed = design action effect (see Section 3.2.3 and 4.3.2)
Ed,dst = design action effect from destabilising actions (see Sections 3.2.2 and 4.3.2)
Ed,stb = design action effect from stabilising actions (see Section 3.2.2 and 4.3.2)
Edh = design action effect from the combined horizontal actions (see Section 4.3.2.2)

2013 Stephen Ferguson 1.7


Formwork Handbook

Edh,dst = design action effect from destabilising horizontal actions (see Section 4.3.2.1)
Edv = design action effect from the combined vertical actions (see Section 4.3.2.2)
Edv,dst = design action effect from destabilising vertical actions (see Section 4.3.2.1)
Ef = modulus of elasticity of the forms or other material (see Section 10.2.3.1)
En = nominal action effect (see Section 3.2.1)
Es = modulus of elasticity of the shore material (see Section 10.2.3.1)
Esh = serviceability action effect from horizontal actions (see Section 4.3.1.1)
Es.max = maximum action effect satisfying serviceability limit states (see Section 3.2.5)
Esv = serviceability action effect from vertical actions (see Section 4.3.1.1)
Eu = earthquake actions (see Section 4.2.4.4)
e = largest end eccentricity of the load or reaction on a member (see Section 8.4.2.1)
e = fixed eccentricity of the load or reaction on a member (see Section 8.3.4.3)
e = expected eccentricity of the load or reaction on a member (see Section 8.3.4.3)
D
Fp = resultant force from concrete pressure (see Section 5.4.1)
f'c = characteristic compressive (cylinder) strength of concrete at 28 days (see Section 9.4.1)
f' ce = early-age characteristic compressive strength of concrete (see Section 9.4.1)
f'cf = characteristic flexural tensile strength of the concrete, in MPa (see Section 9.4.1)
R
fcm = mean grade strength of all results for the grade (see Sections 9.4.1 and 9.5.1)
fcmi = mean in situ compressive strength of concrete at the relevant age (see Section 10.2.2.2)
fy = yield stress used in design (see Sections 3.2.4.2 and 8.4.2.1)
AF
Gc = weight of concrete (see Section 4.2.1.2)
Gf = weight of formwork (see Section 4.2.1.1)
Gser = permanent actions for services, partitions, ceilings, floor treatments etc (see Section 9.4.2)
g = gravity (see Section 5.2)
h = depth below the top of the concrete (see Section 5.2)
hc = depth of concrete pour (see Section 5.2)
hf = vertical form height (see Section 5.2)
T
hh = maximum depth of hydrostatic pressure (see Section 5.2)
hp = height of centre of pressure above the bottom of the form (see Section 5.4.1)
I = second moment of area of the cross section (see Section 8.3.3.2)
I = second moment of area of the uncracked concrete section about the centroidal axis (see
Section 10.2.2.1)
Icr = second moment of area of a cracked section with the reinforcement transformed to an
equivalent area of concrete (see Section 10.2.2.1)
Ief = effective second moment of area of the concrete section (see Section 10.2.2)
K12 = relative stiffness term for slab 1 to 2 (see Section 10.3.1)
K23 = relative stiffness term for slab 2 to 3 (see Section 10.3.1)
K34 = relative stiffness term for slab 3 to 4 (see Section 10.3.1)
K45 = relative stiffness term for slab 4 to 5 (see Section 10.3.1)

1.8 2013 Stephen Ferguson


Chapter 1 - Introduction

k = brace minimum axial stiffness (see Section 8.3.3.2)


k1 = duration of load factor for timber (see Section 4.3.2.4)
kc = assessment factor determined from the number of controlled grade samples (see Section
9.5.1)
ke = member effective length factor (see Sections 8.3.3.2 and 8.4.3)
kn = minimum axial stiffness of a brace number n (see Section 8.3.3.3)
LSD = limit states divisor (see Section 3.2.5)
l = the span between formwork shores (see Section 9.3.2)
l = member length (see Sections 8.3.3.2 and 8.4.2)
l = straight edge length as a span (see Section 3.2.4.1)
l2lap = length the inner member laps inside the outer member (see Section 8.3.5.2)
lc = concrete slab span (see Section 10.2.2)
lf = thickness of the form or other material (see Section 10.2.3.1)
D
ls = length of the shore (see Section 10.2.3.1)
M = maximum bending moment of a beam (see Section 7.4.3)
Mcr = bending moment causing cracking of the section (see Section 10.2.2.1)
Md = design bending moment (see Sections 8.4.2 and 9.4.1)
R
Mds = maximum bending moment at the section (see Section 10.2.2.1)
Md,dst = design moment from forces causing overturning (see Section 3.2.2)
Md,stb = design moment from forces that have a stabilising effect (see Section 3.2.2)
Mn = nominal moment capacity (see Section 3.2.2)
AF
Ms = nominal section moment capacity (see Section 8.4.2)
Muo = ultimate strength in bending, without axial force, at a cross-section (see Section 9.4.1)
m = number of spans (see Section 8.3.3.3)
N1 = notional horizontal action for initial out-of-plumb erection (see Section 4.2.6.1)
N2 = notional horizontal action for braces (see Section 4.2.6.2)
N3 = notional horizontal action to ensure a minimum level of structural integrity (see Section
4.2.6.3)
T
Nbs = design ply bearing force at serviceability limit states (see Section 3.2.4.2)
Nc = nominal member axial compression capacity (see Section 8.4.2)
Nc(kel) = Nc as a function of the member effective length (see Section 8.4.2)
Nc(l) = Nc as a function of the member length (see Section 8.4.2)
Nd = design axial compression force in a member (see Section 8.4.2)
Nomb = elastic flexural buckling load of a braced member (see Section 8.3.3.2)
Nomb(kel) = Nomb as a function of a member effective length (see Section 8.3.3.2)
Nomb(l) = Nomb as a function of a member length (see Section 8.3.3.2)
Np = axial load in a multistorey shore (see Section 10.3.2)
Ns = design axial force at serviceability limit states (see Section 3.2.4.2)
Ns = nominal section axial compression capacity (see Section 8.4.2)

2013 Stephen Ferguson 1.9


Formwork Handbook

n = number of connected parallel columns (see Section 8.3.3.3)


Pch = horizontal concrete pressure acting on a vertical surface or the horizontal component from
concrete pressure acting on an inclined surface (see Section 4.3.1.1)
Pcv = vertical concrete pressure acting on a horizontal surface or the vertical component from
concrete pressure acting on an inclined surface (see Section 4.3.1.1)
Pcx' = component of concrete pressure limited by setting normal to an inclined form (see Section
5.4.2.2)
Pcy = component of concrete pressure limited by setting in the plane of an inclined form (see
Section 5.4.2.2)
pF = probability of failure (see Section 3.2.1)
Qah = horizontal action from construction activity (see Section 4.2.3)
Qgh = horizontal action on guardrails (see Section 4.2.3)
Qgv = vertical action on guardrails (see Section 4.2.3)
Qh = combined effect of horizontal variable actions (see Section 4.2.3)
D
Qm = vertical action from stacked materials and equipment (see Section 4.2.3)
Qser = occupancy live load (see Section 9.4.2)
Qv = combined effect of vertical variable actions (see Section 4.2.3)
Qw = vertical actions from workmen and equipment (see Section 4.2.3)
R
Qxh = other horizontal actions (see Section 4.2.3)
Qxv = other vertical actions (see Section 4.2.3)
R = maximum reaction (see Section 7.4.3)
R = resistance (see Section 3.2.1)
AF
Rc = vertical rate of concrete rise (see Section 5.2)
Rd = design resistance or capacity (Rn) (see Sections 3.2.2 and 7.5.2)
Rh = minimum rate of rise for full height hydrostatic pressure (see Section 5.3.1)
Rn = nominal resistance (see Section 3.2.1)
S = member axial stiffness (see Section 8.3.3.5)
Sc = elastic flexural stiffness of a concrete slab (see Section 10.2.2)
Sf = axial stiffness of any formwork or packing (see Section 10.2.3.1)
T
Sp = effective axial stiffness of a single formwork or multistorey shore (see Section 10.2.3.1)
Ss = axial stiffness of a shore (see Section 10.2.3.1)
Ss = serviceability limit states snow and water actions (see Sections 4.2.4.2 and 4.2.4.3)
Su = ultimate limit states snow and water actions (see Sections 4.2.4.2 and 4.2.4.3)
Sx = plastic section modulus (see Section 8.4.2.1)
s = standard deviation for the grade being assessed (see Section 9.5.1)
T = average ambient temperature (see Section 9.3.2)
T = floor cycle time in days (see Sections 10.1.3 and 10.2)
Tc = concrete temperature at placement (see Section 5.2)
tc = setting time of the concrete (see Section 5.3.3)
tp = thickness of ply (see Section 3.2.4.2)

1.10 2013 Stephen Ferguson


Chapter 1 - Introduction

V = maximum shear force (see Section 7.4.3)


Vc = rate of concrete delivery (see Section 5.3)
W = unfactored construction load on the slab (see Section 9.4.2)
W1 = construction load imposed on the top slab to be shared by multi-storey shoring (see
Section 10.2.3)
Wc = share of W1 transferred to each respective concrete slab by the multistorey shores (see
Section 10.2.2)
Wd = design construction load on the slab (see Section 9.4.2)
WLL = working load limit (see Section 3.2.5)
Wp = share of W1 carried by each respective level of multistorey shoring (see Section 10.3.1.1)
Ws = serviceability limit states wind actions (see Section 4.2.4.1)
Wser = unfactored design service load (see Section 9.4.2)
Wsh = horizontal serviceability limit states wind action (see Section 4.3.1.2)
D
Wsv = vertical serviceability wind action (see Section 4.3.1.2)
Wu = ultimate limit states wind action (see Section 4.2.4.1)
Wult = strength limit states design service load (see Section 9.4.2)
Wuh = horizontal ultimate limit states wind action (see Section 4.3.2.1)
R
Wuv = vertical ultimate limit states wind action (see Section 4.3.2.1)
w = uniformly distributed load (see Section 7.4.3)
Z = section modulus of the uncracked section (see Section 9.4.1)
= concrete pressure reduction factor (see Section 5.4)
AF
= coefficient for simply supported beams (see Section 7.4.3)
= numerical factor which depends on the number of spans (see Section 8.3.3.3)
= reduction factor to take account of joint behaviour (see Section 8.3.3.5)

m = ratio of the smaller to the larger end bending moments (see Section 8.4.2)
= partial load factor (see Section 3.2.1)
d = strength load factor for primary members (see Section 4.3.2.3)
T
p = serviceability load factor for concrete pressure (see Section 4.3.1.1)
= member deflection (see Sections 7.4.3 and 8.3.6)
A = deflection of the slab supported by the multistorey shore at the point where the shore is
located (see Section 10.3.2)
B = deflection of the slab supporting the multistorey shore at the point where the shore is
located (see Section 10.3.2)
l = limiting value of the serviceability parameter (see Section 3.2.4)
o = initial member out-of-straightness (see Section 8.3.6)
s = design serviceability parameter (see Section 3.2.4)
= section parameter (see Section 8.4.2.1)
c = elastic buckling load factor of the whole frame for fully braced falsework (see Section
8.4.4)

2013 Stephen Ferguson 1.11


Formwork Handbook

m = lowest buckling load factor for all compression members for fully braced falsework (see
Section 8.4.4)
ms = elastic buckling load factors for each storey for falsework that can sway (see Section
8.4.4)
= coefficient of static friction (see Sections 7.5.2 and B.1)
= factor for the slab continuity and support conditions (see Section 10.2.2)
= wet density of concrete (see Section 5.2)
Wc = accumulated share of unfactored multistorey construction loads (see Section 10.6.2)
= capacity reduction factor (see Sections 3.2.1, 3.2.4.2, 8.4.2, 9.4.1 and B.1)
= angular change at joints (see Section 8.3.5.2)
p = capacity reduction factor for shore axial stiffness (see Section 10.2.3.1)

References
D
ISO (1998). ISO 2394:1998 General principles on reliability for structures. Geneve, International
Organization for Standardization.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
SA (2002). AS/NZS 1170.0 - 2002 Structural design actions Part 0: General principles. Sydney,
R
Standards Australia.
SA (2003). Amendment No. 1 to AS 3610 - 1995 Formwork for concrete. Sydney, Standards
Australia.
SA (2010). AS 3610 - 2010 Formwork for concrete Part 1: Documentation and surface finish. Sydney,
AF
Standards Australia.
T

1.12 2013 Stephen Ferguson


Chapter 2 - Safe formwork design and construction

2
Safe formwork design and construction
(Procedures, roles, responsibilities, and requirements)
2.1 Introduction
Lessons learnt from past failures are often useful to focus attention on what is important when
considering the procedures, roles, responsibilities and requirements of those involved in formwork
design and construction. In this regard, studies into the cause of construction failures agree
procedural inadequacies enable flaws in the design and/or construction to go undetected, which lead
to failure and collapse(Bragg 1975) and (Hadipriono and Wang 1986).
D
The types of procedural inadequacies identified include: confusion of responsibilities amongst
participants; communication difficulties; inadequate briefing of designers; the lack of design drawings
or inadequate drawings; inadequate checking of designs (particularly those containing novel features);
unapproved modifications of the initial design; or failure to inspect work prior to loading.
Although structural failure and catastrophic collapse may pose one the greatest risks, other risks and
R
hazards also demand attention, such as: falling from heights, strains from manual handling, crushing
by moving plant and equipment, injuries from falling objects, trip and slip hazards, etc.
Safe formwork design and construction is a simple process that involves consultation,
communication, and coordination among participants to:
1. Specify project and construction requirements
AF
2. Prepare a formwork design brief
3. Design and document the formwork
4. Review, validate, check and certify the formwork design
5. Coordinate and supervise the formwork construction; and
6. Inspect and certify the formwork construction.
The purpose of this process is many fold, including:
T
a) Prior to construction:
(i) To enable selection of an appropriate formwork system;
(ii) To identify, assess and, where practicable, eliminate risk in the formwork design;
(iii) To identify, communicate, and make recommendations on minimising risks that
have not been eliminated and remain inherent in the formwork design;
(iv) To confirm the design fulfils the requirements set out in the design brief; and
(v) To identify and eliminate flaws in the formwork design
b) During construction:
(i) To communicate the design details for procurement and construction;
(ii) To identify, assess, and eliminate or, otherwise, minimise new risks in the
formwork design;
(iii) To assess and eliminate or, otherwise, minimise previously identified risks in the
formwork design; and

2013 Stephen Ferguson all rights reserved 2.1


Chapter 2 - Safe formwork design and construction

(iv) To identify and eliminate flaws in construction.


2.1.1 Participants
It is useful to first define the terminology used herein to describe the key participants in this process
and their responsibilities, see Table 2.1.
Table 2.1: Key participants and their roles or responsibilities
Participant Role or Responsibility
Project designer Architectural and structural design of the project under construction.
Construction contractor Construction of the project
Formwork coordinator Coordination of the formwork design and construction
Formwork manufacturer Manufacture of formwork material or components
Formwork supplier Supply of formwork material and components
Formwork designer Design of the formwork
D
Formwork checker Check and certification of the formwork design.
Formwork contractor Construction of the formwork
Formwork supervisor Supervision of the formwork construction
R
Formwork inspector Inspection and certification of the formwork construction
Other trades Those with access to the formwork construction; e.g. steel fixers,
concrete finishers, etc.
AF
These definitions are intended to be generic and are not intended to reflect any particular contractual
arrangement.
A participant may be an individual person or organisation or may refer to multiple persons and/or
organisations. For example, depending on the context: the term project designer may refer to the
project architect or the project structural engineer; the term formwork designer may refer to
geotechnical engineers or formwork engineers, where both participate in and contribute to the
formwork design; or formwork components may be sourced from more than one formwork supplier.
A person or organisation may participant in more than one way. For example: an organisation may
manufacture and supply formwork components; the organisation constructing the formwork may have
T
also designed the formwork; or the organisation responsible for the architectural and structural design
of the project, may also be responsible for the project construction.
Some projects may not involve all participants. For example: situations where formwork is not
fabricated off-site.
2.1.2 Scope
This Chapter sets out the procedures, roles, responsibilities and requirements consistent with safe
formwork design and construction for each participant, namely: project designer (see Section 2.2),
construction contractor and formwork coordinator (see Section 2.3), formwork supplier (see Section
2.4), formwork designer (see Section 2.5), formwork checker (see Section 2.6), formwork contractor
(see Section 2.7), formwork supervisor (see Section 2.8); and formwork inspector (see Section 2.9).
Section 2.10 provides guidance on obligation and competency for those involved in design, checking
and inspecting formwork.

2013 Stephen Ferguson all rights reserved 2.2


Chapter 2 - Safe formwork design and construction

2.2 Project designer


The project designer has an obligation to be aware of, and ensure the design is achievable using,
current construction practice, methods and workmanship (SA 1996).
Work health and safety regulations (SWA 2011) also place obligations on project designers to ensure,
so far as is reasonably practicable, that the structure is designed to be without risks to the health and
safety of persons who:
At a workplace, construct the structure;
At a workplace, carry out any reasonably foreseeable activity in relation to the
manufacture, assembly, use, demolition and disposal of the structure; and
At or in the vicinity of a workplace, are exposed to the structure or whose health and
safety may be affected by an activity related to the structure.
To fulfil these obligations, the project designer must communicate, consult, and work together with
those involved in the construction about potential risks and solutions.
The project designer must provide information in relation to the formwork, which AS 3610.1 2010
(SA 2010) calls project documentation, that:
D
(a) Specifies the requirements associated with the design and construction of the concrete
structure and its elements (SA 2010); and
(b) Provides information in relation to hazards and risks at or in the vicinity of the
construction site, including specifying the hazards relating to the design of the structure
that create a risk to the health and safety of those carrying out the construction work .
R
Where the project documentation is incomplete or unclear, the project designer must respond in
writing to requests from the construction contractor for missing information or clarification.
2.2.1 Project documentation
The project documentation must communicate specific requirements associated with the design and
AF
construction of the concrete structure and its elements; including, where appropriate, instructions and
information relating to concrete surface finish.
Guidance is provided in AS 3610.1 2010 that the project documentation must:
(a) Show dimensioned drawings of general arrangements, plans, elevations, sections and all
necessary details of the concrete structure, elements, special features, pre-camber,
penetrations and mandatory joints to be formed, as well as, locate and detail cast-in items;
and
(b) Specify details relating to surface finish and tolerances, and where relevant: colour
T
control, tonal scale, test panels, surface treatment, critical face of elements, location of
any special measuring points, and repairs.
In addition, the project designer must identify and communicate any situations or loading conditions,
which might arise during construction that are: hazardous; adversely affect the stability, strength, or
serviceability of the partially complete structure (SA 2002). Guidance is also provided in AS/NZS
1170.0 Supplement 1:2002 (SA 2002) that special loading conditions and unusual load paths, which
arise during construction, may need special investigation. For example, investigate construction loads
due to the stacking of materials or the use of equipment, or induced by floor-to-floor propping. The
BCA (ABCB 2013) also requires the magnitudes of actions from construction activity to be
determined.
AS 3610.1 2010 echoes these requirements by requiring the project designer to specify in the project
documentation:
(a) The location and magnitude of the design service loads;
(b) The sequence and timing of concrete placement (if critical);

2013 Stephen Ferguson all rights reserved 2.3


Chapter 2 - Safe formwork design and construction

(c) Any limitations on the magnitude and locations of constructions loads (e.g. stacked
materials);
(d) Loads from the effects of prestress or post-tensioning;
(d) Limitations on the use of the permanent or the existing structure for the restraint of the
formwork;
(e) Minimum stripping times and procedures, or the criteria for the determining minimum
stripping times;
(f) The method of multistorey shoring (e.g. undisturbed or reshoring), the minimum number
of levels, layout and load distribution among the supports relative to the type of
formwork, timing and sequence of its use, the anticipated time between construction of
subsequent floors and the expected ambient temperature during construction of the
permanent structure;
(g) Information about, and any special requirements for, propping or concreting any
composite construction or permanent form systems; and
(h) Information about the foundation that is relevant to the design of the formwork.
2.2.1.1 Location and magnitude of design service loads
D
Knowledge of the design service loads will allow the formwork designer and checker to be aware of
situations where the capacity of the permanent structure may be exceeded during construction and
where direction from the project designer is required.
2.2.1.2 Sequence and timing of concrete placement
R
The sequence and timing of concrete placement may be critical. For example:
The minimum period between pouring successive floors in multi-storey buildings;
The maximum freestanding height of walls above the uppermost slab; and
If columns can be poured at the same time as the slabs they support.
AF
2.2.1.3 Limitations on the magnitude and locations of constructions loads
Where no limitations on the magnitude and location of stacked materials are specified in the project
documentation, the formwork designer is permitted to design for a construction live load of 1.0 kPa
and an additional construction live load from stacked materials of up to 4.0 kPa (SA 1995).
Specifying limitations on the magnitude and locations of construction loads will reduce the risk of
overload.
2.2.1.4 Loads from the effects of pretensioning or post-tensioning
T
The effect of stressing may cause an upward camber, lifting the member, transferring the weight of the
member off the formwork beneath, and redistributing the weight to other parts of the formwork or
structure. Formwork or parts of the structure may be overloaded if the formwork designer is unaware
of such effects.
Stressing may also compress forms and make them difficult to strip.
2.2.1.5 Limitations on the use of the permanent or existing structure for formwork
restraint
It is common practice for formwork designers to assume the permanent structure (newly cast columns
and walls) stabilises the formwork system by providing lateral restraint in the plane of the formwork
soffit. However, this clearly contradicts the guidance provided in the Commentary to AS 3610.1
1995 (SA 1996), specifically that:
(a) If no limitations on the use of the permanent or existing structure for the restraint of the
formwork are specified in the project documentation, the formwork designer cannot
assume that the permanent structure is capable of restraining the formwork assembly and
either the formwork designer shall:

2013 Stephen Ferguson all rights reserved 2.4


Chapter 2 - Safe formwork design and construction

(i) Assume that the permanent structure cannot be used; or


(ii) Obtain written permission from the project designer; and
(b) The project designer must check the capacity of the permanent or existing structure to
resist the applied loads and restrain the formwork.
2.2.1.6 Minimum stripping times and procedures
Stripping affects the surface finish quality, durability and structural reliability of the concrete work.
Premature stripping may cause physical damage to the surface finish, contribute to non-uniformity of
colour and impede hydration. Structurally, early stripping may lead to cracking, increased long term
deformations, overload and possibly collapse.
For practical and economic reasons, stripping times need to be as short as possible, see Chapter 9.
2.2.1.7 Method of multi-storey shoring
The project designer is responsible for specifying the minimum number of floors of formwork and
multi-storey shores and the load distribution between the floors relative to the type of formwork,
timing and sequence of construction (SA 2010). However, this can only be determined and specified
after the construction contractor has informed the project designer of the details of the proposed
D
construction method, schedule, formwork details and shore layout, see Chapter 10.
After the project designer has specified the load distribution between the floors, the formwork
designer is responsible for ensuring the formwork and the shores themselves are not overloaded (SA
1996). Project designer approval for any changes to the proposed formwork shores and layout must be
obtained.
R
2.2.1.8 Requirements for composite construction or permanent form systems
Information and details of composite construction or permanent formwork that serves a structural or
architectural function should be specified.
In particular, details of the support and alignment of composite construction or permanent formwork
should be provided.
AF
Formed concrete tolerances specified in AS 3610.1 2010 do not apply to composite construction or
permanent formwork and the permitted tolerances should be given in the project documentation.
2.2.1.9 Information about the foundation that is relevant to the design of the formwork.
Where information on the bearing capacity and settlement characteristics of foundation material are
known, upon request it should be made available to the formwork designer. Similarly, the formwork
designer should be informed if ground slabs were not designed or are unable to support the load from
formwork shores or multi-storey shoring. Particularly as the latter may be transferring the full weight
of all the slabs above, see Clause 10.1.2.2.
T
2.3 Construction contractor
The construction contractor has an obligation to identify and manage the risks associated with the
construction, including formwork (SA 2002). Managing those risks involves consulting, coordinating,
and sharing information with all those involved with the formwork.
To that end, the construction contractor should:
a) Document a construction plan and work methods (construction documentation) that
eliminate or, otherwise, control the risks; and
b) Appoint a formwork coordinator to coordinate, consult with, communicate with and be
the first point of contact for all participants involved in the formwork.
2.3.1 Construction documentation
The construction documentation should set out specific requirements and details of the construction
method, schedule, equipment, and logistics that affect the formwork. as well as, the requirements of
other trades (e.g. those involved in fixing reinforcement; installing and stressing post-tensioning;

2013 Stephen Ferguson all rights reserved 2.5


Chapter 2 - Safe formwork design and construction

concrete supply, placement and delivery; installation of plumbing and electrical; etc.) and anyone else
needing access to the formwork.
The construction documentation may include a concept of the formwork (system) required.
Guidance is provided in AS 3610.1 2010 that the construction documentation must include, where
applicable:
(a) Details of any planned changes to the project documentation;
(b) Construction method, sequence and schedule;
(c) Relevant details of the plant and equipment to be used in the construction;
(d) Details of the interface between the formwork and other construction equipment and
activities;
(e) Information on construction activities and constraints that affect the formwork;
(f) Information on construction loads the formwork must support and any re-distribution of
loads;
(g) Requirements of access, egress and edge protection that affect the formwork;
D
(h) Documentation for equipment, material and components to be incorporated in the
formwork;
(i) Plans for transporting, handling, moving and reuse of the formwork;
(j) Method and rate of concrete delivery, placement sequence, discharge heights and rate of
rise; and
R
(k) Information on the concrete ingredients or admixtures that have a retarding affect on the
concrete setting and finish, including, but not limited to:
(i) Retarding setting of the concrete;
(ii) Causing excessive blow holes on the concrete surface; and
AF
(iii) Affecting colour control.
In addition, the construction documentation should include details of hazard identification, assessment
and risk control measures, as well as safe work method statements that affect the formwork.
Importantly, the construction documentation should identify any situations or loading conditions that
might arise during construction and that are: hazardous; adversely affect the stability, strength, or
serviceability of the partially complete structure; or must be taken into account in the formwork
design. Such situations or conditions might arise due to the particular chosen construction method,
sequence, equipment, or schedule.
T
2.3.2 Formwork coordinator
The formwork coordinator should possess the ability and authority to ensure the proper execution of
all procedures and that the roles, responsibilities and requirements of those involved are
communicated, clearly understood, and fulfilled.
Accordingly, in choosing a formwork coordinator account should be taken of the following:
a) The formwork coordinator should be competent, and possess qualifications and
experience appropriate for the project;
b) The formwork coordinator should be familiar with the requirements of the formwork
designer, the project designer, construction contractor, formwork contractor, and other
trades;
c) The formwork coordinator should have up-to-date knowledge of the requirements of
relevant authorities, Australian Standards, codes of practice, and work health and safety
regulations;
d) On large construction sites, coordinating the formwork may be a full-time position; and
2013 Stephen Ferguson all rights reserved 2.6
Chapter 2 - Safe formwork design and construction

e) Potential conflicts and additional risks that might arise; e.g. where a person is responsible
for both coordinating the formwork and general construction progress.
The formwork coordinator has a key role in safe formwork design and construction. They must
coordinate, consult, communicate with and be the first point of contact for all participants involved in
the formwork design and construction. Specifically, the formwork coordinator must take the following
steps:
1. Meet with and ensure that each participant understands their role and responsibilities;
2. Prepare the formwork design brief;
3. Ensure the formwork is designed and the design documented;
4. Where more than person or organisation contributes to the formwork design, appoint
someone to act as the overall formwork designer, particularly in regard to the interface
between designs and their compatibility, as well as, overall stability and robustness.
5. Distribute the relevant design information to all participants involved in the construction;
6. Ensure the requirements of the formwork design are understood by all participants
involved in the construction;
D
7. Formally review and validate that the design satisfies the design requirements;
8. Ensure the formwork design is checked and certified;
9. Communicate any feedback and required changes to the formwork designer. Where
changes are required to the design, repeat steps 5 to 8;
R
10. Ensure the formwork material and components supplied and/or fabricated are all
inspected and certified that they comply with the design;
11. Ensure all stages of the construction are monitored to ensure the formwork is handled,
assembled, erected, fixed, stripped, stored and dismantled in accordance with the design;
12. Prior to concrete placement, ensure the formwork is inspected and certified that it
AF
complies with the design; and
13. After concrete placement, ensure the formwork is not removed prematurely.
2.3.2.1 Formwork design brief
Prior to commencing the formwork design, the formwork designer must be properly briefed. To do
this the formwork coordinator must prepare and document a formwork design brief.
The purpose of the formwork design brief is to set out all the information and data relevant to the
formwork design, as well as all the general and specific requirements that the formwork must satisfy
during its working life.
T
Simply specifying The formwork must comply with AS 3610 Formwork for concrete is inadequate.
For example: the design loads given in AS 3610 1995 (SA 1995) may not be sufficient for formwork
of unusual construction or subject to unusual or more adverse loads (e.g. slipform, climbform, and
jumpform), eccentricity for formwork supports higher than 8 m is not considered, no guidance is
provided on the use or design of reshoring multi-storey structures, guidance for calculating the lateral
concrete pressure is limited to cement types and admixtures used in the 1980s, guidance on minimum
stripping times is conditional and based on out-of-date concrete data, etc.
General formwork requirements
General requirements that the formwork must satisfy can be found in the relevant Work Health and
Safety Regulations, Australian Standards, Codes of Practice, project specifications, and/or contract
conditions.
Formwork must be safe, must be fit for its purpose, and must consistently perform as intended
throughout its design working life. As a minimum, the formwork design brief should require:
1. Hazards to be identified and eliminated, or if not reasonably practicable, the risk of injury
controlled;
2013 Stephen Ferguson all rights reserved 2.7
Chapter 2 - Safe formwork design and construction

2. The formwork satisfy stability, strength and serviceability limit states; and
3. Formwork to possess structural integrity; i.e. possess a minimum level of connectivity
and robustness, as well as resist progressive collapse.
These minimum requirements should form the basis of any formwork design brief.
Chapter 3 provides an overview and guidance on these minimum general design requirements that
formwork must satisfy.
Specific formwork requirements
To determine the specific requirements the formwork must satisfy, it will be necessary to carefully
consider all matters that might affect the formwork. Much of this information can be found in the
project, construction, and proprietary documentation. If, in the first instance, the information available
in the project and construction documentation is incomplete or unclear, it will be necessary to request
the missing information or seek clarification from the relevant parties.
It will also be necessary for the formwork coordinator to meet with:
(a) The project designer;
(b) The construction contractor;
D
(c) The formwork contractor;
(d) The formwork and proprietary equipment suppliers, where known;
(e) Site health and safety representatives; and
(f) Related trades (e.g. reinforcement fixer, concrete supplier, concreter, electrician, plumber,
R
etc.).
The purpose of the meeting(s) is to ensure that, in a timely manner, the requirements and
responsibilities of each party are communicated, discussed, understood, agreed, and documented in the
brief. Documentation is essential to assist communication and minimise misunderstanding.
AF
A draft of the formwork design brief should be circulated to the project designer, construction
contractor, site health and safety representatives, related trades, formwork contractor, and formwork
designer for review and comment. The formwork coordinator should take account of all comments and
amend the draft formwork design brief accordingly.
It is important the final formwork design brief is prepared in sufficient time for all subsequent
activities; i.e. with time to design and document the formwork, check the formwork design, construct
the formwork, and inspect the formwork construction.
The completed formwork design brief should be provided to the project designer, construction
contractor, formwork contractor, and formwork designer. Where information or formwork
T
requirements are unclear or confusing, the formwork designer should request clarification and, if
necessary, meet with the formwork coordinator and relevant parties to ensure there is no
misunderstanding. Again the formwork design brief should be amended and reissued.
Without a clear understanding of all the requirements the formwork must satisfy, commencing the
formwork design may be unwise. Should the resulting formwork design not satisfy all the
requirements, redesign may be necessary.
Project, construction and proprietary documentation
In particular, the design brief must include detailed information of all the requirements relating to:
(a) The production of the concrete elements;
(b) The construction plan, method, logistics, equipment and schedule;
(c) Related trades and other users of the formwork;
(d) If known, preferred formwork systems, methods or equipment; and
(e) Existing stock to be incorporated in the design, including quantities available and
condition.
2013 Stephen Ferguson all rights reserved 2.8
Chapter 2 - Safe formwork design and construction

Details of and the requirements for the production of the concrete elements should be set out in the
project documentation.
The requirements and details of the overall construction method, schedule, equipment, and logistics
should be set out in the construction documentation, which should also include the requirements of
related trades that affect the formwork.
Specific details of the particular proprietary construction equipment, any preferred proprietary
formwork systems and existing stock to be used should be set out in proprietary documentation.
Copies of the relevant project, construction and proprietary documentation should be included in the
brief.
2.4 Formwork supplier
Formwork suppliers have obligations under Work Health and Safety Regulations, AS 3610 1995 and
AS 3610.1 2010.
2.4.1 Work health and safety
Formwork is defined under work health and safety regulations as a structure and prefabricated
formwork is defined as plant. Consequently, designers, manufacturers, importers and suppliers of
D
formwork have specific obligations under these regulations to:
1. Ensure, so far as is reasonably practicable, that the formwork supplied is without risks to
health and safety;
2. Register the design of prefabricated formwork 1 and issue the formwork design
registration number to the person with management or control of the prefabricated
R
formwork;
3. Eliminate or minimise risks to the health and safety of those persons:
(a) Using, handling, storing, assembling, and dismantling formwork;
(b) At or in the vicinity of the workplace whose health and safety are exposed to risk
AF
from the formwork; and
(c). Exposed to risks associated with noise, hazardous manual tasks, working in
confined spaces, and hazardous chemicals;
4. Carry out calculations, analysis, testing or examination that may be necessary to eliminate
or minimise risks;
5. With regard to the supply of second-hand prefabricated formwork:
(a) Identify any faults in the prefabricated formwork;
(b) Provide written notice of:
T
(i) The condition of the prefabricated formwork;
(ii) Any identified faults; and
(iii) If appropriate, any restrictions on the use of the prefabricated formwork that
apply until the faults are rectified.
6. Obtain information other participants (e.g. supplier, importer, manufacturer, etc) are
required to provide;
7. Provide adequate information to whom the formwork is supplied. The information
provided must include:
(a) The purpose for which the formwork was designed or manufactured;

1
Registration of prefabricated formwork is not required if the design was started before 1 January 2012 and
completed prior to 1 January 2014; otherwise, unregistered prefabricated formwork must not be supplied. This
requirement may not apply in all States.

2013 Stephen Ferguson all rights reserved 2.9


Chapter 2 - Safe formwork design and construction

(b) The results of any calculations, analysis, testing or examination carried out; and
(c) Any conditions necessary to ensure the formwork is without risks to health and
safety when used for the purpose for which it was designed or manufactured, as
well as when the formwork is used, handled, stored, assembled and dismantled,
including conditions or information regarding:
(i) Hazard identification and risk control measures;
(ii) Installation, commissioning, operation, and maintenance;
(iii) Cleaning, transport, storage and, where capable of being dismantled,
dismantling;
(iv) Systems of work necessary for safe use;
(v) Knowledge, training, skill or qualification necessary for persons;
(vi) Undertaking inspection and testing; and
(vii) Emergency procedures.
Refer to Work Health and Safety Regulations for information on obligations specific to risks
D
associated with noise, hazardous manual tasks, working in confined spaces, and the supply of
hazardous chemicals.
2.4.2 AS 3610 1995 and AS 3610.1 2010
AS 3610.1 2010 defines multiple-use or mass-produced formwork systems and/or components as
proprietary formwork.
R
In addition, AS 3610.1 2010 requires suppliers of proprietary formwork to provide information
(called proprietary documentation) that is required for its correct use. Proprietary documentation
should set out information, data and instructions for the correct and safe use of the proprietary
equipment.
Further guidance for formwork is provided in AS 3610.1 2010 that information provided by the
AF
supplier must include:
(a) Drawings or pictures that clearly identify the formwork;
(b) Adequate information to fully describe its intended use and any limitations thereto;
(c) Instructions for use and, where applicable, maintenance and disposal;
(d) The strength and serviceability limit state capacities in accordance with AS 3610 1995
and/or other relevant Australian Standards;
(e) The working load limit as calculated in accordance with AS 3610 1995.
T
(f) A statement that the formwork depicted in the documentation complies with AS 3610 (all
parts).
(g) Detailed information including, where appropriate:
(i) Part number;
(ii) Dimensions;
(iii) Section properties;
(iv) Weight;
(v) Any permanent camber built into the item;
(vi) Details of any special attachments, e.g., access brackets, hand rail posts standards,
plumbing feet; and
(vii) Locations for tie bolts or support points.

2013 Stephen Ferguson all rights reserved 2.10


Chapter 2 - Safe formwork design and construction

2.5 Formwork designer


The competency of the formwork designer and the level of detail provided in the formwork design
documentation must be commensurate with the level of risk. Guidance on the necessary competence
of the formwork designer is provided in Section 2.10.
Prior to commencing the formwork design, the formwork designer should:
(a) Be properly briefed, see Section 2.3.2.1;
(b) Review the formwork design brief;
(c) Request any missing information or data; and
(d) Seek clarification where information is unclear or confusing.
The formwork designer must take measures to eliminate or, where this is not reasonably practical,
control the risks identified, see Section 2.5.1.
A number of different formwork solutions are likely to fulfil the requirements of the formwork design
brief. Initially, the merit of all candidates should be considered. It may not be obvious early in the
design which formwork solution is best. In this case, each candidate solution should be investigated in
more detail. It may take several iterations of investigation, evaluation and elimination before the best
D
candidate solution becomes apparent.
It is important that the formwork designer provides for access to and egress from the formwork, as
well as access for working and inspecting the formwork. This is a key requirement and should not be
an after-thought. In addition, the formwork design should be based on concepts and details whose
realisation is achievable and can be checked on site (see Section 2.5.2).
R
The formwork designer must document the chosen formwork solution. In addition, the formwork
designer must report on the health and safety aspects of the design.
The formwork designer should provide the completed design documentation to the formwork
coordinator.
AF
2.5.1 Hazard identification, elimination or control
Formwork safety requirements are set out in the relevant state and national Work Health and Safety
(OH&S) Regulations, Codes of Practice, Guidance Notes, Hazard Profiles, Safety Alerts, etc.
Although similar, WH&S Regulations vary from State to State and some significant differences arise.
Prior to commencing any design, formwork designers must first consult the relevant WH&S
regulations to determine their obligations, responsibilities and the design standards and/or codes of
practice to which they must comply.
Ostensibly all WH&S regulations require the elimination or, if this is not reasonably practicable, the
control of risk of injury. In this sense, practicable means capable of being put into practice or action
T
having regard to the:
(a) Severity of the hazard or risk in question;
(b) State of knowledge about that hazard or risk and any ways of removing or mitigating that
hazard or risk;
(c) Availability and suitability of ways to remove or mitigate that hazard or risk; and
(d) Cost of removing or mitigating that hazard or risk.
Put simply, designers have a responsibility to design out hazards and ensure that, in the construction
and use of their designs, others are not subjected to unnecessary risk.
At the design stage, this might be achieved by: first identifying any foreseeable situations, loads or
conditions that are hazardous or adversely affect the stability, strength or serviceability of the
formwork; and, then assessing the associated risk. Formwork designers must consider hazards that
might arise during fabrication, transport, handling, assembly, construction, removal, dismantling,
storage, maintenance, and disposal of the formwork. Formwork designers must also take into account

2013 Stephen Ferguson all rights reserved 2.11


Chapter 2 - Safe formwork design and construction

the effects the design of related products or systems will have on the normal use, maintenance or
operation of the formwork.
Formwork designers must prepare a Safety Report that details the hazards identified during the design,
their assessment, design control measures taken, the hazards that remain in the design, and control
measures to be implemented by others. Later (e.g. prior to completion of the design and during design
checking), this assessment should be reviewed.
The Safety Report together with information on the proper operation and conditions of use of the
formwork must be provided to the formwork coordinator and all other relevant parties; e.g. formwork
supplier/manufacturer, formworker, etc. The level of detail provided should be commensurate with the
level of risk, see Section 2.8 for guidance.
2.5.1.1 Risk Control Measures
Where it is not practicable to eliminate the hazard, risks must be controlled by the highest possible
level of control from a hierarchy of controls; namely (in descending order from highest to lowest):
substitute or change, isolate or separate, engineered controls, administrative controls (instructions and
signs), and personal protective equipment.
Substitute or replace a hazard or hazardous work practice with a less hazardous one; e.g. assemble
D
formwork at ground level rather than at height, use remote release shackles to eliminate the need for
access at height.
Separate or isolate the hazard or hazardous work practice from those at risk; e.g. install screens,
barriers, fences or mark hazardous areas.
Control the hazard by engineering; e.g. install working platforms, edge protection, restraint and fall
R
arrest systems, machine or crane handling, operational controls, warning devices, and emergency
stops.
Administrative controls include compliance with codes of practice and design standards, limiting
exposure, specifying safe work methods, train operators, requiring minimum operator competencies,
supervision, inspections and permits.
AF
Personal protective equipment (PPE) should only be considered when all other control measures are
not practicable. PPE may not be appropriate to control some risks. PPE includes: safety helmets,
boots, gloves, hearing protectors, harness and lanyards, and respirators.
Although formwork designers may only have limited control over the workplace, they must ensure
health and safety of the work place by documenting (as part of the formwork design documentation,
see Section 2.5.3) the identified hazards, risk assessment, and the steps taken to eliminate or control
risks. This information together with information on the proper operation and conditions of use of the
formwork must be provided to the formwork coordinator and all other relevant parties; e.g. formwork
supplier/manufacturer, formworker, etc. The level of detail provided should be commensurate with the
T
level of risk, see Appendix A for guidance.
2.5.2 Practical requirements
The formwork designer must satisfy practical requirements, which include providing:
(a) Means for adjustment and stripping;
(b) Bracing that ensures the formwork can be safely erected and stripped;
(c) Devices to facilitate adjustment of the formwork and permit the controlled movement of
the formwork during stripping;
(d) Details to prevent movement and misalignment at construction joints in the concrete; and
(e) Positive means to prevent any movement would cause formwork components to become
unstable, dislodge or to collapse.
All adjustment devices should be designed such that the anticipated actions cannot dislodge them and
are not subject to uncontrolled movement under load.

