You are on page 1of 144

Technical Report

AP-T313-16

Realising 100-year Bridge Design Life


in an Aggressive Environment:
Review of the Literature
Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Prepared by Publisher

Ahmad Shayan and Aimin Xu Austroads Ltd.


Level 9, 287 Elizabeth Street
Sydney NSW 2000 Australia
Project Manager Phone: +61 2 8265 3300
austroads@austroads.com.au
Ross Pritchard
www.austroads.com.au

Abstract About Austroads


This report details a literature review on durability issues that affect Austroads is the peak organisation of Australasian road
the service life of reinforced concrete structures. transport and traffic agencies.
The review focused on information from sources within Australia and Austroads’ purpose is to support our member organisations to
overseas, including published literature, standards, codes and deliver an improved Australasian road transport network. To
jurisdiction reports. succeed in this task, we undertake leading-edge road and
transport research which underpins our input to policy
Recent amendments to test methods, specifications and codes, development and published guidance on the design,
particularly Section 5.4 (Concrete Durability) of the Australian construction and management of the road network and its
Standard AS 5100.5 (Bridge Design), are also discussed. associated infrastructure.
Austroads provides a collective approach that delivers value
Keywords for money, encourages shared knowledge and drives
consistency for road users.
Reinforced concrete structures; concrete durability; service life
prediction; corrosion of reinforcement; chloride diffusion; stainless Austroads is governed by a Board consisting of senior
steel; AAR; sulfate resistance; durability planning. executive representatives from each of its eleven member
organisations:
 Roads and Maritime Services New South Wales
 Roads Corporation Victoria
ISBN 978-1-925451-40-5  Queensland Department of Transport and Main Roads
Austroads Project No. TS1834  Main Roads Western Australia
 Department of Planning, Transport and Infrastructure
Austroads Publication No. AP-T313-16
South Australia
Publication date November 2016  Department of State Growth Tasmania
Pages 137  Department of Infrastructure, Planning and Logistics
Northern Territory
 Transport Canberra and City Services Directorate,
Australian Capital Territory
© Austroads 2016  Australian Government Department of Infrastructure and
Regional Development
This work is copyright. Apart from any use as permitted under the
Copyright Act 1968, no part may be reproduced by any process  Australian Local Government Association
without the prior written permission of Austroads.  New Zealand Transport Agency.

This report has been prepared for Austroads as part of its work to promote improved Australian and New Zealand transport outcomes by
providing expert technical input on road and road transport issues.
Individual road agencies will determine their response to this report following consideration of their legislative or administrative
arrangements, available funding, as well as local circumstances and priorities.

Austroads believes this publication to be correct at the time of printing and does not accept responsibility for any consequences arising from
the use of information herein. Readers should rely on their own skill and judgement to apply information to particular issues.
Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Summary

This report comprises a detailed literature review on various durability issues that affect the service life of
reinforced concrete structures. It is mainly concerned with collection of information from sources within
Australia and overseas, including published literature, standards, codes and jurisdiction reports. Recent
amendments to test methods, specifications and codes, particularly Section 5.4 (Concrete Durability) of
Australian Standard AS 5100.5-2004 (Bridge Design: Concrete), are also discussed.

The main concerns over durability performance are related to the fact that reinforced concrete bridges in
aggressive environments are designed for a 100-year service life, but they start to deteriorate after only 30
years. This is partly due to inadequate standards and specifications, and partly due to lack of knowledge
regarding the durability issues involved.

Road agencies engage durability specialists to develop detailed durability plans for specific major bridge
structures that are expected to last over 100 years. Examples are provided of such bridges constructed in
some jurisdictions, and the issues that were addressed in each case. A common feature of the
specifications, durability plans, AS 5100.5-2004, and international literature is the use of blended cements for
combating various deterioration problems and achieving long service lives for reinforced concrete structures.
Potentially dangerous deterioration mechanisms, such as alkali-aggregate reaction (AAR) can be avoided.
Existing information shows that realising service lives of beyond 100 years in aggressive environments
requires some of the following actions, where possible:
 an increase in concrete cover thickness
 use of appropriate protective surface coatings
 isolation of concrete from the aggressive environment by appropriate encasement
 installation of cathodic protection or cathodic prevention
 use of corrosion resistant materials, such as appropriate varieties of stainless steel
 use of high performance concrete mixes
 neutralisation of acidic soils
 careful supervision of all aspects of the design and construction phases, which is of paramount
importance.

With respect to delayed ettringite formation (DEF), most specifications and AS 5100.5-2004 (Section 5.4)
allow sulfate contents in the cement of 5% SO3 by mass of cement, which is considered to be too high, and
is suggested to be reduced to a maximum of 3.5%.

However, gaps appear to exist in the available knowledge, which should be addressed through future
research efforts. This literature review has identified the following issues, some of which have been
addressed in this project:
 uncertainties in terms of accuracy of the carbonation rate and chloride diffusion coefficient of concrete,
which need to be addressed through further research
 uncertainties in models for the prediction of time to the appearance of detectable damage and time of
failure
 the long-term tolerance of stainless steel reinforcing materials to high chloride contents in concrete,
particularly that of new low-Ni varieties, which is not well-established and requires further research.

Appropriate concrete ingredients and steel reinforcement materials and concrete mix designs, as well as
testing parameters, have been selected for experimental work to address the concerns mentioned above. A
separate report (Austroads 2016) has been prepared on the results of the experimental part of the project.

Austroads 2016 | page i


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Contents
Summary ......................................................................................................................................................... i
1. Introduction ............................................................................................................................................ 1
1.1 Issues ............................................................................................................................................... 1
1.2 Information Collected for Review ..................................................................................................... 2
2. Durability Issues .................................................................................................................................... 4
2.1 Amendment of AS 5100.5-2004 – Clause 4..................................................................................... 5
2.2 Durability Planning: Main Issues ...................................................................................................... 7
2.3 Protective Measures to Increase Service Life ................................................................................ 13
2.4 Current Practice for Long Service Life Prediction .......................................................................... 16
2.5 In-service Maintenance Phase ....................................................................................................... 17
2.6 Improvement of Current Practice ................................................................................................... 17
3. Definition of Service Life ..................................................................................................................... 19
4. Performance of Reinforced Concrete Materials ............................................................................... 22
4.1 Reinforcing Steel Corrosion ........................................................................................................... 22
4.2 Mitigation of Steel Corrosion .......................................................................................................... 25
4.3 Alternatives to Conventional Reinforcing Steel .............................................................................. 27
4.4 Defects Formed in Concrete at Early Ages .................................................................................... 32
4.5 Long-term Durability Problems of Concrete ................................................................................... 37
4.6 Surface Coating Treatments for Attack Mitigation ......................................................................... 61
5. Prediction of Service Life .................................................................................................................... 64
5.1 Time to Onset of Potential Failure .................................................................................................. 64
5.2 Time to Detectable Failure Due to Steel Corrosion ....................................................................... 73
5.3 Time to Functional Failure Due to Steel Corrosion ........................................................................ 77
5.4 Probability of Failure ....................................................................................................................... 78
6. Tests for The Evaluation of Durability ............................................................................................... 86
6.1 Resistance of Concrete to Chloride Penetration ............................................................................ 86
6.2 Other Transport Properties of Concrete ......................................................................................... 88
6.3 Tests for Preventable Potential Durability Problems ...................................................................... 91
6.4 Corrosion of Stainless Steels ......................................................................................................... 92
7. Summary and Conclusions ................................................................................................................. 93
7.1 Materials for Experimental Work (Austroads 2016) ....................................................................... 94
References ................................................................................................................................................... 96
Appendix A Comparison of Requirements for Concrete Mix Design and Durability by
Four Jurisdictions ......................................................................................................... 123
Appendix B Threshold Values for Steel Corrosion ......................................................................... 130
Appendix C Exposure Conditions of Gateway Bridge Site ............................................................ 133
Appendix D PSTR Requirements for Design Life............................................................................ 134
Appendix E Calculation of Accumulated Corrosion ....................................................................... 135
Appendix F Minimum Design Life of Bridges and Bridge Sub-Assets (as per Roads and
Maritime (2012a) – Gerrinong Upgrade) ...................................................................... 137

Austroads 2016 | page ii


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Tables
Table 1.1: List of documents collected from jurisdictions ............................................................................. 3
Table 2.1: Requirements for concrete structures in aggressive exposure conditions ................................. 8
Table 2.2: Issues addressed in structure durability plans for 100 years or longer service ........................ 12
Table 2.3: The measures for service life in AS 5100.5–Section 4 (draft 2012) or as otherwise
specified .................................................................................................................................... 14
Table 4.1: The maximum acid-soluble chloride ion content in new concrete ............................................ 39
Table 4.2: Influence of humidity on carbonation equation parameters and corrosion rate (CR) ............... 42
Table 4.3: Swedish classification of natural waters containing carbon dioxide ......................................... 43
Table 4.4: List of Australian reactive aggregates ....................................................................................... 45
Table 4.5: Reported cases of AAR-affected structures in Australia up to 2003 ......................................... 46
Table 4.6: Requirements to protect against damage to concrete by sulfate attack from external
sources of sulfate ...................................................................................................................... 57
Table 4.7: Recommendation of BS 8110-1997ǂ for concrete exposed to sulfates .................................... 57
Table 4.8: Sulfate aggressiveness to concrete .......................................................................................... 58
Table 4.9: Recommendations for sulfate resistance .................................................................................. 58
Table 4.10: Requirements for well-compacted cast in situ concrete, 140 mm to 450 mm in thickness,
exposed on all vertical faces to a permeable sulfate soil or fill ................................................. 59
Table 4.11: Types of cement ........................................................................................................................ 59
Table 4.12: The durability concerns of reinforced concrete ......................................................................... 63
Table 5.1: Relationship between safety index and probability of failure .................................................... 81
Table 5.2: Time or temperature-dependent variables used in current service life prediction models ....... 84
Table 5.3: Typical values for mean and coefficient of variation for different parameters .......................... 85
Table 6.1: Test voltage and duration (NT Build 492) ................................................................................. 87
Table 6.2: VicRoads requirements for VPV for various grades of structural concrete .............................. 89
Table 6.3: Sorptivity tests ........................................................................................................................... 90
Table 6.4: Durability requirements for concrete ......................................................................................... 90
Table 6.5: Test methods for assessment of AAR ...................................................................................... 92
Table 7.1: Various concrete mix designs used for various bridge structures ............................................ 94

Figures
Figure 2.1: Durability design process flow chart ......................................................................................... 10
Figure 3.1: Schematic expression of structure service life .......................................................................... 19
Figure 4.1: The E–pH (Pourbaix) diagram of iron, according to chemical equilibriums of species
in an Fe-H2O system ................................................................................................................. 23
Figure 4.2: Volumes of iron and its reaction products ................................................................................. 24
Figure 4.3: Acceptable chloride content according to CEB recommendations ........................................... 38
Figure 4.5: Map-cracking caused by AAR in a large bridge pylon (left) and in a support column of a
water diversion gate in a dam (right) ......................................................................................... 48
Figure 4.6: Parallel cracking caused by AAR in a prestressed railway sleeper .......................................... 48
Figure 4.7: AAR manifested by reaction products around aggregate particles........................................... 49
Figure 4.8: AMBT results for reactive dacite aggregate (top left), non-reactive basalt (top right) and
two slowly reactive gneissic granite aggregates (bottom) by the two test methods indicated .. 52
Figure 4.9: Comparison of AMBT and CPT results for a large collection of aggregates from various
projects ...................................................................................................................................... 53
Figure 5.1: The water viscosity at different temperatures and diffusion coefficient relative to that
at 20 °C ..................................................................................................................................... 67
Figure 5.2: Carbonation degree and the depth measured by phenolphthalein pH indicator ...................... 70
Figure 5.3: Example of relationship between carbonation depth and concrete strength for deck slabs,
beams and piers of two bridges constructed in 1940 and 1976 ............................................... 72
Figure 5.4: Predicted low, median and high estimates of CO2 concentration ............................................. 73
Figure 5.5: Dependence of critical corrosion depth on cover depth and bar diameters ............................. 74
Figure 5.6: Time to crack formation in cover concrete due to carbon steel corrosion; bar size 24
mm, concrete grade 50 MPa, Ccrit = 0.4% of cement mass ...................................................... 76

Austroads 2016 | page iii


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 5.7: Predicted median temperatures and low and high estimates for Sydney, using CSIRO
Mk3.5 climate model for the A1FI emission scenario ............................................................... 76
Figure 5.8: Left: Initial corrosion rate of steel in seawater; Right: Arrhenius plot for the carbon
steel data ................................................................................................................................... 77
Figure 5.9: The principles of a type-dependent reliability analysis ............................................................. 82

Austroads 2016 | page iv


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

1. Introduction
Austroads project TS1834 aims to develop knowledge, based on existing and new experimental data, to
enable specification of materials and practices for achieving greater than 100 years of service life for major
bridge structures in aggressive environments. This report provides a comprehensive review of information
from within Australia and overseas, including published literature and jurisdiction reports. This review was
used as an input to the experimental work program which will produce comparative data on the performance
of materials and methodologies relevant to the long-term performance of bridge structures.

1.1 Issues
Concrete structures can suffer from several types of deterioration, most of which can be resolved and
mitigated against at the design stage. Corrosion of reinforcement in concrete is a long-term, gradual process
and arises due to exposure to aggressive environments, and is the major cause of premature deterioration of
structures in coastal regions or saline soils. Currently, many coastal bridge structures suffer from corrosion
damage at the age of around 30 years, which is far too short, and results in large maintenance and
rehabilitation costs. This has arisen mainly because the level of knowledge on the performance of concrete
materials in aggressive environments was insufficient at the time of construction. For example, Carse and
Bell (2004) stated that only 1% of Queensland bridges would reach a service life of 100 years, i.e. before
major repairs are needed, and 50% would reach 50 years. Bentour et al. (1997) called steel corrosion the
‘bottleneck’ in the durability performance of concrete structures in aggressive environments, as it occurs
before the concrete itself is affected by problems such as sulfate attack.

The specified service life of bridges at present is 100 years (Australian Standard AS 5100), but for very
important structures even this may not be adequate, as shortcomings in the available knowledge and
deficiencies introduced through poor workmanship can result in premature damage. For example, the New
Gateway Bridge in Brisbane, for which a durability plan was developed in 2007 (opened in May 2010), has a
nominal design life of 300 years, but the technical information available then may be insufficient to ensure
such a long service life is achieved. The information which formed the basis of the relevant durability plan
was not drawn from such long-term field data. Instead, models which extrapolate well beyond the age of
available shorter-term data were used to produce durability design information.

It is important to note that bridge elements such as piles, pile caps, columns, etc. are very expensive to
replace (and sometimes impractical) once they suffer major deterioration. Therefore, ensuring their long-term
durability beyond 100 years is crucial.