2013 Stephen Ferguson all rights reserved 2.12


Chapter 2 - Safe formwork design and construction

Additional footings may be required to resist design actions and prevent instability. Accordingly, it
may be necessary to investigate the foundation material to determine its bearing and settlement
characteristics. Importantly, any movement of the footings that occurs before or after the initial set of
the concrete should not have detrimental effects on the cast-in-situ concrete or the capacity of the
formwork.
2.5.3 Formwork documentation
2.5.3.1 - General requirements
AS 3610.1 2010 requires the formwork to be designed and the design to be documented. Formwork
does not comply with AS 3610.1 2010 if it does not have formwork design documentation.
The purpose of the formwork design documentation is to communicate:
(a) The general arrangement, details and operation of the entire formwork construction, as
well as situation specific requirements, conditions and assumptions upon which the
design is based that must exist, apply or be satisfied; and
(b) Identified hazards, control measures, and safe methods of work.
The level of detail provided in the formwork design documentation should be commensurate with the
D
level of risk, see Section 2.10 for guidance.
Depending on the circumstance, the formwork design documentation may originate from a single
source or the formwork coordinator may have to bring together separate documentation from different
sources to make up the formwork design documentation; e.g. formwork documentation from one or
more suppliers for proprietary equipment, and documentation detailing footings from a civil/structural
engineer.
R
It is common practice for some formwork designers to limit the scope of their design and leave part of
the formwork design to others. Formwork design documentation is incomplete where details noted as
by client or by others are not detailed elsewhere.
It is the responsibility of the formwork coordinator to ensure the formwork design and documentation
AF
is complete, and in the latter case, the assembled documentation is compatible and contiguous.
2.5.3.2 Specific requirements of AS 3610.1 2010
To satisfy the requirements of AS 3610.1 2010, formwork documentation must include:
(a) Plans, elevations and sections sufficient to depict the general arrangement and details of
the formwork and to identify and locate all members and connections, including bracing
and footings/soleboards, lifting points and arrangements;
(b) Details of the type, quality and grade of all materials and components;
(c) Details sufficient to fully describe important or unusual features of the design;
T
(d) Copies of referenced proprietary documentation; and
(e) Where proof testing of the formwork assembly or its components is required: test loads,
arrangements, procedures and acceptance criteria.
The formwork design documentation must also set out all assumptions upon which the formwork
design is based and specific requirements that must be satisfied, including:
(a) Site geotechnical and environmental conditions;
(b) The location and magnitude of all permitted loads, including any limitations on when
they may be applied;
(c) Permitted imperfections and acceptance criteria for formwork components and
assemblies; e.g. member out-of-straightness, load and reaction eccentricities, out-of-
plumb erection, etc. (see Chapter 8);
(d) The locations where the permanent or existing structure is assumed to provide restraint to
the formwork, and the direction and magnitude of the loads the structure must resist, or
the stiffness required to effectively restrain the formwork;
2013 Stephen Ferguson all rights reserved 2.13
Chapter 2 - Safe formwork design and construction

(e) Sequence, method and rate of concrete placement and vibration; and
(f) Concrete ingredients and admixtures.
2.5.3.3 Proprietary documentation
Where proprietary equipment or components are used in the design, the design shall take account of
the information contained in the proprietary documentation and copies of proprietary documentation
shall be included as part of the formwork design documentation.
2.5.3.4 Safety report
In addition, the formwork design documentation must report on any foreseeable situations, loads or
conditions, which might arise during the design life of the formwork that are: hazardous; adversely
affect the stability, strength, or serviceability. The formwork design documentation should also detail
the measures taken to eliminate or, where this is not reasonably practical, control the risks
identified(SA 2002), see Section 3.2.
2.5.3.5 Multi-storey shoring
For multi-storey structures, the formwork design documentation must specify the type of multi-storey
shoring (e.g., undisturbed shoring or reshoring) and provide general arrangement drawings detailing:
D
the number of levels, layout, timing, and sequence, as well as the components to be used as shoring.
Guidance on the design of multi-storey shoring is provided in Chapter 10.
2.6 Formwork checker
Prior to the formwork construction, the formwork design and formwork design documentation should
be reviewed, checked, and a formwork design certificate issued.
R
The formwork coordinator should appoint, or agree to the appointment of, a competent person(s) to
check and certify the formwork design and documentation. Guidance on the obligation and
competence of the person(s) checking the formwork is provided in Section 2.10.
The formwork coordinator should provide the person(s) checking the formwork with a copy of the
formwork design brief and formwork design documentation.
AF
The fundamental purpose of the formwork design check is to detect any flaws in the formwork design
and/or formwork design documentation. To this end, the formwork design check must verify that the
formwork design and documentation is complete and satisfies the formwork design brief. Specifically,
that the formwork design documentation complies with the:
(a) Information and data set out in the formwork design brief;
(b) General and specific requirements set out in the formwork design brief; and
(c) General requirements set out in Chapter 3, where they are not included in the formwork
design brief.
T
In addition, the design check should review the identified hazards, as well as the risk assessment and
proposed control measures. The purpose of the review is to detect any foreseeable and unidentified
hazards, as well as flaws in the assessments of risk and selection of control measures.
If the formwork design documentation is found to be incomplete (e.g. refers to details by client or
by others that are not included in the formwork design documentation or does not include sufficient
detail or information):
(a) Checking should not proceed; and
(b) The formwork coordinator should be notified; and
(c) The formwork design documentation should be immediately returned to the formwork
coordinator for completion.
After the design check has been performed, the person checking the formwork design and
documentation shall issue the formwork coordinator with either of the following, in writing, as
applicable:

2013 Stephen Ferguson all rights reserved 2.14


Chapter 2 - Safe formwork design and construction

(a) If the formwork design and documentation complies with all the relevant requirements, a
certificate stating the formwork design and design documentation complies with the
design brief and all relevant requirements, including stating:
Their name(s) and qualifications;
Whether they were or were not involved in the original design,
The list of documents that have been reviewed;
Details of the checks undertaken; and
The specific requirements against which the design has been checked
or
(b) If aspects of the formwork design or design documentation do not comply with all the
relevant requirements, a report containing the details and descriptions of each area of
non-compliance.
The formwork coordinator should provide the formwork designer with a copy of the formwork design
certificate. Otherwise, where aspects of the formwork design do not comply:
D
1. The formwork coordinator should liaise with the person(s) checking the design and the
formwork designer on how best to rectify the aspects of the formwork design, or
formwork design documentation, that are non-compliant. Compliance might require a
completely different design; and
2. The formwork design and formwork design documentation should be amended and the
R
amended documentation re-issued for checking.
Formwork construction should not proceed unless the formwork has been designed, documented,
checked and certified compliant.
2.7 Formwork contractor
AF
The formwork contractor has an obligation to identify and manage the risks associated with the
formwork construction. Managing those risks involves competent people:
(a) Planning the formwork activity;
(b) Consulting, coordinating and sharing information with all those involved with the
formwork, especially the formwork designer;
(c) Consulting with workers and their health and safety representative;
(d) Preparing Safe Work Method Statements (SWMS);
(e) Training workers in matters specific to formwork and falsework activities to be
T
undertaken; and
(f) Supervising the formwork construction and its removal;
2.8 Formwork supervisor
Risks of flaws in the formwork construction can be minimised if it is carefully supervised throughout
and periodically inspected to ensure:
(a) Design data, information, and assumptions about site conditions are valid and applicable;
(b) There are no hazardous situations or loads unforeseen in the design,
(c) The specified materials and components are used;
(d) Unidentified materials and components are not used;
(e) The specified safe work methods are followed;
(f) Hazards identified and remaining in the design have been assessed and controlled; and
(g) The formwork is constructed in accordance with the formwork design.
2013 Stephen Ferguson all rights reserved 2.15
Chapter 2 - Safe formwork design and construction

2.9 Formwork inspector


Prior to concrete placement, the formwork construction should be inspected and formwork inspection
certificate issued.
The formwork coordinator should provide the person(s) inspecting the formwork with a copy of the
formwork design brief and formwork design documentation.
The purpose of inspecting the formwork is to detect any flaws in the construction prior to placing
concrete. This is achieved by comparing the as-built formwork with the formwork as designed and
documented. Flaws may arise when:
(a) Site conditions or loading situations differ from those foreseen in the design;
(b) The design documentation is incomplete or inappropriate; or
(c) The construction departs from the design details.
Where possible, any flaws identified should be rectified. Otherwise, details of any shortcomings, flaws
or disparity should be documented and forwarded to the formwork coordinator, who should seek
approval or direction from the formwork designer.
Any approved alterations to the formwork design or changes directed by the formwork designer
D
should be documented and returned to the formwork coordinator and, where appropriate, checked.
In multi-storey structures, both the formwork and the multi-storey shoring should be inspected.
When the formwork construction and site conditions are all in accordance with the formwork design
documentation, the person inspecting the formwork construction shall certify in writing that the
R
formwork construction complies with the design, including stating:
Their name(s) and qualifications;
Whether they were or were not involved in the design or checking,
The details of formwork construction that has been inspected;
AF
Time, date and details of the inspection undertaken;
The specific formwork design documents against which the construction has been
compared;
If aspects of the formwork construction do not comply with the formwork design documentation, the
person inspecting the formwork shall provide in writing details and descriptions of each area of non-
compliance.
Section 2.10 provides guidance on the necessary competence of the formwork inspector. General
guidance on inspecting formwork can be found in references (CS 2003) and (CS 1999).
T
2.9.1 Approval to load formwork
Concrete placement should not be permitted without approval and permission in writing from the
formwork coordinator.
Approval and permission in writing should not be granted until the formwork coordinator has received
documentation confirming that the formwork has been designed, documented, checked, certified
compliant, inspected, and complies with the design.
2.9.2 Approval to remove formwork
Formwork or multi-storey shoring must not be disturbed, removed, or stripped without approval and
permission in writing.
The formwork coordinator must seek written approval (from the project designer) prior to permitting
the disturbance, removal, or stripping of any formwork. Approval from the project designer may take
the form of a set of criteria that must be satisfied prior to formwork removal. In this case, the
formwork coordinator must verify that all the criteria have been satisfied before permitting formwork
removal.

2013 Stephen Ferguson all rights reserved 2.16


Chapter 2 - Safe formwork design and construction

Premature or unauthorised disturbance or removal of any part of the formwork or multi-storey shoring
may be hazardous and may have detrimental effects on the surface finish, serviceability, strength, or
stability of the concrete structure.
Chapter 9 provides general guidance on criteria for stripping formwork and multi-storey shoring.
2.10 Obligation and competence
The concept of Formwork Risk Level provides a useful framework for specifying appropriate levels of
obligation and competence for situations with different levels of risk.
For formwork design, documentation, checking, and certification, as well as inspecting the formwork
construction, Table 2.2 sets out appropriate levels of obligation and competence for situations with
different levels of risk. A method for selecting the appropriate level of risk for different situations is
set in Appendix A.
Table 2.2 Appropriate levels of obligation and competence for situations with different levels
of risk

Level Formwork design and Formwork design check Formwork inspection and
of documentation and certification certification
D
Risk
Obligation Competence Obligation Competence Obligation Competence

Low Mandatory Experienced Optional Experienced Mandatory Experienced

Moderate Mandatory Experienced Mandatory Qualified Mandatory Experienced


R
High Mandatory Qualified Mandatory Independent Mandatory Qualified

The level of competency required in Table 2.2 increases with risk. Accordingly, the levels of the
competence are, or the approved equivalent of:
ExperiencedA person who has a minimum of 4 years site experience in construction of the
AF
particular type of work.
QualifiedA professional engineer who has qualified as a Member of the Institution of Engineers
Australia (MIEA) and who has a minimum of 4 years experience in the design and construction of the
particular type of work.
CharteredA professional engineer who is a member of the Institution of Engineers Australia with
the status of Chartered Professional Engineer (CPEng) or a person registered on the National
Professional Engineers Register (NPER) or if there is a law that provides for registration of
professional engineers, is a registered professional engineer, and who has a minimum of 4 years
experience in the design and construction of the particular type of work.
T
IndependentA professional engineer who is a member of the Institution of Engineers Australia with
the status of Chartered Professional Engineer (CPEng) or a person registered on the National
Professional Engineers Register (NPER) or if there is a law that provides for registration of
professional engineers, is a registered professional engineer, and who has a minimum of 4 years
experience in the design and construction of the particular type of work, and employed by an
organisation not involved in the original design.
In practice, most formwork is designed by experienced persons who may not possess formal
qualifications. Some authorities (WorkCover NSW 1998) do not require the formwork design be
checked and certified, but rather, require a qualified person to inspect and certify the formwork
construction prior to concrete placement where the level of risk is moderate or high.
Unfortunately, inspections are often called at the last minute at a time when the formwork may be
incomplete. It may be difficult to access all parts of the formwork. Inspectors may be unfamiliar with
project requirements, the formwork brief and design. Furthermore, construction cost and time
pressures may sway inspectors to haste or to compromise and approve formwork or modifications
thereto, which in other circumstances would not have happened and may have delayed concrete

2013 Stephen Ferguson all rights reserved 2.17


Chapter 2 - Safe formwork design and construction

placement. Where inspections are the only form of checking, especially in less than ideal
circumstances, there is a higher risk that flaws in the design or construction will go undetected.
For the majority of formwork (with a moderate level of risk), Table 2.1 requires the design be checked
by a qualified person prior to construction commencing. In this way, design flaws should be detected
prior to construction and subsequent inspection, thereby reducing the onus on inspectors and reducing
risk.
References
ABCB (2013). Building Code of Australia (BCA), Australian Building Codes Board.
Bragg, S. L. (1975). Final report of the Advisory Committee on Falsework. London, Her Majesty's
Stationery Office: 151.
CS (1999). Checklist for Erecting and Dismantling Falsework. Berkshire, The Concrete Society.
CS (2003). Checklis for Assembly, Use and Striking of Formwork. Berkshire, The Concrete Society.
Hadipriono, F. C. and H.-K. Wang (1986). "Analysis of causes of formwork failures in concrete
structures." Journal of Construction Engineering and Management 112: 112 - 121.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
D
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney, Standards
Australia.
SA (2002). AS/NZS 1170.0 Supplement 1:2002 Structural design actions - General principles -
Commentary. Sydney, Standards Australia.
R
SA (2010). AS 3610.1 - 2010 Formwork for concrete Part 1: Documentation and surface finish.
Sydney, Standards Australia.
SWA (2011). Work Health and Safety Regulation 2011. Canberra, Safe Work Australia.
WorkCover NSW (1998). Code of Practice - Formwork. Sydney, WorkCover NSW.
AF
T

2013 Stephen Ferguson all rights reserved 2.18


Chapter 3General designrequirements

3
General design requirements
3.1 Introduction
Formwork should satisfy the requirements set out in the relevant Work Health and Safety Regulations,
Australian Standards and Codes of Practice. To do so, formwork should satisfy fundamental structural
requirements to: be safe, be fit for its purpose, and consistently perform as intended through-out its
design working life.
As a minimum, the formwork design should:
1 Identify hazards and assess the risks;
D
2 Eliminate the hazards, or if not reasonably practicable, control the risk of injury;
3 Satisfy stability, strength and serviceability limit states; and
4 Possess structural integrity; i.e. have a minimum level of connectivity, robustness and
resist progressive collapse.
R
The first two points are discussed at length in Chapter 2. This Chapter provides guidance on
requirements to satisfy stability, strength and serviceability limit states specific to formwork. General
issues regarding structural integrity and the concept of design working life are also discussed.
The philosophy and principles of structural design presented herein are consistent with those set out in
ISO 2394:1998 General principles on reliability for structures (ISO 1998), and AS/NZS 1170.0
AF
Structural design actions Part 0: General principles (SA 2002).
Although not the focus of this text, it is important that the formwork is cost-effective. This Chapter
also discusses aspects related to formwork economy that should be taken into account in formwork
design.
3.2 Stability, strength and serviceability
3.2.1 Limit states design
In the context of structural design, the word state means the condition of a structure.
The fundamental concept of limit states is that a structure can be classified as either satisfactory
T
(serviceable, safe) or unsatisfactory (unserviceable, unsafe) (Gulvanessian and Holicky 1996). Thus,
if the condition of a structure exceeds any limit state, a limit state violation is said to have occurred as
it no longer satisfies the fundamental performance requirements.
Often people make statements such as Ive done it this way thousands of times before and never had
a problem. What does this really mean? It means that, on each past occasion, the load on the
structure was less than its capacity. How much less or whether it will be less on the next occasion, are
all unknowns. The structure may have been serviceable, but was it safe?
If it was possible to test structures to failure a sufficient number of times, it would be possible to use
the data collected to verify the reliability of a structure using purely probabilistic techniques.
Figure 3.1 expresses a simplistic relationship between the distribution of action effects E and
resistance R for a structure, in terms of probability density functions.

2013 Stephen Ferguson all rights reserved 3.1


Formwork & Falsework

D
Figure 3.1: Probability distributions for design action effects and design resistance
In Figure 3.1, Em represents the mean action effect and Rm represents the mean resistance. The shaded
area in Figure 3.1 represents the probability of failure pF, which can be expressed as:
R
p F = P[E > R ] (3.1)

Equation 3.1 represents the probability that the action effects E exceed the resistance R; i.e. E > R. A
situation where E > R constitutes a limit state violation, whether or not this leads to the collapse of part
or all the structure depends on the structures limit states behaviour.
AF
In limit states design, the probability of failure is controlled by separate partial factors for actions and
resistance. Usually, this is expressed as:

E n Rn (3.2)

In Equation 3.2, En is the nominal action effect, Rn is the nominal resistance, is a partial load factor,
and is a capacity reduction factor. The values for the partial factors and methods of determining the
nominal action effect and resistance are specified in the applicable structural Standard. They are
carefully chosen to take account of uncertainties about the probability distributions of the action
T
effects and resistance, as well as the mode and consequence of failure.
Factors affecting the probability of failure, and therefore influencing the choice of partial factors
include: choice of the values of actions; degree of structural integrity; accuracy of structural models
used; quality and durability of materials and equipment; site conditions; environmental conditions;
quality of workmanship; and measures taken to reduce the risk of gross human, design and
construction errors.
In some situations, the choice of partial factors should ensure an even lower probability of failure. For
example where:
(a) The risk of injury, economic, social and environmental losses is greater; or
(b) Collapse occurs suddenly and without warning, rather than where collapse is preceded by
some kind of warning in such way that measures can be taken to limit the consequences.
However, satisfying Equation 3.2 does not guarantee that a structure or part of it will not fail. Failure
may occur due to:

3.2 2013 Stephen Ferguson all rights reserved


Chapter 3General designrequirements

An extremely unfavourable combination of actions, material properties, geometrical


quantities, etc., all of which are associated with ordinary use and other ordinary
circumstances;
Effects of exceptional but foreseeable circumstances and/or actions present only during a
small portion of the design working life and/or with a low probability. For example,
accidental impact or extreme climatic influences;
Consequences of an error, such as lack of information, omission, misunderstanding and
lack of communication, negligence, misuse, etc; and
Influences that are not foreseen.
Formwork cannot be expected to function adequately if exceptional actions or exceptionally low
resistance occur, but measures should be taken to limit the scope of the expected damage. These
measures should not be disproportionate to the original cause.
A practicable approach is to design the formwork for ordinary use in ordinary circumstances and take
the following additional measures.
(a) Design the structure with minimum levels of structural integrity and to avoid progressive
collapse, see Section 3.4;
D
(b) Take protective measures against foreseeable actions. For example, safeguard against
impact by providing additional protection such as bollards; and
(c) Reduce the probability of gross design and construction errors by appropriate quality
assurance and/or quality control measures; e.g. follow the procedures set out in Chapter 2.
R
3.2.2 Stability (limit states)
Formwork must be stable (i.e. resist sliding, overturning and uplift) under extreme and/or frequently
repeated actions.
Stability limit states are concerned with the loss of equilibrium of the formwork or any part of it,
considered as a rigid body, due to overturning, uplift and sliding. For example, the possibility of
AF
suspended slab formwork and falsework overturning, lifting or sliding under extreme wind actions or a
cantilever bearer subject to overturning if only the cantilever is loaded. It is good practice to check
stability limit states first.
Stability limit states are satisfied by ensuring the design action effects of destabilising actions do not
exceed the combined design action effects of the stabilising actions and design resistance.

Ed,dst Ed,stb + Rd (3.3)

where
T
Ed,dst = design action effect from destabilising actions (see Section 4.3.2);
Ed,stb = design action effect from stabilising actions (see Section 4.3.2); and
Rd = design resistance or capacity (Rn).
For example, stability limit states for overturning would be satisfied if the design moment from forces
causing overturning do not exceed the combined effects of the design moment from forces that have a
stabilising effect and any nominal resistance the structure may have; i.e. Md,dst Md,stb + Mn.
The partial load factors and action combinations for stability limit states are set out in Section 4.3.2.
3.2.3 Strength (limit states)
Formwork must resist extreme and/or frequently repeated actions.
Strength limit states are concerned with the failure of the formwork or part of it due to yield, rupture,
fatigue or excessive deformation. Strength limit states also include: instability of the formwork or part
of the formwork due to buckling; the transformation of the formwork or part of it into a mechanism;
and a sudden change of the structural system to a new system (e.g. snap through).

2013 Stephen Ferguson all rights reserved 3.3


Formwork & Falsework

Strength limit states are satisfied by ensuring the design action effect does not exceed the design
resistance for capacity.

Ed Rd (3.4)

where
Ed = design action effect (see Section 4.3.2); and
Rd = design resistance or capacity (Rn).
The partial load factors and action combinations for strength limit states are set out in Section 4.3.2.2.
3.2.4 Serviceability (limit states)
Formwork must perform adequately under all expected actions.
Serviceability limit states are associated with the performance of formwork under conditions of
normal use. Serviceability limit states are satisfied if the serviceability deflection, deformation or
frequency does not exceed the serviceability limit.

s l (3.5)
D
where
s = design serviceability parameter (deflection, deformation or vibration
frequency) determined on the basis of the appropriate combination of actions
(see Section 4.3.1).
R
l = limiting value of the serviceability parameter.
In particular, consideration should be given to the following serviceability limit states:
(a) Concrete surface finish quality;
(b) Concrete positional and dimensional tolerance; and
AF
(c) Elastic behaviour of multiple use equipment.
The partial load factors and action combinations for serviceability limit states are set out in Section
4.3.1.
3.2.4.1 Serviceability limits for formwork deformations
T

Figure 3.2: Cumulative deflections (McAdam 1993)

3.4 2013 Stephen Ferguson all rights reserved


Chapter 3General designrequirements

Formwork usually consists of three layers (form face, secondary and primary beams). Surface
undulations result from the cumulative effects of the combined deflections and dimensional variations
due to permitted formwork material tolerances. Figure 3.2 depicts the magnitude of surface
undulations resulting from the combined deflections of horizontal and vertical formwork framing
members only. If form face deflections were taken into account the total deflection may increase.
In addition to member deflection, variant formwork material and fabrication compounded by
imprecise erection will also detract from the formwork quality and consequently the concrete surface.
For example, deformations in the concrete surface will arise from variations due to:
permitted dimensional tolerances of graded timber;
welding heat deformations during fabrication steel formwork; or
the presence of small gaps between framing members.
AS 3610
AS 3610.12010 (SA 2010) classifies and specifies the requirements for the physical quality and
colour control of the concrete surface finish. There are five classes (1 to 5) of surface finish. Where
colour control is incorporated it is denoted by the suffix C following the surface finish number, e.g.
Class 2C.
D
Class 1 is the highest attainable quality and should only be specified for use in very special cases.
Class 2 has uniform quality and texture and is commonly specified for architectural work. Class 3 is
specified with the intention that the concrete is to be viewed as a whole. Classes 4 and 5 are specified
when the visual quality is not important. Class 4 has good general alignment, while for Class 5 even
alignment is not important.
R
Guidance on the design and detailing of formwork to achieve colour control is beyond the scope of
this text, refer to (CS 1999, CCAA 2006, ACI 2013). However, guidance is provided herein on the
appropriate serviceability deflection limits to use in design.
The stiffness of formwork is important because it affects both the quality of the surface finish and
dimensional accuracy of the concrete. Formwork deformations should not exceed the limits specified
AF
in AS 3610.12010 Clause 3.3.4 and Table 3.3.2. The acceptable surface finish deformation caused
by form face deflection and concrete surface undulations given in AS 3610.12010 Table 3.3.2 are
repeated here in Table 3.1.
Table 3.1: Acceptable form face deformations and surface undulations

Quality of surface finish Class 1 Class 2 Class 3 Class 4 Class 5


Form face deformation Lesser of Lesser of Span/270 Span/270 NA
(not greater than) 2 mm or 3 mm or
span/360 span/270
T
Percentage of readings 95 100 90 100 80 100 70 100 70 100
Surface undulations (mm)
For l = 300 mm, NA NA
(a - b) 1 2 2 4 3 4 5 7

Surface undulations (mm)


For l = 1500 mm, NA NA
(a - b) 2 4 3 6 5 7 8 10

The format for the acceptable deformation for surface undulations is useful for assessing the physical
quality of the concrete surface (see AS 3610.12010 Clause 5.2.2(b) and Figure 3.3 below). However
it is not convenient for the purpose of formwork design and AS 3610.12010 does not provide
guidance on the acceptable limits for surface undulations for a given span.

2013 Stephen Ferguson all rights reserved 3.5


Formwork & Falsework

Figure 3.3: Measuring surface undulations (SA 2010)

In this case, it is useful to consider, the limits for surface undulations expressed in terms of
straightedge length (l) as span to deflection ratios, see Table 3.2.
Table 3.2: Acceptable surface undulations expressed as span to deflection ratios
D
Quality of surface finish Class 1 Class 2 Class 3 Class 4

Percentage of readings 95 100 90 100 80 100 70 100


Surface undulations
For span, l = 300 mm
(a - b) l/300 l/150 l/150 l/75 l/100 l/75 l/30 l/43
R
Surface undulations
For span, l = 1500 mm
(a - b) l/750 l/375 l/500 l/250 l/300 l/214 l/187 l/150
AF
The values expressed in Table 3.2 are intended to result in a concrete surface finish that will comply
with the requirements of Table 3.1, but may be more stringent.
For example, consider using a straightedge that is 1500 mm long to check undulations for a Class 2
surface. Assume readings a and b are taken only 600 mm apart (l = 2 x 600 = 1200 mm) such that (a -
b) = 3 mm and satisfies both the 90 and 100 percentage limits of Table 3.1. When expressed as a span
to deflection ratio, l/(a - b) = 1200/3= 400, which could be interpreted to mean Table 3.1 permits
undulations of up to l/400 rather than the more stringent criteria of l/500 specified in Table 3.2.
By observation, for short spans the requirements for form face deflection are more severe than surface
T
undulation limits for a similar length straightedge (i.e. where l = 300 mm) and will govern design. For
longer spans, where the deflection of secondary and primary members plays a major role, the limits for
surface undulation will govern serviceability limit states for the deflection of secondary and primary
members.
AS 3600
In addition, formwork must also satisfy are the tolerances specified in AS 3600 Concrete structures
(SA 2009), which provides permitted tolerances for plumb, dimensions and surface alignment of
concrete structures and members. In particular, consideration should be given to the:
floor-to-floor plumb tolerance for columns and walls of 1/200 or 10 mm, whichever is the
greater;
deviation from specified height, plan or cross-sectional dimension of 1/200 times the
specified dimension or 5 mm, whichever is the greater; and
deviation from surface alignment, in that the deviation of any point on a surface of a
member, from a straight line joining any two points on the surface, shall not exceed 1/250
times the length of the line.

3.6 2013 Stephen Ferguson all rights reserved


Chapter 3General designrequirements

Recommended serviceability limits for member deflection


Serviceability limits for deflection must satisfy the more severe of the requirements set out in AS
3610.12010 and AS 3600.
Form face deflection and surface undulations will run parallel with the members primarily responsible
for the deformation. Therefore, measurements of form face deflection and surface undulations should
be taken in the direction of the undulation; i.e. with straight-edge parallel to the member span (form
face, secondary or primary member). Thus, these deformations are primarily the result of the
deflection of a single member.
However, measurements for surface alignment to AS 3600 may be taken in any direction and therefore
may take account of the accumulated deflection of two or more members. Thus, to satisfy the AS 3600
surface alignment limit of 1/250 times the length of the line (span), the total deflection (secondary plus
primary member deflection) must be less than span/250.
Table 3.3 shows recommended serviceability limits for deflection for use in the design of formwork
members. They are intentionally more stringent than the acceptable concrete surface deformations and
alignment tolerances permitted in AS 3610.12010 and AS 3600 because they are intended to make
allowance for variant formwork material and fabrication, as well as deflection, and satisfy both AS
3610.12010 and AS 3600.
D
Where light shines across a concrete surface at a flat angle, particularly in the case of smooth glossy
finishes (e.g. anti-graffiti paint), the recommended deflection limits specified in the Table 3.3 may not
be appropriate. Under these conditions, surface imperfections appear exaggerated and less deflection
is desirable. High quality materials must be used and a high standard of workmanship is required.
In addition to controlling deflection, formwork should have sufficient stiffness, mass, or both, to avoid
R
any detrimental effects of vibration on its structural capacity, tolerances and surface finish.
Where formwork acts as only as a working platform or only to provide access and egress, it must
satisfy the serviceability limits set out in the relevant Standard; e.g. AS 1576.1 or AS 1657.
AF
Table 3.3: Recommended serviceability limits for formwork member deflection

Serviceability limit Class 1 Class 2 Class 3 Class 4 Class 5


1 Form face deflection Lesser of Lesser of Span/300 Span/300 NA
1 mm or 2 mm or
span/500 span/300

2 Secondary member Span/750 Span/500 Span/300 Span/300 NA


deflection
T
3 Primary member Span/750 Span/500 Span/300 Span/300 NA
deflection

3.2.4.2 Serviceability limits for multiple-use equipment


For multiple-use equipment, irreversible deformations that result from fatigue and yielding such as
bending, squashing and elongation may have a detrimental effect on strength and might render the
equipment unserviceable.
Accordingly, for multiple-use formwork to remain serviceable, it is important that yielding does not
occur at serviceability limit states. In particular, three areas of concern arise: bending, yielding of
threaded steel form ties and hole elongation due to steel ply bearing failure. In each case, AS 4100
Steel Structures (SA 1998) permits plastic behaviour at ultimate limit states, which in itself is not a
problem, except that the ratio of ultimate to serviceability actions is often lower in formwork than for
normal structures.

2013 Stephen Ferguson all rights reserved 3.7


Formwork & Falsework

For multiple-use members subject to compression or tension, strains at serviceability limit states
should be checked to ensure they remain elastic, namely:

N s An f y (3.6)

where
s = design axial force at serviceability limit states
= capacity factor for tension (see AS 4100 Table 3.4)
An = net area of the cross-section
fy = yield stress used in design

When form ties are fabricated from high strength steel, yielding at serviceability limit states is not an
issue. However, hole elongation is a more serious problem.
For example, it is common that formwork shores are fabricated from steel circular hollow sections.
Height adjustment is achieved by telescoping close fitting sections, which are connected by a shear
D
pin. The ubiquitous adjustable steel prop and shore frame are examples of this arrangement. If these
members were designed in accordance with AS 4100, at ultimate limit states hole elongations up to
60% could be expected (Bridge, Sukkar et al. 2002). In particular, hole elongation in formwork shores
is undesirable because it contributes to unanticipated load redistribution. Thus, limits need to be
placed on the ply-bearing stresses that occur at serviceability limit states.
R
For pin or bolt connections in multiple use steel formwork, in addition to the strength requirements set
out in AS 4100, it is recommended applying the following serviceability requirement that the design
ply bearing force should be limited (Bridge, Sukkar et al. 2002) such that:

N bs 1.6d f t p f y (3.7)
AF
where
Nbs = design ply bearing force at serviceability limit states;
= capacity factor for ply bearing (see AS 4100 Table 3.4)
df = diameter of bolt or pin;
tp = thickness of ply; and
fy = yield stress of steel ply.
T
Satisfying this requirement should limit hole deformations to 2% of hole diameter.
3.2.5 Working load limit (WLL)
AS 3610.12010 requires suppliers of proprietary formwork to publish the strength and
serviceability limit states capacities and working load capacity of proprietary formwork, as calculated
in accordance with AS 36101995 (SA 1995). The requirement to publish both limit states and
working load capacities is intended to minimise the risk of misunderstanding and possible
overloading of the formwork up to its limit states capacity, if only the limit states capacities were
published.
The working load limit (WLL) should satisfy the following conditions for strength and serviceability:

Rn
WLL (3.8)
LSD

and

WLL Es.max (3.9)

3.8 2013 Stephen Ferguson all rights reserved


Chapter 3General designrequirements

where
Rn = strength limit states design resistance or capacity;
LSD = limit states divisor that satisfies LSD 1.5, unless a lessor value is justified
by a rigorous statistical analysis of load and capacity data using probability
methods (SA 2003) 1.
Es.max = maximum action effect satisfying serviceability limit states.
3.3 Structural integrity
Formwork must satisfy minimum structural integrity requirements so that the formwork is not
damaged disproportionally as a consequence of impact or due to human error. In addition, formwork
must resist progressive collapse.
D
R
AF
T
Figure 3.4: Impact damage but not failure (sourced from WorkSafe Victoria)
3.3.1 Robustness
Formwork should be designed so that any damage due to impact or occurring as a consequence of
human error is not disproportionate to the original cause. Figure 3.4 shows an example of a robust
structure. Here the bridge gantry has withstood the impact from the truck tray body, which has
become detached from the truck. There is significant damage, but the gantry did not collapse.
One of the most common causes of formwork falsework collapse is a lack of connectivity and
inadequate bracing (Bragg 1975) and (Hadipriono and Wang 1986). The risk of this occurring can be

1
Prior to the publication of SA (2003). Amendment No. 1 to AS 3610 - 1995 Formwork for concrete. Sydney,
Standards Australia., AS 3610 - 1995 Clause A4.4.4 permitted working load capacities to be derived from the
limit state capacity using a divisor of less than 1.5 (based on the load factors given in AS 3610 - 1995 Table 5.1);
however, calculating the working load capacity using this method was unreliable.

2013 Stephen Ferguson all rights reserved 3.9


Formwork & Falsework

reduced by providing minimum levels of strength, continuity and ductility, i.e. connections should be
designed to be ductile and have a capacity for large deformations under the effect of abnormal actions.
To this end, formwork members and connections should be designed, as a minimum, to resist lateral
loads equivalent to 2.5 percent of the vertical actions, respectively, see Section 4.2.6.3.
3.3.2 Preventing progressive collapse
To reduce the risk of progressive collapse, the designer must identify key structural elements whose
failure would cause the collapse of more than a limited portion of the formwork and then:
(a) If possible, redesign the formwork in such a way that local damage does not lead to
immediate collapse of the whole formwork or a significant part of it; or
(b) If this is not possible, the design should take their importance into account by considering
specified exceptional actions that cover the majority of unforeseen events such as
accidents or similar occurrences.
3.4 Design working life
AS/NZS 1170.0:2002 (SA 2002) defines the design working life of a structure as the minimum
number of years a structure or structural element is to be used for its intended purpose, with required
D
maintenance but without structural repair being necessary.
For permanent structures, the minimum design working life might be considered to be 25 years or
more. For temporary structures such as formwork and falsework, a shorter period may be appropriate.
For example, custom formwork made for particular project may be in use for less than one year and
then scrapped. On the other hand, during its design working life proprietary equipment might be
R
used numerous times over 25 years or more.
For the purpose of determining the magnitude of environmental actions, the design working life of a
particular assembly of formwork components could be considered as the period the particular
assembly is exposed to the environment, see Section 4.2.4.
3.5 Australian Standards
AF
National formwork Standards set out specifications and procedures to ensure that formwork is fit for
its purpose and consistently performs as intended. Currently, as amended, Australian Standard AS
36101995 Formwork for concrete sets out requirements for the design and testing of formwork and
Australian Standard AS 3610 2010 Formwork for concrete Part 1: Documentation and surface finish
sets out the requirements for documentation, surface finish and construction. Sometimes, compliance
with AS 3610 is a requirement of WH&S regulations or AS 3610 is specified as a Code of Practice.
3.5.1 Australian formwork Standards: a brief history
The first Australian formwork Standards were AS CA70-1971 Design and Construction of Formwork
T
and AS CA72, Part 11972 Control of Concrete Surfaces Formwork.
In 1974, these Standards were superseded by AS 15091974 (SAA 1974a) and AS 1510 Part 1
1974 (SAA 1974b), respectively. Essentially, AS 1509 and AS 1510.1 were metric versions of CA70
and CA72 with a few other minor changes.
In 1975, the Standards Association of Australia established a policy of a general unified approach for
the design of all types of structures using limit states design (SAA 1975). In 1984, in keeping with
that policy, Standards Committee BD/43 was formed to write a new Standard that would include limit
states design rules for formwork.
In 1990, AS 3610 Formwork for concrete (SA 1990) was published, together with two Supplements.
AS 3610 set out the requirements for the design, fabrication, erection, and stripping of formwork, as
well as the specification, evaluation and repair of the quality of the formed concrete surface.
Supplement 1 provides additional copies of photographic charts for surface finish and colour control.
Supplement 2 provides a commentary to AS 36101990.
Significantly, AS 36101990 was the first national Standard to introduce limit states formwork design
methods. In fact, AS 3610 1990 set out both permissible stress and limit states methods; however,

3.10 2013 Stephen Ferguson all rights reserved


Chapter 3General designrequirements

the permissible stress methods were only intended for use until other relevant material Standards were
available in limit state format.
Later in 1995, AS 3610 was revised but remained ostensibly unaltered. The current versions of the
Standard and Supplements are AS 36101995, AS 3610 Supplement 11995 (SA 1995) and AS
3610 Supplement 21996 (SA 1996).
In 1997, Standards Australia embarked on a revision of AS 36101995 and a draft (SA 1999) was
issued for public comment in October 1999. Shortly thereafter, potential short comings in the draft
were identified and it was decided that further research was required before the Standard could be
published.
In 2003, Standards Australia issued Amendment No 1 to AS36101995 (SA 2003) and AS 3610
Supplement 2 1996 (SA 2003) as an interim measure to address shortcomings identified in the limit
states action combinations. These shortcomings arose because AS 3610 was written when there was
little statistical data available on the action effects and resistance of formwork. Therefore, the design
methods were specified based on experience and judgement and generally followed the design rules
for permanent structures. Subsequent data became available and a disparity could be demonstrated
(Ferguson 2003) between the reliability of the limit states action combinations in AS 36101995 and
international practice, as well as target reliability indices.
D
In 2005, Standards Australia issued a second draft (SA 2005) for public comment. Subsequently, it
was decided to split the Standard into parts, but to date only one part has been published.
In 2010, Standards Australia published AS 3610.1 2010 Formwork for concrete Part 1:
Documentation and surface finish. The content of AS 3610.12010 supersedes Sections 2, 3 and 5, as
well as part of Clause 4.7, of AS 36101995. Until withdrawn, AS 3610 1995 will coexist with AS
R
3610 2010 Part 1.
3.5.2 Other Australian Standards relevant to formwork
Table 3.4 Applicable Standards for various materials used in formwork

Concrete AS 3600 Concrete structures.


AF
Steel AS 4100 Steel structures
AS/NZS 1554.1 Structural steel welding Welding of steel structures.

Cast steel AS 2074 Cast steels


AS 1998 Welding of steel castings.

Cold-formed Steel AS/NZS 4600 Cold-formed steel structures.