Obviously, longer-term performance information needs to be developed, where possible, for conventional
and alternative materials, such as stainless steel, high performance concrete and corrosion inhibitors, to
enable service lives well in excess of 100 years to be accurately predicted. This is often not feasible and still
some estimation needs to be made.

Austroads 2016 | page 1


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

This project will:


 Discuss all the relevant materials-related issues and processes that can lead to concrete deterioration,
using the available literature.
 identify important issues related to materials and exposure conditions, which arise when the service life is
extended beyond 100 years, so that they can be included in the durability plan for structures in order to
ensure that 100 years would be achieved
 develop some comparative data on performance of selected corrosion resistant reinforcement materials.
This will include appropriate laboratory investigations in future years of this project
 consider and incorporate definite advances in knowledge due to:
– code amendments, i.e. AS 5100.5-2004, with respect to concrete durability issues
– publication and interim findings of Austroads and other relevant international reports
– industry best-practice, i.e. innovations in concrete mix design
 incorporate jurisdictional supplementary information produced since publication or last amendment of the
Guide, and any obvious gaps or shortcomings in this part of the Guide.

1.2 Information Collected for Review


A literature survey was conducted through the ARRB Library to collect relevant information in relation to
various aspects of durability of concrete structures and service life prediction. Google searches were also
conducted for obtaining published information on specific issues as needed. Documents selected for this
review are included in the list of references.

Australian Standard AS 5100.5-2004, Part 5, covers all concrete-related requirements for the different
exposure classifications. Design parameters for achieving long-term durability are covered in Section 4 of the
document. Prescriptive requirements are listed for the various exposure classes in terms of minimum cement
content, minimum strength, maximum water/cement ratio, minimum permeability, and minimum cover
thickness. It should be noted that AS 5100.5-2004 is being amended, and some of the amendments included
in the draft version of the standard are noted in this report.

In addition to the ‘Concrete’ section of the AS 5100.5-2004, road agencies (jurisdictions) were requested to
provide information on their specification for structural concrete. Moreover, where possible, ‘durability plans’
developed for specific structures were obtained from jurisdictions. The latter are considered important as
they reflect the current state of practice in ensuring 100 years of service life for bridges.

At present, durability plans are demanded by the asset owners before the award of construction contracts for
major structures. These plans are necessarily specific to each structure, as unique exposure conditions may
exist in one site and not the others, although many features are often common to most structures.

The plans first define precise exposure classifications for all components of the structure concerned, and
sometimes for different parts of the same element, such as buried, submerged, tidal and atmospheric zones
of piles or columns; the latter even being subdivided into sheltered and exposed areas. Possible
deterioration problems relevant to various structural components, under the given exposure conditions, are
then considered and analysed, and mitigation measures for avoiding each deterioration problem are put in
place to ensure achievement of long-term durability, or the required service life, specified. The list of
documents collected from jurisdictions is given in Table 1.1.

Austroads 2016 | page 2


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 1.1: List of documents collected from jurisdictions

Road agency Contact Specification Durability plan Other documents


Queensland Dr Ross Pritchard MRS70 & New Gateway Bridge Draft of amended AS
Department of MRTS70 5100.5
Transport and Main
Roads (TMR)
Roads and Maritime Dr Radhe Khatri B80 Ballina Bypass Published papers
Services (Roads and Kempsey Bypass
Maritime), NSW Tarcutta Bypass
Hunter Expressway
Banora Point Upgrade
Camden Valley Way
Upgrade
Desalination Plant- Kurnell
Princes Highway Gerringong
Upgrade
VicRoads Mr. Fred Andrews- Section 610 No plan available Published papers &
Phaedonos bridge investigation
report
Department of State Mr. Geoff Mulcahy B10 Sorrell Bridge and Brighton N/A
Growth(1), Tasmania Bypass
Main Roads Western Mr. Didar Cheema 820 No plan available N/A
Australia (MRWA) and Internet
Department of Internet Part 320 N/A N/A
Planning, Transport
and Infrastructure
(DPTI), South
Australia
1 Previously Department of Infrastructure, Energy and Resources (DIER).

Austroads 2016 | page 3


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

2. Durability Issues
Durability of reinforced concrete structures depends on the performance of the hardened concrete, including
the reinforcement material, under the prevailing exposure conditions. Unintentional defects may be built in
the concrete by unsound ingredients, inappropriate concrete mix design, insufficient compaction during
placement, and inadequate early age curing. Even though the selected materials may be sound, poorly
constructed concrete elements would probably deteriorate prematurely as a result of exposure of the
defective elements to aggressive environments (e.g. small cover thickness or poorly compacted concrete).

In reinforced concrete structures, the cover concrete provides protection for the steel bars against the
adverse effects of aggressive agents present in the exposure environment. Therefore, high quality concrete
and sufficient cover thickness would prevent environmental attack on the steel.

Unsound ingredients in concrete or inappropriate combination of them, or their interaction with the
environment, can cause deterioration through various mechanisms, as described later in the report.

To ensure the design life of a mechanically sound structure is achieved, it is essential to use concrete of high
quality that is capable of resisting internal deterioration processes as well as environmental attacks, taking
into account all the mechanisms which may become operative during the service life.

Given proper design of the structure and appropriate concrete formulation, the most important issue during
the construction phase is the execution, which will determine the future performance. As pointed out by
Geiker et al. (1997):
The onsite workmanship, often personalised to the individual worker, determines the
quality of concrete pouring, compaction and curing.

Quality control testing and inspection is essential in ensuring good performance of the constructed facility in
the future. Special care and supervision should be provided for the weak points of the structure, such as
construction joints and rebar-formwork spacers (including type of spacers).

Work plans for casting concrete should be made with consideration of the weather condition, particularly
wind and temperature. Provision should be made for temporary protection of elements from early stage
exposure to the aggressive environment, as they would be quite permeable at very early age.

Provision of mitigation methods for possible deterioration mechanisms and programs of continued
inspection/monitoring and maintenance are very important in extending the service life of concrete
structures, and should be incorporated in the durability plan. Control of corrosion is particularly important
because, in addition to impacting on the maintenance costs, corrosion products can disrupt the bond
between the steel and concrete and result in the loss of function of the steel, which may lead to structural
failure.

Implementation of the requirements of appropriate standards and those of specifically developed ‘durability
plans’ are essential for achieving long service lives.

Austroads 2016 | page 4


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

2.1 Amendment of AS 5100.5-2004 – Clause 4


AS 5100.5-2004, Section 4, ‘Design for Durability’ now clearly points out that:
More stringent requirements would be appropriate for structures with a design life in
excess of 100 years.

Particularly, the durability performance of concrete has been addressed in the amendments by inclusion of
statements such as:
In addition to the minimum requirements of characteristic compressive strength, minimum
cementitious material content and maximum water/cementitious material (binder) ratio
(W/C or W/B), testing for penetrability and absorption of concrete, in accordance with
specified test methods such as VPV (volume of permeable voids: AS 1012.21), RCPT
(Rapid Chloride Permeability Test: ASTM C1202), and sorptivity may be required to
further enhance the durability performance of concrete.

However, the tests and acceptance criteria are yet to be formally specified.

The thickness of the concrete cover, which protects the steel from corrosion, has been emphasised and
various minimum cover depths are specified. The continuity of reinforcement (carbon steel) for the potential
application of cathodic protection in the future is specially addressed. Although not mentioned, this would
also facilitate the in situ investigation of reinforcement corrosion when needed.

Changes are being considered in the definitions of the exposure conditions and ranges of the chemical
compounds in aggressive environments (chloride and sulfate) for each exposure class. For example, the
chloride content for Class B1 has increased from 300 ppm to 2000 ppm, and for Class B2, the range has
changed from 300–2000 ppm to 2000–8000 ppm, which are less conservative. Other main additions and/or
changes are:
 Definitions of exposure for Class C1, the spray zone (for members not containing steel reinforcement),
and Class C2, the tidal zone/splash zone (for reinforced concrete):
– Spray zone is the zone up to 10 m above mean sea level and extending 30 m beyond the high water
mark. Surfaces of structures that are exposed to the spray zone and are not washed by rain are
deemed to be in the spray accumulation zone.
– The tidal/splash zone is the zone from the mean low water springs (MLWS) up to 1 m above the wave
crest level or 1 m above the highest astronomical tide (HAT), whichever is the highest.
 Proportions of supplementary cementing materials (i.e. fly ash; slag; silica fume; and triple blends, i.e. the
combination of the three) to replace general purpose Portland cement (GP cement) in concrete are
specified (AS 5100.5-2004, Table 4.5). However, the formulations for the triple blends are not clear and
need to be explained as to whether the proportions of the additives are based on the total binder content
or the total of the supplementary cementitious materials (SCMs).
 The nominal cover for concrete members reduced by 5 mm for all exposure classes and strength grades
(AS 5100.5-2004, Table 4.13.3.2). Note that the minimum cover is obtained by reducing the nominal
cover values by the applicable tolerances.
 Small sections on self-compacting concrete and delayed ettringite formation (DEF) have been added to
the amended version.

Austroads 2016 | page 5


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

 Changes in acceptance criteria and definitions:


– In addition to strength grades for the various exposure Classes A, B1, B2, C1 and C2, the
corresponding minimum cementitious material content, maximum water to cementitious material ratio
and minimum curing times are specified (AS 5100.5-2004, Table 4.4).
– The minimum compressive strength values at the completion of accelerated curing for Class B2 and
Class C (now C1 and C2) are increased by 1–5 MPa, depending on the exposure class (AS 5100.5-
2004, Table 4.4).
– The least dimension of externally restrained elements requiring control of early age thermal cracking
has increased from 500 mm to 1000 mm. The temperature differential across the concrete member is
not to exceed 20 °C.
 Changes in threshold level of classes:
– The minimum content of sulfates and chloride in soil and groundwater was reduced by half to one-fifth
for the different exposure classes (AS 5100.5-2004, Table 4.8).
– The presence of magnesium ions in the soil in excess of 1000 ppm is considered to increase the
exposure condition by one class.
– The maximum acid soluble chloride content of concrete as cast (AS 5100.5-2004, Table 4.12) is
specified as 2.0 kg/m3 for concrete not requiring protection and cast at ambient environment; 0.6 kg/m3
for reinforced concrete cured at ambient temperature or by heat curing; 0.4 kg/m3 for prestressed
concrete cured at ambient temperature or by heat curing. Note that some jurisdictions specify 0.4
kg/m3 for reinforced concrete.
– The maximum acid soluble sulfate content of concrete as cast (AS 5100.5-2004, Table 4.12) is
specified as 5% by mass of cementitious materials for all types of concrete cured at ambient
temperature, and 4% for heat-cured reinforced or prestressed concrete. Note that these values may
cause the problem of DEF in large cement-rich concrete elements, which may develop hydration
temperatures in excess of 80 °C, particularly if the aggregate is prone to alkali-aggregate reaction
(Shayan et al. 2006).

Several prevention measures against potential deterioration problems are specified, as follows:
 Protecting concrete against chloride contamination in a marine environment has been specially
addressed by requiring that ‘all reinforcement and embedded metallic fixtures shall be protected against
chloride contamination on site, and cleaned off prior to concrete placement’.
 Testing aggregates for the potential risk of alkali-aggregate reaction (AAR) by the accelerated mortar bar
test (RMS T363 or VicRoads RC 376.03) and concrete prisms test (RMS T364 or VicRoads RC 376.04),
and limiting the total alkali content to below 2.8 kg/m 3 (Na2O equivalent).
 Mitigation against delayed ettringite formation by limiting the heat curing temperature to 70 °C for
concrete made with GP cement and 80 °C for concrete made with blended cement. Note that this does
not cover the case of large cement-rich elements mentioned above.
 Use of supplementary cementitious materials (SCM) in concrete to reduce the risk of AAR (AS 5100.5-
2004, Table 4.10), which would also reduce the risk of DEF. However, it should be noted that the levels of
SCMs as percentages of cement replacement specified in the previous version (20–30% for fly ash; 50–
70% for slag; 8–10% for amorphous silica (or silica fume)) were more appropriate than the amended
values of 20%, 50% and 8%, respectively. Experience with some Australian reactive aggregates shows
that they do not adequately respond to 20% or 25% fly ash and require 30% for suppression of AAR
expansion, and it is known that 60–65% of slag is needed to suppress AAR-induced expansion of very
reactive aggregates.

Austroads 2016 | page 6


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

2.2 Durability Planning: Main Issues


A few decades ago, the quality of concrete used to be controlled largely by specifying a given strength grade
and water/cement (W/C) ratio. As pointed out by Gjørv (2009), the simple old water to cement ratio and
water to binder ratio for characterising and specifying concrete quality have lost their meaning, owing to the
fact that modern concretes contain different cementitious materials, reactive fillers and admixtures, or are
even made with various types of processed concrete aggregates. Permeability and transport properties of
concrete have assumed a much more significant role, as most deterioration problems of reinforced concrete
structures are caused by ingress of aggressive agents such as gases (oxygen, CO2 and SO2), moisture, and
soluble salts (chlorides and sulfate ions). Therefore, the rate of chloride ingress, carbonation rate, and water
sorption rate of concrete have a major influence in the deterioration process.

Long-term durability of major concrete bridges and other significant concrete structures cannot be achieved
by addressing single issues in an ad hoc manner. For this purpose, a comprehensive analysis of all the
durability issues is required, taking into account the environmental conditions of the structure concerned and
its required service life. The durability planning process should utilise a variety of Australian and international
standards, knowledge available through relevant published work, physical site inspections, investigations
and testing; and involves interaction between the durability experts, design team and the asset owner. This
process starts in the concept design phase and continues into the site investigations (including air, water and
soil analyses) and construction phases of the project, as well as the post-construction maintenance phase.
Furman (2008) presented a simplified example of the methodology needed for durability planning.

The Concrete Institute of Australia has produced a series of Recommended Practice documents in the
Concrete Durability Series, which includes a detailed document on performance criteria for concrete in
marine environments, titled Z13 (Concrete Institute of Australia (CIA) 2001), and another document on
durability planning, titled Z7/01 (CIA 2014a). It details the requirements for various stages of design,
construction and maintenance of a planned structure in a given exposure environment. This document
provides useful information and should be consulted before developing a particular durability plan.

Another document (CIA 2015) contains a large number of durability performance tests, but informed
technical decisions would be required in order to make an appropriate selection amongst these test method
for incorporation in the durability planning of an individual case.
Road agencies in Australia have established requirements for concrete quality and durability in order to
achieve the design life of their bridges, examples of which are given in Table 2.1. As mentioned in the
previous section, exposure conditions influence the durability of concrete, and the exposure classes
specified in AS 5100.5-2004 have been adopted by some road agencies. The specifications used by most
jurisdictions include requirements for concrete strength, especially the cementitious materials content and
water to binder ratios, as well as other properties like the volume of permeable voids (VPV), specified by
VicRoads, and sorptivity (Roads and Maritime Services 2012a, 2012b). Some agencies also specify the
maximum tolerable crack width for different exposure conditions, especially for B1, B2, C1 and C2 exposure
classes. Note that the jurisdictions have not yet changed the exposure Class C to C1 and C2, hence only
Class C is given in Table 2.1. In addition to the latter, a comparison has been presented between the
approaches used by four jurisdictions in Appendix A.