Timber AS 1720.1 Timber structures Part 1: Design methods,


AS 2082 Timber Hardwood Visually stress-graded for structural purposes
T
AS 2858 Timber Softwood Visually graded for structural purposes

Laminated Veneer AS/NZS 4357 Structural Laminated Veneer Lumber


Lumber (LVL)

Masonry AS 3700 Masonry structures

Aluminium AS 1664 Aluminium structures Limit state design

Plywood AS/NZS 2269 Plywood Structural


AS 2271 Plywood and blockboard for exterior use

The design of all structures, including formwork and falsework, must comply with the requirements of
AS/NZS 1170 Structural design actions Parts 0 to 4. Specifically:
AS/NZS 1170.0:2002 Structural actions Part 0: General principles (SA 2002);
AS/NZS 1170.1:2002 Structural actions Part 1: Permanent, imposed and other actions
(SA 2002a);

2013 Stephen Ferguson all rights reserved 3.11


Formwork & Falsework

AS/NZS 1170.2:2002 Structural actions Part 2: Wind actions (SA 2011);


AS/NZS 1170.3:2003 Structural actions Part 3: Snow and Ice actions (SA 2003); and
AS 1170.4:2007 Structural actions Part 4: Earthquake actions in Australia (SA 2007).
In addition to the actions specified in AS/NZS 1170, loads and load combinations that need to be
considered in formwork design are included in AS 36101995 and updated in Chapter 4.
As formwork can be constructed using combinations of many different materials, formwork designers
need to take account of the information and procedures specified in a range of Standards. Now that all
relevant permissible stress material Standards have been withdrawn, it is appropriate to use the limit
states Standards shown in Table 3.4.
3.6 Economy
Formwork represents a significant cost component of a concrete structure. The overall cost of
formwork is a function of many factors, including: design, materials, transport, storage, handling, as
well as labour availability, skill and productivity. Importantly, greater economies may be achieved
and false economies avoided by analysis of the construction of the concrete structure as a whole rather
than just considering the formwork in isolation.
D
Nowhere is the saying "time is money" more true than in the construction industry. Often, material
economies in the design of the structure (e.g. reducing wall thickness at height) are in fact false
economies, when the effect on construction is taken into account (e.g. delays while system formwork
is modified).
From a formwork perspective, the greatest economies (in labour, material and time) are commonly
R
achieved through repetition and reuse. Repetition leads to higher productivity and opportunities to
automate occur when formwork can be reused many times with little or no change.
Economies also arise on each occasion formwork can be constructed from stock components with little
or no modification or cutting. Project designers might realise these savings by judicious selection of
building dimensions.
AF
Where the quality of the concrete surface finish is important, false economies arise when short cuts in
design and construction of formwork result in non-conformance and costly remedial work.
Economies may also be achieved in formwork material handling. This happens where mechanical and
automated machines can increase productivity by handling large formwork assemblies, negating the
need for them to be dismantled and re-assembled after each use.
References
ACI (2013). Guide to Formed Concrete Surfaces. Farmington Hills, American Concrete Institute.
Bragg, S. L. (1975). Final report of the Advisory Committee on Falsework. London, Her Majesty's
T
Stationery Office: 151.
Bridge, R. Q., T. Sukkar, I. G. Hayward and M. Van Ommen (2002). "The behaviour and design of
structural steel pins."
CCAA (2006). Guide to Off-form Concrete Finishes. Sydney, Cement and Concrete Association of
Australia.
CS (1999). Technical Report 52: Plain formed concrete finishes. Berkshire, The Concrete Society.
Ferguson, S. A. (2003). Limit states design of steel formwork shores Master of Engineering
(Honours), University of Western Sydney.
Gulvanessian, H. and M. Holicky (1996). Designers' Handbook to Eurocode 1: Part 1 Basis of design.
London, Thomas Telford.
Hadipriono, F. C. and H.-K. Wang (1986). "Analysis of causes of formwork failures in concrete
structures." Journal of Construction Engineering and Management 112: 112 - 121.
ISO (1998). ISO 2394:1998 General principles on reliability for structures. Geneve, International
Organization for Standardization.

3.12 2013 Stephen Ferguson all rights reserved


Chapter 3General designrequirements

McAdam, P. S. (1993). Formwork: A practical approach. Brisbane, Stuart Publications.


SA (1990). AS 3610 - 1990 Formwork for concrete. Sydney, Standards Australia.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
SA (1995). AS 3610 Supplement 1 - 1995 Formwork for concrete - Blowhole and colour evaluation
charts. Sydney, Standards Australia.
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney, Standards
Australia.
SA (1998). AS 4100 - 1998 Steel Structures. Sydney, Standards Australia.
SA (1999). DR99481 Formwork for concrete (Draft Australian Standard - Revision of AS 3610-1995).
Sydney, Standards Australia.
SA (2002). AS/NZS 1170.0 - 2002 Structural design actions Part 0: General principles. Sydney,
Standards Australia.
SA (2002a). AS/NZS 1170.1:2002 Structural design actions - Part 1: Permanent, imposed and other
actions. Sydney, Standards Australia.
D
SA (2003). Amendment No. 1 to AS 3610 - 1995 Formwork for concrete. Sydney, Standards
Australia.
SA (2003). Amendment No. 1 to AS 3610 Supplement 2 - 1996. Sydney, Standards Australia.
SA (2003). AS/NZS 1170.3: Structural design actions - Part 3: Snow and Ice actions. Sydney,
Standards Australia.
R
SA (2005). DR05029 Formwork for concrete (Draft Australian Standard - Revision of AS 3610-1995).
Sydney, Standards Australia.
SA (2007). AS 1170.4: Structural design actions - Part 4: Earthquake actions in Australia. Sydney,
Standards Australia.
AF
SA (2009). AS 3600-2009 Concrete Structures. Sydney, Standards Australia.
SA (2010). AS 3610.1 - 2010 Formwork for concrete Part 1: Documentation and surface finish.
Sydney, Standards Australia.
SA (2011). AS/NZS 1170.2: Structural design actions - Part 2: Wind actions. Sydney, Standards
Australia.
SAA (1974a). AS 1509-1974 SAA Formwork Code. Sydney, Standards Association of Australia.
SAA (1974b). AS 1510 Part 1 - 1974 Control of Concrete Surfaces - Formwork. Sydney, Standards
Association of Australia.
T
SAA (1975). AS 1793-1975 Limit State Design Method. Sydney, Standards Australia.

2013 Stephen Ferguson all rights reserved 3.13


Chapter 4 - Actions and action combinations

4
Actions and action combinations
D
R
AF

1. Weight of formwork and falsework (Gf)


2. Weight of concrete (Gc)
3. Vertical actions from workers and equipment (Qw)
T
4 Vertical and horizontal actions on edge protection (Qg)
5. Weight of stacked materials (Qm)
6. Horizontal actions from construction activity (Qah)
7. Accidental impact (Ah)
8. Concrete pressure (P)
9. Serviceability wind (Ws)
10. Ultimate wind (Wu)
11 Snow (Su)
12 Earthquake (Eu)
13. Flowing water (Su)
14. Trapped debris (Su)
15. Other vertical and horizontal actions, eg. thermal, shrinkage, prestress, etc. (Qxv, Qxh)
Figure 4.1 Formwork and falsework design actions

2013 Stephen Ferguson all rights reserved 4.1


Formwork Handbook

4.1 Introduction
Formwork should be designed to resist the effects of all foreseeable actions, including extreme,
frequently repeated and exceptional actions. In determining the magnitude of the design
actions, account should taken be taken of the probabilities of exceedence during the formworks
design working life.
For each separate design situation, the combined effects of simultaneously occurring actions
should be taken into account. Actions that are not spatially fixed should be applied where they
produce the most unfavourable effect. Actions that cannot occur simultaneously should not be
combined.
The magnitude and combination of actions on the formwork may vary during construction. It is
important to consider actions during each stage of construction, namely:
Table 4.1 Stages of formwork construction
Stage 1 Prior to concrete placement, during handling and erection of the
formwork as well as once the formwork is erected
Stage 2 During concrete placement
D
Stage 3 After concrete placement, while the formwork supports the
applied loads

It is important that the formwork designer is briefed by those in control of site activities. Based
R
on knowledge of the construction method, the designer should make a realistic assessment of
combined effects of actions that will act simultaneously.
4.2 Actions
It is important to take account of different types of actions:
AF
(a) Direct actions (e.g. external forces, loads);
(b) Indirect actions (e.g. imposed or constrained deformations due to temperature changes or
differential settlement; or imposed acceleration due to machine excitation or earthquake);
and
(c) Notional actions that are introduced to take account of structural imperfections that have a
significant effect on the behaviour of the formwork.
Direct actions can be classified as either: permanent, variable, or accidental. Permanent actions
act continuously with little variation in magnitude and at a specific location. As the name
T
suggests, the magnitude of variable actions fluctuates with time. Accidental actions are usually
of short duration. Unlike permanent actions, the point of application of variable and accidental
actions can be random.
4.2.1 Permanent actions (G)
4.2.1.1 Vertical actions from weight of formwork ( Gf)
In this text, the notation Gf replaces G used in AS 3610 1995 (SA 1995).
The weight of formwork is a permanent action and should include, where applicable, the weight
of:
(a) Any part of the permanent structure forming part of, or supported by, the
formwork;
(b) Any ancillary structure connected to the formwork;
(c) Forms;
(d) Falsework;

4.2 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

(e) Footings; and


(f) Counter-weights used to provide stability.
4.2.1.2 Vertical actions from weight of concrete ( Gc)
The weight of concrete should be considered as a permanent action (Gc).
The weight of un-reinforced concrete with dense aggregate is 24.0 kN/m3. For reinforced
concrete, add 0.6 kN/m3 for each 1% of reinforcement by volume (SA 2002a). It is common
practice to assume the weight of reinforced concrete is 25.0 kN/m3, but this may underestimate
the weight of heavily reinforced concrete or concrete mixes that contain heavy aggregate (e.g.
concrete made using iron ore as aggregate).
Why in formwork design is concrete treated as a permanent action and not a variable action?
Arguably, since concrete is the main action that formwork is designed to support, it should be a
variable action (say, Qc). However, the way actions vary in time determines whether they are
permanent or variable. Permanent actions are actions that are likely to act continuously
throughout a given reference period and for which variations in magnitude are small compared
with the mean value (ISO 1998). Similarly, variable actions are those for which the variation in
D
magnitude with time is neither negligible in relation to the mean value, nor monotonic.
In general, structural design Standards consider the weight of concrete as a permanent action.
The assumed coefficient of variation for concrete is in the order of 0.10, (Rosowsky, Huang et
al. 1994). Accordingly, a partial load factor of 1.2 is reasonable (Ellingwood, MacGregor et al.
1982).
R
International practice has been to consider the weight of concrete as a dead load (BS 5975 and
ACI 347) probably because designers were familiar with this concept and, conveniently,
permissible stress methods did not differentiate between the effects of permanent and variable
actions, negating the need for any change. Australian (SA 1995), Israeli (SII 1998) and
European (ECS 2004) limit states formwork Standards also consider the weight of concrete a
AF
permanent action.
4.2.2 Concrete pressure (Pc)
During concrete placement, wet concrete behaves as a quasi-fluid and exerts pressure on the
formwork. Initially, concrete behaves hydrostatically and the pressure increases proportionally
with the weight of the fluid concrete head. As concrete sets, increased fluid concrete head has a
reduced effect on concrete pressure.
The behaviour and pressure of wet concrete is discussed in detail in Chapter 5.
4.2.3 Vertical and horizontal variable actions (Qv and Qh)
T
Account should be taken of the most adverse effect of combinations of concurrently acting
imposed vertical and horizontal variable actions:
(a) Combined vertical variable action with the most adverse effect (Qv) will come from
action combinations that include one or more of the following:
Qw = vertical actions from workers, concrete mounding and equipment;
Qm = vertical actions from stacked materials and equipment;
Qgv = vertical actions on edge protection; and
Qxv = other vertical actions
The actions Qw and Qm are not considered to occur concurrently at the same
location and only the action with the most adverse effect need be considered.
(b) Combined horizontal variable action with the most adverse effect (Qh) will come
from action combinations that include one or more of the following:
Qah = horizontal actions from construction activity;

2013 Stephen Ferguson all rights reserved 4.3


Formwork Handbook

Qgh = horizontal actions on edge protection; and


Qxh = other horizontal actions.
To determine the most adverse effect, consider the most adverse combination of magnitude and
position for each set of actions.
4.2.3.1 Vertical actions from workers, concrete mounding and equipment ( Qw)
In this text, the notation Qw replaces Quv and Qc used in AS 3610 1995.
Stage 1 Prior to concrete placement
Prior to concrete placement, for areas of horizontal and sloping formwork that are trafficable,
allow a uniformly distributed vertical action for the weight of workers, which includes an
allowance for their personal tools, Qw1 1.0 kPa.
Stage 2 During concrete placement
During concrete placement, it is necessary to take account of: the weight of workers and their
personal tools, as well as the short-term dynamic effects of discharging concrete out of a skip or
pump, and any associated minor mounding of concrete.
D
R
AF
T
Figure 4.2 Typical construction activity during concrete placement (Stage 2) (Fattal 1983)
AS 3610 recommends considering either of two load situations:
(a) Quv2 1.0 kPa;
(b) Qc 3.0 kPa acting for a 5 minute duration over an area 1.6 m x 1.6 m square at any
location and zero over the remainder.
In addition, AS 3610 1995 specifies a strength limit states load factor = 1.5 for Quv and =
1.0 for Qc. Thus, at strength limit states the magnitude of the factored loads is 1.5Quv2 = 1.5 kPa
and 1.0Qc = 3.0 kPa.
Taking account of the effects of both loading conditions complicates calculations, especially
given the limited area over which Qc applies. A simpler approach is desirable.
It is equally reliable to take account of the effect of a uniformly distributed action of:
(a) For the design of formwork members and supports, Qw2 2.0 kPa; or
4.4 2013 Stephen Ferguson all rights
reserved
Chapter 4 - Actions and action combinations

(b) For the design of multistorey shoring (see Chapter 10), Qw2 1.0 kPa.
For the design of formwork members and supports, this approach is reasonable because at
strength limit states the magnitude of the factored loads is 1.5Qw2 = 1.0Qc = 3.0 kPa. In addition,
designing the formwork to support Qw2 = 2.0 kPa consequently provides a minimum allowance
for stacked materials during Stage 3 of Qm3 = 2.0 kPa, as Qw3 and Qm3 are not considered to act
concurrently at the same location.
Arguments that this approach is more severe, as Qw2 is not limited to an area of 1.6 m x 1.6 m,
do not take account of:
(a) For many members that span up to 1.6 m and whose tributary load width is less
than 1.6 m, there is no change;
(b) The design of members spanning more than 1.6 m is likely to be governed by
serviceability, which does not take account of Qw2; and
(c) The design of formwork supports is governed by the maximum vertical load, which
often occurs during Stage 3 due the effect of stacked materials or subsequent
construction activity.
D
Stage 3 After concrete placement
After concrete placement, allow for a uniformly distributed vertical action Qw3 1.0 kPa over
the concrete surface and trafficable areas of the formwork. This is adequate for the normal
traffic of workers, as well as an allowance for their personal tools, but insufficient for
mechanical equipment or actions from subsequent construction.
R
The magnitude of Qw3 may need to be increased to take account of the actions of the subsequent
construction activity on the newly placed slab.
Floors supporting multistorey shoring
When calculating the action effects in multistorey shoring, the majority of the construction
AF
activity occurs on the uppermost concrete slab (supporting the formwork for the next level to be
constructed) and allowing for a uniformly distributed action of Qw3 = 1.0 kPa should be
adequate.
There is usually less activity on the lower floors, connected by multistorey shoring, and
allowing for a uniformly distributed vertical action of Qw4 = 0.25 kPa should be adequate for
each floor supporting multistorey shoring, see Chapter 10.
Platforms
The design actions for temporary platforms, not part of the formwork, should comply with
AS/NZS 1576.1 (SA 2010).
T
For working platforms attached to formwork and used to provide a working area for workers
and their tools, consider the most adverse effect from either:
(a) a uniformly distributed vertical action of Qw 1.0 kPa; or
(b) a concentrated vertical force of Qw 1.2 kN applied through a 100 mm 100 mm pad at
any point.
For platforms, or parts of the formwork, that will be used to provide access and egress to and
from places of work, account should be taken of crowding by considering the most adverse
effect from either:
(a) a uniformly distributed vertical action of Qw 2.5 kPa; or
(b) a concentrated vertical force vertical of Qw 1.2 kN applied through a 100 mm 100 mm
pad at any point.
4.2.3.2 Vertical actions from stacked materials and equipment ( Qm)
In this text, the notation Qm replaces M used in AS 3610 1995.

2013 Stephen Ferguson all rights reserved 4.5


Formwork Handbook

During construction, it is difficult to prevent the common practice of stacking materials and
construction equipment on the formwork or upon newly placed concrete still supported by
formwork, see Figures 4.3 and 4.4.
Typically, stacked materials include: portable toilets; carpenters tables; column forms; metal
containers; tool boxes; barrels of water; bundles of reinforcement; bricks; sand; metal frames;
steel braces; timber; aluminium beams; and scaffold.
D
R
AF
Figure 4.3 Stacked materials (Stage 1)
T

Figure 4.4 Stacked materials and equipment (Stage 3)

4.6 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

Recommendations for stacked materials in AS 3610 1995


AS 3610 1995 requires the project designer to place limits on the timing, magnitude, and
location of stacked materials. In the absence of specified limits on stacked materials, AS 3610
1995 recommends the design load for stacked materials to be:
Stage 1: M1 = 4.0 kPa;
Stage 2: M2 = 0 kPa (i.e. materials are not stacked on wet concrete); and
Stage 3: M3 = 4.0 kPa.
AS 3610 1995 also considers the load from workers and equipment (Quv) and stacked
materials (M) act concurrently; i.e. the formwork design should take account of Quv + M; i.e.
Stage 1: Quv1 + M1 = 5.0 kPa;
Stage 2: Quv2 + M2 = 1.0 kPa (i.e. materials are not stacked on wet concrete); and
Stage 3: Quv3 + M3 = 5.0 kPa.
Recommendations for stacked materials in the literature
D
Guidance on the appropriate magnitude of stacked materials can be found in literature that
analysed data collected from site surveys that weighed every piece of material and mapped its
location (Ayoub and Karshenas 1994, Karshenas and Ayoub 1994). Ayoub and Karshenas
recommend that for formwork with tributary areas less than 28 m2, the weight of stacked
material can be considered equivalent to an uniformly distributed action of not less than 2.4 kPa.
When considering the design of formwork with a tributary area greater than 28 m2, the authors
R
recommend multiplying the design load by a load reduction factor

2.96
m = 0.44 +
A (4.1)
AF
Where m is not greater than 1.0 and not less than 0.83
Equation 4.1 effectively reduces the magnitude of the action from 2.4 kPa to approximately 2.0
kPa as the tributary area increases from 28 m2 to 56 m2.
Design recommendations
It is important that the formwork designer is briefed by those in control of site activities about
the construction method, equipment, components and plans for stacking materials.
It is reasonable to assume that workers will not traverse across the top of stacked materials. In
T
this case, Qw need not be considered to act concurrently with Qm. However, consideration may
need to be given to combinations where other loads act concurrently with stacked materials (Qm
+ Qxv); e.g. where materials are transported and stacked using mechanical equipment, such as a
fork lift.
It is common practice for materials to be placed on packing (short lengths of timber called
dunnage) that allows space for removing forklift tynes or lifting slings, see Figures 4.3 and
4.4. The effect of the packing is to concentrate the weight of stacked materials under the
packers. Figure 4.3 depicts material bins with legs that act as point loads.
For the design of formwork, the allowance for stacked materials should:
(a) not exceed any limitations specified in the project documentation;
(b) in the absence of any limitations placed on the magnitude of stacked materials in the
project documentation, not exceed 4.0 kPa without seeking approval from the project
designer;
(c) at each stage of construction, be based on a realistic assessment; and

2013 Stephen Ferguson all rights reserved 4.7


Formwork Handbook

(d) satisfy the following minimum criteria:


Stage 1: Qm1 2.5 kPa or the equivalent concentrated line or point loads.
Stage 2: Qm2 = 0 kPa.
Stage 3: Qm3 2.5 kPa.
For the design of formwork components with a tributary area greater than 30 m2, the allowance
for stacked materials may be reduced by a load reduction factor

3
m = 0.44 +
A (4.2)

The value of m cannot be greater than 1.0 nor less than 0.8.
Taking account of tributary area results in an allowance for stacked materials:
a) For tributary areas A 30 m2, Qm 2.5 kPa; and
D
7.5
b) For tributary areas 30 m2 < A < 70 m2, Qm 1.1 + kPa
A
c) For tributary areas A 70 m2, Qm 2.0 kPa.
Prior to concrete placement (Stage 1), the magnitude of stacked materials (Qm1) does not usually
R
govern formwork design. However, the effect of concentrated loads under packers should be
assessed.
After concrete placement (Stage 3), it is common practice to stack material upon recently
poured concrete slabs. At this stage, the formwork should effectively have an inbuilt minimum
reserve capacity to support some stacked materials, given for Stage 2 it has been designed for a
AF
minimum action of Qw2 = 2.0 kPa.
The designer should consider situations where the weight of stacked materials exceeds 2.5 kPa,
previously recommended. In some cases the weight of stacked materials can easily exceed 5.0
kPa (e.g. pallets of bricks).
For economy, the formwork designer may choose to limit the area of formwork (or slab) upon
which materials may be stacked. Any limit must be clearly depicted on the formwork drawings.
The magnitude of the allowance for stacked materials and any limitation placed thereon must
be:
T
(a) consistent with the proposed construction methods;
(b) known and approved by those in control of site activities; and
(c) noted and detailed on the formwork documentation.
For the design of multistorey structures and shoring, the accumulated weight of stacked
materials on several floors will have a significant effect on the number of levels, and magnitude
of the load carried by, multistorey shores and the load shared between supporting slabs. In most
cases, the resulting load will exceed the capacity of the slabs. However, where stacked materials
are permitted allowances should be based on realistic assessments such that Qm 2.0 kPa.
4.2.3.3 Horizontal actions from construction activity ( Qah)
In this text, the notation Qah replaces Quh used in AS 3610 1995.
Construction activities impose horizontal actions on the formwork. Actions may arise from the
individual and combined effects of: concrete pumping systems; the acceleration and
deceleration of trolleys, skips or other vehicles; cable tensions; and the actions of workers and
equipment.

4.8 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

An assessment of the magnitude of the horizontal actions should be made.


Nevertheless, the design horizontal action should be not less than 5 kN nor less than 1 kN/m
distributed uniformly along the edge of the formwork, whichever is the more severe, see Figure
4.5.
D
R
AF
Figure 4.5 Horizontal actions from construction activity acting on formwork a 20 m 10 m
concrete slab.
4.2.3.4 Vertical and horizontal actions on edge protection ( Qgv and Qgh)
Where edge protection is attached to formwork, design edge protection members and
connections to the formwork, as well as the formwork members supporting the edge protection,
to resist the effects of an action acting either: horizontally inwards, horizontally outwards, or
vertically downwards. The magnitude of the action shall not be less than the most adverse of
either:
T
(a) a concentrated force of 0.6 kN acting at any point on the top rail, edge or post; or
(b) 0.35 kN/m distributed uniformly along the top rail or edge.
Toeboards installed on edge protection attached to formwork shall be designed for a
concentrated horizontal force of 0.1 kN acting at any point.
4.2.3.5 Other vertical and horizontal actions (Qxv and Qxh)
In some circumstances, other vertical and horizontal actions arise during construction. Often
the effects of these actions are significant and therefore must be taken into account in the
formwork design.
Other sources of vertical and horizontal actions include:
(a) Manual or mechanical equipment (e.g. forklifts, cranes, etc);
(b) Prestress or post-tensioning; see Chapter 7.
(c) Axial shortening, shrinkage and creep of concrete;
(d) Buoyancy;

2013 Stephen Ferguson all rights reserved 4.9


Formwork Handbook

(e) Temperature change; and


(f) Imposed acceleration due to machine excitation.
4.2.4 Environment actions (wind, snow, water and earthquakes)
4.2.4.1 Wind (Ws and Wu)
Wind acting on formwork and falsework imposes the following actions:
1. Horizontal pressure on walls or side forms;
2. Horizontal frictional drag along soffits or platforms;
3. Horizontal frictional drag on falsework or framework members; and
4. Vertical (upward and downward) pressure on soffit forms.
Formwork should be designed to resist the magnitude of wind actions determined in accordance
with AS/NZS 1170.2 (SA 2002b).
When exposed to strong winds, construction activity will continue until the wind speed reaches
some predetermined threshold. Once the wind speed exceeds this threshold construction
D
activity will cease. Up until construction ceases, it is necessary to take account of action
combinations that include the effects of wind and other actions from concurrent construction
activities.
Working or Serviceability wind speed (Vs)
From a practical perspective, if exposed to strong winds construction is unlikely to proceed
R
beyond a mean site wind speed of 50 km/hr (Force 6 on the Beaufort scale) or 27 knots. This is
consistent with guidance in authoritative references (Tayakorn and Rasmussen 2009) and (ECS
2004) that suggest the value chosen for a working design wind speed (not ultimate) Vs should
not be less than 18 m/s 1.
For ultimate limit state action combinations, the resulting working or serviceability wind
AF
pressure Ws should be multiplied by a limit states partial load factor of 1.5.
Ultimate wind speed (VR)
Normally, before the wind speed reaches ultimate limit states construction will have ceased.
The design wind speed should be based on the annual probability of exceedence specified in
AS/NZS 1170.0 (SA 2002) taking account of the level of risk and region. Table 4.2 is an
example of the annual probabilities of exceedence for ultimate limit states wind, snow and
earthquake events.
The use of levels of risk for formwork in Tables 4.2 and 4.3 replaces the structure importance
T
factor. Refer to Appendix A for guidance on selecting the appropriate level of risk for formwork
in different situations.
Table 4.2 Annual probabilities of exceedence for ultimate limit states
wind, snow and earthquake events
Level
of Earthqua
Wind Snow
Risk ke

Low 1/25 1/25 Not


required
Ordinary 1/100 1/50 Not
required

1
The working design wind speed represents the 3 second gust wind speed for permissible stress design,
which is greater than the mean wind speed measured on site.

4.10 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

High 1/500 1/100 1/500

For the annual probability of exceedance given in Table 4.2, Table 4.3 presents ultimate limit
states regional wind speeds (3 second wind gust speeds).
Table 4.3 Ultimate limit states regional wind speeds, m/s
Level Region
of A (1 to W B C D
Risk 7)
Low 37 43 39 47 53
Ordin 41 47 48 59 73
ary
High 45 51 57 69 88

4.2.4.2 Snow (Ss and Su)


D
The limit states snow actions should be determined in accordance with AS/NZS 1170.3 (SA
2003) considering the appropriate annual probability of exceedence specified in Table 4.2.
4.2.4.3 Serviceability and ultimate water (Ss and Su)
In this text, the notation S replaces Xw used in AS 3610 1995.
R
Where formwork is erected in water, take account of actions by river currents, tides, waves and
flooding. These actions may include: the dynamic pressure of the water; impact from floating
objects; the effects of increased frontal area and head of water due to trapped debris; buoyancy
and uplift.
4.2.4.4 Earthquake (Eu)
AF
It would be unusual to design formwork or falsework to resist earthquakes. Inherently,
satisfying robustness requirements set out in Chapter 3 provides a sufficient resistance.
If it were necessary to design formwork to resist earthquakes, the ultimate earthquake forces
should be determined in accordance with AS1170.4 considering the appropriate annual
probability of exceedence specified in Table 4.2.
4.2.5 Accidental actions (Av and Ah)
Where applicable, formwork should be designed to resist the effects of vertical and horizontal
accidental actions.
T
In this text, the notation A replaces I used in AS 3610 1995.
4.2.5.1 Vertical Impact (Av)
When formwork is erected on sites where overhead or mobile cranes operate, there is some risk
of impact from crane loads landing on top of the formwork. In this situation, in addition to the
weight of the lifted load, an allowance of not less than 25% of the weight of the lifted load
should be applied to the formwork.
4.2.5.2 Horizontal Impact (Ah)
When formwork is erected on sites where overhead or mobile cranes operate, there is some risk
of impact from crane loads, as shown in Figures 4.6 and 4.7. For example, in the situation
depicted in Figure 4.6, the impact of a 2700 kg kibble of concrete traveling at 3 km/hr and
coming to rest in a distance of 25 mm is equivalent to 40 kN (SA 1996).
When formwork is erected adjacent to or bridges access for vehicles, there is some risk of
impact from those vehicles, as shown in Figure 4.8.

2013 Stephen Ferguson all rights reserved 4.11


Formwork Handbook

When formwork assemblies are crane handled, there is risk of collision with another object
during lifting.
D
Figure 4.6 Impact from moving crane load (SA 1996)
R
AF
Figure 4.7 Impact from moving crane load (McAdam and Lee 1997)
T

Figure 4.8 Impact from moving vehicle (SA 1996)


4.2.6 Notional actions (N1, N2 and N3)
Notional actions are applied to idealised perfect structural models to take account of permitted
structural imperfections that have a significant influence on the structural behaviour; otherwise,
the strength and stability of the structure may be overestimated; e.g. N1 for initial out-of-plumb
and N2 for bracing forces.
As structural imperfections are present at all times, notional actions should be considered to act
concurrently and in combination with other actions.

4.12 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

Notional actions are also used to achieve minimum levels of structural integrity; e.g. N2 for
bracing stiffness and N3 for minimum levels of robustness.
In Figures 4.9, 4.10 and 4.11, the forces F1 to F8 represent vertical actions acting on the
falsework at each location. The magnitude of each force F1 to F8 is calculated by first analysing
the structure without any notional actions to determine the combined effect of all other
concurrent actions factored in accordance with the relevant combination under consideration.
Thus, the magnitude of the notional load and its effects will vary depending on the combination
of actions under consideration, see Section 4.3.
4.2.6.1 Notional horizontal forces for initial out-of-plumb erection (N1)
AS 3610 1995 permits falsework, intended to be vertical, to be erected out-of-plumb up to an
inclination of 1 in 200 or a maximum horizontal displacement of 40 mm. This may be taken
into account by analysing a structural model of the formwork, incorporating:
(a) out-of-plumb members; or
(b) vertical members and notional horizontal forces equal to 0.01 times the sum of the
vertical design actions acting at each point of application, see Figure 4.9.
D
The value of 0.01 is consistent with the requirements of other national standards; e.g. (ECS
2004) and (Tayakorn and Rasmussen 2009). It also reflects the results of research that measured
the effects of permitted out-of-plumb tolerances in steel structures and recommends a value for
the notional force of twice the permitted out-of-plumb tolerance, i.e. 2 1 = 0.01 .
200
R
The notional force should be considered to act, in either direction, in combination with direct
actions for serviceability, stability, and strength limit states, see Section 4.3.
AF
T
Figure 4.9 Notional loads required to take account of initial out-of-plumb erection
4.2.6.2 Notional forces for braces that reduce the effective length of compression
members (N2)
For free standing falsework, notional forces (N2) do not apply, because the effect of applying
notional horizontal loads to ensure structural integrity (N3) is sufficient to ensure the bracing is
adequate.
Notional forces (N2) apply to the design of falsework that has a top or intermediate restraint.
Refer to Section 8.3.1 for the difference between top-restrained and freestanding falsework.
Notional forces (N2) are intended to ensure the members that brace compression members to
reduce their effective length have sufficient strength to resist the forces arising in the bracing
due to permitted out-of-straightness and to achieve the minimum brace stiffness needed to be
effective in reducing member effective length.

2013 Stephen Ferguson all rights reserved 4.13


Formwork Handbook

Notional forces (N2) should be applied at each bracing point in a manner that minimises the
residual forces accumulating at points of restraint. For example, in the arrangement shown in
Figure 4.10, within each structure, the forces on each level of bracing are applied in opposing
directions, and the forces on the same level in adjacent structures are considered to act in
opposite directions. The forces in falsework bracing are real and some residual forces can be
expected at points of restraint, see Section 8.3.3.3.
The recommended magnitude of the notional horizontal force at each bracing point is equivalent
to 2.5% of the axial force in the compression member.
Where restraint is provided to more than one compression member, AS 3610 1995 (following
guidance in AS 4100) permits reducing the magnitude of the notional force applied to each
additional member up to a maximum of seven members to 1.25% of the axial force in each
additional compression member.
However, reducing the magnitude of notional force is not recommended where:
(a) Falsework members exceed the out-of-straightness limit permitted in AS 4100 of
l/1000, as is permitted in AS 3610 1995;
D
(b) Multiple columns are connected by a line of braces; or
(c) Columns are braced at multiple points, see Section 8.3.3.3.
The notional force should be considered to act, in either direction, in combination with direct
actions for strength limit states, see Section 4.3.2.2.
R
AF

Figure 4.10 Notional forces to ensure braces have the minimum brace strength and stiffness
T
4.2.6.3 Notional horizontal forces to ensure a minimum level of structural integrity
(N3)
AS/NZS 1170.0 requires all structures to have a minimum level of structural integrity, such that
all parts of the structure are tied together (both in the horizontal and vertical planes) so that the
structure can withstand an event without disproportionate damage.
This requirement is deemed to be satisfied if all parts of the structure are connected to provide
load paths to points of restraint and the members and connections, as a minimum, can resist the
following lateral loads:
(a) formwork and falsework members can resist the effects of a notional horizontal
action equivalent to 0.025 times the applied vertical actions; and
(b) formwork and falsework connections and ties can resist the effects of a notional
horizontal action equivalent to 0.05 times the applied vertical actions.

4.14 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

The applied vertical actions should be determined from, and act simultaneously with, the most
adverse combination of permanent and variable vertical actions [G, Q]. The robustness
horizontal action should be considered to apply at the respective points of application for each
vertical action, see Figure 4.11.
The requirements in AS/NZS 1170.0 for robustness are similar purpose to the tried and proven
BSI 5975 minimum stability requirements, which require falsework structures to resist a
minimum horizontal action of 2.5% of the applied vertical actions (Tayakorn and Rasmussen
2009).
For practical purposes, to comply with the robustness requirements in AS/NZS 1170.0, it is
sufficient for all formwork structures to be designed to resist the most adverse effects that result
from the strength limit states action combinations in Section 4.3.2.2, which include
Combinations 4.11b and 4.12b.
The horizontal notional actions (N3), in Combinations 4.11b and 4.12b, are applied at each point
of application of the vertical actions, see Figure 4.11. The magnitude of the horizontal notional
action (N3) should not be less than 0.025 times the combined factored vertical action at each
point of application. The horizontal notional actions apply in each falsework plane.
D
R
AF
Figure 4.11 Notional horizontal actions to ensure minimum levels of structural integrity

4.3 Action combinations


Verifying that serviceability and ultimate limit states have been satisfied requires first
determining the appropriate combinations of actions for each design situation.
For clarity, some actions (e.g. water, snow, ice and earthquake) have been omitted in the
T
following lists of action combinations. Where these actions may occur, they should be taken
into account in accordance with AS/NZS 1170.0.
In this text, the format and load factors of combinations of actions presented vary from those
presented in AS 3610 1995 Table 4.5.1. They have been updated to comply with AS/NZS
1170.0 and include loads omitted in AS 3610 1995 that need to be taken into account.
Designs based on these equations are deemed to comply with AS 3610 1995 under the
provisions of Clause 1.4, which permits the use of methods of design not specifically
referenced, provided the requirements of the Standard are met.
The notation used in the following sections is intended to express the design action effect that
results from the combined effect of the applicable concurrent actions selected from the listed
set, factored accordingly. The vertical actions and horizontal actions have been grouped
separately, but are considered to act concurrently.
In the following combinations, the prefix is used to indicate the action in question may act in
different directions; e.g. Wsh indicates that the horizontal serviceability wind actions may act in
any direction. Consideration should be given to determining and taking account of the
application of the relevant action in the direction that has the most adverse effects.

2013 Stephen Ferguson all rights reserved 4.15


Formwork Handbook

4.3.1 Action combinations for serviceability limit states


4.3.1.1 Surface finish quality and dimensional control
For the purposes of verifying concrete surface finish quality and dimensional accuracy, the
following possible combinations of vertical and horizontal actions during Stage 2 needs to be
considered, where applicable:

Esv = [Gf, Gc, pPcv] and Esh = [pPch] (4.3)

where
Esv = serviceability action effect from vertical actions;
Esh = serviceability action effect from horizontal actions;
Pcv = vertical component from concrete pressure acting on an inclined surface;
Pch = horizontal concrete pressure acting on a vertical surface or the horizontal
component from concrete pressure acting on an inclined surface; and
p
D
= serviceability load factor for concrete pressure, where:
p = 1.1 for Class 1 and Class 2; otherwise, p = 1.0.
Where the positional accuracy of the formed element is important, consideration should be
given to deformations arising during Stage 2. In particular, combinations that include for the
effects of persistent horizontal actions such as: unbalanced concrete pressure, wind and notional
R
actions:

Esv = [Gf, Gc, Pcv] and Esh = [Pch, Wsh, N1] (4.4)

4.3.1.2 Elastic behaviour in multiple-use formwork


For multiple-use formwork it is important that behaviour at serviceability limit states remain
AF
elastic (reversible). This may be a problem for ductile materials where the ratio of yield to
ultimate strength is less than the load ratio of serviceability limit states to strength limit states;
e.g. mild steel bolts in tension and mild steel plates in bearing.
Inelastic behaviour is most likely to occur when the formwork is most heavily loaded, typically
during Stages 2 and 3. Thus, consideration should be given to combinations of actions that
include for the effects of the combined weight of formwork and concrete plus any concurrently
applied action, such as the weight of workers and equipment, stacked materials, concrete
pressure, wind, and notional loads:
T
Esv = [Gf, Gc, Qv, Pcv, Wsv,] and Esh = [Qh, Pch, Wsh, N1] (4.5)

where
Qv = combination of vertical imposed actions; i.e. [Qw or Qm, Qxv];
Qh = combination of horizontal imposed actions; i.e. [Qah, Qxh];
Wsv = vertical serviceability limit states wind action; and
Wsh = horizontal serviceability limit states wind action.

4.3.2 Action combinations for ultimate limit states strength and stability
For strength and stability limit states, the most adverse combinations of actions that occur
during Stages 1, 2 and 3 must be considered.
During Stage 1, Gc may be zero or represent the weight of reinforcement. At this Stage, the
destabilising effects (uplift, sliding and overturning) of wind are significant, in comparison to
subsequent Stages.
4.16 2013 Stephen Ferguson all rights
reserved
Chapter 4 - Actions and action combinations

Over half of all formwork collapses occur during Stage 2. Thus, it is necessary to consider the
concrete placement sequence (e.g. the possibility of concrete placed only on a single span or
cantilever). It is unlikely that concrete placement would commence or continue during storm
winds, so Stage 2 action combinations including Wu can be neglected. However, as the primary
cause of failure is inadequate bracing, the combinations of vertical and horizontal actions need
careful consideration.
The design of primary members (shores and primary beams) is often governed by the effects of
stacked materials present during Stage 3.
Consideration should be given to the possibility that variable actions might act in the opposite
direction to any permanent actions; e.g. concrete pressure, wind or water actions might act in the
opposite direction to weight of the formwork.
The action combination and partial load factors for accidental impact, reflects the low
probability that impact will occur simultaneously with other horizontal actions and that some
damage is acceptable.
4.3.2.1 Action combinations for stability limit states
D
The basic action combinations for the stability limit state are:
(a) Combinations of vertical and horizontal actions that produce net destabilising effects (Edv,dst and
Edh,dst), include:
Permanent actions only,
R
Edv,dst = [1.35(Gf, Gc)] and Edh,dst = [N1] (4.6)

Permanent, variable and notional actions,

Edv,dst = [1.2(Gf, Gc), 1.5Qv, 1.5Pcv, 1.5Wsv] and Edh,dst = [1.5Qh, 1.5Pch, 1.5Wsh, N1] (4.7)
AF
Edv,dst = [1.2(Gf, Gc), Wuv] and Edh,dst = [Wuh, N1] (4.8)

Permanent and accidental actions,

Edv,dst = [(Gf, Gc), 1.1Av] and Edh,dst = [1.1Ah] (4.9)

where
Edv.dst = net destabilising effect of the combined vertical actions;
Edh.dst = net destabilising effect of the combined horizontal actions;
T
Wuv = vertical ultimate limit states wind action;
Wuh = horizontal ultimate limit states wind action;
Av = vertical actions from accidental impact; and
Ah = horizontal actions from accidental impact.
The above combinations should only include actions that act concurrently and produce a
destabilising effect. They should not include any actions that produce a stabilising effect.
Where applicable, action combinations including the effect of snow, ice and earthquake
actions should be taken into account (SA 2002).
(b) For combinations of vertical actions that produce net stabilising effects (Edv,stb)
Permanent actions only,

Edv,stb = [0.9(Gf, Gc)] (4.10)

2013 Stephen Ferguson all rights reserved 4.17


Formwork Handbook

where
Edv.stb = net stabilising effect of the combined vertical actions;
This combination should only include permanent actions (or parts thereof) that produce a
stabilising effect.
4.3.2.2 Action combinations for strength limit states
The basic combinations for strength limit states are:
(a) Permanent actions only,

Edv = [1.35(Gf, Gc)] and Edh = [N1, N2] (4.11a)

Edv = [1.35(Gf, Gc)] and Edh = [N3] (4.11b)

(b) Permanent, variable and notional actions,

Edv = [1.2(Gf, Gc), 1.5Qv, 1.5Pcv, 1.5Wsv] and Edh = [1.5Qh, 1.5Pch, 1.5Wsh, N1, N2]
D
(4.12a)

Edv = [1.2(Gf, Gc), 1.5Qv, 1.5Pcv, 1.5Wsv] and Edh = [ N3] (4.12b)

Edv = [0.9(Gf, Gc), 1.5Pcv, 1.5Wsv] and Edh = [1.5Pch, 1.5Wsh, N1, N2] (4.13)
R
Edv = [1.2(Gf, Gc), Wuv] and Edh = [Wuh, N1, N2] (4.14)

Edv = [0.9(Gf, Gc), Wuv] and Edh = [Wuh, N1, N2] (4.15)

(c) Permanent and accidental actions,


AF
Edv = [(Gf, Gc), 1.1Av] and Edh = [1.1Ah] (4.16)

where
Edv = design action effect from the combined vertical actions; and
Edh = design action effect from the combined horizontal actions.
As for stability limit states, where applicable, action combinations including the effect of snow,
ice and earthquake actions should be taken into account (SA 2002).
Concrete placement is not likely to proceed during an ultimate wind event; therefore
T
Combinations 4.14 and 4.14 are not considered to act during Stage 2.
4.3.2.3 Strength load factor for primary members
For members and connections critical to structural integrity (i.e. whose failure would cause
structural failure, instability or collapse), the strength limit states design action effect Ed
calculated from action Combinations 4.11a to 4.16, should be multiplied by a strength load
factor d, see Equation 4.17.

d Ed Rd (4.17)

where
Rd = design resistance or capacity, see Chapter 3.
d = strength load factor for primary members

For all primary members and connections critical to structural


integrity and whose failure may cause instability or collapse of the

4.18 2013 Stephen Ferguson all rights


reserved
Chapter 4 - Actions and action combinations

structure (e.g. primary beams, bearers, soldiers, shores, props, form


ties and anchors), d = 1.25; or

For members and connections not critical to structural integrity and


whose failure would have only localised effects and deformations and
not cause overall failure, instability or collapse of the structure (e.g.
secondary beams, joists, studs and form face members), d = 1.0.
The factor d is intended to reduce the probability of failure due to underestimates of the action
effects unique to formwork and falsework structures that are not accounted for by applying
general limit states load factors; i.e. 1.35G or 1.2G +1.5Q and 1.5P . Underestimates may arise
due to phenomena such as unanticipated load distribution (Ikheimonen 1997).
Herein the strength load factor for primary members d replaces the global load factor
introduced in Amendment No 1 to AS 3610 1995 (SA 2003) to take account of unanticipated
load distribution. The value for d has been revised from 1.30 (in Amendment 1) to 1.25, as the
result of calibration with the current combinations of actions in AS1170.0. (i.e. action
combinations 1.2G +1.5Q and 1.35G adopted in this text versus 1.25G +1.5Q specified in AS
D
3610 1995)
4.3.2.4 Duration of load factor for use with AS 1720.1
The resistance of timber, plywood and LVL is dependent upon the duration of loading.
Resistance decreases as load duration increases. Thus, for each design situation, the appropriate
duration of load factor for use with AS 1720.1 (SA 2010) corresponds to the duration of load
R
factor (k1) for the shortest duration action contributing to the combination, see Table 4.4.
Table 4.4 Duration of load factor (k1) for strength
Type of Load Effective duration of Duration of load factor
peak load k1
Formwork weight, Gf
AF
5 months 0.80
Concrete weight, Gc
(a) for plywood 5 hours 0.97
(b) for bearers and joists 5 days 0.94
(c) for supports
5 months 0.80
Imposed actions, Qw and 5 hours 0.97
Qah
Stacked materials, Qm 5 days 0.94
Accidental impact, A 5 seconds 1.00
T
Concrete pressure, P 5 hours 0.97
Wind,
(a) Serviceability wind, 5 hours 0.97
Ws 3 seconds 1.00
(b) Ultimate wind, Wu
Snow, Su 5 days 0.94
Earthquake, Eu 5 minutes 1.00
Water, Su
(a) River currents 5 months 0.80
(b) Tidal action 5 hours 0.97
(c) Flooding
(d) Wave action 5 days 0.94
5 minutes 1.00

For timber products, appropriate design situations can be selected by considering, for each
combination of actions, the magnitude of the quotient of the resulting action effect divided by
the appropriate load duration factor and not merely the magnitude of the action effect alone.