Austroads 2016 | page 7


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 2.1: Requirements for concrete structures in aggressive exposure conditions

Exposure (reference) Cover Concrete requirement Use of SCM Test for durability
(AS 5100.5-2004) 45 mm Standard formwork and Different AAR tests to Roads
B2 — compaction proportions and Maritime or
U 70 mm f’c = 32 MPa recommended for VicRoads methods
f’c = 40 MPa exposure Classes May include VPV
C1 and C2 B2, C1, and C2 as and sorptivity tests
f’c  50 MPa compared to those
for B1 and A
TMR (2010), Tolerance Binder  390 kg/m3, w/b  0.46 At least 20% FA AAR test to Q458
Queensland + 10 mm, f’c = 40 MPa Mixes with 20% fly
MRTS70 –5 mm Binder  450 kg/m3, w/b  0.40 ash are deemed to
B2 for all comply with AAR
f’c = 50 MPa
C exposures requirements.
Chloride diffusion
VicRoads Specification Tolerance Binder = total cementitious 60 GP + 40 Slag Comply with: VPV
Section 610 (VicRoads + 5 mm, material 75 GP + 25 FA test limits
2010a) –0 mm Binder  400 kg/m3; w/b  0.45 90 GP + 10 SF AAR expansion
B2: deck slab for all f’c = 40 MPa 70 GP + 30 FA RC376.03 < 0.10%
C: Piles exposures at 21 days
Binder  450 kg/m3; w/b  0.40 35 GP + 65 Slag or
C: Pile caps, pier Strength gain over
f’c = 50 MPa 30 GP+60 Slag+10
columns SF 3–28 days
Binder  470 kg/m3; w/b  0.36
C: Beams, crown units 90 GP + 10 SF or
f’c = 55 MPa
80 GP + 20 FA
Binder  470 kg/m3; w/b  0.36
f’c = 55 MPa
Roads and Maritime Cement  370 kg/m3; w/c  0.46 Use SCM Sorptivity (RMS
(2012b), NSW, Cement  420 kg/m3; w/c  0.40 T362)
Specification B80 17 mm for GP
B2 8 mm for GB
C concrete
AAR exp. (T363)
< 0.10% at 21 days
Main Roads Western Min. slump 100 mm all concrete 32% GP + 60% Comply with:
Australia (MRWA) Binder  420 kg/m3; GGBFS + 8% SF AAR test limits
(2013), W/C  0.40 GP cement only (WA624.1) < 0.10%
Specification 820: Cement  350–420 kg/m3; Shrinkage < 600 µs
Marine Grade S50M W/C  0.45–0.40 Cl content
All other grades  0.40 kg/m3
Cement SO3  5%
alkali < 2.8 kg/m3
Total soluble salts
in water  1500
ppm
Department of Classes of concrete 20–50 Cement SO3  5%
Planning, Transport Min. cement content Cl  0.40 kg/m3
and Infrastructure 240–460 kg/m3 RCPT < 1500
(DPTI) (2011), SA Max. W/C ratio 0.70–0.40 coulombs
Specification Part 320
AAR exp. < 0.10%
Department of Classes of concrete S15 to S50 60% GP + 40% Comply with VPV
Infrastructure, Energy Min. binder content GGBS test limits
and Resources (DIER) 200–500 kg/m3 75% GP + 25% FA AAR test limit
(2006)(1) Tasmania Max. W/C ratio 0.9–0.36 Silica fume can be Cl  0.40 kg/m3
Specification B10 added Cl in water
< 300 ppm

Austroads 2016 | page 8


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Exposure (reference) Cover Concrete requirement Use of SCM Test for durability
FDOT, USA > 100 mm Cement  444 kg/m3, 18% to 22% FA; Dapp < 0.3
highly aggressive W/C < 0.41; Addition of 22 L/m3 8% SF for driven 10 12 m2/s
(Sagüés & Powers CaNO3 if cover <100 mm piles
1997)
Norwegian code > 100 mm,
NS3473 nominal
Severe exposure, 120 mm
underwater and splash > 60 mm,
zone nominal
above splash zone, up 75 mm
to 6 m in mild marine
exposure
(Kompen 1997)
NPRA, 1996 > 60 mm W/C  0.38
Marine environment  15 mm
(Gjørv 2009)
European code BS EN — W/C  0.45
206: 2013

1 Now Department of State Growth.

Notes:

FDOT = Florida Department of Transportation; NPRA = Norwegian Public Roads Administration; SCM = supplementary
cementitious materials, including Class F fly ash (FA), ground granulated blast iron furnace slag (Slag), silica fume (SF),
etc.; Dapp = apparent diffusion coefficient of chloride ions in concrete.

The SCMs used in concrete shall conform to the relevant Australian Standards listed below, and special
requirements that may be imposed by the road agencies based on local conditions:
 Fly ash – AS/NZS 3582.1-2016; Special requirement by MRTS70: maximum total alkali content 2%, and
maximum available alkali content of 0.5% (Na2O equivalent).
 Slag – AS/NZS 3582.2-2016; Special requirement by MRTS70: maximum total alkali content 1%, and
maximum available alkali content of 0.5% (Na2O equivalent).
 Silica fume – AS/NZS 3582.3-2016; Special requirement by MRTS70: silica fume shall not be used in
decks or slabs.

The jurisdictions’ specifications may not be sufficient for the construction of major structures involving
complex durability issues and different exposure environments for different elements. In such cases,
jurisdictions often engage durability experts to draw up detailed durability plans, such as those already listed
in Table 1.1.

An explicit durability plan is a method of durability design that specifically considers each relevant
deterioration process, the design life and the criteria that define the end of this design life in a quantitative
way (Concrete Society 1996).

Development of durability plans is an involved process. For example, the TMR Queensland Bridge Asset
Management Section has produced guidelines for preparation of durability plans (Department of Transport
and Main Roads (DTMR) 2009) ‘to gather and record the information generated through the asset creation
process, and collate it in a standard format to facilitate consistency of approach to durability planning’. The
process, reproduced from DTMR (2009), is illustrated in the flowchart shown in

Figure 2.1. Similar durability planning schemes have been employed by other agencies, e.g. in planning for
the Kempsey Bypass in NSW and other plans listed in Table 1.1.

Austroads 2016 | page 9


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In practice, each plan would be specific to a particular project as it would need to target specific site
conditions and the associated durability problems, as well as design issues, service requirements and
required service life. Therefore, the plans for different projects would include some common issues and
some specific issues, depending on the exposure condition, services required of the structure, service life,
and the conditions that would be encountered during construction of the works. Post-construction
maintenance requirements could also be different, depending on some of the above factors.

Durability plans are mostly concerned with addressing possible durability problems for the structure
concerned, and providing measures to avoid the problem during the service life of the structure. In general,
durability problems for concrete structures can be divided into two categories, i.e. those affecting the steel
reinforcement and those related to the concrete ingredients. Both degradation mechanisms can adversely
affect the performance of structural members due to deterioration of their design capacity. For reinforced
concrete, corrosion damage is usually caused by carbonation or chloride penetration into the cover concrete,
which protects the reinforcement both physically and by its alkalinity.

Table 2.2 presents the list of durability problems addressed in the durability plans obtained from the
jurisdictions. The issues commonly addressed are reinforcement corrosion due to the ingress of chloride
through the cover concrete, loss of alkalinity due to carbonation of the cover, and degradation of concrete
materials due to exposure to sulfate-rich soils or highly acidic conditions. The Concrete Society (UK) special
publication, CS109 (Concrete Society 1996), discusses developments in durability design and performance-
based specifications of concrete up to the date of publication. The document addresses a number of
deterioration problems, including freezing and thawing which is a problem in cold climates, but not relevant in
most regions in Australia.

The deterioration issues may be present individually in different sites, or combined at the one location. For
example, Ballina Bypass in NSW has various exposure conditions for the different concrete members. For
those in exposure Class B2, concrete of 40 MPa strength grade was used and the specified chloride
diffusion coefficient was 5.5  10-12 m2/s, as determined by the NT Build 443 Test Method (Nordtest 1995).
For the highly aggressive marine environment of the tidal zone, a special class S50 fly ash concrete (25% fly
ash replacement of cement) with a cover thickness of 55 mm, or a S50 high slag cement (65% slag)
concrete with a cover of 35 mm was considered to yield an initiation period of 100 years. A chloride diffusion
coefficient of 2  10-12 m2/s was specified for the S50 concrete (Khatri et al. 2009). It should be noted that the
specified value of the diffusion coefficient was related to a concrete age of 28 days, and would be expected
to decrease significantly with time. This aspect is discussed in Section 5.1. Similarly, the bridge piles of the
New Gateway Bridge in Brisbane were exposed to a high chloride environment in the tidal and splash zones,
and to high chloride plus acid sulfate materials in buried zones, for which different mitigation approaches
were considered.

Austroads 2016 | page 10


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 2.1: Durability design process flow chart

Source: Figure 3, DTMR (2009).

Austroads 2016 | page 11


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 2.2: Issues addressed in structure durability plans for 100 years or longer service
Defects
Steel
formed at Cover concrete Concrete deterioration Other issues
Structure location Main exposure member
early ages addressed
Corrosion Thermal Shrinkage Cl CO2 AAR DEF Sulfate Mg Acids
Sorell Causeway, Seawater
y y y n n n n n n
Tasmania
Brighton Bypass, Acid sulfate
y y y y n y y y n y n
Tasmania soils
New Gateway Bridge, Seawater, acid Future CP
y y y y y y y y y y
QLD soils
Kempsey Bypass, NSW Acid sulfate soil Future CP
y y y y y y y y n y Crack width allowed =
0.3 mm
Princes Highway Acid sulfate soil, Future CP, and Stray
Gerringong Upgrade and Erosion current protection
y y y y y y y y y
Gerringong Upgrade
Durability Plan
Ballina Bypass, NSW Seawater and Bridge design changed
y y y y y y y y
acid sulfate soil to move pier out of water
Camden Valley Way Saline soil and Surface coatings
y y y y y y y y n y
Upgrade, NSW water
Banora Point Upgrade, Tidal water Erosion, future CP
NSW groundwater, y - y y y y y y y system
acid sulfate soil
Hunter Expressway, Soil and Erosion at embankments
y - y y y y y y y y
NSW groundwater
Tarcutta Bypass, NSW Soil and Mild environment
y y y y y y y y n y
groundwater
Sydney Desalination Seawater, brine Corrosion monitoring
y y y y y y y y
Project, NSW discharge

Source: Sorell Causeway: GHD (2002), Gateway Bridge: Maunsell and AECOM (2007), Sydney desalination project: Blue Water (2008), Brighton Bypass, Tasmania: (GHD 2009),
Kempsey Bypass Alliance (2010), Tarcutta Hume Alliance (2009), Hunter Expressway: AbiGroup (2011), Ballina Bypass Alliance (2008), Camden Valley Way Upgrade: Camden
Valley Way Upgrade C2C Alliance (2011), Princes Highway Gerrigong Upgrade: Roads and Maritime Services (2010), Gerringong Durability Report: Roads and Maritime Services
(2012a, 2012b).

Austroads 2016 | page 12


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Durability plans also address defects that can form at an early age in the concrete, for which measures are
found through trials and applied to correct the deficiency before the concrete formulation is used for
manufacturing concrete. In addition to the above, the plans may include measures that would allow
mitigation methods to be implemented in the future, for example provision of connectivity of the steel
reinforcement in the whole structure for the installation of corrosion monitoring devices and application of
cathodic protection systems.

Various manufactured products, such as metallic components (expansion joints, bearings, railings, light
poles and related fittings) used in the construction of the bridge also need to be considered and their
conformity to relevant standards and specifications assessed before being built into the structure (Pritchard
2011).

An important outcome of the durability plans is a list of requirements for the various types of elements in the
bridge. For concrete elements, this would include strength grade, minimum cement content, minimum cover
thickness, maximum water/binder ratio, minimum porosity, maximum chloride diffusion coefficient, maximum
sorptivity, etc.

2.3 Protective Measures to Increase Service Life


Additional measures may be necessary to ensure that the specified design life of bridges is achieved under
aggressive exposure environments. These may include:
 increase in concrete cover thickness
 protection of the concrete surface by application of appropriate surface coatings
 isolating the concrete from the aggressive environment by encasement
 neutralisation of acidic soils by application of lime
 installation of cathodic protection or cathodic prevention
 use of corrosion-resistant materials such as stainless steel
 use of high performance concrete mixes.

Post-construction monitoring and maintenance of structures also play important roles in extending the life of
bridge structures. Andrews-Phaedonos et al. (2011, 2013) presented examples of this approach being used
for two VicRoads bridges, which were considered to have benefited from monitoring. VicRoads has recently
produced a comprehensive guide on the assessment, maintenance and rehabilitation of concrete bridges
(VicRoads 2010b), which will no doubt be valuable in extending the life of concrete bridges.

Current practice for minimising deterioration risks, as specified in Australian standards or jurisdiction guides
is summarised in Table 2.2.