2013 Stephen Ferguson all rights reserved 4.19


Formwork Handbook

The resistance of timber members and connections may be overestimated where the magnitude
of the shortest duration action has an negligible effect or, in the case of variable actions, may
omitted from the action combination. In this case, it would be appropriate to consider the
quotient of the action effect resulting from the combined remaining actions and the duration of
load factor (k1) for the next shortest duration action contributing to the combination.
Design Example.
For the limit states strength design of a formwork soffit bearer, taking account of vertical
permanent and imposed actions and neglecting all other actions, the governing
combination of actions for the design of timber bearers could be one of the following:
1.2G f + 1.5Qm 1.35G f
Stage 1 or
1.2(G f + Gc ) + 1.5Qw 1.35(G f + Gc )
Stage 2 or
1.2(G f + Gc ) + 1.5Qm 1.35(G f + Gc )
Stage 3 or
From Table 4.4, for each action, the effective duration of peak load and duration of load
D
factor for bearers are:
Formwork weight Gf 5 months and 0.80;
Concrete weight, Gc 5 days and 0.94;
Workers and equipment, Qw 5 hours and 0.97; and
R
Stacked materials, Qm 5 days and 0.94.
Taking account of the duration of load factor, the most adverse combination for the limit
states strength design of the timber bearers will be the greater of the following:
1.2G f + 1.5Qm 1.35G f
AF
Stage 1 0.97 or 0.80
1.2(G f + Gc ) + 1.5Qw 1.35(G f + Gc )
Stage 2 0.97 or 0.94
1.2(G f + Gc ) + 1.5Qm 1.35(G f + Gc )
Stage 3 0.94 or 0.94
Any risk of underestimation due to the negligible effect short term variable actions (e.g.
T
workers and equipment) with a high load duration factor or which might be omitted is
avoided by considering combinations for permanent actions only, as well as combinations
for permanent and variable actions.
References
Ayoub, H. N. and S. Karshenas (1994). "Survey Results for Concrete Construction Live Loads
on Newly Poured Slabs." Journal of Structural Engineering, ASCE 120(No. 5, May): 1543-
1562.
ECS (2004). EN 12812.2 Falsework - Performance requirements and general design. Brussells,
European Committee for Standardization.
Ellingwood, B., J. G. MacGregor, T. V. Galambos and C. A. Cornell (1982). "Probability Based
Load Criteria: Load Factors and Load Combinations." ASCE Journal of the Structural Division
108(No ST5 May): 978-997.
Fattal, S. G. (1983). Evaluation of construction loads in multistory concrete buildings.
Washington, D.C., U.S. Dept. of Commerce National Bureau of Standards : For sale by the
Supt. of Docs. U.S. G.P.O.
4.20 2013 Stephen Ferguson all rights
reserved
Chapter 4 - Actions and action combinations

Ikheimonen, J. (1997). Construction Loads on Shores and Stability of Horizontal Formworks


Doctoral Thesis, Royal Institute of Technology.
ISO (1998). ISO 2394:1998 General principles on reliability for structures. Geneve,
International Organization for Standardization.
Karshenas, S. and H. N. Ayoub (1994). "Analysis of Concrete Construction Live Loads on
Newly Poured Slabs." Journal of Structural Engineering, ASCE 120(No. 5, May): 1525-1542.
McAdam, P. S. and G. Lee (1997). Formwork a practical approach. London, E & EF Spon.
Rosowsky, D. V., Y. L. Huang, W. F. Chen and T. Yen (1994). "Modeling concrete placement
loads during construction." Structural Engineering Review 6(2): 71-84.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney,
Standards Australia.
SA (2002). AS/NZS 1170.0 - 2002 Structural design actions Part 0: General principles. Sydney,
Standards Australia.
D
SA (2002). Australian/New Zealand Standard AS/NZS 1170.0:2002 Structural Design Actions
Part 0: General principles. Sydney, Standards Australia.
SA (2002a). AS/NZS 1170.1:2002 Structural design actions - Part 1: Permanent, imposed and
other actions. Sydney, Standards Australia.
SA (2002b). AS/NZS 1170.2: Structural design actions - Part 2: Wind actions. Sydney,
R
Standards Australia.
SA (2003). Amendment No. 1 to AS 3610 - 1995 Formwork for concrete. Sydney, Standards
Australia.
SA (2003). Australian/New Zealand Standard AS/NZS 1170.3:2003 Structural Design Actions
AF
Part 2: Snow and ice actions. Sydney, Standards Australia.
SA (2010). AS 1720.1 - 2010 Timber structures. Part 1: Design methods. Sydney, Standards
Australia.
SA (2010). AS/NZS 1576.1:2010 Scaffolding: Part 1: General requirements. Sydney, Standards
Australia.
SII (1998). SI 904 Formwork for Concrete: Principles, The Standards Institution of Israel.
Tayakorn, C. and K. J. R. Rasmussen (2009). Research Report No R896 Structural Modelling of
Support Scaffold Systems, University of Sydney
T

2013 Stephen Ferguson all rights reserved 4.21


Chapter 5 - Concrete Pressure

5
Concrete Pressure
5.1 Introduction
The behaviour of concrete and its effect on concrete pressure are discussed in detail in this Chapter.
The Chapter starts with a review of the method adopted in AS 3610 1995 (SA 1995) to calculate
concrete pressure.
This is followed by a review of the influence of different factors on concrete pressure and selecting an
appropriate rate of rise so as to avoid cold joints.
The final part of this Chapter presents an overview of the statics of concrete pressure.
D
5.2 CIRIA Report No 108
The concrete pressure exerted on the formwork may be calculated using the method developed by
Clear and Harrison in the CIRIA Report 108 (Clear and Harrison 1985), which was adopted in AS
3610 - 1995.
R
Equations 5.1 and 5.2 were developed from tests involving some 350 sets of data. Figure 5.1 shows a
comparison between measured and calculated pressures using these Equations.
AF
T

Figure 5.1 Comparison between measured and calculated pressures (Clear and Harrison 1985)
Newly mixed concrete comprises a gradation of particles from coarse aggregate down to fine cement
particles suspended in water. At this stage, the concrete exerts a fluid or hydrostatic pressure on the
formwork, which can be determined using Equation 5.1.
When hydration commences, the concrete starts to set. Products form on the surface of the cement
particles enhancing inter-particle bond and restricting inter-particle movement. Once this occurs,
increments of vertical load have a reduced effect on the concrete pressure. Equation 5.2 takes account
of the influence that various factors have on the onset and rate of hydration.

2013 Stephen Ferguson all rights reserved 5.1


Formwork and Falsework

When applied, these equations produce a design pressure envelope similar to that depicted in Figure
5.2
D
Figure 5.2 Concrete pressure envelope (Clear and Harrison 1985)
Figure 5.2 shows the effect concrete setting has on the maximum hydrostatic concrete pressure, which
occurs at a depth hh below the top of the concrete, given in Equation 5.3.
R
The concrete pressure (Pc) at any depth (h) below the top of the concrete can be determined as follows
(Clear and Harrison 1985):
(a) For h < hh,

Pc = gh (5.1)
AF
(b) For h hh,

Pc = ghh (5.2)

where

hh = C1 Rc + C 2 C 3 (h f )
C1 Rc hc
(5.3)
T
In Equations 5.1 to 5.3,
Pc = concrete pressure at a depth (h), in kPa
= wet density of concrete, in kg/m3
g = gravity, m/s2
h = depth below the top of the concrete ( 0 h hc), m
hc = depth of concrete pour, m
hh = maximum depth of hydrostatic pressure, m (see Figure 5.2)
hf = vertical form height, m
Rc = vertical rate of concrete rise up the form, m/hr
C1 = coefficient dependent on the size and shape of formwork,
where the plan width or breadth is greater than 2 m, C1 = 1.0; otherwise, C1 = 1.5.

5.2 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

C2 = coefficient for the effect concrete cement and admixtures have on setting time,
(0 C2 0.6), see Table 5.1
C3 = coefficient for the effect concrete temperature has on setting time,
2
36
C3 = where, Tc = concrete temperature at placement, C.
Tc + 16

In some situations, Equation 5.3 may not have a real solution or the depth hh at which stiffening takes
effect may be greater than the depth of the concrete pour hc. In this case, the concrete pressure Pc
should be determined using Equation 5.1 where h = hc, and Equation 5.2 can be neglected.
Table 5.1 Values of coefficient C2

Group Concrete C2

A Concrete with GP or HE cement


0.30
Basic Concrete with GB cement containing less than 20% fly ash and/or slag
D
Concrete that also includes metakaolin or silica fume.
Concrete with LH, SR, or SL cement
B
Concrete with GB cement. 0.45
Retarded
Concrete
R
Self-compacting concrete (SCC).

C
Any Group A or B concrete (including SCC) with cement containing
Heavily 0.60
greater than 35% flyash or greater than 65% slag.
Retarded
AF
Concrete
For any concrete (including SCC) in Group A or B, increase the value of C2 by adding 0.15 where a
retarding admixture is used in the concrete.

The values in Table 5.1 were taken from AS 3610 - 1995 and have been updated to include
subsequently published guidance (Pallett 2009).
Retarding admixtures include retarders, retarding water reducers, retarding superplasticisers and any
admixture that is used such that it effectively acts as a retarder.
T
Type SR cement is defined on a performance basis and may contain a high percentage of slag
necessitating the use in Table 5.1 of a higher value of coefficient C2 than the 0.30 suggested by Clear
and Harrison.
Equations 5.1 and 5.2 are likely to be conservative for no-fines concrete, underwater concreting,
controlled permeability form fabrics (Arslan 2002), and very permeable forms such as expanded
metal. In the latter case, concrete pressures may be reduced by as much as 50%.
Importantly, the equations have not been proven for temperatures in excess of 30 C or below 5 C
and do not cover: concrete pumped from below; the use of external vibrators attached to the
formwork; revibrating the concrete by deeply immersing internal vibrators; i.e. more than 1.0m. In
these cases, formwork pressures are likely to be higher.
In the case of self-compacting concrete, DIN 18218 Pressure of fresh concrete on vertical formwork
(DIN 2010) provides guidance based on research by Proske (Proske 2002).
Anecdotal evidence suggests that the maximum concrete pressure ever measured is 150 kPa.

2013 Stephen Ferguson all rights reserved 5.3


Formwork and Falsework

5.2.1 Factors influencing concrete pressure


From casual inspection of Equation 5.1, it is clear that concrete density and vertical pour height have a
direct (linear) bearing on the hydrostatic concrete pressure. In Equations 5.2 and 5.3, some of the
factors that influence the rate of hydration and therefore concrete pressure are less obvious and
discussed herein. Form height and the height at which the concrete is discharged are also discussed as
these can affect the maximum concrete pressure.
Factors that affect the maximum concrete pressure are listed in Table 5.2.
Some of these factors are taken into account in Equations 5.1 to 5.3, namely:
= wet density of concrete, kg/m3
hc = vertical pour height, m
C1 = coefficient dependent on the size and shape of formwork
Rc = vertical rate of concrete rise up the form, m/hr
C2 = coefficient for the effect concrete cement and admixtures have on setting time, see
Table 5.1
D
Tc = concrete temperature at placement, C
hf = vertical height of form, m

Table 5.2 Factors affecting concrete pressure (Clear and Harrison 1985)

Concrete Admixtures
R
Aggregate shape, size, grading and density
Cementitious materials
Mix proportions
Temperature at placing
Wet density
AF
Workability
Formwork Permeability/watertightness
Plan shape and area of the cast section
Roughness of the sheeting material
Slope of the form
Stiffness of the form
Vertical form height
Placing Impact of concrete discharge
In air or underwater
Placing method (e.g. lift height or rate of
T
rise)
Vibration

Figures 5.3 to 5.8 are included to demonstrate the sensitivity of concrete pressure to each of these
factors, for pours up to 6 m high. In each case, only one variable changed. Otherwise, each figure
depicts a scenario where concrete with a wet density of 24 kN/m3 is discharged from the top of
formwork that is nominally 200 mm higher than the pour. The nominal rate of concrete rise in the
formwork is 5 m/hr. The nominal coefficient for concrete materials C2 is taken to be equal to 0.45 and
the concrete temperature is assumed to be 20 C. Except for Figure 5.4, where C1 = 1.5 (e.g.
columns), the nominal value for coefficient C1 = 1.0 (e.g. walls).
5.2.2 Plan shape and area of cast section
A common misconception is that deep elements with a large plan area (e.g. 3 m deep raft foundations)
have high concrete pressures. To the contrary, in elements with small plan cross-section shape or area
vibration can be sufficient to mobilise all the concrete; however, in elements with a large plan section

5.4 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

or area, all of the concrete is not mobilised at the same time and less energy is transmitted into the
formwork.
The effect is that the maximum pressures in walls are lower than in columns. For the purposes of
Equation 5.3, a column is defined as a section where both the width and breadth are equal to or less
than 2 m; otherwise, the section is considered to be a wall. The difference in pressure between
walls and columns is addressed in Equation 5.3 by the coefficient C1, which is equal to 1.0 for
walls and 1.5 for columns.

0.0

-1.0 Rc = 1 m/hr
Rc = 2.5 m/hr
Rc = 5.0 m/hr
Vertical Pour Height, hc (m)

Rc = 7.5 m/hr
-2.0
Rc = 10.0 m/hr

-3.0
D
-4.0
R
-5.0

-6.0
0 10 20 30 40 50 60 70 80 90 100 110 120 130
AF
Concrete Pressure, P (kPa)

Figure 5.3: Influence of rate of rise on concrete pressure in walls

0.0

-1.0 Rc = 1 m/hr
Rc = 2.5 m/hr
Rc = 5.0 m/hr
Vertical Pour Height, hc (m)

T
Rc = 7.5 m/hr
-2.0
Rc = 10.0 m/hr

-3.0

-4.0

-5.0

-6.0
0 10 20 30 40 50 60 70 80 90 100 110 120 130

Concrete Pressure, P (kPa)

2013 Stephen Ferguson all rights reserved 5.5


Formwork and Falsework
Figure 5.4: Influence of rate of rise on concrete pressure in columns
The effect of plan shape and area can be seen by comparing the pressures plotted in Figures 5.3 and
5.4, which plot the concrete pressure for a range of rates of concrete rise.
Clearly the pressure in columns (Figure 5.4) is greater than walls (Figure 5.3) and hydrostatic pressure
governs the design of columns more so than for walls.
5.2.3 Concrete rate of rise
The rate at which concrete rises vertically up the formwork is a critical factor. The effect of rate of
rise on concrete pressure is shown in Figures 5.3 and 5.4, for walls and columns respectively.
From Figures 5.3 and 5.4, it can also be seen that:
(a) As the rate of rise increases, the maximum concrete pressure increases; and
(b) The rate of rise has a greater influence compared to other factors.
The rate of rise has a significant effect on the maximum concrete pressure and factors to be considered
when selecting an appropriate rate of rise are discussed in more detail later in the Chapter.
5.2.4 Constituent concrete materials
D
Cconcrete pressure is greater in concrete that takes longer to set (i.e. Groups B and C in Table 5.1).
Coefficient C2 (specified in Table 5.1) takes into account the effects of concrete with different cements
and admixtures. The value of C2 in Table 5.1 should be increased for admixtures that effectively act as
retarders, such as retarding water reducers and any admixture that is used above the recommended
dosage.
R
The effects of different values of coefficient C2 can be seen in Figure 5.5, which plots the maximum
concrete pressure for concrete placed in wall formwork. Figure 5.5 demonstrates that concrete
pressure increases with increasing the values of C2.

0.0
AF
-1.0
C2 = 0.30

C2 = 0.45
Vertical Pour Height, hc (m)

-2.0 C2 = 0.60

-3.0
T
-4.0

-5.0

-6.0
0 10 20 30 40 50 60 70 80 90 100 110 120 130

Concrete Pressure, P (kPa)

Figure 5.5: Influence of coefficient C2 on concrete pressure in walls


5.2.5 Concrete temperature
As with all chemical reactions, the rate of hydration increases with increased temperature. Higher
concrete temperature will increase in the rate of hydration causing a reduction in concrete pressure.

5.6 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

In Equation 5.3, the coefficient (C3) takes account of the effects concrete temperature has on concrete
setting time. Its influence is demonstrated in Figure 5.6, specifically:
(a) As the concrete temperature increases the maximum concrete pressure decreases; and
(b) At low concrete temperatures, changes have a greater effect on the maximum pressure.
The temperature factor C3 is considered sufficiently accurate for concrete temperatures at placing
between 5C and 30C. It would not be prudent to extrapolate the design equation beyond these
values.

0.0

-1.0 Tc = 10 deg C
Tc = 15 deg C
Tc = 20 deg C
Vertical Pour Height, hc (m)

Tc = 25 deg C
-2.0
Tc = 30 deg C
D
-3.0

-4.0
R
-5.0

-6.0
0 10 20 30 40 50 60 70 80 90 100 110 120 130
AF
Concrete Pressure, P (kPa)

Figure 5.6: Influence of concrete temperature on concrete pressure in walls


5.2.6 Vertical form height
The vertical form height hf is important because it limits the maximum possible vertical pour height
and therefore the maximum pressure.
The form height may also dictate the minimum discharge height, which is important because the
impact of falling concrete increases concrete pressure. Figure 5.7 demonstrates how the factor hf
T
should be measured.

Figure 5.7: Measuring vertical form height or concrete discharge height (Clear and Harrison 1985)

2013 Stephen Ferguson all rights reserved 5.7


Formwork and Falsework

Figure 5.8 demonstrates the influence of a range of vertical form heights expressed in terms of height
above the top of the pour.
Concrete may be placed by a static hopper with a tremie pipe, directly to the base of the form. In this
case, the height of the concrete shall be measured from the base of the form to the top of the hopper
(Pallett 2009).
Although, the maximum concrete pressure increases with increasing form height or concrete discharge
height, when compared with the factors previously discussed it has less of an effect on the maximum
concrete pressure.

0.0

-1.0 h' = 0.2m above pour


h' = 0.5m above pour
h' = 1.0 m above pour
Vertical Pour Height, hc (m)

h' = 2.0m above pour


D
-2.0
h' = 3.0m above pour

-3.0
R
-4.0

-5.0
AF
-6.0
0 10 20 30 40 50 60 70 80 90 100 110 120 130

Concrete Pressure, P (kPa)

Figure 5.8: Influence of vertical form height on concrete pressure in walls


5.2.7 Other factors
T
Other factors that affect the maximum concrete pressure include method of vibration and formwork
permeability. In addition, when concrete is placed underwater, concrete pressure is affected.
5.2.7.1 Method of vibration
Equations 5.1 and 5.2 do not cover situations where:
1. Concrete is pumped from below;
2. External vibrators are attached to the formwork; and
3. The concrete is revibrated by immersing internal vibrators more than 1.0 m.
In these cases, formwork pressures are likely to be higher.
5.2.7.2 Formwork permeability
If all other conditions are equal, formwork pressures decrease as the formwork permeability increases.
For example, pressures are substantially lower (up to 50%) for extremely permeable form materials
such as expanded metal or fabric (Pallett 2000).

5.8 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

5.2.7.3 Underwater concreting


When concrete is placed underwater, the effective weight of concrete (density) is reduced by the
weight density of water (e.g. 25 kN/m3 concrete underwater weighs 15 kN/m3). Further guidance can
found in the British Concrete Society publication Formwork a guide to good practice (CS 2012).
5.3 Rate of rise
The rate concrete rises up the form has a significant effect on concrete pressure. In selecting an
appropriate rate of rise upon which to base the formwork design consideration must be given to many
factors including: the proposed method and rate of concrete delivery, the proposed method and
sequence of concrete placement, the duration of concrete placement, formwork economy, formwork
shape and plan area, and any limitations on formwork strength or serviceability. In any case, the
formwork design should be based on a realistic estimate of the maximum rate of rise that could be
expected.
For concrete elements with parallel side formwork, the rate of rise (Rc) can be calculated as follows:

Vc
Rc = (5.4)
Ac
D
where
Vc = rate of concrete delivery, m3/hr; and
Ac = plan area of the concrete element to be cast, m2.
R
For concrete elements where the plan area of the concrete element varies with height, the rate of
concrete delivery must also vary if a constant rate of rise is to be maintained.
5.3.1 Minimum rate of rise for full depth hydrostatic pressure
It is often useful to know the minimum rate of rise Rh at which the concrete pressure will remain
hydrostatic to the full depth of the concrete pour. If the actual rate of rise Rc is greater than Rh,
AF
Equation 5.2 can be neglected.
Equations 5.5 and 5.6 calculate the minimum rate of rise for hydrostatic pressure to occur to a concrete
depth h.
If the difference between the height of the formwork hf and the depth of the concrete h is small, it can
be shown that the concrete will behave hydrostatically to a depth h, if:

h (C 2 C 3 )2
2

Rh (h )
(5.5)
C 1
T
If the difference in height is significant, the equation is much more complex and:

2
((C C ) 2 2h) ((C C ) 2 2h)) 2 4(h 2 (C C ) 2 h )
f
Rh (h )
2 3 2 3 2 3
(5.6)
2C1

5.3.2 Proposed method of concrete placement


Concrete may be placed in many ways. The most common methods are: discharging directly out of the
truck into the formwork; discharging out the truck into a concrete kibble (or skip), which is lifted by a
crane and emptied into the formwork; or discharging out of the truck into the hopper of a concrete
pump and pumped along a pipeline and/or boom into the formwork.
A concrete truck is capable of discharging its contents directly into formwork at approximately 30
m3/hr. The rate at which concrete can be delivered by crane is approximately 20 to 25 m3/hr. If
concrete pumps are used, concrete may be delivered at rates of 50 to 75 m3/hr. If concrete is delivered

2013 Stephen Ferguson all rights reserved 5.9


Formwork and Falsework

concurrently from more than one source (e.g. two pumps) the rates of delivery will potentially increase
proportionally.
On the other hand:
(a) The rate of rise chosen should also be consistent with practical minimum concrete
delivery rates. If the rate of rise is too slow, previously placed concrete may start to set
causing cold joints to form, or concrete yet to be placed may start set causing blockages;
and
(b) When chutes are used to place concrete and avoid segregation the rate of rise may also be
slower than expected.
5.3.3 Proposed sequence of concrete placement
It is good practice to place and vibrate concrete, progressively, in layers (typically 300 to 500 mm
thick). A layer is placed over the whole plan area of the form, before commencing the subsequent
layer. Each layer is placed following the same sequence or path, starting and finishing at the same
location, see Figure 5.9.
D
R
AF
Figure 5.9 Concrete placed in layers
To avoid cold joints (where a layer of concrete sets before the subsequent layer is placed), the time to
place a single layer should be less than the setting time of the concrete. Common practice is to ensure
each layer is placed within in 0.5 hrs. This has a direct relationship on the minimum permitted rate of
rise, specifically:

dl
Rc (5.7)
tc
T
where
dl = thickness of the layer, m; and
tc = setting time of the concrete, hrs.
Based on typical values of dl = 0.5 m and tc = 0.5 hrs, substituting into Equation 5.7 establishes a
practical minimum rate of concrete rise:

0.5
Rc 1.0 m/hr (5.8)
0.5

and Equation 5.4 becomes:

Vc
Rc = 1.0 m/hr (5.9)
Ac

Similarly, a useful expression for the minimum rate of concrete delivery is given by:

5.10 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

Vc 1.0 Ac m3/hr (5.10)

On large pours (e.g. raft foundations) it may not be possible to deliver concrete at a rate that satisfies
the inequality in Equation 5.10. In this case, two options are available:
(a) Divide the concrete element into two or more parts that each satisfy Equation 5.10, and which are
poured on separate days. This can be achieved by introducing vertical construction joints, see
Figure 5.10; or
(b) Place the concrete in layers, which are the full width of the concrete element, but reach full height
in a shorter distance, such that Equation 5.10 is satisfied. In this way, the concrete is progressively
placed from one end of the pour to the other in layers, which might be considered like
parallelograms in elevation, see Figure 5.11.
D
R
Figure 5.10 Vertical construction joints introduced to reduce the area of concrete to be placed.
AF
Figure 5.11 Concrete placed to full height over a shorter distance to avoid cold joints.
In the latter case, the angle and length of the slope as well as the layer thickness is dictated by the
properties and behaviour of the concrete mix, including the internal shear friction (c).
T
5.3.4 Duration of concrete placement
For scheduling reasons, concrete is often placed in vertical elements such as walls and columns late in
the afternoon. At this time of day, the time remaining to the pour the concrete is often limited by the
desire of: workers to go home; or employers to avoid or minimise overtime penalties. This situation is
inconsistent with specifying a slow rate of rise.
5.3.5 Formwork and reinforcement arrangement
The arrangement of the reinforcement and formwork may obstruct or impede concrete placement or
vibration, which may result in a rate of delivery less than what might otherwise be expected. For
example: thin walls and balustrades are narrow; andheavily reinforced elements have little space
between bars.
Consideration should also be given the effects changes in plan area have on the rate of rise; e.g.
tapered walls and columns whose plan area varies with height.

2013 Stephen Ferguson all rights reserved 5.11


Formwork and Falsework

5.3.6 Economy
The direct relationship between rate of rise and the required formwork stiffness and strength provides
an apparent opportunity for economy by choosing a lower rate of rise. This must be balanced to
achieve an overall cost-effective solution. Potential formwork savings may be offset by additional
costs for labour to place the concrete and to deliver concrete in small quantities.
5.3.7 Limitations on formwork strength or serviceability.
It may be necessary to restrict the rate of concrete rise to avoid overloading formwork with a limited
capacity. This often necessary when using proprietary column and wall panel formwork systems.
5.4 Statics of concrete pressure
To design formwork, it is necessary to know the magnitude and shape of the concrete pressure
distribution. At this stage, it is useful to introduce the concept of a concrete pressure reduction factor
( ), such that:

hh
= 1.0 (5.11)
hc
D
Using , Equation 5.1 and 5.2 can be replaced by:

Pc = Pc max (5.12)

where
R
Pc max = ghc (5.13)

5.4.1 Vertical side formwork


AF
For vertical side formwork, it can be shown that the resultant force (Fp) per unit width produced by the
concrete pressure distribution acting at the centre of pressure at a height hp above the bottom of the
form, is given by either:
(a) If = 1.0 , the maximum concrete pressure is hydrostatic, see Figure 5.12(a), and:

Pc max hc
Fp = kN/m; and (5.14)
2

hc
hp = m (5.15)
T
3

(b) If 1.0 , the concrete pressure may be limited by concrete setting, see Figure 5.12(b),
and:

Fp =
Pc max hc
2
[ 2
]
1 (1 ) kN/m; and (5.16)

hc 1 (1 )3
hp = 2
m (5.17)
3 1 (1 )

Equations 5.16 and 5.17 can be considered generic expressions for all values of .

5.12 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

(a) Hydrostatic ( = 1.0 ) (b) Limited by setting( < 1.0 )

Figure 5.12 Concrete pressure distribution on vertical formwork


D
5.4.2 Inclined side or sloping soffit formwork
The situation is slightly more complex with inclined formwork.
5.4.2.1 Hydrostatic concrete pressure
Soffit or lower form
R
First consider the distribution of hydrostatic concrete pressure on an inclined surface AB, shown in
Figure 5.13(a). In accordance with Pascals Law, the concrete pressure distribution is as shown in
Figure 5.13(b).
AF
T
(a)
(b)

Figure 5.13 Hydrostatic concrete pressure distribution on an inclined soffit or lower surface
However, the distribution of hydrostatic pressure could also be considered equivalent to the vector
sum of the orthogonal pressure distribution components as shown in Figure 5.14, namely:

Pc = Pcx + Pcy (5.18)

2013 Stephen Ferguson all rights reserved 5.13


Formwork and Falsework

D
Figure 5.14 Vector components of hydrostatic concrete pressure distribution on an inclined soffit or
lower surface
This is the typical situation for a sloping soffit or the lower form of an inclined wall or column. In this
case, the resolution of the pressure distributions shown in Figure 5.14 is a useful for input into
structural analysis software and often makes hand calculations simpler.
R
Top or upper form
Next, consider the situation of the hydrostatic pressure on a top form of a sloping soffit or the upper
form of an inclined wall or column, as shown in Figure 5.15(a).
AF
T

(a) (b)

Figure 5.15 Hydrostatic concrete pressure distribution on an inclined top or upper surface
The pressure distribution is shown in Figure 5.15(b) and the resolution into orthogonal components in
Figure 5.16.

5.14 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

D
Figure 5.16 Vector components of hydrostatic concrete pressure distribution on an inclined top or
upper surface
5.4.2.2 Concrete pressure limited by setting
Soffit or lower form
R
First assume that the maximum concrete pressure shown in Figure 15.13(a) is limited by concrete
setting. Intuitively, the pressure diagram might be expected to be as shown in Figure 5.17, which is
often how it is presented in the literature.
AF
T
Figure 5.17 Incorrect concrete pressure distribution (limited by setting) on an inclined soffit or lower
surface
However, by considering the vector sum of the orthogonal pressure components shown in Figure 5.18
it can be shown that the distribution in Figure 5.17 is incorrect. The distribution shown in Figure 5.17
does not take account that the vertical component of concrete pressure is not limited by setting.
The correct concrete pressure distribution is as shown in Figure 5.19.

2013 Stephen Ferguson all rights reserved 5.15


Formwork and Falsework

Figure 5.18 Vector components of concrete pressure distribution (limited by setting) on an inclined
D
soffit or lower surface
In Figure 5.19, the magnitude of the component of concrete pressure normal to the bottom of the
inclined form (Pcx') is given by the expression:

Pcx ' = Pc max ( sin 2 + cos 2 ) (5.19)


R
Furthermore, the magnitude of the component of concrete pressure in the plane of the inclined form
Pcy' increases from zero at a depth of hh = hc to a maximum value at the full depth of the concrete at
point A given by the expression:

Pcy ' = Pc max sin cos (1 )


AF
(5.20)

The latter is significant as it results in an axial tension/compression force in the form, which is often
neglected in the literature and as a consequence may cause underestimation.
The validity of this solution can be seen by considering as approaches 90 degrees (closer to a
vertical form), the concrete pressure Pcx' (in Equation 5.19) approaches Pcmax and Pcy' (in Equation
5.20) approaches zero. Conversely, as approaches 0 degrees (closer to a horizontal form), Pcx'
approaches Pcmax, which approaches zero, and Pcy' (in Equation 5.19) also approaches zero.
T

Figure 5.19 Correct concrete pressure distribution (limited by setting) on an inclined soffit or lower
surface

5.16 2013 Stephen Ferguson all rights reserved


Chapter 5 - Concrete Pressure

Top or upper form


In a similar way, consider the orthogonal vector components of the concrete pressure distribution on a
top form of a sloping soffit or upper form of an inclined wall or column as shown in Figure 5.20.
D
Figure 5.20 Concrete pressure distribution limited by setting on an inclined top or upper form
In this case, the orthogonal distribution diagram is shown in Figure 5.21.
R
AF
T
Figure 5.21 Vector components of concrete pressure distribution limited by setting on an inclined top
or upper form
In either case, common errors can be avoided and hand calculations simplified by considering the
vector components shown in Figures 5.14, 5.16, 5.18 and 5.21, as appropriate.
References
Arslan, M. (2002). "Effects of drainer formwork on concrete lateral pressure." Construction and
Building Materials 16: 253-259.
Clear, C. A. and T. A. Harrison (1985). Report 108 - Concrete pressure of formwork, CIRIA.
CS (2012). Formwork A guide to good practice. Berkshire, The Concrete Society.
DIN (2010). DIN 18218 Pressure of fresh concrete on vertical formwork. Berlin, German Standards.
Pallett, P. (2000). "Construction joints and stop ends with Hy-Rib and ggbs concrete." Concrete
34(No. 10, Nov/Dec): 37-40.

2013 Stephen Ferguson all rights reserved 5.17


Formwork and Falsework

Pallett, P. (2009). "Concrete groups for formwork pressure deterimination." Concrete vol 43(no 2,
March): pp 44-46.
Proske, T. (2002). "Self-Compacting Concrete - Pressure on formwork and ability to deaerate."
Darmstadt Concrete 17.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
D
R
AF
T

5.18 2013 Stephen Ferguson all rights reserved


Chapter 6 - Side Formwork

6
Side Formwork
6.1 Introduction
The term "side formwork" covers a broad range of applications including slab edge, beam side,
wall and column formwork. The discussion herein will focus on design of wall formwork, but
principles apply equally to other types of side formwork.
D
R
AF
T

Figure 6.1 Wall formwork with secondary horizontal walers and primary vertical soldiers
Side formwork usually consists of: a form face, secondary members (horizontal walers or
vertical studs), primary members (vertical soldiers or horizontal walers), and form ties as shown
in Figures 6.1 and 6.2.

2013 Stephen Ferguson all rights reserved 6.1


Formwork and Falsework

D
R
AF
Figure 6.2 Wall formwork with secondary vertical studs and primary horizontal walers
The concrete pressure from freshly placed concrete is transferred via the one-way action of the
form face and framing members to the form ties and/or bracing, see Figure 6.3.
T

Figure 6.3 Load distribution through side formwork (McAdam and Lee 1997)

6.2 2013 Stephen Ferguson all rights


reserved
Chapter 6 - Side Formwork

The most common method of resisting the concrete pressure is to balance the force on opposing
faces using form ties in tension, see Figure 6.4.
D
R
Figure 6.4 Double-sided wall formwork (bracing not shown)
In some situations, form ties are not or cannot be used. In this case, the concrete pressure must
be resisted by bracing alone. Such a situation is typical for single-sided formwork, as shown in
AF
Figure 6.5.
T

Figure 6.5 Single-sided wall formwork


Irrespective of the method used to resist concrete pressure, side formwork requires bracing to
resist destabilising imposed actions and accidental impact. In addition, when formwork is

2013 Stephen Ferguson all rights reserved 6.3


Formwork and Falsework

inclined or opposing forms are not parallel, the formwork must be designed to resist out-of-
balance concrete forces.
In general, the principles governing the design of side formwork are adequately covered in the
literature (CS 2012) and apply equally to formwork designed using permissible stress or limit
states methods; however, some important aspects are often misunderstood and warrant attention.
In addition, in some cases, confusion exists about the correct application of limit states design
methods and this also warrants further explanation.
6.2 Form ties
6.2.1 Types of form ties
Form ties are tension members and usually consist of a rod or bar with a connector at each end.
Form ties can be fixed length or adjustable and some types of ties can be recovered and reused.
Common types of form ties are shown in Figure 6.6, including Bar ties, She-bolts, Coil ties, and
Snap ties.
D
R
AF
T

Figure 6.6 Common types of form ties

6.4 2013 Stephen Ferguson all rights


reserved
Chapter 6 - Side Formwork

Bar ties are adjustable and consist of a coarse threaded rod with nuts and plates. They can be
fully recovered by isolating the threaded rod from the concrete using a sleeve and cone
assembly. Parts of She-bolts and Coil ties can be recovered and reused, leaving only a threaded
rod cast-in the wall. Snap-ties are not intended to be recovered and reused. Form ties with non-
recoverable ferrous parts, such as She-bolts Coil ties and Snap ties, should be designed so that
no part of the tie remains within the concrete cover zone.
6.2.2 Form tie capacity
In determining the design capacity of form ties it is important to:
(a) Take account of the mode of failure ductile or brittle;
(b) Consequence of failure;
(c) Ensure form ties are more reliable than the members they join;
(d) Take account of wear and tear that may impair the strength of reusable sections;
and
(e) If reusable, ensure elastic behaviour at serviceability limit states.
In the past, the design capacity of form ties was established from tests using the ultimate
D
strength method (SAA 1974a). The permissible load of a form tie was established by dividing
the ultimate strength determined from tests by a load factor. The value of the load factor was
2.0 for non-reusable sections and 3.0 for reusable sections. Importantly, when determining the
ultimate strength from a limited number of tests any differences between the structural
properties of the materials tested and guaranteed minimum properties must be taken into
R
account.
6.2.2.1 Comment on the current design rules for tension members resisting
concrete pressure and other threaded tension members
Studies into the cause of falsework failures suggest that often the failure of side formwork is a
trigger for the collapse of falsework (Hadipriono and Wang 1986). In addition, researchers
AF
measuring the load in form ties found them to vary from predicted values. Thus, the
consequence of form tie failure is important and an increase in design actions or reduction in
capacity is necessary to take account o unanticipated load redistribution and the mode of failure.
Currently, AS 3610 1995 (SA 1995) permits the design capacity components to be established
by calculation or testing in accordance with Appendix A. In addition, AS 3610 2010 (SA 2010)
requires all threaded components to be free of wear, deformation or corrosion that might impair
strength.
A major short coming of testing in accordance with AS 3610 1995 Appendix A, is that it does
not distinguish between non-reusable and reusable sections, nor mode of failure (ductile or
T
brittle), nor whether the component tested is a member or connector, where the latter requires a
greater level of reliability.
To some extent this shortcoming is addressed in that AS 3610 1995 requires the design action
in tension members resisting concrete pressure, and other threaded tension members, to be
increased by 20%. This factor is intended to take account of the mode and consequence of
failure; i.e. sudden failure without warning and the risk of progressive collapse.
Unfortunately, the current requirement to increase the design action of tension members by 20%
can be easily overlooked because it is applied to the actions and not the capacity. An
undesirable consequence of introducing this load factor, versus reducing capacity, was
published capacities for some form ties ostensibly increased from earlier published data.
Amendment No 1 to AS 3610 1995 (SA 2003) introduced a global load factor to take account
of unanticipated load distribution. For members whose failure could cause collapse, the
magnitude of the global load factor is 1.3; otherwise, for all other members the magnitude of the
global load factor is 1.0. The global factor applies to form ties and is in addition to the 20%
increase in design actions.