Austroads 2016 | page 13


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 2.2: The measures for service life in AS 5100.5–Section 4 (draft 2012) or as otherwise specified

Durability issue Protective measure Recommendation Notes


Chloride- and Limit chloride content Reinforced concrete (ambient The limit may increase when
carbonation-induced in concrete mix curing): 0.6 kg/m3 using stainless steel
corrosion of Prestressed concrete and reinforcement
reinforcement reinforced concrete cured by
steam: 0.4 kg/m3
Use required grade of B1: 45 mm, 32 MPa, 40 mm, 40 Cover thickness for well-
concrete and cover MPa, 35 mm,  50 MPa; compacted concrete in rigid
thickness B2: 55 mm, 40 MPa, 45 mm,  50 steel formwork cover may be
MPa; reduced by 5–10 mm;
B2: 45 mm, 32 MPa, 40 mm, 40 Spun or rolled member cover
MPa, 35 mm,  50 MPa; reduces by 10–15 mm
C1: 60 mm,  50 MPa; C2: 70 mm,
 55 MPa
Use stainless steel May use nominal cover 30 mm and High tolerance against
lower grade concrete corrosion, depending on type
of stainless steel
Others Crack control, using controlled –
permeability formwork, the addition
of corrosion inhibiting admixtures in
the concrete; surface coating
Reinforcement Joint preparation – GHD (2009)
corrosion, localised
Thermal cracking at Control temperature Temperature differential across the Environment temperature
early age rise concrete member being affects the concrete
constructed shall not exceed 20 °C temperature
during the curing period
Drying at early age Evaporation limit When evaporation rate exceeds VicRoads (2010a),
0.50 kg/m² per hour, the contractor Specification Section 610
shall take precautions to minimise
evaporative moisture losses
AAR Testing aggregate for The aggregates classified as –
AAR reactivity reactive by the prism test shall not
be used
Limit the total alkali Na2O equivalent  2.8 kg/m3 –
Using SCM-blended Fly ash 20%; Slag 50%; Silica Some aggregates may require
cement Fume 8% 25–30% fly ash and 65% slag
for AAR suppression
DEF Limit concrete < 70 °C, where the cement type is Allowed limit of up to 5% SO3
temperature during GP; in cement is considered too
the curing period < 80 °C, where the cement type is high
GB
Soft water attack Surface preparation Use SCMs GHD (2009)
Sulfate attack Restriction on sulfate Acid-soluble SO3/ cementitious –
content material:
5.0% for reinforced concrete; 4.0%
for prestressed concrete
Magnesium Consider increasing If Mg2+ > 1000 ppm, an exposure –
exposure classification of one higher class
classification should be adopted
Acidic soil Consider increasing pH > 5.5: B1; pH = 4.5–5.5: B2; pH For low permeability soils, the
exposure =4–4.5: C1 for high permeability corresponding class is one
classification if pH > 4 soils lower
Additional protective Exposure Class C2 –
measures if pH < 4

Austroads 2016 | page 14


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Durability issue Protective measure Recommendation Notes


Freezing-thawing Concrete grade 40 MPa for occasional exposure; –
50 MPa for frequent exposure ( 
25 cycles per annum)
Air entraining in 8% to 4% for 10 mm to 20 mm Nominal size aggregate
concrete aggregate
6% to 3% for 40 mm aggregate

Source: AS 5100.5-2004.

Overseas, the Federal Department of Transport (FDOT) in the USA has placed restrictions on the chloride
diffusion coefficient, cover (> 100 mm), and using corrosion inhibitors, as specific requirements for structures
having a design service life of 100 years. Note that the cover thickness was estimated for the penetration of
chloride at the corner of square elements, as the worst case scenario (Sagüés & Powers 1997). Norwegian
code NS3473 (Standards Norway 2003), based on a survey of chloride distribution at various parts of
bridges, specifies cover thickness according to the exposure conditions, e.g. nominal 120 mm with a
minimum 100 mm for the splash zone and underwater zone. Note that the Norwegian code takes the
standard deviation into consideration.

It should be noted that the value of the standard deviation of various concrete properties, and those for
tolerances in the construction work, must be considered carefully as this can have a significant effect on the
long-term durability. These should be included in the specifications for quality control. For example, AS
5100.5-2004 allows a range of –5 mm and +10 mm for the specified cover thickness for slabs, beams, walls
and columns.

For a structure with an expected service life of 300 years, the current prediction models require very thick
concrete cover, but this may not be practical in some situations, where a thinner cover concrete is more
appropriate. Moreover, as the integrity of the cover concrete cannot be guaranteed in the very long-term, due
to inadequacies in prediction models, using stainless steel as an alternative to carbon steel seems to be
inevitable for very long service lives. This issue was considered and implemented in the case of the New
Gateway Bridge in Brisbane (Maunsell & AECOM 2007). In this case, the aim was to achieve a service life of
300 years rather than 100 years. In this regard, Connal and Berndt (2009) summarised the actions taken to
implement measures for achieving the design life of 300 years, as directly quoted below:

Selection of good quality concrete. This is not a cost impost but rather business as usual
and can often be based on the current concrete specifications of the State road
authorities. The extra requirement is the need to define target chloride diffusion and
carbonation coefficients, and achieve them.

Selection of greater cover to reinforcement. In this case 55 mm cover was adopted for
50 MPa superstructure concrete when the AS 5100.5 code would have permitted 45 mm.

Use of a ternary concrete mix and selective use of stainless steel as surface
reinforcement in the splash zones.

Provision of electrical continuity for reinforcement in substructure elements in more


aggressive environments, to enable future cathodic protection to be installed if necessary.

Good detailing to enable compaction of concrete, along with good vibration and
subsequent curing during construction, to ensure a dense layer of cover concrete.

It takes a commitment from owners, designers and builders to achieve the goal of a very long life. This
commitment was demonstrated by all involved in the Gateway Bridge project, down to the men placing,
finishing and curing the concrete. This commitment achieves results when supported by an adequate level of
supervision and guidance, plus active and independent surveillance.

Austroads 2016 | page 15


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

2.4 Current Practice for Long Service Life Prediction


Currently, several service life prediction models, e.g. Life-365, are used for prediction of time when chloride
content in the close vicinity of reinforcing steel increases to a level that can initiate steel corrosion. This is
actually the carefree service life instead of the effective service life. This type of approach underestimates
the effective life span of bridges or prescribes impractically large cover thickness for the bridge components.
This approach has been extended to include the period taken for the steel to undergo active corrosion and
develop a sufficient amount of rust products to cause cracking of the cover concrete (Liu 1996; Liu & Weyers
1998; Bhargava et al. 2005) (note: these models have deficiencies which are discussed in Section 5).
Therefore, more recent models of service life predict the sum of the carefree period plus the time to cracking
of the cover concrete. At this time, remediation measures may need to be implemented to minimise the rate
of deterioration.

In reality, the cracked concrete would remain functional for many years after crack development, until steel
section loss increases to a level where the mechanical properties of the element are below an acceptable
safety factor for load bearing. Very few models consider the whole spectrum of deterioration processes. This
approach was used by Shayan and Xu (2006), the experimental work for which was based on Austroads
Publication AP-R245 (Austroads 2004).

A number of service life prediction models have been presented and discussed in Austroads publications
AP-T06 (Austroads 2000a), AP-T07 (Austroads 2000b), AP-T11 (Austroads 2002), and AP-R245 (Austroads
2004). Further discussion of some of the models is presented in Section 5. Aldridge et al. (2013) discussed
some issues related to service life prediction, and stated that a reasonable prediction can be made using a
known time-dependent chloride diffusion coefficient and surface chloride concentration of concrete.

There are several uncertainties or errors in the theoretical formulations and test data used in some of these
models, e.g. Austroads publication AP-R245 (Austroads 2004) highlighted possible errors in a corrosion
model used to predict time to cracking, which will be discussed later. Another issue is that the current models
have not considered the combined effects of deterioration mechanisms. For example, chloride diffusion is
commonly considered for intact concrete, not when another crack-inducing mechanism, such as alkali-
aggregate reaction, is also operative.

The uncertainties in the various models result in difficulties in service life predictions, as experienced by
Safehian and Ramezanianpour (2013), who found predicted service lives varying from 15 years to 66 years.
They stated that each model needs to be first calibrated for the new sites. This makes the use of such
models cumbersome and less practical. Tang et al. (2015) also used various models, including Life-365 and
DuraCrete (1999), and found that different models gave different service life predictions. As mentioned
earlier, this occurs because of interactions between mechanical stresses and chemical/environmental
stresses operating on the concrete under field conditions, which are not built into the models.

Alexander and Thomas (2015) reviewed the current developments in service life prediction and performance
testing, and stated that a combination of performance testing and predictive models can provide a
reasonable prediction of service life. The chloride ingress profiles of 25-year old concrete blocks of various
binder compositions, which had been exposed to marine conditions in this period, were given as examples of
practical application of the performance approach from which predictive parameters can be derived. They
also stated that predictive models can be used to derive performance parameters for given concrete mixes
under given exposure conditions.

Austroads 2016 | page 16


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

As mentioned earlier, most predictive models rely largely on chloride diffusion into concrete for service life
prediction; and build-up of chloride concentration at the steel reinforcement to corrosion initiation levels is the
main factor determining the service life. However, in practice, concrete structures perform well for some
decades after this stage. Deterioration of concrete after corrosion initiation would depend on the rate of
corrosion and the consequent cracking of concrete and loss of steel section. Xu and Shayan (2016)
investigated the relationship between rust growth and cracking of the cover concrete, and showed that the
corrosion rate depended on the chloride content of concrete, and the time to cracking depended on the cover
thickness and steel bar diameter, as well as the quality of the concrete. As discussed by Shayan and Xu
(2006), the final stage of service life of elements affected by corrosion-induced damage would depend on the
safety factor set for the deteriorating element.

2.5 In-service Maintenance Phase


It would be expected that with adequate design and quality of construction, costly repair operations in the
future can be avoided or minimised. However, procedures for regular inspection, maintenance, and remedial
works need to be developed at early stages of the durability plan. Consideration should be given to access
difficult areas for inspection and to the extent and frequency of inspection, as well as choice of inspection
methods. Some modern bridges, e.g. the Oresund Bridge crossing the Oresund strait between Sweden and
Denmark, have been designed to allow operation of under-bridge inspection units.

The monitoring may include making mock-up concrete specimens and testing their properties at different
stages, as well as extracting concrete samples from the structure concerned at appropriate times. Modern,
embedded devices for monitoring steel corrosion may also be considered, if the long-term reliability and
serviceability of the devices is established. Austroads publication AP-T06 (2000a) discusses the advantages
and disadvantages of the various corrosion monitoring techniques. Andrews-Phaedonos et al. (2011, 2013)
have used such techniques to monitor the corrosion behaviour of two bridges subjected to tidal waters, with
the aim of detecting the onset of corrosion activity and appropriate intervention times.

Traditionally, patch repair techniques and various surface coatings have been used to prevent corrosion, but
this has been associated with varying degrees of success and is generally considered an expensive option in
the long-term. One of the practical remedial options would be the application of cathodic prevention for new
structures, and cathodic protection (CP) if the reinforcing steel is close to the corrosion initiation stage, has
started to corrode, or detectable signs of failure are observed. Ideally, the structure and the concrete should
initially be designed to allow the application of a CP system. Song and Shayan (1998) summarised the
causes of corrosion, its detection and prediction of deterioration in a state-of-the-art report, and Song and
Shayan (1999) have elaborated on the various techniques of corrosion prevention that could be used for
reinforced concrete structures. Guidelines on design and construction practices to mitigate corrosion of
reinforcement in concrete structures have also been published by the American Concrete Institute (2003).

2.6 Improvement of Current Practice

2.6.1 Code and/or Specification Requirements

Although the present codes or some road agency specifications specify certain performance-based
properties as requirements for durability, and there are laboratory test methods available for this purpose, the
interpretation of the test results with respect to long-term field performance needs to be clarified through in-
depth research. For example, NT Build 443 (Nordtest 1995) uses very high chloride concentrations for
relatively young concrete specimens to determine their chloride diffusion coefficient. The long-term chloride
diffusion of field concrete is sometimes assumed as one-tenth of this value (e.g. durability plan for Sorrell
Bridge, Tasmania). The basis for this assumption is not robust, and the relevance of the results to long-term
chloride ingress into field concrete under ambient seawater conditions is not established and requires
detailed studies. Some of these issues will be discussed in later sections of this report.

Austroads 2016 | page 17


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

2.6.2 Prediction of Service Life

Estimation of the time to detectable damage is an important parameter in the prediction of service life and
should be included in all the prediction models. As mentioned earlier, most models used to consider only the
carefree portion of service life as the service life, and more recently the component of time to cracking has
been added to it. As stated above, the existing models appear to include uncertain assumptions, which will
be discussed in Chapter 5. This aspect has also been explored at ARRB Group through an Austroads
funded study (Austroads 2004), and the findings will also be discussed in Section 5.

2.6.3 Use of Stainless Steel

It has been known that stainless steels can tolerate much more severe environments than can carbon steels.
For this reason, they have been considered as the final solution to chloride-induced corrosion and
deterioration of reinforced concrete structures in marine or other salt-laden environments.

As discussed later, the extent of tolerance of stainless steel to chloride contamination still needs to be
established for specific conditions, particularly for the low-Ni stainless steels, which are more likely to be
used in the future due to lower cost. Stainless steel reinforcement bars are more likely to be used in
structural members with relatively thin cover. However, such thin cover concrete will be completely
carbonated within the first 100 years, and airborne chloride ions will accumulate in the cover concrete.
Therefore, the stainless steel must remain tolerant to corrosion under these conditions.

Unfortunately, service records for stainless steel as concrete reinforcement subjected to aggressive marine
conditions is limited to around 70 years (Rostam 2005), and longer-term data for various types of stainless
steel is lacking. Particularly, long-term corrosion resistance of stainless steel bars in carbonated concrete,
with and without chloride contamination, needs to be investigated and established in the near future.
Moreover, new low-Ni varieties, which have been marketed for their lower cost, need to be investigated to
establish the level of chloride contamination that they can tolerate in the long-term.

Austroads 2016 | page 18


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

3. Definition of Service Life


The service life of a structure is the period during which it performs the required functions with minimal
maintenance cost. It consists of three stages: time to onset of potential failure (dormant period), time to the
point of detectable failure, and time to the point of functional failure (Figure 3.1).

Figure 3.1: Schematic expression of structure service life

Onset of potential failure


100%
Condition (required performance)

Point of detectable failure

Point of
Minimum Acceptable Performance functional
n% failure

Failed Condition

t1 t2 t3
Time
Source: Miah and Hinton (2007).

The performance of a highway bridge depends on several factors, including the durability of structural
components in a given exposure environment; durability of replaceable components, such as bridge
bearings and deck joints; and load carrying capacity for the required traffic loading. This report focuses on
the first issue, the durability of structural components. Design issues for various bridge components are
outside the scope of the report.

The structural components of a reinforced concrete bridge, such as deck slab, girders, columns, piles, etc.,
are directly exposed to the environment, which can be aggressive to the materials that comprise the bridge
components, and can deteriorate with time. It would be desirable for these components to resist the adverse
effects of the exposure environment during their design/service life, and for the quality of the components to
be assured at the construction stage. The demand by road agencies for longer service life of bridges has
been increasing with increased requirements for traffic loading. Fortunately, this has been made possible
due to improvements in construction technology. Currently, the design service life of major bridge structures
is 100 years (AS 5100.5-2004 and AS 3600-2009), but much longer periods could be demanded for some
bridges. The Department of Transport and Main Roads Queensland required 300 years of service life for the
New Gateway Bridge in Brisbane (Maunsell & AECOM 2007). It should be noted that different components of
the bridge structure can have different design lives for example the components listed in Appendix D for the
new Gateway Bridge in Brisbane.

Austroads 2016 | page 19


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The challenge of such severe requirements is that there is no reliable information on the behaviour of
concrete and reinforcement materials in service for 100 years. On the contrary, severe failures have been
detected for some components of modern bridges in service for only 30 years. These unpredicted
deterioration problems highlight the need for, and the importance of, consideration of the deterioration
mechanisms, testing methods, prediction models, and prevention methods before the stage of construction.