2013 Stephen Ferguson all rights reserved 6.5


Formwork and Falsework

Herein the strength load factor for primary members d, replaces the global load factor
introduced in Amendment No 1 to AS 3610 1995 and the value has been recalibrated from 1.3
to 1.25 to be consistent when using the current combinations of actions in AS1170.0. (i.e. 1.2G
+1.5Q and 1.35G which supersede the combination 1.25G +1.5Q in AS 3610 1995), see
Section 4.3.2.3.
6.2.2.2 Recommendations on the design of tension members resisting concrete
pressure and other threaded members
For formwork members resisting concrete pressure in tension, and other threaded tension
members, increase the design forces calculated using the strength limit states action
combinations (see Section 4.3.2.2):
(a) By the strength load factor for primary members, d = 1.25, to take account of
unanticipated load distribution (see Section 4.3.2.3); and
(b) By a further 20%, except for members whose capacity has been reduced to take
account of mode and consequence of failure.
6.2.3 Serviceability limit states
Another shortcoming of the current design method is that it does not ensure elastic behaviour at
D
serviceability limit state.
Excessive movement or elongation at serviceability limit states is undesirable. Deformations at
serviceability limit states should be checked to ensure that are acceptable. Extremely ductile
materials may be unsuitable for use as formwork ties.
R
Guidance on ensuring elastic behaviour of multiple-use equipment is provided in Chapter 3. A
simple and more conservative approach may be to ensure elastic behaviour at ultimate limits
states, which negates the need to check serviceability limit states.
6.2.4 Form tie identification
A potential problem arises when form tie components from different manufacturers are mixed,
AF
i.e. nuts from manufacturer A used with threaded bar from manufacturer B.
Purchasing stock from one source or marking matching components may reduce the risk of
unintentional mixing.
Instances of two nuts being used to increase the capacity of suspect form ties are also cause
for concern. When two nuts are locked together, the 2nd nut carries the load. This practice does
not increase capacity and is not recommended. If the loaded nut fails, the load will be
transferred to the other nut which may also fail. The purpose of lock nuts is to prevent
unintentional loosening. In this case, the lock nut is placed on the bolt first followed by the full
nut, not vice-a-versa as is common practice.
T
6.2.5 Precautions when using form ties
Form ties are intended to be used in tension. The tensile capacity of form ties is eroded when
subject to shear and bending.
Due to the chemical properties of high tensile steel form ties, welding is not recommended.
Unless special procedures are followed, the strength and ductility of the form tie bar may be
impaired by welding.
In addition, high tensile form ties should not be hot dip galvanised due to the risk of brittle
fracture. Similarly, the risk of brittle fracture increases when high tensile steel under load comes
into contact with wet concrete. Accordingly, some suppliers of high tensile fasteners do not
recommend using steel with a tensile strength greater than 800 MPa when cast in concrete.
6.3 Double-sided wall formwork
6.3.1 Balanced concrete pressure
The most common and economical method of wall formwork is double-sided. The concrete
pressure on each face is balanced by form ties, as shown in Figure 6.7.
6.6 2013 Stephen Ferguson all rights
reserved
Chapter 6 - Side Formwork

D
Figure 6.7 Form ties balance concrete pressure on double-sided formwork
R
In the situation depicted in Figure 6.8, the pressure on the lower face is greater than on the upper
face. The form ties balance the horizontal concrete pressure on the upper face leaving an
imbalance on the lower face, which must be resisted by the formwork, bracing and their
connections to avoid instability.
AF
T

Figure 6.8 Form ties balance concrete pressure on double-sided inclined formwork

2013 Stephen Ferguson all rights reserved 6.7


Formwork and Falsework

6.3.2 Limit states design of wall form face and framing members
A concrete pressure envelope can be determined using Equations 5.1 to 5.3. For serviceability
limit states the action combination given in Equations 4.3 to 4.5 apply. For stability limit states,
the action combinations given in Equation 4.7 applies. For strength limit states, the action
combinations in Equations 4.12a and 4.13 apply.
The deformations and action effects in form face and framing members can be determined from
the pressure at the appropriate height intervals. For stability and strength limit states, the action
effects in primary beam (e.g. vertical soldiers or horizontal walers), braces and form ties should
be multiplied by factor d = 1.25 to take account of unanticipated load redistribution.
Form ties are often capable of exerting much higher forces on soldiers and walers than they can
resist in bearing, web yielding, or web buckling.
For timber formwork, often bearing capacity perpendicular to the grain will govern design. For
example, the characteristic bearing capacity of LVL products is of the order of 12 MPa;
therefore, twin LVL soldiers (50 mm apart) have a limit states bearing capacity of
approximately 50% of the limit states tensile capacity of commonly used high tensile form ties,
when exerted by a bearing plate 130 mm wide by 100 mm high.
D
For cold-form steel formwork members, often web buckling or web crippling will govern
design.
6.3.3 Unbalanced concrete pressure
Figure 6.9 shows a plan of side formwork. Where the opposing sides are not parallel, out-of-
balance effects arise. Axial forces are generated in the plane of the formwork. Instability might
R
be avoided by creating a shear connection between adjacent sides, as shown in Detail 1.
AF
T

Figure 6.9 Out-of-balance effects when opposing side forms are not parallel
Figure 6.10 shows various arrangements of inclined and tapered formwork with out-of-balance
effects. In each case, there is an imbalance of concrete pressure that results in an unbalanced
vertical action. Often the vertical action can be balanced by a restraint (tie or support) at the
base of the formwork. In addition, significant axial forces are generated in the formwork and it
is necessary to ensure that the formwork members are capable of resisting the combined effects
of bending, shear and axial tension or compression, as appropriate.

6.8 2013 Stephen Ferguson all rights


reserved
Chapter 6 - Side Formwork

(a)
D
R
AF
(b)
T

(c)
Figure 6.10 Out-of-balance effects on inclined and tapered wall formwork

2013 Stephen Ferguson all rights reserved 6.9


Formwork and Falsework

6.4 Single-sided wall formwork


In some situations, it is not possible to use form ties to balance concrete pressure. However,
considerable additional work (cost) is involved in bracing single-sided formwork. So much so, it
is recommended that single-sided formwork should only be adopted after careful consideration.
The action effects of pouring concrete in the single side formwork in Figure 6.5 are depicted in
Figure 6.11. A common mistake is to underestimate the magnitude of uplift to be resisted at the
base of the formwork. The connections anchor the formwork and brace should be designed for
the appropriate forces.
D
R
AF
Figure 6.11 Action effects on single-sided formwork

6.5 Bracing
Side formwork nearly always requires bracing to provide stability and alignment.
6.5.1 Bracing for alignment
T
Bracing is required at intervals along the length of side formwork to ensure transverse
alignment, in particular at the top of the formwork. The spacing of the bracing is a function of
the formwork stiffness, straightness and continuity.
Formwork may have sufficient stiffness to satisfy strength and serviceability requirements
across the short span between form ties, but along its length it may be slender and require
bracing at regular intervals.
In addition, horizontal framing members may not be straight and bracing may be required to
straighten the formwork. The sides of column formwork are often slightly twisted or prone to
twisting under load. Therefore, bracing is required to hold the formwork square.
Larger concrete elements, with long sides, may be formed by joining a series of formwork
panels together. Without bracing, the joints may be misaligned causing steps and/or angular
misalignment in the concrete face.

6.10 2013 Stephen Ferguson all rights


reserved
Chapter 6 - Side Formwork

6.5.2 Bracing for stability


6.5.2.1 Robustness
Bracing may also be necessary to ensure a minimum level of stability. This can be achieved by
designing the bracing members and connections to resist 2.5% of the weight of the formwork
and concrete, as well as imposed actions, see action Combinations 4.11b and 4.12b in Section
4.3.2.2.
6.5.2.2 Imposed actions
In addition to any out-of-balance effects from concrete placement, bracing an its connections are
also required to resist the appropriate combinations of destabilising effects of imposed actions.
For example: the weight of workers and equipment on cantilever platforms, minimum
horizontal actions associated with construction activity and wind actions.
Figures 6.12 and 6.13 depict examples of the imposed actions on side formwork, other than
concrete pressure. Refer to Chapter 4 for guidance on the appropriate combination of actions in
different situations.
D
R
AF
T

Figure 6.12 Imposed and notional actions on side formwork.


6.5.2.3 Accidental actions
In situations where there is a foreseeable risk of impact (e.g. from crane load, see Figure 6.13),
bracing should be designed to resist such an event in a manner that is proportional to the risk. It
may be acceptable that the formwork is damaged, but catastrophic collapse that endangers
workers should be prevented. The Commentary to AS 3610 1995 (SA 1996) provides useful
guidance on dealing with such an event.

2013 Stephen Ferguson all rights reserved 6.11


Formwork and Falsework

D
R
AF
Figure 6.13 Bracing for accidental impact
6.5.3 Bracing anchors
Where it is necessary to fix bracing to concrete using masonry anchors, deformationcontrolled
T
anchors, including self-drilling anchors and drop-in (setting) impact anchors, should not be used
AS 3850 2003 (SA 2003). Deformationcontrolled anchors are not suitable because they have
no additional load capacity after the initial setting process, fail without warning and are highly
sensitive to installation procedures.
Load-controlled anchors are preferred because they behave elastically until they first begin to
slip, after which they exhibit ductile load behaviour. The maximum load applied to these
anchors should be limited to 65% of the first slip load.
AS 3850 2003 provides useful guidance on assessing anchor capacity. AS 3850 recommends
that anchors that rely sole on chemical adhesion should not be used unless each fixing is
individually proof tested.
Masonary anchors should be used in accordance with the manufacturers recommendations.
Typically, the concrete in which masonary anchors are fixed should have a minimum strength of
at least 15 MPa and be at least three days old.

6.12 2013 Stephen Ferguson all rights


reserved
Chapter 6 - Side Formwork

References
CS (2012). Formwork A guide to good practice. Berkshire, The Concrete Society.
Hadipriono, F. C. and H.-K. Wang (1986). "Analysis of causes of formwork failures in concrete
structures." Journal of Construction Engineering and Management 112: 112 - 121.
McAdam, P. S. and G. Lee (1997). Formwork a practical approach. London, E & EF Spon.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney,
Standards Australia.
SA (2003). Amendment No. 1 to AS 3610 - 1995 Formwork for concrete. Sydney, Standards
Australia.
SA (2003). AS 3850:2003 Tilt-up constructionrequirements. Sydney, Standards Australia.
SA (2010). AS 3610 - 2010 Formwork for concrete Part 1: Documentation and surface finish.
Sydney, Standards Australia.
SAA (1974a). AS 1509-1974 SAA Formwork Code. Sydney, Standards Association of
D
Australia.
R
AF
T

2013 Stephen Ferguson all rights reserved 6.13


Chapter 7 - Soffit Formwork

7
Soffit Formwork
7.1 Introduction
Figure 7.1 shows the general arrangement of simple slab formwork, which is indicative of most soffit
formwork.
D
R
AF
T
Figure 7.1 Simple suspended slab formwork(McAdam and Lee 1997)

The attention of this Chapter is focused on aspects associated with the design of soffit forms. Chapter
8 addresses the design of the falsework that supports the soffit forms. In particular, this Chapter
discusses:
(a) The load path for vertical loads through the formwork;
(b) Loading patterns;
(c) Load distribution in both horizontal and sloping soffit forms;
(d) Sloping soffit formwork; and
(e) Unbalanced concrete pressure and discontinuous formwork.

2013 Stephen Ferguson all rights reserved 7.1


Formwork and Falsework

7.2 Load path for vertical loads through soffit formwork


Typically:
The form face material is subject to a uniformly distributed pressure and spans one-way
between the secondary members (joists, studs, etc);
The secondary members are subject to a uniformly distributed load and span simply
supported between primary members or continuously over several primary members
(bearers); and
The primary members are subject to point loads, from the secondary members, and span
simply supported between supports or continuously over several supports (props, shores,
etc), see Figure 7.2.
D
R
AF
Figure 7.2 Load distribution through soffit formwork (when viewed from underneath)

7.3 Loading Patterns


7.3.1 Stage 1 - Prior to concrete placement
T

Figure 7.3: Line or point loads that arise from stacked materials during Stage 1 (SA 1996)

7.2 2013 Stephen Ferguson all rights reserved


Chapter 7 - Soffit Formwork

Prior to concrete placement, the soffit forms may experience heavy loading from materials stacked on
the formwork. Importantly, the loads may be point or line loads rather than uniformly distributed and
the consequential detrimental effects need to be considered.
Figure 7.3 depicts bundles of reinforcement placed on top of the soffit forms. To facilitate
disconnecting and removing the crane chains it is common practice to land reinforcement and other
loads on timber spreaders or "gluts". Doing so effectively causes the form face directly under the glut
to be loaded with a line load or the secondary members by point loads.
7.3.2 Stage 2 - During concrete placement
During concrete placement, concrete is generally placed, starting from one side of a slab pour,
progressively across the formwork. Thus, it is possible that during concrete placement only one span
of a continuous joist or bearer may be loaded, see Figures 7.4 and 7.5. In general, all spans of
continuous beams will not be fully loaded until the concrete front has passed.
D
R
AF
Figure 7.4: Adverse partial loading of multiple span bearer (McAdam and Lee 1997)
T

Figure 7.5: Importance of direction of pour (McAdam and Lee 1997)

2013 Stephen Ferguson all rights reserved 7.3


Formwork and Falsework

7.3.2.1 Stability limit states


Rarely are secondary or primary beams connected to their supports in a way that would prevent uplift,
which can occur when only one span of a continuous span beam is loaded, see Figure 7.4. This is an
important situation to consider when determining the strength and stability of formwork members, as
well as the assembly as a whole. In particular, in the case of cantilever formwork failing to prevent
uplift may cause overturning, see Figure 7.5.
7.3.2.2 Strength limit states
Multiple span bearers and joists should be designed for the most adverse action effects arising from
the loading of any valid arrangement of one or more spans. This is the reason that all joists and bearers
spanning two or more equal spans should be designed to resist the:
(a) Bearing and shear forces for the central support of for a two-span arrangement; and
(b) Bending moments that arises mid-span of a single simply supported span or over the
central support of a two-span arrangement.
Serviceability limit states
For form deflection, it is common practice to assume that deformations are elastic and the interim
D
larger deformations on the loaded spans of a partially loaded joist or bearer will not be present when
all spans are loaded. In this case, the interim deflections can be ignored and only the deflections
associated with the final loading arrangement need be considered, as this is the deflected shape the
concrete will retain when set.
7.3.3 Stage 3 - After concrete placement
R
After concrete placement, concrete quickly gains strength and it is often assumed that the slab is
sufficiently stiff to span between closely spaced supports. In this situation, any subsequent loads (such
as stacked materials) need only be taken into account in the design of primary members and their
supports; i.e. need not be taken into account in the design of conventional form face and secondary
members. However, this assumption may not be valid in all situations and must be verified, in
particular, for longer spanning members, e.g. long span permanent formwork.
AF
7.4 Analysis of soffit form members
7.4.1 Point loads vs UDL
Soffit form primary beams (bearers) are subject to a series of point loads from each secondary beam
(joist), see Figure 7.6.
T
Figure 7.6: Point loads from secondary beams (joists) acting on the primary beam (bearer)
A comparison of the effect of point loads versus an equivalent uniformly distributed load on a
continuous primary beam (Ikheimonen 1997) shows:
(a) When calculating beam reactions (shore loads), in most cases, point loads can be replaced
by a uniformly distributed load without large errors; and
(b) Deflections, bending moments, and shear stresses due to point loads could be higher than
if replaced by an equivalent uniformly distributed load.
In practice, it is much simpler to analyse the action effects in primary beams by replacing point loads
with an equivalent uniformly distributed load. In nearly all cases, the application of the strength load

7.4 2013 Stephen Ferguson all rights reserved


Chapter 7 - Soffit Formwork

factor for primary members (d = 1.25) will take account of any underestimation of shear force and
bending moment that might arise by replacing point loads with a uniformly distributed load.
7.4.2 Lateral buckling of beams
In conventional formwork, it is common practice to use members whose height to width ratio is
approximately 2 or less; e.g. common LVL sections 95x47, 95x65, 130x77 and 150x77. Although less
efficient as beams, these members are less likely to roll or fall over and are less susceptible to lateral
buckling than are narrow members, see Figure 7.7.
D
Figure 7.7: Narrow timber beams required lateral restraint (McAdam and Lee 1997)
R
In most situations, it is assumed the form face and secondary beams provide effective lateral restraint
to the secondary and primary beams, respectively. Consequently, in most conventional formwork
design situations the lateral buckling of beams need not be considered. However, connection details
between members should be checked to ensure they are consistent with this assumption. In particular,
lateral buckling should be considered when using slender members (whose height to width ratio
AF
exceeds 2) and when channel sections are used.
7.4.3 Simply supported beams
Often the action effects in beams can be based on one of three simple beam load cases, namely: a
simply supported beam with a uniformly distributed load on one, two or three or more spans. In this
case, the action effects for the maximum beam reaction, bending moment, shear and deflection can be
represented by the following equations:

Maximum reaction, = (7.1)

Maximum bending moment, = 2


T
(7.2)

Maximum shear force, = (7.3)

4
Maximum deflection, = (7.4)

In Equations 7.1 to 7.4,


(a) The coefficients for are given in Figure 7.8;
(b) For secondary and primary beams, the magnitude of the uniformly distributed load w
should be calculated from the most adverse combination of actions and after taking
account of the continuity of the members supported by, and the reaction, on the beam
under consideration.
In some situations, a more rigorous analysis is required and the use of beam analysis software would
be appropriate.

2013 Stephen Ferguson all rights reserved 7.5


Formwork and Falsework

D
R
AF
T

Figure 7.8: Coefficients for beam action effects

7.6 2013 Stephen Ferguson all rights reserved


Chapter 7 - Soffit Formwork

7.5 Sloping soffit formwork


For analysis purposes, it is conservative to neglect any effect reinforcement will have on preventing
concrete flow down a slope.
7.5.1 Vertical falsework
If concrete remains at rest on a sloping soffit because of friction between the form face material and
the wet concrete, only vertical forces are applied to the falsework, see Figure 7.9.
D
R
Figure 7.9 Concrete and formwork held at rest on a sloping soffit by friction
Figure 7.10 shows the concrete and form face are prevented from sliding together because of the
friction between the form face material and the joists. Similarly, friction prevents sliding between the
joists and bearers, as well as between bearers and wedges.
AF
T

Figure 7.10 Action effects on sloping soffit formwork


However, if the concrete does not remain stationary, additional forces parallel to the form surface will
arise due to changes in momentum at the commencement, surge, or cessation of concrete flow that
must be taken into account.
If friction between the layers of framing members is insufficient to resist sliding, movement will occur
and a mechanical connection is required. In this case, the stabilising effects of friction are ignored and
the mechanical connection designed to resist the entire destabilising effects causing sliding.

2013 Stephen Ferguson all rights reserved 7.7


Formwork and Falsework

7.5.2 Stability limit states Sliding


To satisfy the assumption that the concrete and formwork remain at rest due to friction and therefore
satisfy stability limit states:

Rd Ed,dst (7.5)

where

Rd = (0.9Gf + 0.9Gc)cos() (7.6)

Ed,dst = (1.35Gf + 1.35Gc)sin() (7.7)

In Equation 7.6, is the design coefficient of static friction. In Appendix B, Table B1 provides
guidance on appropriate values for the static friction coefficients () for use in the limit states design
of temporary structures with a capacity factor, = 0.8.
Destabilising combinations of actions including imposed actions need not be considered as these
actions cannot be applied to wet concrete.
D
Substituting Equations 7.6 and 7.7 into 7.5 yields

1.88 tan ( ) (7.8)

or
R

arctan (7.9)
1.88

Equations 7.8 and 7.9 are useful for verifying that stability limit states for sliding have been satisfied.
AF
Example 1 Verification of stability limit states for soffit formwork sliding
For sloping soffit formwork constructed from film faced plywood and softwood joists and bearers,
determine the maximum slope that friction resistance alone will satisfy stability limit states.
From Table B1, film-faced plywood sliding on softwood timber perpendicular to the grain =0.1.
Softwood sliding on softwood perpendicular to the grain = 0.3.
0.1
Therefore, to satisfy stability limit states, for film faced plywood, arctan 3.0 ; and for

1 . 88
0.3
T
softwood timber, arctan 9.0

1.88
For slopes greater than 3.0, the fasteners holding the plywood to the joists would have to resist
sliding. Similarly for slopes greater than 9.0, to resist sliding the joists would have to be fixed to the
bearers and, likewise, the bearers would have to be fixed to the wedges.
7.5.3 Stability limit states Overturning
It may be possible for narrow joists or bearers running across the slope to overturn. In this situation,
the friction forces form a destabilising couple that is balanced by the stabilising component of the
vertical action effects. To satisfy stability limit states,

Ed,stb Ed,dst (7.10)

Ed,stb = (0.9Gf + 0.9Gc) cos() (b/2) (7.11)

Ed,dst = (1.35Gf + 1.35Gc)sin()(d) (7.12)

7.8 2013 Stephen Ferguson all rights reserved


Chapter 7 - Soffit Formwork

In Equations 7.11 and 7.12, the variables b and d refer to the width and depth of the joist shown in
Figure 7.11.

Figure 7.11 Destablising and stablising action effects on joists running across the slope

Substituting Equations 7.11 and 7.12 into 7.10 yields

3.0 tan ( )
b
(7.13)
D
d

Equation 7.13 is useful for verifying that stability limit states for overturning have been satisfied.
Example 2 Verification of stability limit states for joists overturning
Given that 9 is maximum slope at which softwood joists will satisfy stability limit states for sliding,
R
determine the minimum width depth ratio for softwood joists running across the slope that satisfies
stability limit states for overturning.

Therefore,
b
d
( )
3.0 tan 9 0.5
AF
For slopes up to 9, softwood timber joists or bearers running across the slope with a depth to width
ratio equal to or less than 2 to 1 will satisfy stability limit states for overturning.
Overturning can also be prevented by lateral bracing.
7.5.4 Out-of-vertical falsework
If the falsework supporting sloping soffit formwork is designed out-of-vertical, as shown in Figure
7.12, it must be capable of resisting the combined actions of both axial forces and destabilising forces
in the plane of the soffit.
T

Figure 7.12 Concrete held at rest on a sloping soffit by friction, with out-of-vertical falsework

2013 Stephen Ferguson all rights reserved 7.9


Formwork and Falsework

7.6 Unbalanced concrete pressure - Discontinuous soffit formwork


For nominally level soffit formwork, the fluid concrete is contained within an area by edge forms or
stop ends. It is usually assumed that the concrete exerts an equal and opposite pressure on the
opposing edge forms or stop ends and the soffit is continuous; therefore, these forces are balanced and
there is no effect on the falsework.
Initially during placement, concrete behaves as quasi-fluid. When placed against a form edge, the
concrete exerts pressure on the edge form which is fixed and braced to the soffit. The horizontal force
from the edge form is transferred into the soffit and balanced by friction between the wet concrete and
form face material. On larger pours, concrete pressure is exerted on previously placed concrete is
transferred by friction between the setting concrete and form face material and balanced in a similar
manner.
However, if the soffit formwork is discontinuous, the falsework must resist horizontal forces from
lateral concrete pressure. Figures 7.13 and 7.14 show examples of situations where the soffit
formwork is discontinuous.
D
R
AF
T

Figure 7.13 Examples of discontinuous soffit formwork where horizontal forces from lateral concrete
pressure are transferred to the falsework (CS 1995)

7.10 2013 Stephen Ferguson all rights reserved


Chapter 7 - Soffit Formwork

To avoid horizontal forces being transferred to the falsework in the situations shown in Figure 7.13 (c)
and 7.13 (d), the soffit formwork must be continuous and have sufficient capacity to balance the
opposing horizontal forces in tension.
The falsework must also be designed to resist horizontal forces when concrete is cast on soffit
formwork erected adjacent to but not connected to an existing or previously constructed structure or
where the edge form is not connected to the soffit form, see Figure 7.14.
D
Figure 7.14 Examples of horizontal forces being transferred to falsework, when concrete is cast on
formwork adjacent to but not connected to an existing structure (CS 1995)

References
CS (1995). Formwork A guide to good practice. Berkshire, The Concrete Society.
R
Ikheimonen, J. (1997). Construction Loads on Shores and Stability of Horizontal Formworks.
Department of Structural Engineering. Stockholm, Royal Institute of Technology: 161, 76.
McAdam, P. S. and G. Lee (1997). Formwork a practical approach. London, E & EF Spon.
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney, Standards
AF
Australia.
T

2013 Stephen Ferguson all rights reserved 7.11


Chapter 8 - Falsework

8
Falsework
8.1 Introduction
Formwork falsework is often heavily loaded. In addition, when falsework structures are tall and
slender their capacity is sensitive to detrimental second-order effects. Furthermore, permitted
tolerances for falsework structures and members are greater than for permanent structures, and
falsework connections are semi-rigid with complex behaviour, or rely on friction. Thus, the capacity
of falsework is difficult and complex to determine, and easily over-estimated.
D
To simplify falsework design, designers make assumptions upon which the safety of the falsework
depends. Designers often base their design on assumptions about:
1. The magnitude of loads acting on the falsework, including their presence or otherwise,
and their path through the structure to points of restraint, for example;
R
(a) The soffit formwork acts as a diaphragm and all horizontal loads are transferred to
the permanent structure;
(b) No allowance for stacked materials;
(c) No allowance for multi-storey loads; and
AF
(d) Side forms do not induce any side loads into the falsework (i.e. internally tied).
2. The strength and stability of the falsework structure, for example:
(a) The top of the falsework is restrained by the permanent structure (i.e. the falsework
is not freestanding);
(b). The falsework is fully braced and will not sway;
(c) The falsework shores are loaded concentrically or within a specified minimum
T
eccentricity;
(d) The falsework is not subject to differential settlement or axial shortening; and
(e) The falsework is built from components that are undamaged (i.e. in as new
condition) or whose imperfections are within permitted tolerances.
All the design assumptions should be stated in the formwork documentation. However, in practice this
is not always the case.
Assumptions that simplify design should be conservative; i.e. tend toward overestimating loads and
underestimating capacity. However, in practice this is not often the case. For example, all the
assumptions listed above are common place and, yet, all risk either underestimating loads or
overestimating the falsework capacity. The risk of failure increases if the design assumptions are
inappropriate for, or not applicable to, the actual formwork construction.
In practice, the onus for checking that the design assumptions are valid and applicable to the situation
on site falls to others; e.g. those responsible for checking the formwork design, supervising the
formwork construction, and inspecting the completed formwork.

2013 Stephen Ferguson all rights reserved 8.1


Formwork and Falsework

In situations where the formwork design is split between two or more parties, one party must take
overall responsibility to ensure the design assumptions for each part of the design are consistent, valid
and applicable. For example, if the falsework and formwork are designed by different parties, one
party must ensure the falsework design assumptions are consistent with the formwork design and vice
versa.
To reduce the risk of falsework failure, in addition to the formwork and falsework designer(s), it is
necessary for all those involved in co-ordinating, checking, supervising and inspecting formwork and
falsework to have knowledge of the:
1. Loads and load combinations that should be taken into account in the design; and
2. Aspects of the falsework design that have a major influence on strength and stability.
Those involved in designing the formwork and falsework, as well as checking the design require the
greatest depth of knowledge and understanding. Those supervising and inspecting the construction
need not have the same level of understanding as the designer, but must be able to determine if the
design assumptions are valid for the situation at hand. Those responsible for co-ordinating and
constructing the formwork need only a basic understanding.
D
Section 8.2 provides an overview of the load situations encountered at each stage of construction,
which should be taken into account in the falsework design as they may be critical to falsework safety.
Section 8.3 provides guidance on aspects that should be taken into account in determining the
behaviour and action effects that result from the vertical and horizontal loads. Section 8.4 provides
guidance on calculating the capacity of falsework members; in particular, how to take account of the
R
combined effects of axial forces and bending moments in falsework shores.

8.2 Falsework design actions


To determine the most adverse design situations, for each stage of construction, consideration should
be given to all valid combinations of foreseeable direct, indirect, and accidental design actions (see
AF
Chapters 4 and 7).
For falsework design, in addition to the vertical loads, it is particularly important that proper account is
taken of horizontal loads, namely:
(a) Horizontal loads from construction activity and other sources, see Section 4.2.3;
(b) Wind actions, see Section 4.2.4;
(c) Notional loads for initial out-of-plumb erection and bracing forces, see Section 4.2.6; and
T
(d) Minimum notional horizontal loads that ensure structural integrity, see Section 4.2.6
It is important to remember that after analysis of the structure, the resulting strength limit states action
effects in all the primary members should be amplified by the strength factor for primary members,
d = 1.25, see Section 4.3.2.

8.2.1 Stage 1 - Prior to concrete placement


Prior to concrete placement, falsework structures are typically vulnerable to destabilising horizontal
actions such as wind and accidental impact. This is accentuated where the stability of falsework relies
on friction for connectivity; i.e. where the formwork soffit forms are intended to act as a diaphragm to
transmit horizontal loads. At this stage of construction, without significant vertical actions, the
friction between the formwork components may be insufficient for the soffit form to act as a
diaphragm and transmit the loads to points of restraint.

8.2 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

8.2.2 Stage 2 - During concrete placement


Nearly half of all falsework failures occur during concrete placement (Hadipriono and Wang 1986).
At this stage, the falsework is vulnerable to combinations of vertical and destabilising horizontal
actions.
The design must take account of destabilising horizontal actions which may arise from: unbalanced
lateral concrete pressure, construction activity, wind, water, accidental impact, out-of-vertical
members, out-of-plumb erection, differential settlement, temperature changes, imposed acceleration
from equipment or due to machine excitation, and bracing out-of-straight members, see Chapters 4 and
7.
The sequence of concrete placement may also affect stability; e.g. uplift and overturning if cantilevers
are loaded first, see Chapter 7.

8.2.3 Stage 3 - After concrete placement


Often, during this stage, the falsework experiences it maximum vertical loading. This occurs due to
additional load from construction activity on the completed slab. In particular, additional loads may
D
arise from: the operation of equipment; multi-storey loading; stacked materials; post-tensioning of the
slab; axial shortening, shrinkage and creep of concrete; and premature removal of formwork and
falsework.
At this stage of construction, the permanent structure (newly cast slab) usually provides additional
lateral restraint that was not present during concrete placement. Consequently, although the falsework
R
may experience is maximum vertical load, the degrading effects of horizontal actions may not be as
great and this stage of construction may not govern design.

8.3 Factors influencing falsework behaviour, stability, and strength


AF
Investigations into falsework collapses (Bragg 1975; Hadipriono and Wang 1986) identify inadequate
bracing as the primary cause of falsework failure. Inadequate bracing may result if the design
assumptions are inappropriate (e.g. see Section 8.1 paragraphs 1(a) and (d), as well as 2(a) to (e)).
Falsework bracing serves several purposes. It may be necessary:
(a) For practical purposes to maintain stability during, and facilitate, erection;
(b) To reduce sway, so as to satisfy serviceability limit states;
(c) To transmit design actions (direct and indirect) from points where they arise to anchorage
or reaction points at the foundation or permanent structure; and/or
T
(d) To restrain and reduce the effective length of compression members.
The stability and strength of the falsework will depend on whether the falsework is adequately braced.
Falsework structures are nearly always intended to be fully braced frames that do not sway; therefore,
are subject to primarily axial loads. Often, for economic reasons, designers omit or minimise the
amount of bracing without considering (or understanding) the consequence. If the bracing is
inadequate:
1. Falsework will sway;
2. Axial forces in bracing will increase;
3. Bending moments will arise in the vertical members that significantly reduce capacity;
and
4. Any assumptions about column effective length may no longer be valid.

2013 Stephen Ferguson all rights reserved 8.3


Formwork and Falsework

To avoid overestimating the stability of the falsework and underestimating the action effects in
members and connections requires careful consideration of how the falsework is restrained and braced,
the consequential effects on falsework behaviour, and how this behaviour can be properly taken into
account in a structural model.
Section 8.3 looks at the different ways falsework can be restrained and braced, as well as the
effectiveness of each.

8.3.1 Falsework restraint


The stability, strength and behaviour of the falsework will depend on whether the falsework is
effectively restrained or freestanding.
Falsework can be considered to be: freestanding (unrestrained), top restrained or partially restrained,
as shown in Figures 8.1, 8.2 and 8.3 respectively. Figures 8.1, 8.2 and 8.3 depict falsework with
braces that have pin connections.
In Figures 8.1, 8.2 and 8.3, the initial shape of the falsework is depicted on the figure on the left hand
side and the possible deformed shape under purely vertical loads is depicted in the figure(s) on the
D
right-hand side. Predicting the deformed shape is useful in estimating the effective length of the
vertical members and in turn their capacity.
Sway occurs where one end of a vertical member is displaced horizontally relative to another. In all
the cases in Figure 8.1, 8.2 and 8.3, sway has occurred. Assumptions that these arrangements are fully
braced frames subject to only axial loads would not be valid.
R
AF
T

8.4 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

8.3.1.1 Freestanding

a) (b)
D
R
(c) (d)
AF
T

(e) (f)

Figure 8.1 Unrestrained or freestanding falsework

2013 Stephen Ferguson all rights reserved 8.5


Formwork and Falsework

8.3.1.2 Top restraint

(a) (b)
D
R
AF
(c) (d) (e)
T

(f) (g) (h)

Figure 8.2 Top restrained falsework

8.6 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

8.3.1.3 Intermediate restraint

(a) (b)
D
R
(c) (d) (e)
AF
T

(f) (g) (h)

Figure 8.3 Intermediate restraint

2013 Stephen Ferguson all rights reserved 8.7


Formwork and Falsework

8.3.1.4 Requirements for formwork to be considered top restrained

If permitted by the project designer, the surrounding permanent structure is often assumed to provide
restraint to the falsework structure. This is achieved if the formwork on top of the falsework is
laterally restrained in all directions and rotationally by the permanent structure, which means the
following must apply:
(a) The formwork provides load paths laterally and rotationally (i.e. members and connections of
sufficient strength and stiffness) from the point of application of the horizontal loads to the
points of restraint, see Figure 8.4;
(b) The connection from the formwork on top of the falsework to the permanent structure has
sufficient strength and stiffness to transmit the accumulated horizontal loads; and
(c) The permanent structure has sufficient strength and stiffness to provide restraint.
It must be justified to assume that the formwork soffit acts as a diaphragm with sufficient lateral and
rotational strength and stiffness to transfer the horizontal forces to points of restraint.
D
R
Figure 8.4 Designation for the load paths required to provide full lateral and rotational restraint

Rarely, does the surrounding permanent structure provide lateral restraint in all directions and
AF
rotational restraint. Examples of different levels of restraint are shown in Figures 8.5 and 8.6.
T
(a) (b) (c)

(d)
(e)

Figure 8.5 Differing levels of top restraint provided by surrounding walls

8.8 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

(a) (b)

(c) (d)
D
Figure 8.6 Differing levels of top restraint provided by columns

It is important to verify, if it is permitted to use the permanent structure to provide restraint, that the
permanent structure is structurally adequate and sufficiently stiff to provide the required degree of
restraint. For example, tall slender concrete columns may be too flexible to provide effective restraint.
R
It is important that the designer identify or provide for load paths from the point of application of
horizontal actions to points of restraint. The members and connections between the point of
application of the horizontal loads and points of restraint should be capable of transferring the
accumulated horizontal action, without detrimental effects or displacements that would degrade the
level of restraint.
AF
Where top restraint in any particular direction (or rotationally) is not provided, the falsework capacity
should be based on it being freestanding in that direction (or free to rotate) and the falsework should
be designed as an unrestrained freestanding structure whose members and connections are capable of
bracing the vertical members to provide stability, resist vertical loads, and provide a load path from the
point of application of the horizontal loads to a point of restraint.

8.3.2 Falsework - Sway or fully braced frames


A member can be considered a sway member if the transverse displacement of one end of the member
T
relative to the other end is not prevented. This can occur if a member is not braced or not braced
effectively. Due to the translation of the ends of sway members, second order effects (e.g. bending
moments) are introduced, which must be taken into account.
All the frames depicted in Figures 8.1, 8.2 and 8.3 contain sway members. Similarly, the tower on the
left of Figure 8.7 has sway members. For all members to be considered fully braced, it would be
necessary to brace the tower as shown on the right of Figure 8.7, providing the bracing used is fully
effective (see the following Sections).
The amount of sway, and falsework buckling capacity, is sensitive to the length of screw jack
extension at the top and bottom of the falsework. Whenever possible, the screw jack extension should
be minimised.

2013 Stephen Ferguson all rights reserved 8.9


Formwork and Falsework

D
Figure 8.7 Freestanding falsework with sway and fully braced members

Similarly, Figure 8.8 shows different bracing configurations for falsework with multiple columns
connected by horizontal bracing. The bracing arrangement on the right in Figure 8.8 is recommended
R
to avoid the bracing from concentrating the vertical action effects due to horizontal forces into the first
two columns.
Both falsework structures shown in Figure 8.8 contain sway members. For all the members to be
considered fully braced, it would be necessary to effectively brace the falsework, including the top and
bottom screw jacks, in a manner similar to that shown in Figure 8.9.
AF
T

Figure 8.8 Freestanding falsework with multiple column bracing

8.10 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

D
Figure 8.9 Freestanding falsework with fully braced members

Figures 8.7, 8.8 and 8.9 depict freestanding falsework. To avoid sway members in falsework that is
R
top restrained, it is also necessary to brace the top and bottom screw jacks in manner similar to that
shown in Figure 8.10; otherwise bending moments are introduced, which must be taken into account.
Failure to do so may result in overestimating the capacity of the falsework.
AF
T

Figure 8.10 Top restrained and fully braced falsework

2013 Stephen Ferguson all rights reserved 8.11


Formwork and Falsework

8.3.3 Falsework bracing


This Section explains aspects that affect the design of bracing, especially bracing intended to reduce
the effective length of falsework shores.

8.3.3.1 Problems with existing design criteria for falsework bracing

Unfortunately, the design criteria for falsework bracing (restraint) systems specified in the relevant
material Standards may not be adequate when applied to formwork falsework, because:
(a) AS 3610 1995 (SA 1995) permits the use of compression members with a larger out-
of-straightness than is permitted in material Standards (e.g. AS 4100);
(b) The forces that arise in braces that reduce the effective length of compression members
increase with member out-of-straightness.
(c) When more than one compression member is connected by a line of horizontal bracing,
the minimum brace stiffness required to reduce the effective length of all the compression
members increases non-linearly;
D
(d) In addition, when compression members are braced at multiple points along their length,
the minimum brace stiffness required to reduce the effective length of the compression
member increases non-linearly;
(e) In some cases, the connection of horizontal bracing members behaves, not as a pin joint,
R
but more like a semi-rigid joint; and
(f) Often braces are connected eccentrically from node points and/or in a manner that
reduces the effective axial stiffness of the bracing.
In this Section, we investigate the influence each of these factors.
AF
8.3.3.2 Concept of effective length

Most designers are familiar with the concept that the elastic buckling load of a braced member (Nomb).
The magnitude of the elastic buckling load is dependent on the end restraint provided by the
surrounding framework.

2 EI
N omb ( k e l ) = (8.1)
(k e l )2
T
where
E = Youngs modulus of elasticity;
I = second moment of area of the cross section;
k el = member effective length, which is the product of:
ke = member effective length factor; and
l = member length
For members with idealised end restraints the value of the member effective length (ke) is given in
Figure 8.11. Unfortunately, the effective length of members in braced frames, like falsework, is more
complex to estimate.

8.12 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

D
Figure 8.11 Effective length factors for members with idealised end restraints (SA 1998)

Another familiar concept is the need to brace members along their length to reduce effective length
and increase axial capacity. Figure 8.12 depicts an axially loaded compression member with a single
R
central brace. In Figure 8.12, the brace is idealised as a spring, this reflects the knowledge
(Timoshenko and Gere 1961) that for the brace to be effective it must have a minimum axial stiffness
(k) such that:

4 N omb
k (8.2)
AF
l

If the inequality in Equation 8.2 is satisfied, the member effective length factor ke = 0.5 and the elastic
buckling capacity increases four fold.

4 2 EI
N omb = (8.3)
l2
T

Figure 8.12 Braced column with an initial out-of-straightness imperfection o

2013 Stephen Ferguson all rights reserved 8.13


Formwork and Falsework

8.3.3.3 Minimum brace axial stiffness and forces in braces intended to reduce effective
length of falsework shores

Member out-of-straightness

In a straight ideal column, braced to reduce effective length, there are no forces in the braces. To be
effective the braces must merely satisfy minimum axial stiffness requirements similar to Equation 8.2.
However, Figure 8.12 also depicts that real members have geometric and material imperfections
(represented by an initial out-of-straightness o). Under load, these members will deflect immediately,
which introduces real forces into the brace and reactions at the end restraints.

Single column with single brace

Figure 8.12 shows the classic situation of a single column with an intermediate brace that is intended
to reduce the effective length of the column. In this situation, the current method in AS 4100 1998
Steel structures (SA 1998), and adopted in AS 3610 1995, is to design column braces to resist an
axial force of 0.025Nd.
However, the methods in AS 4100 and AS 3610 1995 differ. In AS 4100, the structure is designed
D
to resist the greater of either: the design actions and notional forces or the application of the bracing
force 0.025Nd. AS 3610 1995 requires the bracing force to be considered to act in conjunction with
other design actions and notional forces. In effect, the bracing force acts as a notional load, see
Chapter 4.
By following the requirements in AS 4100 to design column braces to resist an axial force of 0.025Nd,
R
it is implicit that 0.025Nd exceeds the magnitude of the brace forces, Equation 8.2 will be satisfied and
brace stiffness need not be checked.
Studies (Clarke and Bridge 1994; Trahair 1999) to determine the value of the brace design force Nb
found:
AF
(a) the force in the brace is directly proportionally to the initial out-of-straightness imperfection;
and
(b) that for columns complying with the permitted out-of-straight tolerances in AS 4100 (l/1000),
the design criteria of 0.025Nd is conservative.
Trahair suggests, for columns with a permitted out-of straightness of l/1000, a brace force design
criteria between 0.005Nd to 0.015Nd.
Based on these findings and the knowledge that AS 3610 1995 permits falsework members to be
out-of-straight up to l/300 compared with the lower limit of l/1000 permitted in AS 4100, it might be
T
argued that an appropriate design criteria for falsework bracing intended to reduce the effective length
of columns be increased by a factor of 1000/300 and should fall between from 0.015Nd to 0.045Nd.
Use of the design criteria of 0.025Nd for formwork bracing (see Section 4.2.6.2) might be seen to be
consistent with a permitted out-of-straight tolerance of l/600, because, in this case, the range of
bracing force criteria suggested by Trahair, factored by 1000/500, would fall between from 0.008Nd to
0.025Nd.