Various components of a bridge have different required or guaranteed service lives, and have vastly different
replacement or repair requirements when the deterioration becomes so severe as to affect the function of the
bridge. The main structural components of major bridges, including piles, pile-caps, columns, crossheads,
girders, beams, deck units and abutments, can be very difficult and costly to repair, and consequently require
very long service lives, being more than 100 years or even 300 years (as in the case of the New Gateway
Bridge). AS 5100.5-2004 specifies a 100-year service life for major bridge structures.

In general, the three stages of reinforced concrete service life are represented by:
1. Penetration of carbon dioxide (atmosphere) or chloride ions (mainly from seawater and contaminated
soils for bridges in Australia). These agents penetrate into concrete, and result in breakdown of the
passivation film on the steel surface after reaching the steel bar in the concrete. Material deterioration
includes concrete being attacked by aggressive environments or defects generated in improperly
fabricated elements. The time needed for the aggressive agents to reach the surface of steel bars and
break down the passive film corresponds to the onset of potential failure. This is the time that most of the
present service life prediction models focus on and take as the service life, in spite of the fact that the
deterioration would only start at this time.
2. Active corrosion of steel when moisture and oxygen are available. In chloride-contaminated concrete, the
rate of corrosion increases with a further increase in chloride content. The build-up of rust around the
steel bars will eventually result in concrete cracking. At this stage, concrete shows visible defects such as
scaling, cracking due to ingredients of concrete such as AAR and delayed ettringite formation which can
enhance the onset of corrosion. This corresponds to ‘point of detectable failure’.
3. Further progress of deterioration, enhanced by large cracking and delamination, which expose the steel
directly to the aggressive agents. This may lead to severe loss of steel section or total detachment
between reinforcement and concrete. At this stage, the affected concrete shows large areas of
delamination and spalling. This time corresponds to the ‘point of functional failure’.

The main durability concerns for reinforced concrete are: 1) steel reinforcement corrosion and the resulting
component failures, and 2) chemical and physical deterioration of the concrete materials caused by chemical
incompatibility and environmental conditions.

In relation to the former, a number of approaches are available to prevent or delay the corrosion process and
its consequent damage, including the use of corrosion-resistant steel and/or corrosion inhibitors, and a high
performance concrete. A reasonable estimation of the corrosion initiation time, progress of corrosion and
time to cracking, as well as time to unacceptable loss of function, would then be required for accurate
prediction of service life. These aspects have been the subject of modelling studies. Earlier models only
considered the corrosion initiation time as the service life, which is very conservative. In practice, the latter
phases are important components of service life prediction and should be included in the service life
prediction models. The objective of the design phase is to select appropriate parameters in order to prolong
these periods.

Regarding the second concern, prevention of deterioration of the concrete materials would require:
 all the individual ingredients of concrete to be selected through appropriate testing
 appropriate concrete mix designs to be developed and verified through trials for the relevant exposure
class
 appropriate construction techniques (including placement, compaction, curing), close site supervision and
quality control to be employed.

Not only are the concrete deterioration mechanisms detrimental to structural integrity, but they often enhance
the process of steel corrosion.

Austroads 2016 | page 20


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

To realise a service life of at least 100 years for a given structure, very careful planning is required, which
should include investigations of the site and its exposure conditions for different components of the proposed
structure; design investigations on bridge components so that the elements are appropriate for loading
conditions; materials investigations, including appropriate tests and their acceptance limits; robust quality
control processes for concrete and other materials; and provision of appropriate site supervision practices,
including proven methodology for delivery, placement, compaction, finishing and curing. The Concrete
Institute of Australia (CIA) has prepared a useful Recommended Practice document in the Concrete
Durability Series, titled, Good Practice Through Design, Concrete Supply and Construction, Z7/04 (CIA
2014b), which covers many of the issues related to the concrete ingredients, supply and manufacture of
reinforced concrete, and construction issues.

In the following sections, concrete deterioration problems known to occur in Australia will be discussed,
including the latest developments and mitigation options. An example of models used to predict service life
will also be presented.

Austroads 2016 | page 21


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

4. Performance of Reinforced Concrete Materials


Deterioration of reinforced concrete involves reactions which can occur within the concrete, between its
ingredients, or as a result of its interaction with the environment. The reactions can attack the cement and
aggregate components or the reinforcing steel, or sometimes both. They can cause various forms of defects
in the concrete and in the steel reinforcement under the given environmental conditions. Some of the
deterioration mechanisms can coexist and even enhance each other, which is a fact that must be taken into
consideration in the prediction of service life.

The environmental or exposure conditions are highly significant with respect to the durability performance of
concrete structures, and may be different not only in different geographical locations, but even for different
elements of the same structure, i.e. several localised microenvironments can exist in one site.

Marine conditions prevailing in coastal areas, as well as those in saline groundwater, are amongst the most
aggressive environments. Another highly aggressive exposure condition for reinforced concrete is that of
acid sulfate soils, which generate very acidic conditions when disturbed or drained. High levels of chloride
and sulfate salts in these environments are the cause of concrete deterioration, which is promoted in warmer
climates.

It follows from the above that concrete structures need to be designed for each specific exposure condition,
and that detailed site investigations should be carried out as the first step in design of the various elements.
The Roads and Traffic Authority (2005) prepared Guidelines for the management of acid sulfate materials:
acid sulfate soils, acid sulfate rock and monosulfidic black ooze.

The evaluation of durability and assessment of service life of a newly planned structure should be based on
thorough analysis of the local conditions, and the resistance of concrete materials to deterioration under the
prevailing conditions.

In Australia, AS 5100.5-2004 includes five exposure classifications, each requiring specified values of
concrete compressive strength and cover thickness. Road agencies specify requirements for durability of the
structure, e.g. minimum cement content as well as characteristic strength, cover thickness, porosity and
water absorption, maximum chloride diffusion coefficient, etc., but these have not been included in AS
5100.5-2004.

4.1 Reinforcing Steel Corrosion


Steel corrosion is an electrochemical process in which the iron atoms change to Fe 2+ ions by losing electrons
in an anodic reaction (Equation 1):

Fe → Fe2+ + 2e– 1

The free electrons pass though the steel to cathodic sites where they are consumed by the cathodic reaction
(Equation 2). In neutral or alkaline solutions, the cathodic reaction involves the consumption of oxygen
dissolved in the solution, i.e. oxygen reduction:

O2 + 2H2O + 4e– → 4OH– 2

The hydroxyl ions react with Fe2+ ions to form Fe(OH)2 and ultimately rust, e.g. Fe2O3·nH2O. Under
anaerobic conditions, such as underwater portions of piles or columns, the dissolved Fe 2+ ions diffuse away
into the solution without leaving any visual sign of corrosion.

Austroads 2016 | page 22


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Steel in concrete elements is protected by the concrete alkalinity (pH > 12.5), which forms a protective
passive film (thin rust layer) on carbon steel bars. The carbon steel will corrode when the passive film is
destroyed, which will happen in two situations: where the concrete alkalinity has been neutralised by
carbonation (pH < 9), and where sufficient amounts of chloride ions penetrate the concrete to the surface of
steel bars and dissolve the passive film (Figure 4.1); where the equilibrium relationships amongst various
phases of iron (Fe) are presented in a plot of the potential (V, measured in relation to the standard hydrogen
electrode, SHE) against the pH, known as the Pourbaix Diagram. Two vertical blue lines in the diagram show
the pH of new Portland cement concrete, which is around 13–13.5; and that of the carbonated concrete, in
which the pH decreases to around 9. Steel in carbonated concrete is seen to fall in the zone of corrosion,
when the potential is in the range of –0.60 V to 0.8 V, which is normally encountered in concrete structures.
At high pH, steel falls in the passive (protected) zone. The dotted red lines represent increasing chloride
concentrations in the system (concrete), which indicate that at high chloride concentrations, steel falls in the
zone of pitting corrosion, which is more rapid than general uniform corrosion.

Figure 4.1: The E–pH (Pourbaix) diagram of iron, according to chemical equilibriums of species in an Fe-H2O
system

Source: The effects of chloride are cited from figures in Nilsson et al. (1996).

The steel corrosion produces rust, which has a larger volume compared to the corroded steel (Figure 4.2).
The volume increase due to the reinforcing bar corrosion will cause the concrete cover to crack (Liu 1996;
Shayan & Xu 2003). Severe corrosion has caused complete delamination of the cover concrete, and
considerable reduction in the cross-section of the steel which can adversely affect the load bearing capacity
of the element concerned (Shayan, Xu & Al-Mahaidi 2004). Some research shows that corrosion caused by
salt spray deposited on steel bars reduces its tensile ductility (Apostolopoulos & Koutsoukos 2008).

Austroads 2016 | page 23


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 4.2: Volumes of iron and its reaction products

Source: Mansfeld (1981) – cited by American Concrete Institute (ACI) (2001), ACI 222R-01.

The relative volume increase  due to rebar corrosion in concrete can be expressed by a simple relation
(Shayan & Xu 2003):

 steel 3
  dV / Vsteel  (  1)
rust

where

V = volume

 = density

 = the molecular mass ratio of Fe to that of rust

For typical rust Fe(OH)2 and Fe(OH)3,  = 0.6215 and 0.5226, respectively. Densities of steel and rust are
7.86 and 3.46 g/cm3, respectively; the latter as determined by Shayan and Xu (2003), which give  = 2.66
and 3.35 for the two forms of rust, respectively. According to the E–pH diagram, Fe(OH)3 is the main type
that may exist on reinforcing steel in concrete, i.e. pH = 9–13.5, while corrosion potential is about –350 mV
relative to copper-copper sulfate electrode CSE.

Under the condition of oxygen deficiency, such as the underwater part of a pier, the corrosion reaction can
proceed in the form of a large macro-cell, whereby the steel in and above the splash zone acts as the
cathode to receive oxygen, while the underwater part acts as the anode to give up electrons, and the Fe 2+
ions would diffuse out of the porous concrete. This was observed on a pier at a location of 6.8 m below the
water surface (Bolin et al. 1997).

Austroads 2016 | page 24


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

4.2 Mitigation of Steel Corrosion


Using corrosion inhibitors in the concrete mix is the most convenient way to protect the steel from corrosion.
Corrosion inhibitors can be divided into three types: anodic, cathodic and mixed, depending on whether they
interfere with the corrosion preferentially at the anodic or cathodic sites or whether both are involved
(Ramachandran 1984).

A brief review of the application of various corrosion inhibitors in concrete is presented in Song and Shayan
(1999). The review shows that not all inhibitors performed satisfactorily in all applications.

Broomfield (2007) presented a concise summary of the results of the field application of various inhibitors,
and whether they perform satisfactorily as corrosion inhibitors for newly constructed structures (i.e. used as
an admixture in fresh concrete) or as corrosion inhibitors for the rehabilitation of hardened concrete
subjected to, or at risk of, reinforcement corrosion. This included the cathodic and anodic inhibitors, and
those which inhibit both the cathodic and anodic reactions.

Based on the composition of the inhibitors, they are classed as inorganic inhibitors (such as zinc oxide,
nitrites, phosphates, etc.) and organic inhibitors, such as those based on amines and esters, and a subgroup
based on aminoalcohols that have high vapour pressure, which are called vapour phase or volatile inhibitors.
All of these inhibitors have been used in experiments of field applications to protect the steel reinforcement in
concrete.

The most commonly used anodic inhibitor is calcium nitrite (Ca(NO3)2). The NO2- ions compete with Cl- for
ferrous ions to produce ferric oxide, FeOOH, which is a stable oxide and thus the iron surface is protected.

The shortcomings of this approach are that calcium nitrite is costly, and that it is effective only when the Cl-
/NO2- ratio < 1. It was reported that insufficient nitrite concentration may increase the corrosion activity (de
Rincón et al. 2002).

Although some nitrite inhibitors, such as sodium nitrite (NaNO 2), cause significant set retardation in concrete
and reduce the compressive strength, calcium nitrite is more compatible with concrete and, in fact, acts as a
set accelerator and increases the early age compressive strength (Ann et al. 2006). These authors also
found that calcium nitrite reduced the corrosion rate and increased the chloride threshold level.

To overcome the high cost of calcium nitrite inhibitors, the less costly calcium nitrate has been introduced
into the market. Justnes (2004, 2006a, 2006b) conducted a comparative study of calcium nitrite and calcium
nitrate, and found that the effectiveness and mode of action of the two materials in preventing/inhibiting steel
corrosion are the same, but the calcium nitrate has slower kinetics (which is not an issue as the corrosion
process is slow). It was also stated that the nitrate provides a greater buffering capacity as an inhibitor. In
addition, Justnes (2006a) also elaborated on other useful functions that calcium nitrate could play in
concrete, and that unlike calcium nitrite, it is harmless to the environment. Østnor and Justnes (2011)
showed that 3–4% of these inhibitors was sufficient to prevent reinforcement corrosion.

The commonly used cathodic inhibitors are bases which increase the pH of the pore solution, e.g. NaOH,
Na2CO3 and NH4OH. However, addition of sodium ions in concrete may increase the risk of AAR, and the
presence of soluble alkalis in mixing water may disturb the chemical balance of cement hydration, e.g.
presence of a small amount of NaOH will cause a great reduction in solubility of Ca(OH) 2 due to the common
ion effect (Way & Shayan 1989; Sarkar & Xu 1993).

Zinc oxide (ZnO) is considered a cathodic inhibitor in water because it precipitates Zn compounds at the
cathode, where the alkalinity is high. In concrete, the Zn compounds may precipitate everywhere, including
on both cathodic and anodic areas of steel. The compounds formed can passivate the steel and can reduce
concrete porosity (de Rincón et al. 2002).

ZnO has been shown to be effective in controlling corrosion, especially when it is used with NaNO2 as a
mixed inhibitor (Saraswathy & Song 2007).

Austroads 2016 | page 25


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

More complex inorganic compounds, such as sodium mono-fluoro-phosphate, have been used as migrating
corrosion inhibitors for hardened concrete (Cortec Corporation 2011). This compound was found to be useful
only for fully carbonated concrete, as the presence of Portlandite causes precipitation of the fluoro-apatite
and limits the penetration of the material (Chaussadent et al. 2006).

In a research project in the USA, sponsored by the Joint Highway Research Advisory Council, Goodwin et al.
(2000) investigated the corrosion inhibiting effects of two new inhibitors based on alkali metal and ammonium
salts of an alkenyl-substituted succinic acid, and compared them to two unidentified commercial inhibitors.
The new inhibitors prevented rust formation on the steel over the 24 months of monitoring, and performed
better than the commercial inhibitors. However, verification of long-term effectiveness of the inhibitors
requires much longer times, and test conditions closer to field situations.