Multiple columns with single line of braces

A more common situation in falsework is when multiple parallel columns are connected by a line of
bracing, as modelled in Figure 8.13.
Guidance provided in Clause 6.6 of AS 4100 and Clause 4.4.6 of AS 3610 1995 suggests that the
first brace should be designed to transfer a force of 0.025Nd and all subsequent braces, up to a
maximum of seven, be designed to transfer a force of 0.0125Nd.

8.14 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

Thus in Figure 8.13, the design force accumulates along the line of bracing reaching a maximum of
0.025 + (4 0.0125) = 0.075 N d . There appears no justification for this rule in the literature, other than
it is reasonable to assume that the imperfections are not uniform.
D
R
Figure 8.13 A series of parallel out-of-straight columns restrained by a line of bracing

An analysis of the arrangement in Figure 8.13 demonstrates that, if the imperfections are uniform,
then:
(a) the force in the bracing increases proportionally to the number of columns; and
AF
(b) to be effective the brace stiffness increases non-linearly, see Figure 8.14.
T

Figure 8.14 Brace stiffness multiplier for a series of parallel 48.3CHS4.0 columns 3 metres long and
restrained by a line of bracing at mid point

2013 Stephen Ferguson all rights reserved 8.15


Formwork and Falsework

Figure 8.14 depicts the brace stiffness multiplier (a) such that:

k n an k (8.4)

where
kn = minimum axial stiffness of brace number n
an = brace stiffness multiplier given in Figure 8.14
k = minimum axial brace stiffness for a single column, see Equation 8.2
In practice, all braces would have to satisfy the maximum value of kn.
Of concern is that the current criteria, to design braces to resist a design force is based on the brace
stiffness required for a single column, does not take into account the non-linear increase in brace
stiffness required for braces that connect multiple columns.
D
Single column with multiple braces

Another variation on the single column with a central brace is when a column is braced at multiple
points along its length, as shown in Figure 8.15.
R
AF
T

Figure 8.15 Single column braced at multiple points along its length.

From first principles it can be shown (Timoshenko and Gere 1961) that the required brace stiffness for
a single column with multiple braces is given by the expression

mN omb
k (8.5)
l

8.16 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

in which m is the number of spans and is a numerical factor which depends on the number of spans,
see Table 8.1. For more spans, the magnitude of the factor asymptotes to 0.250.

Table 8.1 Values for the factor in Equation 8.8

m 2 3 4 5 6 7 9 11

0.500 0.333 0.293 0.276 0.268 0.263 0.258 0.255

For practical purposes, the brace stiffness required to reduce the effective length of a column increases
with the number of spans along the column length. For example, the brace stiffness required to ensure
a 6 metre high column has an effective length of 2 metres is 1.5 times the brace stiffness required if
the same column is 4 metres high with effective length of 2 metres.
Again concerns arise as to whether the current criteria, for a single column with a central brace, to
design braces to resist a design force of 2.5% of the axial force in the column will result in braces
D
whose axial stiffness is sufficient to provided required axial stiffness to adequate brace columns at
more than one point along their length.

8.3.3.4 Brace connection behaviour

The capacity and behaviour of connections of horizontal and diagonal braces vary from one falsework
R
system to the next.
Depending on the falsework, the connection of horizontal and diagonal braces to vertical shores is
considered to behave as either: a pin, which is free to rotate; or a semi-rigid joint. Rarely are
falsework brace connections rigid.
It would be conservative to neglect the rotational stiffness of brace connections and model them as
AF
pin-ended:
(a) If the falsework was intended to behave as a fully braced frame; or
(b) In the absence of any published technical data on the characteristic stiffness and capacity
of a falsework connection,
Researchers who have investigated the semi-rigid properties of horizontal brace connections have
found the joints have an initial looseness, the bending stiffness about the horizontal axis is
significantly greater than the bending stiffness about the vertical axis, and the rotational stiffness about
T
the horizontal axis is tri-linear (Tayakorn and Rasmussen 2008). The rotational joint stiffness varies
significantly for different falsework systems.
Researchers who have investigated the sway stiffness of proprietary scaffold structures needed to take
into account the semi-rigid joint properties of the horizontal brace to column connection (Godley and
Beale 1997; Godley and Beale 2001).
Figure 8.16 depicts a semi-rigid connection that is typical of many modular scaffold systems
commonly used for falsework. Commonly, these systems rely on a wedge detail for positive fixing,
but do not incorporate a locking device to guarantee the connection remains fixed.
Taking account of the semi-rigid behaviour of brace connections complicates falsework design. Semi-
rigid behaviour will introduce bending moments in the horizontal and vertical members, and change
the effective length of the vertical compression members.
Comparatively, falsework with diagonal bracing is more stable than falsework that relies solely on
semi-rigid connections for stability. The behaviour and stability of falsework with both diagonal
bracing and semi rigid connections will depend on the comparative influence each; however, the effect
of semi-rigid connections will be to increase the capacity of the falsework.

2013 Stephen Ferguson all rights reserved 8.17


Formwork and Falsework

D
Figure 8.16 Typical horizontal brace (ledger) to column (standard or shore) connection.
R
During installation, there is anecdotal evidence that hammering in the second end may cause the first
end to come loose. Without any locking device, this type of connection might be susceptible to
loosening during stress reversals. Some authorities require that joints undergo cyclic testing with
acceptance criteria limiting increases in rotation. The cyclic testing will also establish any difference
between positive and negative rotation performance. Figure 8.17 presents a plot of hysteresis loops
AF
typical of the results of cyclic tests.
T

Figure 8.17 Hysteresis loops for horizontal brace (ledger/transom) to column (standard) connection
(ECS 1997)

To establish design values for rotational stiffness it is necessary to take account of the variability of
the connections and the quality of workmanship. AS 4084 Supplement 1 1993, which provides
commentary to the Australian Steel Storage Racking Standard, is relevant because, similar to
falsework, steel storage racks are tall slender heavily loaded structures. AS 4084 Supp. 1 recommends
a capacity reduction factor of 0.67 be applied to the joint spring constant established by testing.
Usually, steel storage rack connections are fitted with locking devices to prevent disengagement. For

8.18 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

falsework connections without any locking device; a capacity reduction factor of less than 0.67 might
be justified.

8.3.3.5 Brace axial stiffness

If horizontal and diagonal braces connect to a vertical member in the same plane and their centre lines
intersect at the same point (without any eccentricity), their axial stiffness can be expressed as:

EA
S= (8.6)
l

where
S = member axial stiffness;
E = modulus of elasticity;
A = member cross-sectional area; and
D
l = member length.

Braces connected with scaffold tube and couplers

Figures 8.18 and 8.19 show scaffold couplers and how they are used to connect scaffold tube lacing
R
and bracing to vertical scaffold tube compression members.
AF
T

(a) Double coupler (forged). (b) Scaffold tube lacing connection using a
double coupler (pressed steel).
(Sourced from http://www.doughty-
engineering.co.uk/cgi- (Sourced from http://www.doughty-
bin/trolleyed_public.cgi?action=showprod_T24901) engineering.co.uk/shop/22/index.htm)

Figure 8.18 Scaffold tube double coupler

2013 Stephen Ferguson all rights reserved 8.19


Formwork and Falsework

D
(a) Swivel coupler (forged) (b) Scaffold tube diagonal brace connection
using a swivel coupler (pressed steel)
Sourced from http://www.doughty-
engineering.co.uk/cgi- (Sourced from http://www.doughty-
R
bin/trolleyed_public.cgi?action=showprod_T24801) engineering.co.uk/shop/22/index.htm)

Figure 8.19 Scaffold tube swivel coupler


AF
When scaffold tube connected by scaffold couplers is used as horizontal lacing or diagonal bracing, its
effective axial stiffness is reduced. This can be dealt with in the following way (ECS 2004):

EA
S= (8.7)
l

Where

= reduction factor introduced to take account of the joint behaviour, eccentric


T
connections, and out-of-plane bracing (caused by the offset inherent in the
coupler connection).

For horizontal bracing connected using double couplers, = 20 ; and

For diagonal bracing connected using swivel tube couplers, = 35


(providing the distance between the horizontal lacing connection (node
point) and the diagonal bracing connection is a maximum of 150 mm).

Proprietary diagonal bracing

In addition to the semi-rigid properties of proprietary brace connections, Godley and Beale determined
the effective axial stiffness of proprietary diagonal bracing. This was necessary because the
eccentricity and construction of the end connections significantly reduced the axial stiffness of the
bracing member, see Figure 8.20.

8.20 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

Figure 8.20 An example of eccentric end connection of a diagonal brace

For example, the effective axial stiffness of a 48.3 mm diameter tube proprietary diagonal brace 3.2 m
long with an cross sectional area of approximately 460 mm2 was equivalent to a member with a cross
D
sectional area of 10.4 mm2. In terms of reduction factors, this would be equivalent to = 44 .

Where it may be conservative to neglect the rotational stiffness of member connections and model
them as pin-ended, it is not conservative to neglect the reduction in axial stiffness of lacing and
bracing members. The stability and capacity of falsework may be over-estimated if the design of
proprietary brace connections reduces the axial stiffness of bracing members. It is especially
R
important to take account of the effect of reduced brace axial stiffness given the increased requirement
for brace stiffness that could be expected in falsework structures that typically consist of a series of
parallel columns braced at multiple points along their length.

8.3.3.6 Plan bracing (diagonal bracing in the horizontal plane)


AF
In addition to horizontal and diagonal bracing in vertical planes, for tall slender falsework structures, it
may be necessary to provide plan bracing to maintain the orthogonal arrangement of the falsework and
prevent buckling about a non-orthogonal axis.
Plan bracing may also be required to provide a load path for horizontal loads to points of restraint.

8.3.4 Falsework base plates and screw jacks

8.3.4.1 Eccentricity
T
Due to the nature and conditions of working on construction sites, accidental or unintentional end
eccentricities occur at the base and head of falsework shores. For example:

Despite the best intentions, bearers may be placed eccentric to the shore centreline, see
Figure 8.21(a);

Eccentricities may be implicit in the arrangement or design of the formwork, see Figure
8.21(b);

Irregular or stiffness variations in the bearing surfaces under formwork shores might
cause eccentricities, see Figure 8.22; and

Eccentricities arise when formwork shores are erected out-of-plumb, see Figure 8.23.

2013 Stephen Ferguson all rights reserved 8.21


Formwork and Falsework

(a) Continuous bearer positioned eccentric to shore centreline (McAdam 1993)


D
R
AF

(b) Discontinuous bearers lapping on shore (SA 1996)


T
Figure 8.21 Examples of eccentric loading

8.22 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

D
Figure 8.22 Examples of irregular or variable stiffness bearing surfaces (SA 1996)
R
AF

Figure 8.23 Example of eccentric reaction (McAdam 1993)


T
8.3.4.2 Detrimental effect of eccentric loads or reactions

The effect of eccentric loads or reactions is to introduce bending moments in screw jacks and
falsework shores that significantly degrade their axial capacity. In addition, unless braced, eccentric
loads will cause the falsework members to sway and, in some situations, cause instability.

2013 Stephen Ferguson all rights reserved 8.23


Formwork and Falsework

D
R
Figure 8.24: Column strength curves for eccentrically loaded shore in new condition.

Figure 8.24 compares the design axial capacity versus member effective length for a 48.3CHS4.0
G250 steel formwork shore out-of-straight l/1000 and loaded eccentrically: 5 mm, 15 mm and 25 mm.
AF
Figure 8.24 demonstrates that the axial capacity of the shore is markedly reduced by small
eccentricities.

8.3.4.3 Minimum eccentricity

Figure 8.25 provides some guidance on choosing an appropriate eccentricity that should be taken into
account for members typically encountered in formwork construction. Where appropriate, larger
eccentricities should be considered.
It is necessary to determine the situation(s) where eccentric actions/reactions have the most adverse
effect. In some cases:
T
(a) The maximum eccentricity may not occur under maximum load (see Figure 8.25 (c), (e)
and (j)). In these situations, the combined effect of partial load and maximum
eccentricity, may be more adverse then full load and less eccentricity; and
(b) The effect of eccentricity is more detrimental in one direction than another (see Figure
8.26).
Unless the applied forces on the members are at a fixed eccentricity, the design eccentricity should
include provision for an unintentional eccentricity not less than 5 mm.

8.24 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

e" e">5

N* N*

(a) (b)

e" for N* will vary e" for N* will vary


between e"1 & e"2 between e"1 & e"2
e"1 e"2 e"1 e"2

N1* N2* N*
D
(c) (d)

e" for N* will vary e" for N * will vary


between e"1 & e"2 between e"1 & e"2
e"1 e"2 e"1 e"2

N1* N2* N1* N2*


R
(e) (f)
AF
stiff portion of stiff portion of
base plate of base plate of
member member

N* N*
e" e"

(g) (h)

e' for N*will vary


T
between e'1 & e'2

e" e'1 e'2


N* N1* N2*

(i) (j)

Figure 8.25 Illustrations of eccentricities of actions and reactions

2013 Stephen Ferguson all rights reserved 8.25


Formwork and Falsework

D
R
(a) The effect of eccentric actions is more severe when bearers are parallel to frames
AF
T

(b) The effect of eccentric actions is less severe when bearers run perpendicular to frames

Figure 8.26 Effects of eccentric actions can be more severe in one direction

8.26 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

8.3.4.4 Rotational stiffness

The capacity of falsework may be overestimated where a falsework base jack or base plate that bears
on a surface is considered to be a fixed (i.e. full moment) connection.
It is not clear how to take account of the fixity of connections at the base of falsework. Research on
different falsework and scaffold systems, from different sources, all achieve good correlation between
the results of tests and analysis of structural models. However, in each case the assumption about base
fixity differs; e.g. some researchers model the base as a pin connection, while others model a
rotational spring or fixed connection.
In the absence of data on rotational stiffness or fixity, for falsework intended to behave as a fully
braced frame, it would be conservative to model the falsework base connection as pinned and
eccentrically loaded, see Figure 8.41.

8.3.5 Spigot connections


The connection of falsework standards using spigots may be a cause of failure, see Figure 8.27.
D
R
AF
T
Figure 8.27: Falsework failure at a spigot joint during testing (Tayakorn and Rasmussen 2009)

Spigot joints are a possible source of failure due to eccentricity, angular change, and bending
weakness in vertical falsework members. The possible detrimental effects of spigot joints can be
reduced by careful consideration of their location.
Connections in compression members should be located:
(a) At points of minimum bending, see Figures 8.28(a) and (b); and
(b) In a manner and position where the detrimental effects of angular change are prevented or
minimised, see Figure 8.28.
Figure 8.28(a) depicts the preferred option for locating connections in compression members, with the
height staggered and two horizontal braced points above and below. Where this is not possible,

2013 Stephen Ferguson all rights reserved 8.27


Formwork and Falsework

equalise, as much as is practical, the length of the compression member above and below the
connection, as shown in Figure 8.28(b).
Figure 8.28(c) is an example of bad practice. The connections are located in a position and in a manner
that permits and accentuates the detrimental effects of angular change, as well as where bending
moments may arise.
D
R
AF
(a) Good practice (b) Good practice (c) Bad practice

Figure 8.28: Examples of good and bad practice in locating connections in compression members

8.3.5.1 Eccentricity

Tolerances at joints in shores are also a potential source of eccentricities, see Figure 8.29.
T

Figure 8.29: Eccentricities arise at spigot joints

8.28 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

8.3.5.2 Angular change at joints

When tubular members such as the vertical falsework members are joined using telescoping
components (e.g. spigots, base plates, screw jacks and prop inners) that have inbuilt clearances for
assembly, angular imperfections may occur, see Figure 8.30. The effects of this imperfection can be
minimised by increasing the lap or detailing close fitting collars and spacers to reduce the diametral
clearance.
D
R
Figure 8.30 Angular imperfections at joints
AF
The angular imperfection can be calculated using Equation 8.8 (ECS 2004)

d d 2 od
tan o = 1.25 1id (8.8)
l
2 lap

where,
d1id = internal diameter of outer member, mm;
T
d2od = external diameter of inner member, mm; and
l2lap = length the inner member laps inside the outer member, mm.

8.3.5.3 Structural model

Structurally, spigot connections in falsework standards resist bending but do not to transfer axial loads.
This can be conceptually modelled as shown in Figure 8.31 with the ends of the falsework standard
free to rotate and the spigot connected to the falsework standards with pin-ended stiff links capable of
transmitting only lateral forces.

2013 Stephen Ferguson all rights reserved 8.29


Formwork and Falsework

D
Figure 8.31 Spigot structural model (Tayakorn and Rasmussen 2009)

8.3.6 Out-of-straight compression members


R
In a compression member, the effect of an initial out-of-straightness is to introduce additional bending
stresses that reduce the axial capacity of the member. This is commonly called the P- effect.
A compression member with an initial out-of-straightness is depicted in Figure 8.32.
AF
T

Figure 8.32 Initially out-of-straight slender compression members.

In Figure 8.32, o is the initial out-of-straightness and is the deflection occurring under load. Where
the initial out-of-straightness (o) is within specified tolerances, design Standards implicitly take
account of these initial imperfections, thereby negating the need for designers to consider their effects.
AS 3610 permits the use of out-of-straight compression members that exceed the specified tolerances
in design Standards (e.g. AS 4100). To avoid over-estimating the capacity of the compression
member, the effects of additional out-straightness must be taken into account explicitly, see Section
8.4.2.

8.30 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

Figure 8.33 demonstrates the detrimental effects of additional out-of-straightness. The capacity of a
steel 48.3CHS4.0 formwork shores (a section commonly used in falsework) calculated in accordance
with AS 4100, which permits an initial out-of-straightness, o = L/1000, is compared to the capacity of
the same section with an initial out-of-straightness, o = L/300, which is permitted in AS 3610 - 1995.
D
R
AF
Figure 8.33: Column strength curves for shores complying with the different out-of-straightness
tolerances permitted in AS 4100 and AS 3610 - 1995.

8.3.7 Differential settlement and axial shortening


Differential settlement of the foundation material (foundation stiffness) or differential axial shortening
(axial stiffness) may cause load redistribution between adjacent falsework shores. The load will be
redistributed from the less stiff to the stiffer shore; i.e. the shore with the least amount of settlement or
T
axial shortening. If adjacent shores are connected by horizontal or diagonal bracing, additional axial
forces and bending moments are likely to arise in the bracing.
The capacity of the shore carrying the greater than expected load or the bracing capacity may govern
design.

8.3.7.1 Differential settlement

Most commonly differential settlement occurs where falsework bears on a foundation with uneven
compaction. Differential settlement will also occur in the situations depicted in Figures 8.34 and 8.35.
In Figure 8.34, neglecting differential settlement and the continuity of the primary and secondary
beams, the shores are equally spaced and could be expected to carry approximately the same load.
However, due to differential settlement, the shore resting on the rigid concrete foundation will carry a
greater share of the load than the adjacent shore bearing on an elastic soil foundation.

2013 Stephen Ferguson all rights reserved 8.31


Formwork and Falsework

Figure 8.34 Differential settlement due to the presence of concrete foundations


D
Similarly, in Figure 8.35, due to differential settlement, load will be redistributed from the shores at
the centre of the beam (where the beam deflection is greatest), to the shores closer to the beams
supports (where the beam deflection is the least).
R
AF
T
Figure 8.35 Load distribution due to beam flexural stiffness

On uniformly compacted soil, it would be reasonable to expect the effects of differential settlement to
be minor and therefore taken into account by the strength load factor for primary members d = 1.25,
see Section 4.3.2.3.
Where uniform settlement does not occur, any expected difference in settlement must be explicitly
taken into account. Where the location of differential settlement is not fixed, it should be considered to
occur where it would have the most detrimental effect.
For the situations depicted in Figures 8.34 and 8.35, differential settlement can be taken into account
by analysing models of the structure with:
(a) For Figure 8.34, vertical spring restraints whose stiffness models the behaviour of the
foundation material, or
(b) For Figure 8.35, the falsework supported by the beam.

8.32 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

8.3.7.2 Differential axial shortening

Where the axial stiffness of adjacent shores differs and the stiffness of the primary beam permits, load
redistribution may occur such that a shore with a greater axial stiffness will carry a greater share of the
load than an adjacent shore that is less stiff and would otherwise shorten.
In the example depicted in Figure 8.36, part of the falsework is supported by a previously poured slab
that acts as a rigid foundation. As a result, the adjacent shores that differ significantly in length carry
differing loads. The shorter shore, which is more heavily loaded, is three to four times as stiff as the
tall lightly loaded shore. A significant share of the load is redistributed from the less stiff tall shore to
stiffer short shore.
D
R
AF
T
Figure 8.36 Load redistribution due to differential axial shortening

8.3.8 Knee Buckling


Research (Ikheimonen 1997) has shown that knee buckling can occur when compressive stresses
between the top of the shore and the underside of the bearer are high and the bearer is not adequately
restrained to prevent overturning, see Figures 8.37 and 8.38.
In Figure 8.37, the buckling load of shore and bearer acting together may be considerably lower than
that of the shore itself.

2013 Stephen Ferguson all rights reserved 8.33


Formwork and Falsework

Figure 8.37 Euler buckling and knee buckling, (Ikheimonen 1997)

In the case shown in Figure 8.38, where hinges develop between the bearer and screw jack, the
D
capacity of the falsework is less than for falsework with top restraint.
R
AF

Figure 8.38 Knee buckling of formwork frames with extended screw jacks (Ikheimonen 1997)
T
Knee buckling arises because the modulus of elasticity of timber varies with angle and direction of the
annular growth ring gradient. Consequently, when a timber is loaded perpendicular to the grain, the
deformation is non-uniform and the surface is distorted, see Figure 8.39.

Figure 8.39 Non-uniform deformation of timber loaded at right angles to the grain (Ikheimonen
1997)

8.34 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

Importantly, tests show that knee buckling can occur at bearing stresses less than permitted in design
standards. Thus, merely ensuring that a limit states bearing violation will not occur is insufficient to
prevent knee buckling. In addition, timber that is out-of-square may cause an initial set or out-of-
straightness, which makes the formwork prone to knee buckling. Furthermore, eccentric loading may
also increase the risk of knee buckling.
Knee buckling may be prevented by restraining the bearer from overturning; e.g. where the bearer is a
close fit inside a "U-head screw jack. Alternatively, knee buckling may be prevented by laterally
bracing the top of the shore.
D
R
Figure 8.40 Models for Euler and knee buckling (Ikheimonen 1997)

In the absence of any preventative measures, account should be taken of the lower buckling load
arising from knee buckling, see Figure 8.37. In Figure 8.40, to take account of knee buckling, an
elastic hinge is introduced to model the non-uniform deformation of the bearer.
AF
8.4 Analysing falsework structures
Falsework is often a heavily loaded structure with slender members. For a slender member subject to
an axial load, deformations increase instability and detrimental second-order effects. Second order
effects arise from loads acting on the falsework and its members in their displaced and deformed
configuration.
A plastic analysis takes into account material non-linearity, which is not appropriate for falsework. It
is more appropriate for sway structures with high bending moments and small axial loads. An elastic
analysis is appropriate for falsework structures that are intended to be braced frames, primarily subject
T
to axial loads, and to behave elastically at all times.
A first-order elastic analysis that ignores second-order effects will underestimate the action effects in
the structure and over-estimate stability. Accordingly, AS 4100 requires that second-order effects be
taken into account.
An approximation of second-order effects is possible by amplifying first-order moments.
Alternatively, a second-order elastic analysis that accounts for geometrical non-linearity provides a
conservative estimate of ultimate load. A more accurate analysis is possible using an advanced
analysis (Clarke, Bridge et al. 1992), but is not yet in general use.

8.4.1 Structural Models


Readily available software packages make the geometrical elastic second-order analysis of falsework
structures feasible and practical. In this way, it is simple to explicitly take account of the effectiveness
of bracing, joint rotational stiffness, axial shortening, differential settlement, the behaviour of spigot
connections, etc.

2013 Stephen Ferguson all rights reserved 8.35


Formwork and Falsework

A more accurate assessment and understanding of the behaviour and action effects will result from 3D
structural models of falsework. Where the falsework arrangement and loading is regular, analysis of a
series of representative 2D frames in both directions is a reasonable and practical approach. In some
situations, uncertainty associated with the accuracy of a 2D analysis and the importance of the
falsework, may warrant more rigour and performing a 3D analysis.
Superposition of amplified first order moments with the moment diagram and action effects
(determined from second-order analysis) is a useful method of taking account of the effects of
imperfections (such as: end eccentricity of reactions at the base and loads at the top of falsework and
member out-of-straightness greater than permitted in design standards). This avoids complicating
frame models with imperfections that may be random in nature and whose effect is difficult to model
accurately.

8.4.2 Calculating falsework member capacity


For practical reasons of robustness and telescoping, compact hollow sections are the most common
column sections used in formwork falsework. Accordingly, the effects of local, distortional and
flexural-torsional buckling can be safely neglected.
D
Falsework compression members are subject to the combined effects of axial compression and
bending; therefore they must satisfy each of the following criteria:

Member axial compression d N d N c (k e l ) (8.9)


R
Nd c M
Member axial compression and bending d + m d 1.0 (8.10)
N c (l ) M s
AF
Nd M
Section axial compression and bending d + d 1.0 (8.11)
N s M s

In Equations 8.9, 8.10 and 8.11,

d = strength load factor for primary members, d = 1.25, see Section 4.3.2.3;
Nd = strength limit states design axial compression in the member, determined after
T
taking account to second-order effects;
= capacity reduction factor for the particular action effect;
Nc(kel) = nominal member capacity as a function of the member effective length;
ke = member effective length factor;
l = member length;
Nc(l) = nominal member capacity as a function of the member length l;
Md = maximum strength limit states design bending moment along the length of a
member, determined after taking account to second-order effects;
Ns = nominal section axial capacity, see Equation 8.14
Ms = nominal section moment capacity;

8.36 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

cm = factor for a braced member subject to end moments only that takes account of
unequal bending moments, given in Equation 8.12;

cm = 0.6 0.4 m 1.0 (8.12)

In Equation 8.12,
m = ratio of the smaller to the larger bending moment at the ends of the member.
If the moments are equal and in opposite directions, m = 1 ;
If the moments are equal but in the same direction, m = 1 ; or
For unequal end moments, 1 < m < 1 .

8.4.2.1 End eccentricity and member out-of-straightness

Equation 8.13 provides a conservative estimate of the nominal member axial compression capacity of
an out-of-straight and eccentrically loaded falsework member (Ferguson 2003). Equation 8.13 takes
D
account of the load-moment interaction and amplifies the first-order effects of member out-of-
straightness and eccentricity to approximate second-order effects (Timoshenko and Gere 1961).
Equation 8.13 is a useful method of taking account of the second-order effects of end eccentricity and
out-of-straightness without the need to include them in 2D or 3D structural models. The value of
Nc(kel) can be substituted into Equation 8.9.
R
Equation 8.13 can also be used to determine the value of the nominal member capacity as a function of
the member length Nc(l), used in Equation 8.9. This is achieved by substituting the elastic buckling
load as a function of member length Nomb(l) for the elastic buckling load as a function of member
effective length Nomb(kel), and recalculating cm based on a bending moment diagram that superimposes
the bending moments due to out-of-straightness and eccentricity with those from the structural
AF
analysis of the 2D or 3D model.

N s + (1 + )N omb (k e l ) N s + (1 + )N omb (k e l )
12
2

N c (k e l ) = N s N omb (k e
l ) (8.13)
2 2

In Equation 8.13,
Ns = nominal section capacity, given in Equation 8.14 for compact sections;
Nomb(kel) = member elastic buckling capacity as a function of member effective length, see
T
Equation 8.1; and

= section parameter given in Equation 8.15.

N s = An f y (8.14)

In Equation 8.14,
An = net area of the cross-section, in mm2; and
fy = yield stress of the member, in MPa.

( o + cm e )An
= (8.15)
Sx

2013 Stephen Ferguson all rights reserved 8.37


Formwork and Falsework

In Equation 8.15,

o = permitted initial out-of-straightness, in mm;


cm = factor for a braced member subject to end moments only that takes account of
unequal bending moments, given in Equation 8.12;
e = largest end eccentricity (e1 or e2), in mm, see Figure 8.41; and
Sx = plastic section modulus, in mm3.
D
R
AF
(a) Eccentricities on the same side.
T

(b) Eccentricities on the opposite side.

Figure 8.41: An eccentrically loaded pin-ended strut.

8.4.2.2 Example

To demonstrate taking account of additional out-of-straightness and end eccentricities, consider the
following example:

8.38 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

Determine the suitability of a 1.5 m long falsework shore whose section is 48.3x 4.0 CHS G250,
given the results of a second-order elastic analysis and rational buckling analysis show a design
axial force Nd = 40 kN, negligible bending moments, and an effective length of 1.5 m.
Assume equal opposite end eccentricities of 5 mm and a permitted out-of-straightness of l/300.

The member must satisfy Equation 8.9, d N d N c (k e l ) .

Substituting into Equation 8.15, o = 1500/300 = 5 mm, e = 5 mm, cm = 0.60 - 0.4(-1.0) = 1.0, An = 557
mm2 and Sx = 7,870 mm3, gives:

=
(5 + 1.0 5) 557 = 0.71
7870

Substituting into Equation 8.13, = 0.83, Ns = 250(557) = 139 kN and Nomb(1500) = 121 kN, gives:

139 + (1 + 0.71) 121 139 + (1 + 0.71) 121


12
2

N c (1500 ) = 139 121 , in kN
D
2 2

therefore
Nc(1500) = 58.2 kN
R
Substituting into Equation 8.9, d = 1.25, Nd = 40 kN, = 0.9, Nc(1500) = 58.2 kN
1.25(40) 0.9(58.2)
50.0 52.7, OK
AF
By way of comparison and to demonstrate the degrading effects of end eccentricity and additional out-
of-straightness, the capacity in axial compression of a concentrically loaded 48.3x 4.0 CHS G250 with
an effective length of 1.5 m and out-of-straight less than l/1000 is 79.5 kN versus 52.7 kN calculated
herein.

8.4.3 Falsework member column effective length


Accurately determining the effective length of falsework compression members is critical to avoid
over-estimating falsework capacity.
In frames that are fully braced, the effective length factor for compression members ke = 1.0, see
T
Figures 8.9 and 8.10. In sway frames, the effective length factor for compression members ke 1.0,
see Figure 8.42 to 8.46.
The previously mentioned software packages that can be used for the geometrical elastic second-order
analysis of falsework structures also provide the ability to carry out a rational elastic buckling analysis
of the falsework that will determine the falsework buckling load factors, buckling mode shapes and
member effective lengths.
Making an accurate initial estimate of the member effective length is useful for preliminary design and
set out of the falsework. To this end, BSI 5975 (BSI 2008) provides the guidance for commonly
encountered situations. Initial estimates should be verified by checking the member effective lengths
determined by a rational buckling analysis.

8.4.3.1 Estimates of column effective length in sway frames

BSI 5975 provides the following guidance for estimating effective length of falsework compression
members.

2013 Stephen Ferguson all rights reserved 8.39


Formwork and Falsework

Top restrained falsework

For a top restrained frame with central pinned bracing, as shown in Figure 8.42, the effective length of
the compression member kel = l.
D
Figure 8.42: Effective length of members in top restrained frames with central pinned bracing

For a top restrained frame with pinned bracing at the base, as shown in Figure 8.43, the effective
R
length of the compression member kel = 0.85[l (x/2)].
AF
T
Figure 8.43: Effective length of members in top restrained frames with pinned bracing at the base

Freestanding falsework

For a freestanding frame with central pinned bracing, as shown in Figure 8.44, the effective length of
the compression member kel = l (x/2).

8.40 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

Figure 8.44: Effective length of members in freestanding frames with central pinned bracing

For a freestanding frame with pinned bracing at the base, as shown in Figure 8.45, the effective length
of the compression member kel = 2l x.
D
R
AF
Figure 8.45: Effective length of members in freestanding frames with pinned bracing at the base

For a freestanding frame with pinned bracing, as shown in Figure 8.46, the effective length of the
compression member is the greater of, kel12 = l2 + 2 l1; kel3 = l3; or kel45 = l4 + 2 l5.
T

Figure 8.46: Effective length of members in freestanding frames with pinned bracing

2013 Stephen Ferguson all rights reserved 8.41


Formwork and Falsework

8.4.4 Falsework frame buckling


For fully braced falsework (e.g. Figures 8.9 and 8.10) the buckling load factor of the whole frame (c)
is the lowest buckling load factor (m) of all compression members, where for each member

N omb
m = (8.16)
Nd

For falsework that can sway, the elastic buckling load factor (c) for the whole frame is the lowest of
all the elastic buckling load factors for each storey (ms), where for each storey

N omb
l

ms = (8.17)
N
l d
D
For example, Equation 8.17 would be used to calculate the elastic buckling load factor of the top and
bottom row of screw jacks shown in Figure 8.46.
Using Equations 8.16 and 8.17, an estimate of the elastic buckling load factor for the whole frame can
R
be determined based on initial estimates of column effective length, using guidance similar to shown
in Section 8.4.3.1.
It is useful to compare the expected elastic buckling load factor for the whole frame (c) (calculated in
accordance with Equations 8.15 and 8.16) with the elastic buckling load factor determined by a
rational buckling analysis. The elastic buckling load factor determined by a rational analysis (c)
AF
should be greater than or equal to the lowest elastic buckling load of all the braced members (m) and
the lowest elastic buckling load factor for each storey (ms), as applicable.
When the elastic buckling load factor determined by a rational buckling analysis is less than the
expected elastic buckling load factor, it is indicative of one of the following:
(a) Underestimating member effective lengths;
(b) A previously unidentified member buckling first;
(c) The whole frame buckling in an unexpected manner; or
T
(d) Errors in the structural model.

References
Bragg, S. L. (1975). Final report of the Advisory Committee on Falsework. London, Her Majesty's
Stationery Office: 151.
BSI (1995). BS 5975:1996 Code of practice for Falsework. London, British Standards Institution.
BSI (2008). BS 5975:2008 Code of practice for temporary works procedures and ther permissible
stress design of falsework. London, British Standards Institution.
Clarke, M. J. and R. Q. Bridge (1994). Lateral bracing force and stiffness requirements for axially
loaded columns. Australasian Structural Engineering Conference, Sydney, Australia, The Institution of
Engineers Australia.
Clarke, M. J., R. Q. Bridge, et al. (1992). "Advanced Analysis of Steel Building Frames." Journal of
Constructional Steel Research 23: 1-29.

8.42 2013 Stephen Ferguson all rights reserved


Chapter 8 - Falsework

ECS (1997). prEN 12810-2 Facade scaffolds made of prefabricated elements - Part 2: Methods of
particular design. Brussels, European Committee for Standardization.
ECS (2004). BS EN 12812 Falsework - Performance requirements and general design. Brussells,
European Committee for Standardization.
Ferguson, S. A. (2003). Limit states design of steel formwork shores, University of Western Sydney.
Godley, M. H. R. and R. G. Beale (1997). "Sway stiffness of scaffold structures." The Structural
Engineer 75(No. 1): 4-12.
Godley, M. H. R. and R. G. Beale (2001). "Analysis of large proprietary access scaffold structures."
Proceedings of the Institution of Civil Engineers, UK 146(No. 1): pp 31-39.
Hadipriono, F. C. and H.-K. Wang (1986). "Analysis of causes of formwork failures in concrete
structures." Journal of Construction Engineering and Management 112: 112 - 121.
Ikheimonen, J. (1997). Construction Loads on Shores and Stability of Horizontal Formworks.
Department of Structural Engineering. Stockholm, Royal Institute of Technology: 161, 176.
D
McAdam, P. S. (1993). Formwork: A practical approach. Brisbane, Stuart Publications.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney, Standards
Australia.
R
SA (1996). AS/NZS 4600 - 1996 Cold-formed Steel Structures. Sydney, Standards Australia.
SA (1998). AS 4100 - 1998 Steel Structures. Sydney, Standards Australia.
Tayakorn, C. and K. J. R. Rasmussen (2008). Research Report No R893 Scaffold Cuplok Joint Tests,
AF
University of Sydney
Tayakorn, C. and K. J. R. Rasmussen (2009). Research Report No R896 Structural Modelling of
Support Scaffold Systems, University of Sydney
Timoshenko, S. P. and J. M. Gere (1961). Theory of Elastic Stability. New York, McGraw-Hill.
Trahair, N. S. (1999). "Column Bracing Forces." Australian Journal of Structural Engineering
Transactions vol SE2(no 2&3): pp 163-168.
T

2013 Stephen Ferguson all rights reserved 8.43


Chapter 9 Stripping Criteria

9
Stripping Criteria
9.1 Introduction
The term stripping refers to the removal of formwork after the concrete has set.
Stripping affects the surface finish quality, durability and structural reliability of the concrete work.
Premature stripping may cause physical damage to the surface finish, contribute to non uniformity of
colour and impede hydration. Structurally, early stripping may lead to cracking, increased long term
deformations, overload and possibly collapse.
D
For practical and economic reasons, stripping times need to be as short as possible Therefore, this
Chapter focuses on minimum stripping times that satisfy structural requirements. The guidance in this
Chapter may be inappropriate for architectural concrete and does not fully address or take account of
all matters relating to curing or protection of the exposed concrete surface.
The Chapter starts by explaining different methods for stripping horizontal forms. This is followed
R
with guidance based on the requirements set out in AS 3610 1995 Formwork for concrete (SA
1995) and AS 3600 2009 Concrete structures (SA 2009), as well as guidance on calculating the
minimum early-age concrete strength required for stripping that conforms to the requirements in AS
3600.
9.2 Methods for stripping horizontal forms
AF
9.2.1 Single-stage stripping
In single-stage stripping the forms and shores are removed over large areas, allowing the concrete to
span between the permanent supports in the manner intended in the project design, see Figure 9.1.
Once the forms and shores are removed, the concrete will carry its own weight and any superimposed
construction loads; e.g. workman and equipment, stacked materials, etc.
T

Figure 9.1 Single-stage stripping (SA 1995)

2013 Stephen Ferguson all rights reserved 9.1


Formwork Handbook

9.2.2 Two-stage stripping


For economic reasons, some formwork systems allow the forms to be removed before the shores (two-
stage stripping), see Figure 9.2.

Figure 9.2 Two-stage stripping leaving undisturbed shores (SA 1995)


D
Alternatively, a similar result can be achieved by backpropping. The term back-propping refers to
the procedure of installing additional shores (backprops) prior to removing small areas of the
formwork, thereby preventing the concrete from carrying load, as shown in Figure 9.3.
R
AF
T
Figure 9.3 Two-stage stripping by back-propping (SA 1995)
Either way, the benefit of two-stage stripping is the early recovery of formwork without letting the
concrete carry its own weight or any construction load, because until the formwork shores or
backprops are eventually removed the weight of the concrete and any superimposed construction load
is carried by the shores.
9.3 Minimum stripping times
For economic and practical purposes, stripping times need to be as short as possible.
The project designer should specify the minimum stripping times
In the absence of any specified stripping times, AS 3600 and AS 3610 provide guidance on minimum
stripping times. However, much of the guidance in AS 3610 is impractical and out-of-date.
Structurally, formwork should not be stripped until the concrete has attained sufficient strength and
stiffness to support its own weight and any superimposed loads, safely and without damage or
detriment to its intended use (SA 2009).

9.2 2013 Stephen Ferguson all rights reserved


Chapter 9 Stripping Criteria

The minimum stripping time is usually determined by calculating the minimum early-age
characteristic concrete strength the in-situ concrete is required to attain taking into account the pace of
construction, imposed loads, ambient temperatures and early-age strength gain characteristics of
concrete.
Post-tensioned concrete is often initially partially stressed after 2 or 3 days; however it should not be
stripped until fully stressed.
Where colour control is specified, it is advisable to strip forms on different elements at the same age
and as early as is permissible.
9.3.1 Development of concrete strength with age
The development of concrete strength with age varies with the type of cement, grade specified and
ambient temperature, see Figures 9.4 and 9.5, as well as Table 9.1.
D
R
AF
Figure 9.4 Typical development of concrete strength with age (Guirguis 1998)
T

Figure 9.5 Typical compressive strength development of Portland cement (SA 1996)

2013 Stephen Ferguson all rights reserved 9.3


Formwork Handbook

Table 9.1 Early-age mean strengths for normal-class concrete (SA 1997)

Average 7-day compressive


Grade designation
strength
(MPa)

N25 12
N32 16
N40 20
N50 25

Compared to normal-class concrete, the rate of strength gain is reduced in concrete that incorporate
higher contents of supplementary cementitious material such as fly ash; e.g. Green Star products.
9.3.2 Minimum stripping times for vertical forms
D
The minimum stripping times for vertical forms (side forms for footings, walls, columns slabs and
beams) may be based on achieving the minimum average concrete compressive strength specified in
Table 9.2. Extra care is needed if vertical formwork is stripped within 18 hours after casting.
R
Table 9.2 Minimum compressive strength of concrete
for stripping vertical forms (SA 1996)

Surface finishes Average compressive strength


MPa
AF
Classes 1, 2 and 3 5
Classes 4 and 5 2

9.3.3 Minimum stripping times for horizontal forms and removal of shores
AS3600 permits formwork to be removed when the concrete member has attained sufficient strength
to safely support, without detriment to its intended use, its own weight and any currently or
subsequently imposed actions, providing:
T
a) Formwork supports are removed to a planned sequence that will not subject the concrete
structure to any unnecessary deformation, impact, or eccentric loading; and
(b) Concrete is cured for the specified period or attains the minimum strength specified in
Table 9.3, for the applicable exposure classification.
Curing is achieved by the application of water to, accelerated curing of, or the retention of
water in, the freshly cast concrete. It should commence as soon as practicable after the
finishing of any unformed surfaces has been completed.
Where formwork is stripped before the end of the specified curing period, exposed
surfaces shall be cured until at least the end of the specified curing period.