Ormellesse et al. (2006) found that similar organic compounds reduced the penetration of chloride due to a
pore blocking effect, but the effect on the corrosion rate was not clear.

Organic inhibitors have been produced and used in structures in the past two decades. Cortec Corporation in
the USA (Cortec Corporation 2011) has demonstrated that a type of migrating corrosion inhibitor (MCI)
including amino-alcohol-based and amino carboxylate-based products can provide better protection than
calcium nitrite for the reinforcing steel in a cracked beam exposed to 3.5–6.0% NaCl solution.

Saraswathy and Song (2007) showed that there was an optimum amount of inhibitor, which was about 2%
by mass of cement, whereas a higher dosage showed a negative effect. Of course, the amount would vary
with the type of inhibitor, e.g. calcium nitrate is used at 3–4% by mass of cement (Østnor & Justnes 2011).

The most uncertain issue regarding the long-term effect of the corrosion inhibitors is that they will diffuse out
from concrete while the chloride concentration increases with time. Therefore, the inhibitors should have an
important feature that they can provide protection at all concentration levels, or at least there is no negative
effect when the ratio of the effective group to chloride ions is low.

Broomfield (2007) has listed the following general issues regarding the corrosion inhibitors, and states that
relevant information is needed in order to enable them to be used confidently:
 Field data are mostly short-term and poor, with no clear documentation of the amount used or applied,
the amount retained in concrete, or whether it reached the reinforcement and reduced the corrosion rate.
 Independent verification of penetration of the inhibitor into hardened concrete is lacking and sometimes
difficult to do.
 Definitive ratios for the inhibitor/chloride to achieve a given low corrosion rate have not been stabilised.
 For migrating inhibitors, quantitative data on penetration as a function of concrete permeability is lacking.
 Well controlled, long-term field trials on the performance of inhibitors is lacking.

Broomfield (2007) considered that proven inhibitors such as calcium nitrite can be used as admixtures in
concrete to effectively inhibit steel corrosion, provided it is backed up by proper assessment and whole-of-life
costing. He considered that applications to hardened concrete should be done on a trial basis with
appropriate long-term monitoring.

A new technology which would address the leaching problem has been presented by Kupwade-Patil et al.
(2012), using nanoparticles deposited directly onto the reinforcement bars. The so-called electrokinetic
nanoparticle treatment involves deposition of 24 nm nanoparticles via capillary pores onto the steel surface
under the influence of an electric field. This treatment significantly reduced the corrosion rate and reduced
the porosity of concrete in the vicinity of steel by 74%. It also expelled the chloride ions from the paste
surrounding the steel. This technology is very new and needs extensive field studies at this stage.

Austroads 2016 | page 26


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

4.3 Alternatives to Conventional Reinforcing Steel


To reduce reinforcing steel corrosion, the available alternatives to uncoated carbon steel bars are epoxy-
coated steel or galvanised steel, which isolate the steel from moisture and chloride in concrete. Another
alternative is the selective use of stainless steel bars in the most vulnerable zones of the structure, such as
the tidal and splash zones. A brief summary on the use of these alternative reinforcing bars is presented in
ACI (2001).

4.3.1 Epoxy-coated Bars

The main known problems associated with using epoxy-coated steel bars are de-bonding of the coating
before placement in the structures (Sagüés & Zayed 1989; Manning 1996), presence of pinholes in the
coating, development of defects in the coating during onsite handling, and adhesion of concrete to the bars
(Lee et al. 1996). The latter authors found that the thickness of the epoxy coating varies on the surfaces of
the ridges and valleys on deformed bars, and that the coating is susceptible to chloride solutions, where
more defects develop compared to chloride-free water.

With the experience of using epoxy-coated steel bars, standard practice has been established for successful
application. ASTM A775 (2007) specifies that:
1. the coating thickness should be in the range of 130 to 300 microns
2. bending of the coated bar around a standard mandrel should not lead to formation of cracks
3. the number of pinhole defects should be no more than six per metre
4. the damaged area on the bar should not exceed 2%.

Although the intention of epoxy coating reinforcement bars is to provide a barrier between the steel and the
concrete, the integrity of the epoxy-coated bars could not be guaranteed, because of damage that may occur
as a result of exposure to ultraviolet light and handling activities on construction sites.

Lawler and Kruass (2011) reported a survey on bridge decks older than 35 years, some of which contained
epoxy-coated steel bars and some uncoated steel bars. The decks containing epoxy-coated bars were still in
good condition, except a few locations where the thickness of the coating on the bars was found to be less
than 7 mm (180 µm). However, the decks containing uncoated bars showed more than 5% delaminated
area.

Rostam (2005) and Markeset et al. (2006) presented examples of failure of epoxy-coated reinforcement in
saline environments. A number of cases of failure of the epoxy-coated steel bars in the 1990s in Florida have
detracted from the use of epoxy-coated reinforcement bars in Europe, but it seems to be still practised in
North America (Broomfield 2007).

In structures with long design life, long-term durability of the epoxy coating would be essential during
placement and compaction of concrete, as well as under the harsh exposure conditions of concrete (high
alkalinity, high humidity, elevated temperature in warm climates, and loading-unloading cycles for some
members). This type of information is not available. McDonald and Pfeifer (1995) suggested that tests on
long-term durability of epoxy coating products should be performed and should be made standard practice.

Darwin et al. (2007) tested corrosion performance of epoxy-coated steel bars, and showed that although
corrosion occurred in some damaged spots of the coating, the epoxy-coated steel had significantly less
corrosion than its uncoated counterpart, and the performance was further improved by the use of corrosion
inhibitors in concrete. They suggested that the chloride ion at the steel bar tends to concentrate on some
areas so that the corrosion started more quickly in those areas than in other areas, which seem to have been
spared. This effect contributed to less corrosion of the coated bars.

Berke and Hicks (1998) found that mechanically damaged epoxy-coated bars performed well when calcium
nitrite was used in the concrete. However, if calcium nitrite is to be used in a structure, there would be no
reason to use the more expensive epoxy-coated bars rather than uncoated bars.

Austroads 2016 | page 27


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

From the above, it appears that epoxy-coated bars should not be the first choice against corrosion mitigation
in aggressive environments.

4.3.2 Galvanised Steel Bars

Galvanised steel is particularly appropriate for protecting concrete subjected to carbonation, because zinc
remains passive to much lower levels of pH than does carbon steel.

In fresh concrete, zinc dissolves under high pH conditions with the evolution of hydrogen (H 2) as the cathodic
reaction. When zinc-coated (galvanised) steel is used in concrete, a porous layer of concrete can form
around the reinforcing bar, if steps are not taken to prevent it. Boyd and Tripler (1968) stated that a small
amount of chromate salt may be added to fresh concrete to prevent hydrogen evolution, and that calcium
nitrite has been used to prevent hydrogen evolution from galvanised formwork used in precasting.

Modification of the galvanising material and using low-alkali cement may also improve the performance of
galvanised steel in aggressive environments. Bellezze et al. (2006) showed that Zn-Ni-Sn-Bi galvanised
bars, embedded in low-alkali cement concrete, showed high resistance to chloride-induced corrosion
(threshold chloride being 4.02% by cement mass), whereas they were vulnerable to corrosion in high-alkali
concrete. They also showed that the bars with conventional galvanising coat, i.e. Zn-Pb, had a chloride
threshold of 1.73% by cement mass.

Galvanised bars are more tolerant to site handling than epoxy-coated bars, but mixed use of galvanised and
mild steel in a structure would result in the loss of galvanising effect (Broomfield 2007). Moreover,
electrochemical techniques such as chloride extraction cannot be applied to structures incorporating
galvanised steel. It appears that for major structures exposed to aggressive environments, galvanised steel
reinforcement would not be the first choice either, although it may perform satisfactorily in smaller structures,
under mild to moderate conditions.

4.3.3 Stainless Steel Bars

Stainless steel is an alloy containing several elements in addition to iron, and there are many types of
stainless steel with different properties, depending on their compositions. Stainless steel gains its corrosion
resistance from a passive surface layer of chromium oxide, only microns thick. Should the surface be
damaged, the chromium oxide layer will heal using free oxygen from the surrounding atmosphere. This
reaction will take place even at very low oxygen levels. Thus, stainless steel has good resistance to
corrosion and reinforcing bars made of stainless steel have been produced in various sizes in forms which
are ready to be used in construction.

The types of stainless steel used as reinforcement are the austenitic and duplex stainless steels, the former
being non-magnetic, whereas the latter are ferritic and have magnetic properties.

The superior corrosion resistance of stainless steel compared to mild steel has been demonstrated in several
studies (Cox & Oldfield 1996; Flint & Cox 1988; Hewitt & Tullmin 1994; Nurnberger et al. 1993; Treadaway
1978; Treadaway et al. 1989; Zoob et al. 1985).

Sørensen et al. (1990) reported that corrosion attack on stainless steel was more localised than on carbon
steel, and that the critical chloride concentration for corrosion initiation was more than ten times higher for
stainless than for carbon steel. They also found that welding reduced the critical chloride concentration of the
stainless steel due to the combined effects of oxidation and insufficient compaction of the concrete around
the weld. The debris generated from welding acted as initiation points for corrosion attack.

Austroads 2016 | page 28


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Rasheeduzzafar et al. (1992) found that stainless steel-clad carbon bars (duplex) provided similar corrosion
resistance to solid stainless steel after seven years of exposure to chloride-contaminated concrete in the
environment of Eastern Saudi Arabia, and neither showed any sign of corrosion. Manufacturer data states
that the duplex stainless steel, which has higher chromium contents, is more resistant to corrosion than
austenitic stainless steel (ANCON 2016). Note that the Durability Report for the New Gateway Bridge
(Maunsell & AECOM 2007) has used chloride threshold values of 0.5% by concrete mass for Type 316L/LN
(austenitic) and 0.35% for LDX 2101 (duplex), which seems to differ from the manufacturer’s information.

Nurnberger (1996) compared the behaviours of three types of steel, being unalloyed steel, ferritic chromium
stainless steel (1.4003), and austenitic Cr-Ni-Mo stainless steel (1.4571), before and after welding, in
concrete under alkaline conditions, carbonated conditions, and carbonated plus chloride-contaminated
conditions. They found that the latter condition was most aggressive and severely corrosive to unalloyed
steel. Unwelded ferritic chromium steels tolerated conditions of carbonated and chloride-contaminated
concrete, whereas the welded specimens underwent corrosion, particularly in the welded zone. Pitting
corrosion, rather than general corrosion, was observed. The austenitic stainless steel was resistant to all
exposure conditions used, and it was suggested as offering high security both in unwelded and welded
conditions.

Chauveau and Demelin (2007) determined that the pitting potentials of stainless steels 304L and 310L in a
medium of pH 8 and chloride concentration of 0.6 mole/L (2.13% Cl in solution), was 164 mV and 230 mV
(CSE), respectively. According to research results, austenitic steel 304L may be safely used for a chloride
content of up to 5% by weight of cement, but in the presence of welding scale on the steel surface, a lower
chloride content of 3.5% should be assumed (Gjørv 2009). Pickling is a process in which the scales are
dissolved in acid solution, thereby increasing the chloride threshold of the stainless steel.

Randström et al. (2010) tested austenitic and duplex stainless steels in solutions of various chloride contents
(expressed as a percentage of chloride in solution) and determined the corrosion initiation threshold chloride
by potentiometric titration. They found some scatter in the data and concluded that rather than a single
value, it is better to define the threshold as a probability distribution. Nevertheless, the data showed minimum
threshold values of close to 3% for grade 304L, 5.8% for both 316L and LDX 2101 (duplex), 4.5% for 2304
(duplex), and about 0.3% for carbon steel, clearly indicating the much greater corrosion tolerance of the
stainless steels.

Bertolini and Gastaldi (2009) investigated the corrosion resistance of the low-Ni stainless steel in chloride-
contaminated environments. Bertolini and Gastaldi (2011) also demonstrated that high contents of
manganese or a high percentage of ferritic phase in stainless steel caused a significant decrease in the
corrosion resistance of the alloy in both alkaline solution and in concrete.

Houska and Holsing (2012) presented the results of Randström et al. (2010) on the critical chloride threshold
of stainless steel in solution as well as in concrete in the form of probability graphs, which showed that with a
probability of 90%, the chloride threshold will be larger than about 5.8% for stainless steel grades 316, LDX
2101 and 2304. They also showed that the chloride threshold was reduced by about 50% at 40 °C compared
to ambient temperature.

Sørensen et al. (1990) found that the threshold chloride content for corrosion initiation depended on the type
of stainless steel, pH of surrounding concrete, and temperature. The threshold was higher at higher pH and
lower temperature. Markeset et al. (2006) quoted studies from Italy which showed that the chloride threshold
of stainless steel (type 316 or similar) was above 10% by cement mass at pH of 13.9, at both 20 °C and
40 °C. At pH of 12.75, the chloride threshold for the stainless steel was 4% by mass of cement, which
increased to 6% at pH of 13.0. They also stated that the stainless steels remained passive in carbonated
concrete, with a pH of around 9. Studies from the UK (Cox & Oldfield 1996; Flint & Cox 1988; Treadaway et
al. 1989) showed that the stainless steels of 316 and 2205 grades were virtually immune to corrosion.

Austroads 2016 | page 29


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

A clear example of the high performance of stainless steel in concrete structures subjected to extremely
aggressive environments is the condition of the stainless steel reinforced concrete pier at Progreso in
Mexico, reported by Knudsen and Skovsgaard (1999). The old pier built in 1937–41 incorporated stainless
steel reinforcement and has not undergone any major repairs during its lifetime. Despite the poor quality
materials used in construction and the harsh marine environment, corrosion has not taken place in this
structure. In contrast, the adjacent new pier, constructed much later in the 1960s, used ordinary carbon steel
and has been completely destroyed due to chloride-induced corrosion.

Stainless steel can also corrode under certain exposure conditions. ASTM A955 (2012) states that the
chemical composition of the stainless steel alloy shall be selected for suitability to the application involved by
agreement between the manufacturer and the purchaser, because the desired corrosion resistance may not
be provided by all stainless steels.

Freire et al. (2011) tested the corrosion behaviour of stainless steel in solutions with pH varying from 9 to 13.
They demonstrated that at pH 9 in the presence of chlorides, the corrosion resistance of stainless steel
dropped and pitting activity increased, which was attributed to defects formed in the passive film; however,
these defects seem to have been repaired with longer exposure time.

Nickel (Ni) is the expensive element in stainless steel that has increased its price in recent years. To make
stainless steel reinforcing bars more affordable, new types of stainless steels have been manufactured which
contain lower contents of Ni.