9.4 2013 Stephen Ferguson all rights reserved


Chapter 9 Stripping Criteria

Table 9.3 Minimum strength and curing requirements for concrete (SA 2009)

Minimum Minimum average compressive


Exposure Minimum initial curing strength at time of stripping
forms or removal from moulds
classification requirement
(MPa) (MPa)

A1 20 Cure continuously for at


15
A2 25 least 3 days

B1 32 20

B2 40 Cure continuously for at 25

C1 50 least 7 days
32
C2 50
D
In addition, AS3600 requires, where applicable:
a) For reinforced beam and slab soffit formwork where control samples are available:
(i) A minimum of 3 days between casting and before commencement of stripping; and
R
(ii) The member to remain uncracked taking account of the appropriate characteristic
strength of the concrete determined from the average strength of control samples
taken, cured, see Sections 9.4 and 9.5.
b) For stripping soffit forms between undisturbed shores under normal class reinforced
concrete slabs:
AF
The minimum period of time after casting and before commencement of stripping forms
only (i.e. leaving the shores undisturbed) should not be less than specified in Table 9.4.

Table 9.4 Minimum times for stripping of forms


between undisturbed shores (SA 2009)

Period of time before


Average ambient
stripping normal-class
temperature over the
concrete with specified early-
T
period (T)
age strength
C
Days
T > 20 4
20 T > 12 6
12 T > 5 8

The values in Table 9.4 only apply providing both the following criteria are satisfied:
l 280

A
dc (d c + 100)
where
l = span between formwork shores, in mm; and

2013 Stephen Ferguson all rights reserved 9.5


Formwork Handbook

dc = overall depth of the concrete section, in mm.


B the imposed construction load is not greater than 2.0 kPa.
The periods in Table 9.4 shall be increased if the average temperature over the period is
less than 5C. Increase the periods by half a day for each day the daily temperature was
between 2C and 5C; or by a whole day for each day the daily average temperature was
below 2C.
c) For removing formwork supports under slabs or beam not supporting structures above:
(i) Calculations based on known or specified early-age strengths that demonstrate the
concrete has gained sufficient strength so that the degree of cracking or
deformation that will occur, then or subsequently, is not greater than that which
would occur if the design serviceability load were applied to the member when the
concrete has attained its required design strength, see Section 9.4; or
(ii) In the absence of early-age strength data, the period of time after casting the
concrete is not less than that given in Table 9.5.
D
Table 9.5 Minimum times before removal of supports from slabs
and beams not supporting structures above (SA 2009)

Average ambient Period of time before removal


temperature over the of formwork supports from
R
period (T) reinforced members
C Days

T > 20 12
20 T > 12 18
AF
12 T > 5 24

The values in Table 9.5 only apply providing the imposed construction load is not
greater than 2.0 kPa.
The periods in Table 9.5 shall be increased if the average temperature over the
period is less than 5C. Increase the periods by half a day for each day the daily
temperature was between 2C and 5C; or by a whole day for each day the daily
average temperature was below 2C.
T
d) For removal of multistorey shoring under slabs and beams:
(i) Calculations that demonstrate that the magnitude of cracks and deflections in all
supported and supporting floors and beams, under the current and subsequent
imposed loads, will not impair the strength or serviceability of the completed
structure; and
(ii) A minimum elapsed time of 2 days after placing of concrete before removal of any
shores directly or indirectly supporting the concrete.
(e) For removal of forms and formwork shores under prestressed concrete slabs and beams:
The strength of the concrete in the member and the number of tendons stressed are such
as to:
(i) Provide the necessary strength to carry its own weight and any currently or
subsequently imposed actions; and
(ii) Meet the associated serviceability and other limit state requirements.

9.6 2013 Stephen Ferguson all rights reserved


Chapter 9 Stripping Criteria

9.4 Calculating the minimum early-age strength for stripping


9.4.1 Stripping the forms only
For reinforced concrete, soffit forms may be stripped from between undisturbed supports (shores,
columns, walls, etc) providing the concrete remains uncracked. The minimum concrete strength
required before stripping can be determined conservatively by assuming:
(a) The concrete is un-reinforced;
(b) After the forms are removed, the concrete spans one-way continuously between
undisturbed rigid formwork supports; and
(c) The concrete is subject to a uniformly distributed load.
In practice, formwork supports may only be considered rigid if supported on an extremely stiff
foundation; e.g. ground slab or raft. In multistorey construction, when the formwork supports do not
extend to the ground, this assumption may not be valid unless the props are supported on an extremely
stiff floor; e.g. plant room slab. Slabs and beams connected by multistorey shoring share load and,
therefore, may be already under stress prior to stripping, which is not taken into account in the
following method.
D
The analysis assumes the concrete member, of depth dc metres, is un-reinforced. Furthermore, after
the forms are removed the concrete spans one-way continuously between undisturbed rigid formwork
supports and is subject to an imposed vertical action from construction activity, Qv, and the self-weight
of the concrete, Gc.
R
The imposed action from construction activity (Qv) is that determined from the most adverse
combination of concurrent imposed actions acting directly on the concrete surface, including where
applicable:
(a) Workers and equipment, Qw;
(b) Stacked materials applied directly to the concrete surface, Qm; and
AF
(c) Other vertical loads, Qx.
If construction for a subsequent level has commenced prior to formwork removal, an allowance for the
weight of formwork erected on top of the concrete (Gf) and an allowance for imposed actions from
workers and equipment on top of the formwork (Qw) would need to be taken into account in
determining the most adverse combination of concurrent imposed actions (Qv).
During construction, the appropriate ultimate design action combination is the most adverse of:

Wd = 1.2Gc + 1.5Qv , kPa (9.1)


T
Wd = 1.35Gc , kPa (9.2)

Where Qv < 0.1Gc ; Equation 9.2 governs.


Consider a slab with a depth dc metres and width b metres, that spans continuously one-way over rows
of undisturbed shores spaced equally l metres apart. The maximum design bending moment (above
the shores) is Md, and maybe determined using Equation 9.3:

Wd bl 2
Md = , kNm (9.3)
10

To satisfy strength limit states

M d M uo (9.4)

2013 Stephen Ferguson all rights reserved 9.7


Formwork Handbook

where, Muo is the ultimate strength in bending. To prevent a slab from cracking, that is un-reinforced
for negative bending above the shores

M d M uo. min = f ' cf Z 10 9 (9.5)

In Equation 9.5,
= capacity factor for bending (0.6);
fcf = characteristic flexural tensile strength of the concrete, in MPa; and
Z = section modulus of the uncracked section (Z = bdc2/6), in m3
dc = depth of the slab, in metres
Substituting for Md into Equation 9.5 yields

Wd bl 2 10 6 0.6 f cf bd c 10
' 9 2

(9.6)
10 6
D
or

2
W l
f '
cf d , MPa (9.7)
1000 d c
R
At 28 days, assuming standard curing

f cf' = 0.6 f c' , MPa (9.8)


AF
Therefore, Equation 9.7 can be written in terms the characteristic compressive strength fc as

4
l
2
W
f d
c
'
, MPa (9.9)
600 dc

Assuming the relationship between tensile and compressive strength holds true for concrete less than
28 days old, Equation 9.9 can be applied to determine the required early-age characteristic
compressive strength at the time of stripping. Importantly, the early-age characteristic compressive
T
strength should not be confused with the early-age mean grade strength of all the results, f cm .
To avoid any confusion, the notation fce will be adopted for the early-age characteristic compressive
strength in Equation 9.10 and rewritten as:

4
l
2
W
f d
'
ce
, MPa (9.10)
600 dc

Issues relating to determining fce are discussed later.


Design Example
What is the minimum early age characteristic compressive strength that concrete in a 150 mm thick
flat suspended slab must reach before the forms could be removed between undisturbed shores spaced
continuously at 2.1 m centres?
If Qv = Qm = 5.0 kPa, the most adverse design action would be:

9.8 2013 Stephen Ferguson all rights reserved


Chapter 9 Stripping Criteria

Wd = 1.2Gc + 1.5Qv

Wd = 1.2 0.15 25 + 1.5 5.0 = 12.0 kPa

The early age characteristic compressive strength must reach

4
l
2
W
f d
'
ce

600 dc
2 4
12 2.1
f ce' 15.4 MPa
600 0.15

9.4.2 Stripping formwork supports under reinforced concrete


To determine the most appropriate criteria for stripping formwork supports under flat reinforced
D
concrete slabs up to 300 mm thick, Beeby 2000 investigated the most severe effects of bending, shear,
deflection and cracking. The results of the research demonstrate that cracking governs, which is
consistent with the requirements of AS 3600.
For flat reinforced concrete slabs, formwork may be stripped providing the following inequalities are
satisfied.
R
W
1.0 (9.11)
Wser

and
AF
0.6
W f'
ce (9.12)
Wser f ' c

In Equations 9.11 and 9.12, W is the unfactored construction load on the slab, given by

W = Gc + Qv , kPa (9.13)

where
T
Gc = weight of the concrete slab, in kPa; and
Qv = sum of the most adverse concurrent construction loads on the slab, in kPa.

In Equations 9.11 and 9.12, Wser is the unfactored design service load on the slab, given by

Wser = Gc + G ser + Qser , kPa (9.14)

where
Gc = weight of the concrete slab, in kPa;
Gser = permanent actions for services, partitions, ceilings, floor treatments, etc; and
Qser = occupancy live load.
Nevertheless, it is prudent to also ensure that strength limit states are not violated and Equations 9.15
and 9.16 should also be satisfied.

2013 Stephen Ferguson all rights reserved 9.9


Formwork Handbook

Wd
1.0 (9.15)
Wult

Wd Rue
(9.16)
Wult Ru

In Equations 9.15 and 9.16,


Wd = strength limit states construction load on the slab, given by the most adverse of
combination from either Equation 9.1 or 9.2; and
Wult = strength limit states service design load, which is usually given by the most adverse
of:

Wult = 1.2(Gc + G ser ) + 1.5Qser (9.17)

Wult = 1.35(Gc + G ser ) (9.18)


D
Rue = early-age design capacity; and
Ru = design capacity
Essentially, Equation 9.11 prevents the load on the slab exceeding the unfactored design service load.
R
Equation 9.12 limits the extent of cracking to that implicit in the concrete design. In doing so, they
satisfy the requirements specified in AS 3600. Equation 9.15 prevents the load on the slab exceeding
the service strength limit states design load and Equation 9.16 prevents the early-age loading
exceeding the early-age design capacity.
Importantly, Equations 9.11, 9.12, 9.15 and 9.16 are not applicable to the first stage of two stage
stripping, where only the forms are removed leaving all the formwork supports undisturbed and
AF
supported on a rigid foundation.
Design Example
What is the maximum construction load a N32 flat reinforced concrete suspended slab could support if
stripped when the characteristic early age strength reaches 20 MPa. The slab is 250 mm thick and
designed for an imposed floor live load of 3.0 kPa and dead load of 1.5 kPa for services, partitions and
ceilings?
Assume the density of reinforced concrete is 25 kN/m3.
T
0.6
W f'
ce
Wser f ' c

(25 .25) + Qv
0.6
20

(25 0.25) + 1.5 + 3.0 32
6.25 + Qv
0.75
10.75

Therefore,

Qv (0.75 10.75) 6.25 .

Qv 8.06 6.25 1.81 , kPa

9.10 2013 Stephen Ferguson all rights reserved


Chapter 9 Stripping Criteria

W
By observation the unfactored load inequality 1.0 is also satisfied.
Wser
Wd
Check the strength limit states inequality 1.0 is satisfied
Wult

Wd 1.2 6.25 + 1.5 1.81


=
Wult 1.2 (6.25 + 1.5) + 1.5 3.0

Wd 10.21
= = 0.74 1.0 , OK.
Wult 13.80

Wd Rue
Check the strength limit states inequality is satisfied, by assuming strength governed by
Wult Ru
0.6
f ' ce
D

c
f '

Wd 10.21
= = 0.74 0.75 , OK.
Wult 13.80
R
9.4.3 Stripping formwork supports under prestressed concrete
Strength rather than cracking may govern the minimum early-age concrete strength required before
stripping formwork supports under prestressed concrete slabs. In this case, only Equations 9.11, 9.15
and 9.16 need be satisfied.
AF
9.5 Assessment of concrete strength at early age
The early-age characteristic compressive strength fce is determined by sampling, curing and testing
concrete from which a lower bound 95% confidence limit is chosen as the nominal characteristic
strength.
9.5.1 AS 3600
Clause 17.6.2.8 of AS 3600 requires:
a) Taking control test-samples of each concrete grade placed on any one day at a minimum
T
frequency of one sample for each 50 m3, or part thereof;
b) Storing and curing the samples under conditions similar to those of the concrete in the
work; and
c) Testing at least two samples from each grade for strength at the desired time of stripping;
d) Assessing the early-age strength of the concrete on the basis of the average strength of the
samples tested at that age.
Where control samples have been taken, cured and tested in accordance with AS 3600, the mean
compressive strength of the specimens at that age can be determined. In order to obtain an estimate of
the characteristic compressive strength at that age, the mean strength must be reduced to take account
of the variability and uncertainty of the test methods.
For example, AS 1379 (SA 1997) requires when assessing production control that:

f cm f ' c + k c s (9.18)

where,

2013 Stephen Ferguson all rights reserved 9.11


Formwork Handbook

f cm = mean grade strength of all results for the grade

kc = assessment factor determined from the number of controlled grade samples

s = standard deviation for the grade being assessed.


As a guide, for controlled grades, typical values for kc vary from 3.2 for 4 or less samples to a
minimum of 1.25 for 15 or more samples, see Table 7 in AS 1379. The value for the standard
deviation (S) is calculated, except when the number of sample test strengths is less than 5,
where the value should not be less than 3 MPa.
9.5.2 Other methods of assessing early-age compressive strength
Economies of early stripping may warrant adopting more accurate methods of assessing concrete
strength. In this regard, guidance can be found in the Concrete Institute of Australias Current Practice
Note 22 Non-destructive Testing of Concrete (CIA 2008).
References
Beeby, A. W. (2000). ECBP Task 4 Report - Early Striking and Backpropping (Report BR 394).
D
London, BRE.
CIA (2008). Current Practice Note 22 - Non-destructive Testing of Concrete.
Guirguis, S. (1998). Cements - Properties and Characteristics. Sydney, Cement & Concrete
Association of Australia.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
R
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney, Standards
Australia.
SA (1997). AS 1379 - 1997 Specification and supply of concrete. Sydney, Standards Australia.
SA (2002). AS/NZS 1170.0 - 2002 Structural design actions Part 0: General principles. Sydney,
AF
Standards Australia.
SA (2002a). AS/NZS 1170.1:2002 Structural design actions - Part 1: Permanent, imposed and other
actions. Sydney, Standards Australia.
SA (2009). AS 3600-2009 Concrete Structures. Sydney, Standards Australia.
T

9.12 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

10
Multistorey Shoring
10.1 Introduction
In multistorey construction, rarely does one floor have sufficient strength to support the weight
imposed by the construction of the next floor. Overloading can be avoided by sharing the construction
load down to a rigid foundation or between a sufficient numbers of suspended floors. This is achieved
by installing multistorey shores between the floors.
In practice, often one or two floors of formwork are used in conjunction with several floors of
multistorey shores. Figure 10.1 depicts an example from a recent project showing how multistorey
D
shoring was specified.
R
AF
T

Figure 10.1 Example of multistorey shoring with one floor of formwork and three floors of shores
The project designer is responsible for specifying the minimum number of floors of formwork and
multistorey shores and the load distribution between the floors relative to the type of formwork, timing
and sequence of construction (SA 2010). However, this can only be determined and specified after the
construction contractor has informed the project designer of the details of the proposed construction
method, schedule, formwork details and shore layout.
After the project designer has specified the load distribution between the floors, the formwork
designer is responsible for ensuring the formwork and the shores themselves are not overloaded (SA
1996). It is necessary to obtain project designer approval for any changes to the proposed formwork
and multistorey shore layout.

2013 Stephen Ferguson all rights reserved 10.1


Formwork Handbook

10.1.1 Significance of multistorey shoring design


Of prime importance is avoiding the under-design of the multistorey shoring. Too few floors of
multistorey shoring or too few shores on each floor would result in overload, unacceptable cracking
and risk structural failure. When considered in isolation, the risk of under-design might warrant a
conservative approach and simply specifying: longer curing periods prior to stripping and loading,
more floors of formwork and multistorey shoring, and higher capacity shores. However, there is a
significant economic benefit to (without increasing the floor cycle time) reduce the number of floors
of formwork and multistorey shores, as well as to reduce the number of shores on each floor. Doing so
reduces the cost of construction and shortens the construction period.
Examples of construction cost savings that result from minimising multistorey shoring, include: less
labour; reduced quantity of formwork and shoring; reduced cost of perimeter screens covering fewer
floors; and reduced construction delay as subsequent trades can follow closer to the wet head
without obstruction from the presence of multistorey shores. Over the construction period of a high
rise building, the cost benefits of having one less floor of multistorey shoring, or reducing the floor
cycle by one day can be measured in the many tens, if not hundreds, of thousands of dollars.
The rigour and effort invested in the design of the multistorey shoring system should be commensurate
D
with the consequence of failure and potential cost benefit.
10.1.2 Current design guidance in Australian Standards
10.1.2.1 AS 3600 2009
The Australian Standard for concrete structures, AS 3600 2009 (SA 2009), requires the number of
R
floors of formwork and multistorey shores or the load distribution between the floors to be calculated,
but provides no guidance on how this is achieved.
In addition, AS 3600 2009 specifies:
(a) Where backpropping is used, the procedure shall comply with AS 3610 1995;
(b) Before removing supports from under a storey, all supported floors above shall be
AF
checked by calculation for cracking and deflection under the resulting loads; and
(c) No formwork supports or multistorey shoring shall be removed within 2 days of the
placing of any slab directly or indirectly supported by the supports or shoring.
10.1.2.2 AS 3610 1995
The Australian Standard for formwork, AS 3610 1995 (SA 1995), sets out rules to determine the
minimum capacity of multistorey shoring, specifically:
(a) Where the lowest level of multistorey shoring is seated on a rigid foundation, the
minimum capacity of the shoring shall not be less than the sum of the total weight of the
T
suspended floor systems and imposed construction loads for all the levels above the
shoring; or
(b) Where the lowest level of multistorey shoring is seated on a suspended floor structure, the
minimum capacity of the shoring shall not be less than twice the sum of the weight of
heaviest single supported floor above the lowest level of multistorey shoring and the
imposed construction loads on that floor.
Implicit in this requirement is that the lowest level slab should also be designed to carry twice the sum
of the weight of heaviest single supported floor above the lowest level of multistorey shoring and the
imposed construction loads on that floor
For economic reasons, in practice, rarely is this ever followed.

10.2 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

10.1.3 Guidance provided in the literature


10.1.3.1 Problems with past practice
Assuming infinite shore stiffness underestimates the load carried by the upper floors
Of particular concern, is the knowledge that estimates of the load distribution between floors by
assuming the axial stiffness of the shores is infinite will underestimate the load carried by the upper
floors (Beeby 2000). This is significant because the upper floors are the youngest, least capable and at
the greater risk of overload.
Figure 10.2 compares the load distribution, with one or two levels of reshores and equal slab stiffness,
based on site measurements and predictions assuming infinite shore stiffness and that the shore layout
is equivalent to a uniformly distributed load. A much larger proportion of the load was found to be
supported by the uppermost slab and a smaller proportion was transmitted to lower slabs by the
reshores.
When reshoring additional floors of multistorey shoring have little effect
Another significant finding was that by taking account of the effective axial stiffness of the shoring,
little benefit is gained by increasing the number of floors of reshores.
D
R
AF
T
Figure 10.2 Indicative discrepancy in load distribution (Beeby 2000)
Thinning out multistorey shores on lower levels is counter-productive
The load carried by the upper floors will be further increased if the number of multistorey shores on
lower levels is reduced, see Figure 10.3. Although justified by the reduction in force carried in the
shores on lower levels, the consequential reduction in axial stiffness of the shoring will increase the
share of the load carried by, and may contribute to the overload of, the floors above.
Conclusions
Based on these findings, the share of the load carried by the uppermost slab supporting the formwork
will be larger than previously thought and such that it may often be significantly overloaded in
particular, when:
(i) Multistorey shoring is sparse and heavily loaded;
(ii) Floor to floor heights vary, as shore stiffness is directly proportional to length; or

2013 Stephen Ferguson all rights reserved 10.3


Formwork Handbook

(iii) Shores are made from less stiff materials; e.g. Aluminium is approximately one third the
stiffness of steel, but some alloys have a comparable strength to steel.
D
R
Figure 10.3 Example of thinning multistorey shoring
10.1.3.2 Recent guidance predicting load distribution
There is evidence and guidance in the literature (Beeby 2001, Moss 2003, Park, Hwang et al. 2011)
that a reasonable estimate of the share of the construction load carried by each floor connected by
AF
multistorey shores can be determined by taking account of the:
(a) Method and sequence of stripping and shoring;
(b) Flexural stiffness of each floor at time of loading;
(c) Effective axial stiffness and layout of the formwork and multistorey shores on each floor;
and
(d) Preload in multistorey shores.
There is also guidance on assessing the ability of slabs and beams to carry the construction loads
T
(Beeby 2001) by considering the:
(a) Magnitude and location of the construction loads, including loads from multistorey
shoring;
(b) Magnitude and location of the in-service design loads; and
(c) Floor construction cycle time relative to early-age concrete strength gain.
Other factors that influence the multistorey load distribution include: concrete shrinkage, creep, and
ambient temperature change (McAdam and Behan 1990).
The ACI Guide for Shoring/Reshoring of Concrete Multistorey Buildings (ACI 2005) also provides
useful information.
10.1.4 Guidance provided herein
This Chapter explains the factors that influence the design of multistorey shoring and introduces and
discusses several methods for estimating the load distribution between the floors connected by
multistorey shoring and the load in the multistorey shores themselves. In addition, explanations are
provided on the different methods to determine the load distribution for reshoring and undisturbed

10.4 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

shoring. Special situations that need to be addressed are highlighted and the Chapter concludes with a
discussion on assessing the capacity of slabs to carry loads at an early-age.
10.2 Factors influencing the design of multistorey shoring
10.2.1 Method and sequence of stripping and shoring
There are two methods of multistorey shoring: reshoring or undisturbed shoring.
In Figure 10.1, it is not apparent, without more information, whether the project designer intended
specifying reshoring or undisturbed shoring.
10.2.1.1 Reshoring
Reshoring is the most common method adopted for multistorey shoring; primarily, because it
minimises the load distributed to the lower floors, and therefore minimises the amount of shoring
required.
Figure 10.4 shows the sequence of construction with one floor of formwork and two floors of reshores.
If additional formwork is available, it is not necessary to wait until at least Day 3 to strip and recycle
the formwork, forming the new slab could commence on Day 1 using the additional formwork.
D
Reshoring is characterised by:
1. Removing large areas of formwork thereby allowing the slab to relax, support its self-weight
plus any imposed construction load, and span as-designed between its permanent supports, as
shown in Figure 10.4 on Day X; and then
R
2. Installing new shores under the slab.
Reshoring causes the newest (uppermost) slab at an early-age to carry its self-weight and any
construction loads imposed at the time the formwork is removed, and any addition construction loads
imposed up to the time of reshoring. After reshoring, all slabs in the system carry the entirety of their
own self-weight and a share of any subsequent construction load; e.g. the load from the newly placed
slab.
AF
For typical floors on a multistorey structure, if reshoring is adopted, the uppermost (youngest and least
capable) slab, connected by multistorey shoring, carries the greatest share of load from the
construction of the next floor.
Post-tensioning has a similar effect to reshoring; i.e. leaving the slab carrying its own weight without
disturbing the formwork shores.
The Commentary to AS 3610 1995 (SA 1996) warns that reshoring is a hazardous operation and AS
36101995 does not provide any guidance, other than to warn that it requires close attention to the
early development of concrete strength.
T

2013 Stephen Ferguson all rights reserved 10.5


Formwork Handbook

D
Figure 10.4 Multistorey shoring with one floor of formwork and two floors of reshores
10.2.1.2 Undisturbed shoring
Undisturbed shoring describes the situation when the slabs remain supported continuously.
R
The simplest approach is to have multiple levels of formwork shores, which remain undisturbed until
it is time for the lowest level to be removed and recycled to be used again, see Figure 10.5.
AF
T
Figure 10.5 Multistorey shoring with three floors of undisturbed formwork
Figure 10.5 shows the sequence of construction with three floors of formwork. If additional formwork
is available, it is not necessary to wait until Day 3 to strip and recycle the formwork, and forming the
new slab could commence on Day 1 with the additional formwork.
Undisturbed shoring also applies where:
(a) For reasons of economy, some formwork systems allow the forms to be removed, leaving
the shores undisturbed (two-stage stripping), as shown in Figure 9.2; or

10.6 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

(b) Stripping and backpropping 1, where shores are installed prior to removing small areas of
the formwork, thereby preventing the slab from relaxing, as shown in Figure 9.3.
Compared to reshoring, using undisturbed shoring on typical floors will generally result in the lowest
(oldest and most capable) slab, connected by multistorey shoring, carrying the greatest share of the
load (Grundy and Kabaila 1963). All slabs carry a portion of their own self-weight plus a share of the
self-weight of each of the slabs above and any construction loads. Only the lowest slab in the system
carries the entirety of its own self-weight.
10.2.2 Flexural stiffness of each floor at time of loading
The elastic flexural stiffness of each concrete slab (Sc) is a function of its effective second moment of
area, modulus of elasticity, span and support conditions, see Equation 10.1. Importantly, the modulus
of elasticity varies with concrete age.

E cj Ief
Sc =
lc
(10.1)
D
where
Ecj = mean modulus of elasticity of the concrete at the relevant age;
Ief = effective second moment of area of the concrete section;
= factor for the slab continuity and support conditions; and
R
lc = concrete slab span.
10.2.2.1 Effective second moment of area
For reinforced concrete, the second moment of area will change with the onset of cracking. Long-term
effects of creep and shrinkages can be neglected because of the relatively short construction time.
AF
For a rigid foundation, such as a thick concrete raft, the second moment of area approaches infinity.
For beams with reinforcement ratios Ast/bd 0.005, AS 3600 specifies an effective second moment of
area given by:

3
Mcr
Ief = Icr + ( I Icr) I
Mds (10.2)

where
T
Icr = second moment of area of a cracked section with the reinforcement
transformed to an equivalent area of concrete;
I = second moment of area of the uncracked concrete section about the
centroidal axis;
Mcr = bending moment causing cracking of the section; and
Mds = maximum bending moment at the section, based on the construction load.
For beams with reinforcement ratios Ast/bd < 0.005, Ief is still calculated by Equation 10.2, but cannot
exceed 0.6I.
Alternatively, the following method (Bischoff and Scanlon 2007) has been used to determine the
second moment of area of flat plate slabs with low reinforcement ratios (Park, Hwang et al. 2011):

1
The term backpropping is often used to describe any type of multistorey shoring. Herein backpropping
refers to only the shores installed as shown in Figure 9.3. It does not refer to reshores.

2013 Stephen Ferguson all rights reserved 10.7


Formwork Handbook

Icr
Ief = I
2
Mcr Icr
1 1
Mds I (10.3)

An average value of Ief for the beam or slab is required. This can be determined from the values at the
critical sections based on the averaging procedure specified in AS 3600 (Clause 8.5.3.1).
10.2.2.2 Concrete modulus of elasticity
The concrete modulus of elasticity increases with age. Unless determined by testing in accordance
with AS 1012.17, the mean modulus of elasticity (within a range of 20%) is specified in AS 3600 as:

Ecj = ( 1.5) (0.043 f cmi ) ,when fcmi 40 MPa (10.4a)

Ecj = ( 1.5) (0.024 f cmi + 0.12 ) ,when fcmi > 40 MPa (10.4b)
D
where
Ecj = mean modulus of elasticity of the concrete at the relevant age;
= density of concrete (for normal weight concrete 2400 kg/m3); and
R
fcmi = mean value of the in situ compressive strength of concrete at the relevant
age.
Without testing, the mean modulus of elasticity of concrete determined using Equations 10.4a or 10.4b
may vary up to 20% of the actual value.
If the concrete modulus of elasticity is determined using Equations 10.4a or 10.4b:
AF
(a) It would be prudent to determine the sensitivity of load distribution to variations in the
concrete modulus of elasticity by calculating the effects assuming 0.8Ecmj to 1.2Ecmj;
(b) At serviceability limit states, a load distribution based on the mean modulus of elasticity
should be acceptable;
(c) At ultimate limit states, adopt the most adverse load distribution resulting from assuming
the concrete modulus of elasticity at either end of the range of 0.8Ecmj to 1.2Ecmj.
10.2.2.3 Span and support conditions
The flexural stiffness of a concrete slab will depend on its span, continuity and support conditions; e.g.
T
simple supported, continuous, fixed, etc.
For typical floors, the span and support conditions vary from bay to bay on each floor, but, generally,
are the same from floor to floor. Accordingly, the multistorey load distribution may vary from span to
span. For example, corner and edge slabs, which are less stiff due to their edge support conditions,
with the same span and same shore layout as the internal slabs, will have a higher shore to slab
stiffness ratio and hence allow a greater amount of the construction load down to the lower slabs.
For one-way slabs, beams and simple two-way slabs, it is possible to develop simple expressions for
lc for use in Equation 10.1. However, this is not practical for more complex slab and beam
arrangements.
10.2.3 Effective axial stiffness of shores
From Day 1 (the first day after a slab is poured) construction loads are applied to the uppermost
(newest) slab culminating in the pouring of the next floor on Day T. During this time, the multistorey
shoring arrangement (similar to that depicted in either Figure 10.4 or 10.5) can be represented by an
idealised model shown in Figure 10.6.

10.8 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

In Figure 10.6, multistorey shores are represented by springs. The variables Sp1 to Sp2 represent the
effective axial stiffness of each respective level of multistorey shores. The flexural stiffness of each
concrete slab sharing load is represented by the variables Sc1 to Sc3, respectively. The construction load
imposed on the top slab to be shared by the multistorey shoring (see Section 10.2.4) is represented by
the variable W1. This may comprise both point and uniformly distributed loads. The variables Wp1 and
Wp2 represent share of W1 carried by each respective level of multistorey shores and Wc1 to Wc3
represents the share of W1 transferred to each respective concrete slab by the multistorey shores.
This arrangement is repeated in Figure 10.10 and more complex arrangements can be seen in Figures
10.11 to 10.13.
D
R
Figure 10.6 Idealised model of one floor of formwork (not shown) and two floors of multistorey
shoring
10.2.3.1 Single shore
The effective axial stiffness of a single formwork or multistorey shore (Sp) given in Equation 10.5
AF
takes account of:
(a) The reduction in stiffness caused by formwork, packing or other material between the top
and/or bottom of the shore and the concrete; and
(b) The reduction in stiffness caused by the inclination and other shore imperfections.

p
Sp =
1 1
+
Ss Sf (10.5)
T
where
p = capacity reduction factor for shore axial stiffness that takes account of the
detrimental effects of imperfections;

In the absence of tests to confirm the actual stiffness of formwork or


multistorey shores, p 0.80;
Ss = axial stiffness of a shore, see Equation 10.6; and
Sf = axial stiffness of any formwork or packing between the top or bottom of the
shore and the concrete, see Equation 10.7.

Es As
Ss =
ls (10.6)

where

2013 Stephen Ferguson all rights reserved 10.9


Formwork Handbook

Es = modulus of elasticity of the shore material;


As = cross-sectional area of the shore; and
ls = length of the shore.

Ef Af
Sf =
lf (10.7)

where
Ef = modulus of elasticity of the forms or other material between the top and/or
bottom of the shore and the concrete;

The range of the mean modulus of elasticity perpendicular to the grain falls
between (BS 2008), for:
(i) For softwood timber, 230 MPa to 530 MPa; and
D
(ii) For hardwood timber, 630 MPa to 1330 MPa.
Af = bearing area of the forms or other material; and
lf = thickness of the form or other material.
The presence of softwood forms between the shores and concrete can reduce the effective axial
stiffness of a steel shore to approximately 30% of its axial stiffness based on the shore section only
R
(Fang, Zhu et al. 2001).
In addition, Fang et al. found the presence of imperfections, such as inclination, reduces the theoretical
shore stiffness and also needs to be taken into account. Failure to do so may lead to underestimation of
the load carried by the uppermost slabs.
AF
The effective axial stiffness of shores will also be less where there is high-strutting to the floor above;
e.g. entrance foyers, plant rooms, etc. Aluminium and timber shores may have significantly less axial
stiffness than steel shores.
10.2.3.2 Multiple shores
On any one level, the effective axial stiffness of the multistorey shores as a group is a function of the
axial stiffness of each shore and the shore layout. This can be demonstrated by observing the change in
load distribution in the following four design examples; in particular, the share of the load carried by
the uppermost slab, see Table 10.1.
Table 10.1 Load distribution in multistorey shoring taking account of shore axial stiffness and layout
T
Load Design Example
Ratio 1 2 3 4
Wc1/W1 0.236 0.404 0.423 0.496
Wc2/W1 0.251 0.250 0.273 0.278
Wc3/W1 0.257 0.188 0.173 0.140
Wc4/W1 0.257 0.159 0.131 0.087

Design Example 1 Multistorey load distribution assuming infinite shore stiffness


This design example calculates the load distribution assuming infinite shore stiffness.
Consider the load distribution for the situation depicted in Figure 10.7. In this situation, 35
number 3.25m tall steel shores are spaced evenly at 1.2 m in one direction and 1.8 m in the

10.10 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

other. Each shore has a cross-sectional area of 574 mm2. The shores are in direct contact with
the 250 mm thick concrete slabs, which are continuous on all four sides and span 8 m in both
directions. The characteristic compressive strength of the concrete in each slab at the time of
loading is, from top to bottom: 20 MPa, 25 MPa, 32 MPa, and 32 MPa.
D
R
Figure 10.7: One level of formwork and three levels of reshores all on a 5 x 7 grid
Neglecting cracking and assuming infinite shore stiffness, an analysis of this arrangement shows
the share of the load (applied to the top floor) carried by each floor is 0.236, 0.251, 0.257 and
0.257, top to bottom.
Design Example 2 Effect of axial shore stiffness on multistorey load distribution
AF
This design example calculates the load distribution taking account of the actual shore stiffness.
In the absence of any specific guidance on the effective axial stiffness of the shores, it is
possible to take account of shore imperfections by adopting a reduced cross-sectional area for
each shore (as per Equation 10.5); i.e. the effective cross-sectional area of each shore, Ase = p
574 mm2 = 0.80 574 mm2 = 459 mm2.
Thus, neglecting cracking and taking account of the shore stiffness, reanalysing shows the share
of load (applied to the top floor) carried by each floor is 0.404, 0.250, 0.188 and 0.159, top to
bottom.
T
The share of load carried by the uppermost slab increased by a factor of 1.71 (0.404/0.236).
Design Example 3 Effect of a thinned shore layout on multistorey load distribution
This design example compares the effect on load distribution of taking account of the shore
stiffness and thinned shore layout.
On the lower two levels, the shores are thinned out from a grid of 5 x 7 to a grid of 5 x 4, as
shown in Figure 10.8. Reanalysing the arrangement shows the share of the load (applied to the
top floor) carried by each floor is 0.423, 0.273, 0.173, and 0.131, top to bottom.
The share of load carried by the uppermost slab increased by a factor of 1.79 (0.423/0.236).
The increase in load carried by the uppermost slab depends on revised shore layout because
removing shores near the centre of the slab increases the load on the uppermost slab more so
than removing those near the slab's supports.

2013 Stephen Ferguson all rights reserved 10.11


Formwork Handbook

D
Figure 10.8: As per Figure 10.7 except the lower two floors of reshores are on a 5 x 4 grid
Design Example 4 Effect of the presence of soft wood formwork on load distribution
R
This design example takes account of softwood timber formwork remaining in place between
the top of the shores and the concrete slab; i.e. the shores were relaxed allowing the slab to
deflect before being reset without removing the formwork.
Consider the formwork is constructed from F14 17 mm plywood, 95x65 LVL joist at 400 mm
centres and 150x77 LVL bearers spanning 1.2m between and bearing on shores with a 150x150
AF
end plate.
Using Equation 10.6, the axial stiffness of a single formwork shore is:

200000 574
Ss =
3250

Ss = 35.3 kN/mm

Based on a mean modulus of elasticity perpendicular to the grain of 400 MPa (BS 2008) and an
effective bearing area calculated at the centre of each member assuming a 45 load distribution,
T
the effective compressive axial stiffness of the formwork, considering the stiffness of the
plywood, joist and bearer, is given by:

1
Sf =
1 1 1
+ +
400 3( 65 + 17) 1800 400 3( 75 + 65) 65 400 ( 150 + 150) 75

17 95 150
Sf = 39.3 kN/mm

Using Equation 10.5, the effective axial stiffness of a single formwork shore taking into account
the presence of the formwork is:

0.8
Sp =
1 1
+
35.3 39.3

10.12 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

Sp = 14.9 kN/mm

Taking account of the presence of formwork between the top of the shores, the share of the load
(applied to the top floor) carried by each floor is 0.496, 0.278, 0.140, and 0.087.
The share of load carried by the uppermost slab increased by a factor of 2.10 (0.496/0.236).
It is useful to represent the effective axial stiffness of a level of shores as a multiple of the elastic
flexural stiffness of the 28-day slab flexural stiffness, see Section 10.3.1.
For example, the results from Design Examples 1 to 4 have been calibrated using the relative stiffness
method presented in Section 10.3.1.3 and the effective axial stiffness each level of shores relative to
the 28-day slab flexural stiffness presented in Table 10.2.
Table 10.2 Axial stiffness of each level of shores relative to the 28-day slab flexural stiffness

Stiffness Design Example


Ratio 1 2 3 4
Sp1/Sc28 3.6 3.6 1.9
D
Sp2/Sc28 3.6 2.1 1.1
Sp3/Sc28 3.6 2.1 1.1
R
In the worst case, neglecting the effective axial shore stiffness could be underestimate the load carried
by the uppermost slab by 50%
10.2.4 Preload in multistorey shores
It is normally assumed that shores (reshores and backprops) are installed with zero preload. The effect
of excessive, or predetermined, tightening of the shores is beneficial as it increases the load in slab
AF
below and decreases the load in the slab above; i.e. move load from the slab above to the slabs below.
When reshoring, preloading will result in a more even distribution of the share of load between the
supporting slabs (BRE 2004). Accordingly, there may be merit in taking preload into account in
determining the load distribution between slabs; however, in practice, the magnitude of preload in
shores is difficult to control and measure.
10.2.5 Construction loads
To ascertain the load on each floor at the relevant time (see Figure 10.9), it is necessary to take
account of:
T
(a) The self-weight of concrete floors, Gc (see Section 4.2.1.2);
(b) The imposed construction loads on each floor, Qv; and
(c) The load transferred to each floor through the multistorey shores.

2013 Stephen Ferguson all rights reserved 10.13


Formwork Handbook

D
R
Figure 10.9: Different multistorey shoring load situations
10.2.5.1 Imposed construction load
The imposed construction load on any floor may include:
(a) The weight of the formwork and multistorey shores, Gf (see Section 4.2.1.1); and
AF
(b) Imposed actions from:
(i) Workers and their equipment on the formwork, Qw;
(ii) Stacked materials, Qm (see Section 4.2.3.2); or
(iii) Construction equipment that may be present, Qx (see Section 4.2.3.5).
Loads Qw and Qm are not considered to act concurrently at the same location.
For the weight of workers and their equipment (Qw), it is recommended (SA 1995) to allow, acting
concurrently:
T
1.0 kPa acting on the soffit formwork, prior to or during concrete placement;
1.0 kPa acting on the uppermost concrete slab; and
0.25 kPa for all other slabs carrying a share of the load.
The weight of construction material or equipment stacked on soffit formwork or on any of the slabs
carrying a share of the load is likely to increase the number of floors of shoring and must be taken into
account. This may be achieved economically by limiting the magnitude of the load from stacked
materials and locations where stacking material is permitted.
Additional multistorey shoring may be required to share the load imposed by construction equipment,
such as loading platforms, forklifts, elevated working platforms, perimeter screens, personnel and
material hoists, formwork hoists, and concrete pump booms. Where the location of the construction
equipment is fixed, specific shoring may be introduced to share the imposed load; otherwise, the load
from construction equipment should be considered to act concurrently with other construction loads
and in the most adverse location.