Bertolini et al. (2000) found that chloride-contaminated and carbonated concrete did not increase the
corrosion rate of high- Ni austenitic stainless steel containing 17–18% chromium, even at chloride contents
up to 6% by mass of cement. Galvanic coupling of corroding carbon steel with stainless steel grade 316L did
not generate any significant galvanic current, indicating that the corrosion was not aggravated. Bautista et al.
(2006) tested a low-Ni variety of stainless steel (204Cu) in carbonated media, and in non-carbonated but
chloride-contaminated media, and showed that its performance was only slightly lower than that of traditional
austenitic stainless steels when used in highly aggressive solutions.

García-Alonso et al. (2007) monitored the corrosion current density (icorr) of stainless steel bars (8 mm) in
mortars containing 0, 2 and 5% chloride by cement mass. They showed that in the mortar without chloride,
the corrosion rate (icorr) of the low-Ni stainless steel was almost the same as that of carbon steel, and it
remained low in the mortars with chloride.

However, welded low-Ni stainless steel, placed with the weld in the mortar containing 2% Cl-, showed
increased corrosion rates from 0.1 to 0.5 µA/cm2 after about 150 days of testing. They observed that the
passive film of stainless steels was quickly restored once it was broken, and the low-Ni stainless steels
(HSS2 (Ni 1.5%) and HSS3 (Ni 2.3%)) behaved similarly to AISI 304 and AISI 316 steels in the 2% Cl-
mortar. However, passive film restoration did not happen for the welded HSS1 (Ni 0.22%) and HSS3.

Bertolini et al. (1996) showed that the [Cl-]/[OH-] threshold value for stainless steels (304, 304L, 316, 316L)
was 1–4% in a solution of pH 9, compared to 2 and 8 in solutions with pH values of 12.6 and 13.9,
respectively. This result strongly suggests that stainless steel in carbonated concrete will have a much lower
chloride threshold than that in non-carbonated concrete.

Monticelli et al. (2014) found that low-Ni austenitic stainless steels, which are less expensive than high-Ni
stainless steel varieties, are comparatively slightly less resistant to pitting corrosion. However, they are more
resistant to pitting corrosion in carbonated, chloride-contaminated alkali-activated fly ash-based concrete
than in normal Portland cement concrete.

Austroads 2016 | page 30


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Despite the superior corrosion resistance of stainless steel as reinforcement in concrete structures, its use
has been limited compared to carbon steel due to its higher price (by a factor of 5–8 times). For this reason,
stainless steel should be used selectively for areas most prone to chloride-induced corrosion, and its use can
only be justified on a life-cycle-cost basis (Cramer et al. 2002; Knudsen et al. 1998). Val and Stewart (2003)
demonstrated that using stainless steel bars would add a 10% premium to the total project cost compared to
carbon steel bars. However, this was expected to reduce the cumulative costs by 50% over the 120 years of
bridge life. The reduction in life-cycle-cost for structures with longer design life (e.g. 300 years) would be
expected to be even greater, and could easily pay back the additional initial expenditure, compared to the
use of carbon steel.

Selective use of stainless steel appears to be a sound choice for corrosion mitigation of structures in highly
aggressive conditions. Further long-term data needs to be developed in relation to the corrosion behaviour of
stainless steel in very high chloride environments, e.g. 1% or greater by concrete mass, in order to develop
even more confidence in the use of this material. Further data would also be beneficial to be obtained on the
behaviour of stainless steel in carbonated concrete subjected to medium to high chloride contamination.

4.3.4 Carbon Steel Bar Coupled with Stainless Steel Bar

Due to the higher material cost, stainless steel reinforcing bars are used selectively in the zones where
chloride ingress would be a major concern. A practical example of the selective use of stainless clad
reinforcing bars was reported by McDonald et al. (1995) for the deck of the bridge in Trenton, New Jersey,
which was built in 1984. The Oregon Department of Transportation built a bridge in a highly aggressive
marine environment, using highly alloyed stainless steel reinforcing bars in its concrete structures (Scranton
Gillette Communications 2002). The bridge is expected to provide maintenance-free service for about 120
years. The stainless steel was utilised in the most critical structural elements and accounted for only 13% of
the total bridge cost. Other examples of the application of stainless steel as reinforcement in concrete
structures in Europe, North America and Australia are given by Cutler et al. (1998).

The combined use of carbon steel and stainless steel in the same structure could result in a galvanic effect
due to differences in the chemical potentials of the two types of steels, and concern may exist that this effect
would lead to increased corrosion of the carbon steel in the structure. Knudsen et al. (1998) and a report by
the Force Institute (1999) showed that the fear of galvanic corrosion, arising from joining stainless steel and
carbon steel in the same structure, is unfounded, and a study in Australia has confirmed this (Shayan & Xu
2005).

Pérez-Quiroz et al. (2008) showed that, in passive state, the coupling of carbon steel with stainless steel has
little effect on the corrosion behaviour of either material, and the galvanic current was negligible. García-
Alonso et al. (2007) showed the current density of carbon steel, HSS3 stainless steel and the coupled steels
in a mortar containing 2% chloride (per cement mass) was about 10, 0.09 and 0.8 µA/cm 2, respectively. They
considered the galvanic current density to be negligible compared to that of the active corroding carbon
steel, which is similar to the conclusion reached by Shayan and Xu (2005).

The latter work showed that the galvanic effect would increase the corrosion of carbon steel by 10% in a very
high chloride environment (1.97% by mass of cement, or 0.27% by mass of concrete). The measured
galvanic current density was 0.4 µA/cm 2 after three years, and it was slightly higher for the duplex stainless
steel than the austenitic stainless steel. This current density is equivalent to a corrosion rate of 4 µm/year,
which may not be negligible for a structure exposed to an aggressive environment for more than 100 years.
However, it should be noted that the corrosion activity was concentrated in the lapped area where the two
steels were welded. Shayan and Xu (2005) suggested that in replacing the corroding carbon steel with
stainless steel in damaged areas (such as tidal and splash zones), the joint should be placed well above the
splash zone, where the chloride content is much lower. In this case, the galvanic effect would be negligible.

Austroads 2016 | page 31


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

4.4 Defects Formed in Concrete at Early Ages


Deterioration problems in concrete can arise from the following causes:
 Use of unsound materials in concrete, e.g. unsound cement or aggregate.
 Use of incompatible materials, although individual components could be sound, e.g. combination of high-
alkali cement and reactive aggregate, where the latter could safely be used with low-alkali cement.
 Improper mix design, i.e. incorrect proportions of ingredients, e.g. high water content, insufficient cement
content.
 Improper manufacture of concrete, e.g. poor compaction, excessive heat curing.
 Insufficient moist curing of concrete, e.g. early age drying out of freshly placed concrete.
These issues can cause physical defects or chemically-induced cracking and deterioration in the concrete,
depending on the factors involved. Some of the defects occur early in the life of the concrete, such as
settlement cracks and shrinkage cracks, which could easily be prevented through adjustments at the time of
trial mixes. Others may take many years to cause defects, such as alkali aggregate reaction, and another
group occur due to interaction with an aggressive environment over a long period of time, such as corrosion
of reinforcement and sulfate attack. Another problem, delayed ettringite formation, is triggered by an
unsuitable manufacturing process at an early age (excessive heat curing temperature), but its effect is a
chemical reaction that manifests itself after many years. Brief accounts of some of the more significant
problems that affect concrete structures are given below.

4.4.1 Thermal Cracking

Thermal cracking may occur at an early age of concrete if the hydration heat could not be dissipated at a
sufficient rate. This can cause thermal gradients from the core of the component to the edge, and when the
temperature differential exceeds 20 °C, the thermal stress exceeds the tensile strength of the newly set
concrete and the concrete cracks. Although this is a typical problem for mass concrete (Moser 2007), it may
happen for components with relatively large volume when concrete contains a large amount of cement.

The ACI 201.2R-10 report (ACI 2008) states that marine structures often involve thick sections and rather
high cement factors. Such concrete may need to be treated as mass concrete; that is, concrete in which the
effect of heat of hydration needs to be considered.

Reducing the amount of cement is the most effective way to reduce heat in concrete, and partial replacement
of cement with fly ash has been used since the 1950s with success. The amount of fly ash for reducing the
hydration heat depends on the chemical composition and should be tested before construction. Thomas et
al. (1995) showed that heat evolution decreased with increasing level of fly ash replacement, and the level of
ash required to produce a given reduction in heat output increases with calcium content of the ash. Riding et
al. (2008) showed that fly ash reduces cracking risk because of the decrease in the heat of hydration of the
cementitious materials and, to a lesser extent, the increased early age creep.

The advice note in the Design Manual for Roads and Bridges in the UK (Highways Agency 1987) points out
that thermal cracking is a result of restraint of the component and the low early strength of the immature
concrete, and that a sufficient amount of reinforcement should be provided to control the crack width and
spacing, so that the component is serviceable. The formulas for these purposes are provided in this note,
e.g. the amount of steel used in concrete for prevention of cracking is to satisfy the equation (Equation 4):

Austroads 2016 | page 32


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

As f y  Ac f ct* 4

where

As = area of reinforcement in a given direction around the perimeter of a section

Ac = area of effective concrete

fy = tensile strength of reinforcement

tensile strength of immature concrete, which is taken as 0.12(fcu)0.7 (fcu is the


f ct* =
characteristic cube strength)

The design for construction should evaluate the thermal gradient in the component. The temperature
distribution in concrete can be expressed as:

T   2T  2T  2T  Qh 5
c  k  2  2  2  
t  x y z  t

where

c = specific heat

 = density

=
T temperature

k = thermal conductivity

Qh = the hydration heat which depends on the amount of cement and the cement
composition

The thermal stress can be estimated by:

E 6
  Kr T
1
where

Kr = degree of restraint

= modulus of elasticity
E
 = creep factor

 = coefficient of thermal expansion

T =
the temperature change

Austroads 2016 | page 33


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Note that the equations for thermal distribution and stress in the durability report by Maunsell and AECOM
(2007) were not correctly presented.

Possible methods of mitigation would involve choosing the type of cement to reduce heat evolution, including
using sulphate-resisting Portland cement and avoiding high early strength cement, partial replacement of
cement with supplementary cementing materials such as fly ash and slag, and not specifying a too low water
to cement ratio. Using set retarders in the mix can also reduce the intensity of heat release.

Special measures can be taken if the concrete is poured in very warm weather, including cooling coarse
aggregate and using ice as part of the mixing water, flushing the mix with liquid nitrogen, and other
treatments.

4.4.2 Delayed Ettringite Formation

Ettringite is a hydrated calcium-tri-sulfo-aluminate (CaO·Al2O3·3CaSO4·32H2O), also termed the AFt phase,


which forms as an early hydration product of Portland cement at ambient temperature, and is responsible for
the very early strength development of the paste. It forms from the hydration reaction of calcium aluminate in
the presence of calcium sulfate at normal temperature. The reaction is a ‘through solution’ process rather
than a solid transformation process. However, if the curing temperature exceeds about 70–75 °C, or if the
heat of hydration of thick, cement-rich elements rises above this range (as observed by Meland et al. 1997;
Stark et al. 1992; Stark & Bollmann 1997), then a large amount of the sulfate ions would remain in solution,
as well as being retained in the calcium silicate hydrate (CSH) phase. Therefore, ettringite does not form
early in the hydration process, i.e. ettringite is destabilised at such high curing temperatures. Diamond and
Ong (1994) and Shayan (1995a) showed that addition of alkali hydroxide to concrete for AAR testing has
similar effects on ettringite stability and concentration of sulfate ions in the concrete pore solution.

Consequently, instead of ettringite, a less hydrated, sulfate-poor variety, calcium-mono-sulfo-aluminate


(AFm) forms in the concrete. It is believed that, with subsequent exposure to a humid environment, the
sulfate ions which had remained in the concrete pore solution, or been retained by the CSH, will react with
the sulfate-poor phase present in the hardened cement paste to form ettringite (hence, delayed ettringite
formation, DEF). Famy (1999) showed that when delayed ettringite forms within the hardened cement paste,
due to the reaction of mono-sulfo-aluminate with SO42 ions in the pore solution, then paste expansion can
occur due to this crystallisation. Direct precipitation of colloidal, gel-like ettringite in confined spaces and
crack tips is also believed to be involved in the expansion process (crystallisation pressure). This is different
from harmless precipitation of ettringite in open spaces.

Results of laboratory work by Scrivener and Taylor (1993), Scrivener and Lewis (1997), and Famy et al.
(2002) indicate that the sulfate content of the CSH phase increases at high curing temperatures, which could
later be released under moist conditions to enhance DEF. Johansen et al. (1994), and Johansen and
Thaulow (1999) proposed that a homogenous paste expansion takes place as a result of the reaction of the
released sulfate with the aluminate phases in the hardened paste, and this causes gaps around the
aggregate particles at the paste-aggregate interfacial zones, which later get filled with ettringite; the latter not
contributing to the expansion. The issue of DEF-induced cracking has become controversial and so far two
international symposia have been dedicated to the topic in order to discuss viewpoints from various countries
(Erlin 1999; Scrivener & Skalny 2002), but differences in opinions still exist, largely due to different interest
groups (cement suppliers, aggregate suppliers, concrete manufacturers and asset owners) trying to shift the
blame to other groups.

Lawrence (1999) showed that higher alkali content of cement gave larger mortar bar expansion, but could
not explain this observation. This is very likely due to the increased concentration of sulfate ions in the pore
solution of concrete in the presence of elevated alkali hydroxide (Diamond 1994; Shayan 1995a), which
enhances the possibility of DEF taking place. Ettringite can subsequently form when alkali is leached out of
concrete or is consumed by reactions such as AAR. Shayan et al. (1993) and Michaud et al. (1997) showed
that formation of AAR products encourages ettringite formation.

Austroads 2016 | page 34


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Overseas research work has reported cases where deterioration of steam-cured elements or laboratory
specimens has been attributed to DEF; for example, Heinz and Ludwig (1986) and Heinz et al. (1989).
Tepponen and Eriksson (1987) attributed the deterioration of Finnish concrete sleepers to DEF, but Shayan
and Quick (1994) found that AAR was the cause of cracking of these sleepers, and that the conclusions of
Tepponen and Eriksson (1987) were not justified. Similar cases in Texas were presented by Lawrence et al.
(1999). They documented studies of deterioration of cast-in-place and precast concrete elements, in which
completely insufficient attention had been paid to the possibility of AAR being involved, but instead DEF had
been claimed to have been the cause of significant cracking.

Shayan and Quick (1992a, 1992b) found that AAR was the main cause of cracking of Australian concrete
sleepers, and that DEF precipitation in the cracks had occurred as a consequence of AAR-induced
microcracking, rather than the DEF being the cause of cracking. Shayan (1993a) demonstrated that the
granitic aggregate used in Australian concrete sleepers caused significant expansion and cracking in
concrete blocks of the same cross-section as the concrete sleepers (300 mm cubes) without the steam
curing process. The reactivity of these rocks was also independently verified (Shayan 1993b). Oberholster et
al. (1992) also found AAR to be the main cause of cracking of South African concrete sleepers, and that only
those sleepers containing reactive quartzite or granite had cracked, whereas those containing a non-reactive
dolerite had not. Visual manifestations of DEF were also present only in those sleepers which had exhibited
AAR, and not in those containing the non-reactive dolerite.