10.14 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

10.2.6 In-service design floor load


During their service life, structural floors are designed to carry their self-weight and imposed
permanent and variable actions.
During construction, under the effects of construction loading the structure should not suffer any
detrimental effects nor be subject to any greater risk of failure than would occur if the in-service
design floor load were applied once the concrete has reached full strength.
The project designer is responsible for specifying the in-service design loads. For floors in concrete
multistorey buildings, the un-factored design service load follows the format given by Equation 9.14
and the strength limit states design load follows that given by the most adverse of Equations 9.16 or
9.17.
10.2.7 Floor construction cycle time
The timing of construction activities, and consequently loading, relative to the rate of gain of early-age
concrete strength is of critical importance. World-wide, floor cycles commonly vary between 2 days
(Ferguson 2000) to 2 weeks. In Australia, compliance with AS 3600 effectively limits the minimum
floor cycle to 4 days (see Figures 10.4 and 10.5), as shores must remain undisturbed for 2 days after
D
placing concrete.
During a floor-to-floor cycle of T days, critical load situations arise that must be assessed, namely:
1. Day 1 to Day T-1 (other than the day of concrete is placed), additional construction loads
may occur at anytime (see Figure 10.7), including:
R
(a) Stacked material on newly placed or intermediate floors; and
(b) Loads due to commencing construction of the next floor; e.g. erecting formwork or
perimeter screens.
2. Day 3 to Day T-1, due to removal of, disturbing, or installation of formwork shores or
multistorey shores; and
AF
3. Day T, casting a new slab.
For each situation, assess and determine:
(a) The early-age concrete modulus of elasticity for each floor sharing the construction load;
(b) The construction load on each floor and, if applicable, formwork erected for the next
floor, see Figure 10.7.
10.2.8 Other factors
Other factors that influence the multistorey load distribution include: concrete creep and ambient
T
temperature change.
These factors should be taken into account when assessing the load distribution and capacity of slabs
to carry construction loads.
10.2.8.1 Concrete creep
Concrete creep causes a continuous process of load redistribution upwards as the new slab gains
strength (stiffness). The magnitude of redistribution may cause the load in the top floor to increase by
10 to 20% (McAdam and Behan 1990, Duan and Chen 1995).
10.2.8.2 Ambient temperature change
Concrete strength gain
The rate of gain of strength of concrete is reduced in cold temperatures. At an early age, reduced
concrete strength impairs the capacity of slabs to carry load and reduces the flexural stiffness of slabs.
Accordingly, the load distribution and multistorey shoring required will vary with ambient
temperature and may require change over the construction period.

2013 Stephen Ferguson all rights reserved 10.15


Formwork Handbook

Multistorey shore load


A rise in temperature, relative to that on the day of the pour of a new slab, will cause the supporting
shores to expand. The expansion is restrained by the slabs and columns and the load in the slab below
increases, while the load in the slab above decreases (McAdam and Behan 1990). A drop in
temperature will have the reverse effect.
10.3 Methods for calculating load distribution in multistorey shoring
There are several methods recommended in the literature for calculating the load distribution in
multistorey shoring. Three methods are discussed herein, namely:
1. Relative stiffness method;
2. Slab shore interaction method; and
3. Finite element analysis.
All the methods are suitable for both reshores and undisturbed shores.
10.3.1 Relative stiffness method
D
The Guide to Flat Slab Formwork and Falsework (CSG 2003) presents simplified formulae to
calculate the load in each level of shoring based on the flexural stiffness of the slabs and effective
axial stiffness of the level of shores.
Unfortunately, the use of this method has limitations:
(a) It is appropriate for one-way and two-way reinforced concrete slabs with beams, as well
R
as beams that behave elastically; e.g. slabs up to 350 mm thick and not heavy stiff beams.
(b) It is useful for reshoring, but less so for undisturbed shores;
(c) If the effective axial stiffness of each floor of shores is not known, which is usually the
case, application of this method relies on making an accurate assumption or estimate of
the effective axial stiffness of each level of shores relative to the flexural stiffness of the
AF
slabs; e.g. in terms of a multiple of the flexural stiffness of the slabs at 28 days; e.g. as per
the example in Section 10.2.3.2 where Sp = 0.8Sc28; and
(d) The actual load in individual shores is not known only the total load in the level of shores,
which may lead to underestimating the load in the heaviest loaded shores.
For the purposes of simplifying the formulae provided in Guide to Flat Slab Formwork and
Falsework, it is useful to define a "relative stiffness term" for slab 1 to 2, K12, and for slab 2 to 3,
K23, and so on as:

Sc1 Sc1
T
K12 = 1 + +
Sc2 Sp1 (10.8)

Sc2 Sc2
K23 = 1 + +
Sc3 Sp2 (10.9)

Sc3 Sc3
K34 = 1 + +
Sc4 Sp3 (10.10)

Sc4 Sc4
K45 = 1 + +
Sc5 Sp4 (10.11)

10.16 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

10.3.1.1 One level of multistorey shores


D
Figure 10.10: One level of multistorey shoring
For systems with a single level of multistorey shoring (see Figure 10.10), the load in the multistorey
shoring, Wp1 is given by:

W1
Wp1 =
R
K12 (10.12)

In Equation 10.12, W1 is the construction load on the top slab to be shared by the multistorey shoring.
W1 will be the sum of the construction load from the formwork shores and any imposed construction
load acting on the top slab.
AF
The share of W1 transferred to the top slab is given by:

Wc1 = W1 Wp1 (10.13)

The share of W1 transferred to the second (bottom) slab is given by:

Wc2 = Wp1 (10.14)


T

2013 Stephen Ferguson all rights reserved 10.17


Formwork Handbook

10.3.1.2 Two levels of multistorey shores


D
Figure 10.11: Two levels of multistorey shoring
R
For systems with two levels of multistorey shoring (see Figure 10.11), the load in the top level of
multistorey shoring, Wp1, and the load in the second level of multistorey shoring Wp2, is given by:

W1
Wp1 =
S c1
AF
S c2
K12
K23 (10.15)

Wp1
Wp2 =
K23 (10.16)

The share of W1 transferred to the top slab is given by Equation 10.13.


The share of W1 transferred to the second slab is given by:
T
Wc2 = Wp1 Wp2 (10.17)

The share of W1 transferred to the third (bottom) slab is given by:

Wc3 = Wp2 (10.18)

10.18 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

10.3.1.3 Three levels of multistorey shores


D
R
Figure 10.12: Three levels of multistorey shoring
Based on the same theory, for systems with three levels of multistorey shoring (see Figure 10.12), the
loads in the respective level of multistorey shoring, Wp1, Wp2, and Wp3, are given by:

W1
AF
Wp1 =
S c1
S c2
K12
S c2
S c3
K23
K34 (10.19)

Wp1
Wp2 =
T
S c2
S c3
K23
K34 (10.20)

Wp2
Wp3 =
K34 (10.21)

The share of W1 transferred to the top slab is given by Equation 10.13. The share of W1 transferred to
the second slab is given by Equation 10.17. The share of W1 transferred to the third slab is given by:

Wc3 = Wp2 Wp3 (10.22)

The share of W1 transferred to the fourth (bottom) slab is given by:

Wc4 = Wp3 (10.23)

2013 Stephen Ferguson all rights reserved 10.19


Formwork Handbook

10.3.1.4 Four levels of multistorey shores


D
R
AF
Figure 10.13: Four levels of multistorey shoring

Based on the same theory, for systems with four levels of multistorey shoring (see Figure 10.13), the
loads in the respective level of multistorey shoring, Wp1, Wp2, Wp3, and Wp4, are given by:

W1
Wp1 =
S c1
S c2
T
K12
S c2
S c3
K23
S c3
S c4
K34
K45 (10.24)

Wp1
Wp2 =
S c2
S c3
K23
S c3
S c4
K34
K45 (10.25)

10.20 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

Wp2
Wp3 =
S c3
S c4
K34
K45 (10.26)

Wp3
Wp4 =
K45 (10.27)

The share of W1 transferred to the top slab is given by Equation 10.13.


The share of W1 transferred to the second slab is given by Equation 10.17.
The share of W1 transferred to by the third slab is given by Equation 10.22.
The share of W1 transferred to the fourth slab is given by:

Wc4 = Wp3 Wp4


D
(10.28)

The share of W1 transferred to the fifth (bottom) slab is given by:

Wc5 = Wp4 (10.29)


R
10.3.2 Slab shore interaction method
The Guide to Flat Slab Formwork and Falsework (CSG 2003) includes a spreadsheet for calculating
the loads in multistorey shoring and the connected flat plate slabs. The spreadsheet takes account of
the interaction of slab flexural stiffness, shore axial stiffness and preload based on methods for
estimating load distribution that have been previously investigated (Liu, Chen et al. 1985, El-Shahhat,
AF
Rosowsky et al. 1992) and refined by Beeby (Beeby 2001). Useful commentary on its use can be
found in the Early age construction loading (BRE 2004).
Using this method, for a given multistorey shore layout, the axial load in each multistorey shore (Np)
can be expressed as:

Np = Sp ( A B) (10.30)

In Equation 10.30,
Sp = Effective axial stiffness of each individual multistorey shore, see Equation 10.5;
T
A = Deflection of the slab at the top of the multistorey shore at the point where the shore is
located; and
B = Deflection of the slab at the bottom of the multistorey shore at the point where the shore
is located.
The deflection of the slab at the top of each shore (A) is equal to the sum of the deflections that arise
from the point load from each individual formwork or multistorey shore on the level above and any
other construction loads acting downward on the top of the supported slab, less the reduction in
deflection that arises from the sum of the effects of the point load from each individual multistorey
shore on the same level as the shore in question, acting upward on the underside of the supported slab.
The deflection at the bottom of each shore (B) is equal to the sum of the deflections in the slab
supporting the shore that arise from the point load from each individual multistorey shore on the same
level as the shore in question and any other construction loads acting downward on the slab supporting
the shore, less the reduction in deflection that arises from the sum of the effects of the point load from
each individual multistorey shore on the level below the shore in question (where present), acting
upward on the underside of the slab supporting the shore in question.

2013 Stephen Ferguson all rights reserved 10.21


Formwork Handbook

Considering the axial load in every shore on each level of multistorey shoring, results in a series of
complex simultaneous equations that can be solved using matrix methods.
The accuracy of this method depends on the suitability of the deflection coefficients used to calculate
the slab deflection at each shore location. For two-way slabs, using simple methods for estimating slab
deflections (e.g. (Scanlon and Suprenant 2011)) may result in inaccurate estimates of load distribution,
because although these methods may provide a reasonable estimate of the maximum deflection for a
two-way slab, without adjustment, they can underestimate the deflection near the slab edges and
therefore the load in those shores.
The advantages of this method are that it is suitable for:
1. Situations where the layout of formwork shores and/or multistorey shores differs from
floor to floor, see Figure 10.14; and
2. One-way slabs and beams, as well as simple rectangular two-way flat slabs; and
3. Programming in computer spreadsheets or worksheets that, once written, allow for rapid
and efficient analysis of different multistorey shoring arrangements.
D
R
AF

Figure 10.14: The layout of formwork shores and multistorey shores differ from floor to floor
10.3.3 Finite element analysis
Historically, the emphasis of researchers has been to seek simple methods for determining the
T
multistorey load distribution. In principle, there is nothing wrong with such an approach providing the
simplified methods did not tend to underestimation in a manner that would increase the risk of failure
beyond acceptable limits. However, there is evidence that, for reshoring in particular, the load carried
by the upper floors may have been significantly underestimated.
Work Health and Safety Regulations place clear obligations on project designers to minimise risks to
the health and safety of those involved during construction. In doings so, it is reasonable that workers
constructing the structure should not be put at any greater risk from structural failure than those who
will later occupy the building. Accordingly, the level of rigour and level of reliability appropriate for
the design of the structure to resist multistorey loading should be no less than for the design of the
structure to resist the loadings it will experience during its working life.
There can also be a significant economic benefit to adopt more rigorous methods of analysis that
optimise multistorey shoring, see Section 10.1.1. As stated earlier, the rigour and effort invested in the
design of the multistorey shoring system should be commensurate with the consequence of failure and
potential cost benefit.

10.22 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

Given the ready availability of software for finite element analysis (FEA) and its growing use for the
design of multistorey structures, it would seem reasonable that the load distribution and effects of
multistorey shoring also be analysed using FEA.
The advantages of this approach include:
1. Direct comparison of the in-service and construction design action effects;
2. More closely simulates the actual structural arrangement, than idealised models; and
3. Most importantly, provides an increased level of certainty.
Figure 10.15 shows an example of the output from a finite element analysis of the accumulated
deflection on a portion of the lowest level (4th floor) slab supporting undisturbed multistorey shoring.
D
R
AF
Figure 10.15: Slab deflections under staged construction loads from undisturbed multistorey shoring

10.4 Analysis methods for reshores vs undisturbed shores


10.4.1 General
T
When loads are applied to, or removed from, a floor connected by multistorey shoring, load is shared
between the connected floors. The load in the floors and multistorey shores accumulate with each
loading event. For each event, the share of the load in the floors and multistorey shores can be
determined based on the relative stiffness of the floors and shores at the time of each event.
Loads may be applied to any floor connected by multistorey shoring, the most common case is when
loads are applied to the top floor, and the load is shared between the floors below, see Figure 10.6.
This may occur on any day of the floor cycle. However, if an intermediate floor is loaded (e.g. with
stacked materials), the distribution of this load will change the share of load carried by the slabs above
and below the floor.
In Figure 10.16, the intermediate floor is shown carrying an additional construction load (W2) from
material stacked on the floor prior to pouring the next slab. The load distribution for W2 must be
determined based on the time the load was installed and before considering the distribution of W1.
Thus, it is necessary to sequentially determine the distribution from each load event using the stiffness
of the concrete and shore layout at the time of loading.

2013 Stephen Ferguson all rights reserved 10.23


Formwork Handbook

D
Figure 10.16 Idealised model with both the top and intermediate floor loadings
10.4.1.1 Floor numbering
Note on Day 1 of any floor cycle, the newly poured slab has some stiffness and is identified as slab 1
R
in the multistorey arrangement (e.g. as shown Figure 10.7) and the floors below are renumbered. The
numbering applies for all load events up to and including the casting of the next floor. The day after
the next floor is poured, what was slab 1 becomes slab 2 and so forth.
10.4.2 Reshoring
As part of the reshoring process, the formwork shores under the slab are removed and the uppermost
AF
slab is allowed to relax and span as-designed. At that time, load is no longer shared to the lower floors.
Therefore, each time reshores are installed the analysis recommences, for each subsequent loading
event the loads are distributed and accumulate, until the formwork shores are removed after pouring
the next slab.
10.4.3 Undisturbed shoring
When analysing undisturbed multistorey shoring, it is necessary to determine the load distribution in
the slabs and multistorey shoring that accumulates from each loading event throughout the time each
slab participates as part of the multistorey shoring arrangement. This starts the day after a slab is
poured and continues until the day the slab is the lowest slab connected by multistorey shoring and the
T
multistorey shores it supports are removed.
For a multistorey structure, it is necessary to sum the calculated load distribution from each load event
slabs connected by multistorey shoring experience, in order, and taking account of the stiffness of the
slabs and shores on each occasion loads are applied or shores removed, from the start to the finish of
construction.
For typical floors, the loads in the slabs and shores will converge and follow a pattern. Particular
attention must be paid to non-typical situations, see Section 10.5.
In the case of undisturbed shoring, removing the lowest level of shores, as shown on the left hand side
of Figure 10.17, can be simulated by loading the floor previously supported by the removed shores
with point loads whose magnitude is equivalent to the load carried by the shores before they were
removed. This situation is represented by an idealised model shown on the right hand side of Figure
10.17.

10.24 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

D
R
Figure 10.17 Idealised model of load effect of stripping the lowest level of undisturbed multistorey
shoring (reproduced from second image from the left in Figure 10.5).

10.5 Special situations to consider


10.5.1 Unloaded multistorey shores
AF
Typically, multistorey shores carry compression only. There are situations where an analysis of the
multistorey shoring system will show a level of shores in tension. This situation may arise when:
(a) After installing reshores and:
(i) Loads that were present on the top slab, before reshoring, are removed; or
(ii) The floor supporting the reshores, or a floor below, is loaded;
(b) A floor connected by multistorey shoring, other than the lowest level, is significantly
stiffer than the floors below (e.g. a floor with a thicker slab or shorter spans).
T
This is not a problem providing the tension does not exceed the accumulated compression or preload
in the shores. The latter case may be indicative of how shoring beneath very stiff slabs may be
ineffective.
Nevertheless, it is good practice to install multistorey shores with a minimum preload to minimise the
risk of unloaded shores falling.
10.5.2 Onset of cracking
During construction, at some point, in slabs sharing multistorey loading subject to increasing loads,
cracking will take place and reduce the flexural stiffness of the slab. It may be prudent to be aware of
the magnitude of the load that will cause the onset of cracking, so any change in flexural stiffness can
be introduced at the appropriate stage.
10.5.3 Foundations settlement
Where multistorey shores are supported directly on soleboards or temporary footings on a soil
foundation, the settlement characteristics of the foundation need to be taken into account. Any
settlement will have a detrimental effect on the axial stiffness of the shores.

2013 Stephen Ferguson all rights reserved 10.25


Formwork Handbook

10.5.4 Props not directly over each other


Where formwork shores or multistorey shores on one level are not directly above the shores on the
level below, less load is transferred to the lower level than if the props on different levels were aligned
(Beeby 2001).
10.5.5 Shores at the centre of the slab carry more load than those closer to the
supports
The load carried by multistorey shores is greatest for the shores in areas where the slab deflection is
greatest; e.g. the centre of any span. Conversely, the load in shores adjacent to rigid supports such as
columns and walls will carry the least load. As a guide, for the cases analysed in Section 10.2.3.2 the
heaviest load shore carries between 1.5 to 1.6 times the load carried by the shores for each level (Wp)
divided by the number of shores on each level.
Only analysis using the slab shore interaction method or finite element analysis will predict the load in
each shore. Care should be exercised when selecting multistorey shores to ensure sufficient capacity
to avoid overload.
10.6 Acceptance criteria for early-age loading
D
Whichever method is used to estimate load distribution in multistorey shoring, Equations 9.11, 9.12,
9.15 and 9.16 provide useful criteria for assessing early-age loading, see Chapter 9.
10.6.1 Uncertainty
In assessing early-age loading due to construction loads from multistorey shoring, it is recommended
R
consideration be given to the uncertainty associated with the accuracy of the calculated load
distribution and the possible variation in magnitude of the construction load.
Possible sources of variations in the estimated load in multistorey shores may include, but are not
limited to:
(a) Inaccuracies in the simulated model of the structure, including:
AF
(i) Neglecting concrete creep;
(ii) Neglecting axial shortening of the columns;
(iii) Inaccuracy in estimating concrete modulus of elasticity;
(iv) Variations between calculated and the actual shore axial stiffness;
(v) Variations between estimated and actual foundation settlement characteristics;
(b) Changes in ambient temperatures;
(c) Level of site control over:
T
(i) The location of and magnitude of stacked materials;
(ii) Variations in preload of shores at installation;
(iii) Unintentional removal or relocation of shores;
(iv) Type of shores used versus that specified; and
(v) Variations in shore layout.
10.6.2 Serviceability limit states
When assessing acceptance criteria at serviceability limit states, the following combination of actions
are recommended:
For reshoring,

W = Gc + Qv + Wc, kPa (10.31)

10.26 2013 Stephen Ferguson all rights reserved


Chapter 10 Multistorey Shoring

Note: prior to the installation of reshores below the uppermost slab, any construction loads applied to
the floors below connected by multistorey shoring will be shared and should be included in Qv.
For undisturbed shoring,

W = Wc, kPa (10.32)

In Equations 10.31 and 10.32,


Gc = weight of the slab or beam being assessed;
Qv = share of imposed construction loads prior to installation of reshores; and
Wc = share of multistorey construction loads, accumulated up until that point of time,
acting on the slab or beam being assessed.
10.6.3 Ultimate limit states
When assessing acceptance criteria at ultimate limit states, it is recommended that the multistorey
construction load be factored as a variable action imposed on the slabs rather than a combination of
permanent and variable actions, using the following combination of factored actions.
D
For reshoring,

Wd = 1.2Gc + 1.5(Qv + Wc), kPa (10.33)

For undisturbed shoring,


R
Wd = 1.5Wc, kPa (10.34)

10.6.4 Acceptable overload


The acceptance or otherwise of the magnitude of construction load from multistorey shoring is the
AF
responsibility of the project designer.
A discussion and guidance on issues associated with loading a slab to above the design service load in
the UK is presented in Appendix E of the Guide to Flat Slab Formwork and Falsework (CSG 2003).
References
ACI (2005). ACI 347.2R-05 Guide for Shoring/Reshoring of Concrete Multistorey Buildings.
Farmington Hills, American Concrete Institute.
Beeby, A. W. (2000). ECBP Task 4 Report - Early Striking and Backpropping (Report BR 394).
London, BRE.
T
Beeby, A. W. (2001). "Criteria for the loading of slabs during construction." Structures & Buildings
146 May 2001(2): 195-202.
Beeby, A. W. (2001). "The forces in backprops during construction of flat slab structures." Strucutres
& Buildings 146 August 2001(3): 307-317.
Bischoff, P. and A. Scanlon (2007). "Effective Moment of Inertia for Calculating Deflectiosn of
Concrete Membes Containing Steel Reinforcement and Fiber-Reinforced Polymer Reinforcement."
ACI Structural Journal 104(No. 1): 68-75.
BRE (2004). Early age construction loading. London, The Concrete Centre.
BS (2008). BS EN 338:2008 Structural timber: Strength classes. London, British Standards.
CSG (2003). Guide to Flat Slab Formwork and Falsework. Berkshire, Concrete Society on behalf of
Concrete Structures Group.
Duan, M. Z. and W. F. Chen (1995). Effects of Creep and Shrinkage on Slab-Shore Loads and
Deflections during Construction. Project Report: CE-STR-95-24, Purdue University.

2013 Stephen Ferguson all rights reserved 10.27


Formwork Handbook

El-Shahhat, A. M., D. Rosowsky and W. F. Chen (1992). "Improved Analysis of Shore Slab
Interaction." ACI Structural Journal 89(No 5 Sept - Oct): 528-537.
Fang, D.-P., H.-Y. Zhu, C.-D. Geng and X.-L. Liu (2001). "On-Site Measurement of Load
Distribution in Reinforced Concrete Buildings during Construction." ACI Structural Journal 98(No.
2): 157-163.
Ferguson, S. A. (2000). "A 2-day cycle using timber formwork." Concrete Vol. 34(No. 3, March): 22-
26.
Grundy, P. and A. Kabaila (1963). "Construction Loads on Slabs with Shored Formwork in
Multistorey Buildings." ACI Journal Proceedings V 60(No 12 Dec): 1729-1738.
Liu, X. L., W. F. Chen and M. D. Bowman (1985). "Construction loads on supporting floors."
Construction International(December): 21-26.
McAdam, P. S. and J. E. Behan (1990). Multi-storey Formwork Loading (Technical Paper 7). Sydney,
Concrete Institute of Australia: 18 pages.
Moss, R. M. (2003). Best practice in concrete frame construction: Practical application at St George
Wharf. London, BRE Centre for Concrete Construction.
D
Park, H.-G., H.-J. Hwang, G.-H. Hong, Kim, Yong-Nam and J.-Y. Kim (2011). "Slab Construction
Load Affected by Shore Stiffness and Concrete Cracking." ACI Structural Journal 108(No. 6): 679-
688.
SA (1995). AS 3610 - 1995 Formwork for concrete. Sydney, Standards Australia.
R
SA (1996). AS 3610 Supplement 2 - 1996 Formwork for concrete - Commentary. Sydney, Standards
Australia.
SA (2009). AS 3600-2009 Concrete Structures. Sydney, Standards Australia.
SA (2010). AS 3610.1 - 2010 Formwork for concrete Part 1: Documentation and surface finish.
Sydney, Standards Australia.
AF
Scanlon, A. and B. A. Suprenant (2011). "Estimating two-way slab deflections." Concrete
International(July): 29 - 34.
T

10.28 2013 Stephen Ferguson all rights reserved


Chapter 11 Concrete Finishes Identifications of defects

11
Concrete Finishes Identification of defects
11.1 Introduction
The intent is to supplement those parts of Australian Standard AS 3610.12010 (SA 2010) that refer
to the inspection, evaluation of the quality of finish and defects visible following removal of the forms.
There often is a difference between the expectations of the architect and what is achievable on site. In
many situations the initial reaction to a surface defect is to focus on the quality of the formwork when
site practices may have resulted in the observed problem.
11.2 Blowholes
D
The evaluation of blowholes can be very subjective when comparing an actual surface with the
photographic charts published in AS 3610.12010 and AS 3610 Supplement 1 (SA 1995). It is also
important to remember that the comparison of the charts with the surface being evaluated must be at a
viewing distance of not less than 2 m when determining if the number of blowholes compares with
required finish Class.
R
Where the number and size of blowholes exceeds that permitted for the finish Class repair often can
effected by filling holes with a mortar of matching colour. However, if the specified finish is textured
such as with sawn boards it is virtually impossible to produce a good repair and particular attention
needs to be paid to the placing and compaction of the concrete. The effect of blowholes in a sawn
board finish is illustrated in Figure 11.1. In this instance the blowholes shown in Figure 11.1 do not
AF
detract from the effect of the sawn board finish, but an increase in their number would result in an
unacceptable appearance that could not be effectively repaired.
T

Figure 11.1 Acceptable blowholes

11.1
Chapter 11 Concrete Finishes Identifications of defects

11.3 Face steps


Typically, face steps permitted by AS 3610 for Class 1, 2 and 3 are within a range of 1 mm to 5 mm.
Unacceptable face steps can be reduced by grinding but a consequence is that the colour of the repair
compared to the concrete on either side may fall outside the tonal range permitted for the specified
Class. An example of the result of grinding is shown in Figure 11.2.
Unintended consequences of grinding the concrete surface can be an increase in the number of
blowholes, which then may fall outside the range permitted for the specified finish Class, or exposed
aggregate, which may be unacceptable if colour control is specified. Little or no effort has been made
to colour-match the mortar used to repair the form tie holes in Figure 11.2.
D
R
Figure 11.2 Face step reduced by grinding
Repair of face steps when carried out properly can result in an acceptable surface finish when the
concrete surface is new and when aged can be almost impossible to detect. The photographs in Figures
AF
11.3 and 11.4 are examples of face step repairs that were considered acceptable for Class 2; i.e.
without colour control.
T

Figure 11.3 Acceptable quality repair of face steps

11.2
Chapter 11 Concrete Finishes Identifications of defects

D
Figure 11.4 Acceptable quality repair of face steps
Face steps of less than 5 mm can be difficult to measure with any degree of accuracy on the job. A
simple means is to use widow packers that have stated thickness of 1.5 mm and 3.2 mm. An example
of measuring a rebate in a concrete surface is shown in Figure 11.5. Here the blue packers used are
R
marked with a thickness of 1.5 mm. Some face steps can be so large that repair to fit within the limits
for Class 3 is difficult if not impossible.
AF
T

Figure 11.5 Measuring face steps

11.3
Chapter 11 Concrete Finishes Identifications of defects

11.4 Honeycombing
Areas of concrete surface that are coarse and stony are described as honeycombing. Honeycomb
defects often are initially blamed on poorly sealed formwork joints. However, insufficient fine
material in the mix or incorrect aggregate grading, as well as poor practices during mixing, placement
and compaction of the concrete can result in surface problems. The photograph in Figure 11.6 shows
honeycombing, which could be the result of either: grout leakage from the formwork joint; or concrete
that has not been adequately mixed when delivered to the formwork and poorly compacted. The
absence of darker concrete around the edges (typical of loss of water or grout) would suggest the
latter. Apart from the surface appearance there has to be concern of the extent of the voids in the
concrete and the detrimental effect on durability regardless of a satisfactory surface repair.
D
R
Figure 11.6 Honeycombing along the bottom edge of a concrete beam
Placing and compaction problems with textured forms can be difficult if not impossible to repair. The
photograph in Figure 11.7 is the result of poor compaction where repairing the surface to replicate the
sawn board finish will be very difficult.
AF
T

Figure 11.7 Poor compaction with board finish

11.4
Chapter 11 Concrete Finishes Identifications of defects

11.5 Debris contanimation


Forms can be easily and economically cleaned before the concrete is placed but it can be difficult
and expensive to repair the concrete surface following removal of the forms. Some typical examples
are shown in Figure 11.8.
D
R
AF
T

Figure 11.8 Debris left when formwork not cleaned


Such cleaning matters can be readily identified before the concrete is placed. However, there is
another possible problem even when the form surface has been cleaned of all debris. This results from
debris such as scraps of tie wire being left on a soffit form for some period of time before being
removed. The photograph in Figure 11.9 shows rust staining of the form surface that has subsequently
imprinted on to the concrete surface. In addition the concrete shows shading resulting from reinforcing
mesh also being left on the form for a period of time. These stain and shading marks are of no

11.5
Chapter 11 Concrete Finishes Identifications of defects

consequence if the soffit is to be covered by a suspended ceiling, plastered or painted; however, if the
soffit is to be left bare (i.e. colour control specified) then such marking can be objectionable. Similar
tie wire stain marks can be seen in the earlier photograph that showed rubbish left in the trough of the
beam form.
D
Figure 11.9 Rust stains left on the forms
11.6 Other common defects
R
The photograph in Figure 11.10 shows damage to a concrete wall surface that became evident after
removal of forms. The damage probably is the result of rain water leaking down between the form face
and the concrete before the concrete had set. If rain is expected steps must be taken to prevent a water
build up on the top exposed surface of a freshly poured wall. Water blasting can remove the unsightly
ridges resulting in a textured finish, assuming that is acceptable for the project.
AF
T

Figure 11.10 Concrete surface damaged after removal of the forms


The photograph in Figure 11.11 shows a concrete surface that has been contaminated after the forms
have been removed. This is the result of poor construction practices at higher levels of the building.
Depending on the type of contamination a repair of the surface can be difficult or it may be impossible
to completely clean the marking resulting in a need to paint the concrete as the only effective remedy.

11.6
Chapter 11 Concrete Finishes Identifications of defects

D
R
Figure 11.11 Concrete surface contaminated after removal of the forms
AF
Other stains can adversely mark a wall surface and repair can present some problems. If the intended
finished surface is to have some form of texture that may not be a problem. The photograph in Figure
11.12 shows unsightly white staining of the textured surface which can be repaired by water blasting,
if that does not change the specified finish.
T

Figure 11.12 Concrete surface stains

11.7
Chapter 11 Concrete Finishes Identifications of defects

Dirty faces of wall forms will result in objectionable appearance of the concrete following removal of
the forms. Problems such as the marking in the photograph in Figure 11.13 can be very time
consuming and costly to effectively to remove. Where a wall is specified as an exposed wall care must
be taken to ensure that the forms are adequately cleaned before being erected.
D
R
Figure 11.13 Dirty faces of wall forms result in objectionable appearance
AF
The final photograph in Figure 11.14 shows the result of inadequate erection procedures, in particular
inadequate sealing between the soffit form and concrete wall, accepting that the stair soffit off form
finish was generally Class 2.
T

Figure 11.14 Poor formwork sealing results in objectionable appearance

11.8
Chapter 11 Concrete Finishes Identifications of defects

References
SA (1995). AS 3610 Supplement 1 - 1995 Formwork for concrete - Blowhole and colour evaluation
charts (Supplement to AS 3610 - 1995). Sydney, Standards Australia
SA (2010). AS 3610 - 2010 Formwork for concrete Part 1: Documentation and surface finish. Sydney,
Standards Australia.
D
R
AF
T

11.9
Appendix A Formwork Importance

A
Appendix A Formwork Importance
A.1 Level of Risk
In structural design, structures or structural elements in different situations need to achieve the
appropriate degree of reliability. Therefore, it is necessary that account should be taken of:
The consequence of failure.
The risk of failure should be lower where the risk of injury, economic, social and environmental
losses is greater.
D
The cause and mode of failure.
The risk of sudden collapse should be lower than where collapse is preceded by some kind of
warning in such way that measures can be taken to limit the consequences.
Factors affecting the risk of failure.
Factors such as: choice of the values of actions; degree of structural integrity; accuracy of
R
structural models used; quality and durability of equipment; site conditions; environmental
conditions; quality of workmanship; and measures taken to reduce the risk of gross human,
design and construction errors.
The expense, level of effort and procedures necessary to reduce the risk of failure.
To this end, the concept of level of risk is similar to the philosophy of structure importance adopted in
AF
AS/NZS 1170.0:2002 Structural design actions Part 0: General principles (SA 2002), is useful.
Table A.1 Level of risk for formwork in different situations

Level of Risk Consequence of failure Situations


Low Formwork, whose Formwork to the sides of shallow footings and slabs.
failure poses a risk to Formwork for the side of small walls and columns, up to 2 m high.
few people and has
Horizontal formwork that supports concrete whose soffit is less than 3 m
small or negligible
above the lowest surrounding ground level and whose plan area is less than
economic, social or
16 m2, providing the concrete to be placed has a volume of not more than 2.5
environmental
m3.
T
consequences
Formwork in areas where access is prevented, such that few people would be
put at risk in the event of its failure.
When construction sites have been vacated*.
Ordinary Formwork not in other All formwork not in other Levels of risk.
Levels of risk
High Formwork, whose Formwork on the perimeter of high-rise buildings in populated areas.
failure poses a risk to Formwork lifted over busy streets.
people in crowds or has
Bridge formwork spanning over major arterial roads.
great economic, social
or environmental Formwork in environmentally sensitive areas.
consequences. Suspended or cantilever formwork, e.g. formwork supported off cantilever
needles, climbform, slipform, jumpform, etc
Formwork that would otherwise have an ordinary level of risk, and whose:
(a) mode of failure is sudden and without warning; or
(b) design, construction or materials are new, novel or unusual.

2013 Stephen Ferguson all rights reserved A.1


Formwork Handbook

A low, ordinary or high level of risk for formwork is comparable with structure importance level 1, 2
or 3 in AS/NZS 1170.0, respectively. The concept of levels of risk based on consequence of failure
provides a useful framework for specifying the annual probability of exceedence of design events and
also the obligation and competence required to verify the formwork design and inspect the formwork
construction, see Sections 2.8 and 4.2.2.
Table A.1 provides guidance on selecting the appropriate levels of risk for formwork in different
situations. Where there is a choice (or doubt) between levels of risk, the highest level applies.
Amendment No 5 to AS/NZS 1170.0 2002 specifies that construction equipment, such as formwork
falls into Structure Importance Level 2 and, accordingly, sets out the relevant annual probabilities of
exceedence for ultimate limit states design. Such a broad approach would unnecessarily penalise
formwork that would otherwise have a low level of risk and possibly tend to underestimation in the
case of the formwork has a high level of risk. Accordingly, the three tier approach presented herein
has merit.
References
SA (2002). AS/NZS 1170.0 - 2002 Structural design actions Part 0: General principles. Sydney,
Standards Australia.
D
R
AF
T

A.2 2013 Stephen Ferguson all rights reserved


Appendix B - Coefficients of Static Friction

B
Coefficients of Static Friction
B.1 Introduction
For reasons of economy and speed of construction, instead of relying on positive connections, the
stability of temporary structures often relies only on friction. Such structures include: demountable
grandstands, platforms, stages and towers; scaffold; falsework; and formwork. In addition, during
construction (albeit for a short period) the stability of permanent structures often depends on friction, e.g.
during the erection of precast concrete buildings.
D
Investigations and surveys into the collapse of temporary structures used during construction identified
the lack of adequate provision for lateral and longitudinal stability as a primary cause of failure (Bragg
1975) and (Hadipriono and Wang 1986). The same studies highlight the danger of using a multiplicity of
unconnected elements on top of each other, relying only on friction for structural integrity.
R
Unfortunately, there is little guidance in the literature on appropriate values for the coefficient of friction
between different materials. The information available differs from source to source and its origin is not
often known. Prompted by the need for reliable data, the UK Health and Safety Executive funded
research at the University of Birmingham to establish practical values of the coefficient of friction for
commonly used materials in temporary works (Pallett, Williamson et al. 2000).
AF
Based on data from 957 tests performed at the University of Birmingham, calibrated limit states design
resistance values for coefficients of static friction for use with a capacity reduction factor = 0.8 to
achieve a target reliability (safety) index of 4.5 are shown in Table B1.

References
Bragg, S. L. (1975). Final report of the Advisory Committee on Falsework. London, Her Majesty's
Stationery Office: 151 pp.
BS (1996). BS 5975:1996 Code of practice for Falsework. London, British Standards Institution.
T
Ferguson, S. A. and R. Q. Bridge (In preparation). "Proposed static friction coefficients for use in the
limit states design of temporary structures."
Hadipriono, F. C. and H.-K. Wang (1986). "Analysis of causes of formwork failures in concrete
structures." Journal of Construction Engineering and Management 112: pp. 112-121.
Pallett, P., S. Williamson, et al. (2000). "Friction resistance in temporary works." Concrete 34(No. 3,
Mar): 15-17.
Pallett, P. F., N. J. S. Gorst, et al. (2002). "Friction Resistance in Temporary Works Materials."
CONCRETE 36(No. 6, Jun): 12-15.

2013 Stephen Ferguson all rights reserved B.1


Formwork and Falsework

Table B1 Nominal design coefficient of static friction for use in limit states design

SURFACE 1
Steel Alum. Timber Plywood Concrete
Soft wood Hard wood Good Film Film
Plain Plain Galv. Prop. Prop. Proprietary Combi Cast
SURFACE 2 one faced faced
Unrusted rusted painted waling Parallel Perp Parallel Perp beam ply faced face
side Finnish quality

Steel
Plain unrusted

Plain rusted

Galvanised

Proprietary painted
D 0.25

0.3

0.2

0.25
0.3

0.35

0.2

0.45
0.2

0.2

0.2

0.25
0.25

0.45

0.25

0.55
0.2

0.35

0.2

0.3
0.25

--

0.3

0.35
0.3

--

0.35

0.35
0.35

0.4

0.35

0.3
0.35

--

0.35

0.4
0.4

0.35

0.3

0.4
0.25

0.25

0.15

0.3
--

0.15

--

0.15
--

0.15

--

0.15
0.05

0.15

0.05

0.15
--

--

--

--

Aluminium

Timber
Proprietary waling

Softwood

Hardwood
Parallel
Perpendicular
Parallel
R
0.2

0.25
0.3
0.35
0.35

--
--
0.4
0.2

0.30
0.35
0.35
0.3

0.35
0.35
0.3
0.25

0.3
0.3
0.3
0.3

0.45
0.4
0.3
0.3

0.4
--
0.3
0.3

0.3
0.3
0.35
0.3

0.35
--
0.35
0.2

0.35
0.3
0.3
0.2

0.2
0.2
0.2
0.2

0.2
0.15
--
0.2

0.2
0.2
--
0.05

0.15
0.1
0.15
--

0.55
0.5
0.4

Plywood
Good one side

Combi ply faced


Perpendicular

Proprietary beam

Film faced Finnish

Film faced quality


0.35

0.4

0.25

--

--

0.05
--

0.35

0.25

0.15

0.15

0.15
--

--
AF
0.35

0.3

0.15

0.05
0.4

0.4

0.3

0.15

0.15

0.15
0.3

0.2

0.2

0.2

0.2

0.05
0.35

0.35

0.2

0.2

0.2

0.15
--

0.3

0.2

0.15

0.2

0.1
0.35

0.3

0.2

--

--

0.15
--

0.3

0.3

--

--

0.15
0.3

0.35

0.2

--

--

0.15
0.25

0.2

0.3

0.2

0.2

0.15
--

--

0.2

--

--

--
--

--

0.2

--

--

--
0.15

0.15

0.15

--

--

0.1
0.5

--

0.25

0.2

0.2

0.15

Cast face -- -- -- -- -- 0.55 0.5 0.4 0.5 -- 0.25 0.2 0.2 0.15 --
Hardened
Concrete

Soil

B.2
Trowelled face

Granular
0.4

--
0.55

--
0.2

--
0.45

--
0.35

--
0.75

--
T
0.55

--
0.5

--
0.5

--
0.45

--
0.25

--
--

--
--

--

2013 Stephen Ferguson all rights reserved


--

--
--

--
Appendix C Recommended Reading

C
Appendix C Recommended Reading
C.1 Introduction
To obtain a comprehensive understanding of formwork design, in addition to the references cited at
the end of each Chapter, the following references are recommended reading (sorted geographically,
alphabetically and chronologically).
References
American
D
1. Hurd, M. K. (1995). Formwork for Concrete. Farmington Hills, American Concrete Institute.
Australian
2. CCAA (2006). Guide to Off-form Concrete Finishes. Sydney, Cement and Concrete
Association of Australia.
R
3. McAdam, P. S. and G. Lee (1997). Formwork a practical approach. London, E & EF Spon.
4. EWPAA (1993). Plywood in Concrete Formwork Manual. Brisbane, Engineered Wood
Products Association of Australasia.
European
AF
5. BS 5975:2008 Code of practice for temporary works procedures and the permissible stress
design of falsework. London, British Standards Institution.
6. CS (1999). Checklist for Erecting and Dismantling Falsework. Berkshire, The Concrete Society.
7. CS (2003). Checklist for Assembly, Use and Striking of Formwork. Berkshire, The Concrete
Society.
8. CS (2012). Formwork A guide to good practice. Berkshire, The Concrete Society.
9. CSG (2003). Guide to Flat Slab Formwork and Falsework. Berkshire, The Concrete Society on
behalf of Concrete Structures Group
T
10. DIN (2010). DIN 18218 Pressure of fresh concrete on vertical formwork. Berlin, German
Standards.
11. ECS (2004). BS EN 12812 Falsework - Performance requirements and general design.
Brussells, European Committee for Standardization.

2013 Stephen Ferguson all rights reserved C.1

Potrebbero piacerti anche