Shayan and Ivanusec (1996) showed that in laboratory specimens, which had the potential for either AAR or
DEF, or both, only those with AAR potential developed deleterious expansion, and evidence of additional
DEF expansion was noted only when the specimens had the potential for AAR expansion.

Diamond (1994) suggested that growth of ettringite in pre-existing microcracks could cause expansion,
whereas Johansen et al. (1994), Scrivener and Taylor (1993), and Glasser et al. (1995) suggested that it did
not. However, Deng and Tang (1994) and Xie and Beaudoin (1992) stated that if microcracks are filled with a
solution which is supersaturated with respect to ettringite, then crystallisation of ettringite could cause
expansive forces through hydrostatic pressure in the confined areas. This supports the experimental
observations made by Diamond (1994, 1996) and Shayan and Ivanusec (1996) of additional DEF-induced
expansion in AAR-affected specimens, which they suggested was due to ettringite formation in microcracks.
Sahu et al. (1998) did not support the mechanism of uniform paste expansion, neither did they entirely
support the ‘crystallisation pressure’ hypothesis, but concluded that localised (non-uniform) paste expansion
could lead to gaps being generated in concrete, and crystallisation of ettringite therein could further enhance
the expansion.

The tendency of various cements to cause DEF as a result of heat curing has been related to its
composition, with SO3, aluminate and alkali contents being the major factors (e.g. Hobbs 1999; Kelham
1996, 1997, 1999; Lawrence 1999). The presence of a slowly soluble anhydrous CaSO4 in the clinker, which
is alleged to release sulfate at later ages, has been blamed for causing DEF expansion (Marusin 1995; Hime
& Marusin 1999; Kurdowski 2002). Hertfort et al. (2003) strongly refuted Kurdowski’s conclusions, claiming
they were erroneous. This would also apply to similar statements made by Marusin (1995) and Hime and
Marusin (1999). Moreover, Kelham (1999), Michaud and Suderman (1999), and Klemm and Miller (1999)
found that anhydrous CaSO4 in clinker dissolved rapidly and did not cause DEF expansion, whereas added
alkali sulfates did, which also invalidates both Kurdowski’s conclusions and assertions made by Hime and
Marusin (1999).

Mehta (2000) stated that the presence of interconnected microcracks and water are necessary components
of any sulfate-related distress in concrete, requiring the presence of microcracks as a prerequisite for the
development of DEF distress. Shayan and Quick (1992a) had stated that AAR was the mechanism that
caused microcracking in the concrete railway sleepers they studied, and then allowed DEF distress to
develop as a secondary cause. This is different from the mechanism stated by Johansen et al. (1994), which
attributed the microcracking to DEF-induced uniform paste expansion.

Austroads 2016 | page 35


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

It should be noted that the mere observation of ettringite in cracks is not sufficient to conclude that an
expansive DEF process has taken place in the concrete. Scrivener and Lewis (1997) have shown that
formation of XRD-detectable ettringite was unrelated to the expansion of the element concerned. Both
expansive and non-expansive heat-cured mortars contained XRD-detectable ettringite, and even for those
that expanded, increasing amounts of ettringite were detected after expansion had ceased. They related the
observed expansion to ettringite formation within the CSH gel (i.e. paste expansion), and not in the cracks.
Stark and Seyfarth (1999) also found that the extent of damage to heat-cured concrete was unrelated to the
amount of ettringite detected in the element, being about 3% in the cases they examined.

It appears that misunderstanding of this fact led Collepardi (1997) to attribute a case of AAR damage of cast
concrete pedestals to DEF distress, just because ettringite was found in the cracks. In fact, Petrov and
Tagnit-Hamou (2003) have shown that pre-existing microcracks decrease DEF expansion by providing
space for ettringite deposition outside of the solid paste.

In some cases of coastal structures in Australia, multiple deterioration problems have been observed
(Shayan & Morris 2005, 2006), where AAR products, mats of ettringite and corrosion products are all
present. The main cause of the problem in these cases was identified to be AAR. This mechanism caused
concrete cracking, which facilitated precipitation of ettringite and exposed the reinforcing steel to the high
chloride environment, which induced steel corrosion.

Materials other than ettringite have also been found to fill existing cracks and be associated with
deterioration of concrete elements. Shayan (1985) showed that the warping of many large concrete cladding
panels on a multi-storey building was caused by dimensional instability of the concrete when exposed to
wetting and drying cycles, which caused microcracking, and then filling of the cracks with lime, which
prevented closure of the opened cracks (i.e. an overall expansion). Repeated cycles of wetting and drying
caused progressive cracking and filling, which led to incremental expansion, and ultimately to warping of the
panels. Ettringite formation could be involved in a similar role.

In Australia, no field case of DEF alone has yet been identified, and symptoms of DEF have always been
associated with AAR damage. However, in two extensive laboratory experimental programs (Shayan, Xu &
Tagnit-Hamou 2004; Shayan et al. 2006), deleterious expansion was noted in steam-cured concrete (80 °C)
containing non-reactive coarse aggregate, only when high aluminate, high sulfate and high alkali contents
were simultaneously present in the concrete. These conditions are often not met using currently
manufactured Australian cements. Even in this case, to prove that DEF was the cause of expansion, it
remains to be shown that the sand component did not react deleteriously, given the high alkali content of the
concrete.

In relation to prediction of DEF in concrete, attempts have been made by some researchers to relate the
potential of some cements for DEF expansion to their chemical composition (e.g. Kelham 1999; Zhang et al.
2002) However, equations developed for this purpose do not appear to be applicable to other cases. In
relation to detection of concrete mixes susceptible to DEF, test methods have been developed involving wet
and dry cycling of concrete prisms at elevated temperature. Such a test method, originally proposed by Scott
and Duggan (1987) for assessing alkali reactivity of aggregate, is now used for assessing the DEF potential
of concrete mixes. A modification of this test was more recently proposed by Tagnit-Hamou and Petrov
(2004). This test involves an initial heat curing and seven days of storage in lime-saturated water, followed
by thermal cycling (50–10 °C). This test can be used to determine the maximum temperature at which DEF
would not occur in the given concrete.

Prevention of DEF damage in new structures made with Portland cement is different from that of AAR
damage, and involves control of cement composition, such as sulfate, aluminate and alkali contents, in
addition to controlling the curing temperature to below 70 °C. However, the use of sufficient amounts of
supplementary cementitious materials as cement replacement in concrete would eliminate the risk of both
these deleterious mechanisms (Shayan et al. 1993).

Austroads 2016 | page 36


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In Australia, concrete specifications used by various jurisdictions, as well as AS 5100.5-2004 Clause 5.4,
allow 5% Portland cements with SO3 contents as high as 5% to be used, which is considered to be too high,
given that the average sulfate content of the Australian cements is around 2.6%, expressed as %SO 3 by
mass of cement. It is suggested that this value be reduced to a maximum of 3.5%. Blended cements could
tolerate a larger amount of SO3 without causing a durability issue.

4.5 Long-term Durability Problems of Concrete

4.5.1 Carbonation of Concrete

Carbon dioxide (CO2) in air dissolves in water. Its solubility depends on the partial pressure of CO 2 in air, as
expressed by Henry’s law (Equation 7):

p  kH c 7

where

p = partial pressure

c = concentration of solute

kH = Henry’s constant

For CO2 at 25 °C, kH = 3.4  10-2 mol/L·atm-1. In alkaline water, the dissolved CO 2 converts to carbonic acid,
H2CO3 (Equation 8):

CO2 + H2O → H2CO3 8

Carbonation is a process in which CO2 gas diffuses into concrete and dissolves in the pore solution where it
reacts with calcium hydroxide ions generated by cement hydration to form calcium carbonate (Johnstone &
Glasser 1992). At very high CO2 concentration, the CO2 can also attack anhydrous cement grains
(Constantinou & Scrivener 1997). The chemical reaction for carbonation can be expressed by Equation 9:

H2CO3 + Ca(OH)2 →CaCO3 + 2H2O 9

In carbonated regions, the pH of the pore solution decreases, being that for calcium carbonates (pH about 9
or lower), which is lower than the equilibrium pH developed through the cement hydration process involving
formation of Portlandite and C-S-H. These products dissociate to provide more calcium and hydroxyl ions in
the pore solution, when these ions are depleted, e.g. by carbonation, and the reaction cycle continues. So
long as solid Ca(OH)2 is present in the paste, the pH of the pore solution would not fall below 12.5. In fact,
the PH of the pore solution of Portland cement concrete is higher than 12.5 (which is the equilibrium pH for
saturated Ca(OH)2 solution) and can approach values around 13.5–14 due to the presence of alkali metals
(K and Na) in the cement which convert to alkali hydroxide as a result of cement hydration.

Reinforcing steel is protected under the high alkalinity of the concrete, due to the formation of a thin iron oxy-
hydroxide layer which protects the steel from corrosion. However, the steel will lose its passivation in the
carbonated concrete (Figure 4.3) and corrode when moisture and oxygen are available, which is one of the
major long-term durability concerns of reinforced concrete in atmospheric exposure conditions.

Austroads 2016 | page 37


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 4.3: Acceptable chloride content according to CEB recommendations

Source: Comite Euro-International du Beton (1992) – cited by ACI (2001).

4.5.2 Chloride-induced Steel Corrosion – the Threshold Value

In concrete structures exposed to high chloride environments, such as marine conditions or saline soils,
chloride ions penetrate into concrete and may reach the depth of the steel reinforcement after many years.

When the amount of chloride ions reaches a certain level, the steel starts to corrode. This amount is
commonly called the critical chloride content or threshold concentration. The value of the critical content
depends on the concrete type and usage. For reinforced concrete, the critical chloride content is 0.2 to 0.4%
by cement mass depending on the moisture content of the environment and whether or not the concrete was
carbonated (Comite Euro-International du Beton 1992).

The ACI (2001) report concludes that the threshold for uncoated carbon steel reinforcing bar is 0.20%
chloride by cement mass, citing the laboratory work of Clear (1976) and Pfeifer et al. (1987), and field
investigations by Stratfull et al. (1975) and Chamberlin et al. (1977). A comprehensive data review by Angst
and Vennesland (2007) demonstrated a very wide range of threshold values expressed in various ways. In a
review of critical chloride content (Ccrt) in reinforced concrete, Angst et al. (2009) summarised the values of
Ccrt from many studies, which varied very widely depending on whether the steel bars were under outdoor
exposure conditions, in solution or in concrete, and also within each group, depending on the experimental
conditions. The value of Ccrt was as high as 2.5% Cl/cement mass in concrete incorporating supplementary
cementitious materials. The method of corrosion detection could also affect the results.

It has been reported that the actual threshold may depend on the ratio of [Cl-]/[OH-] in the concrete pore
solution, i.e. the pH of the solution is an important factor. The pH varies according to the type of cementing
materials used. Hausmann (1967) showed that at [Cl-]/[OH-] > 0.61, rebar corrosion would be initiated in
concrete. Petterson (1994) showed that the ratio was 2.5 to 6 in a solution where oxygen supply was limited.
A summary of the threshold values according to the survey by Hurley and Scully (2002) is presented in
Appendix B.

The rate of steel corrosion would increase rapidly when the concrete develops cracking. Hwang et al. (1994)
demonstrated that in concrete with 0.1 mm cracks, corrosion of steel occurred even though the [Cl-]/[OH-]
ratio was only 0.023, and there was no clear relation between corrosion rate and concrete type. This
highlights the great significance of ensuring that reinforced concrete structures in aggressive environments
remain free of crack-inducing mechanisms such as AAR.

Austroads 2016 | page 38


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The new amendment to AS 5100.5-2004, Clause 4 (draft 2012) has lowered the acceptable chloride content
in new structural concrete from 0.8 to 0.6 kg/m3. VicRoads’ current specification Section 610 limits the
chloride content to lower than 0.15% by cement mass, which is also equivalent to 0.6 kg/m3 for concrete with
a cement content of 400 kg/m 3. Note that the unit used by AS 5100.5-2004 is mass per concrete volume.
The conversion of this to mass percentage in cement depends on the cement content of concrete, e.g. 0.6
kg/m3 is equivalent to 0.12% by cement mass for concrete containing 500 kg/m 3 cement, and 0.19% for
concrete with 320 kg/m3 cement content, which are below and above the VicRoads’ limit of 0.15%.

Table 4.1 presents the maximum allowable chloride content for new structures. The highest value is that of
BS 8110-1997 (now replaced by BS EN 1992-1-1:2004+A1:2014), i.e. 0.30% by mass of cementitious
material.

The threshold chloride content should be greater than the maximum allowable chloride content, and 0.4% by
cement mass has been commonly used as the criterion for the carbon steel.

Table 4.1: The maximum acid-soluble chloride ion content in new concrete

(AS 5100.5) (ACI 222


(Roads and
Cl- content (VicRoads 610.07) Committee (BS 8110)ǂ
Maritime B80.5)
Concrete per Cl- / Cementitious Report) Cl- / Cementitious
Cl- content per
concrete material Cl- / Cementitious material
concrete volume
volume material
Prestressed B1: 0.4 kg/m3
0.4 kg/m3 0.10% 0.08% 0.10%
concrete B2 & C: 0.3 kg/m3
0.10%, wet
Reinforced B1: 0.4 kg/m3 conditions
0.6 kg/m3 0.15% 0.30%
concrete B2 & C: 0.3 kg/m3 0.20%, dry
conditions
Post–
tensioning — 0.3 kg/m3 0.03% — —
grout

ǂ Now replaced by BS EN 1992-1-1:2004+A1:2014.

In Tasmania, DIER’s Specification B10 (DIER 2006) requires that the total acid-soluble chloride ion content
of the concrete from all mix sources shall not exceed 0.4 kg/m3 of concrete, and the same value is specified
by WA Specification 820 (MRWA 2013). The chloride contents allowed by SA DPTI Specification Part 320
(DPTI 2011) are the same as those of VicRoads.

As discussed in Section 4.3.3, the threshold chloride content of stainless steel depends on the steel type and
is much larger than that of carbon steel, by an order of magnitude. Data available are largely those obtained
through tests conducted in synthetic solutions, and more recently in concrete.

4.5.3 Corrosion Rate Increase with Chloride Content

The rate of steel corrosion increases with the increase in chloride content. For example, Liu (1996)
expressed test results by ln(icorr)  0.618ln[Cl], where icorr is corrosion current density (mA/ft2), and [Cl] is the
mass of chloride per mass of concrete (lb/yd3). DuraCrete (1999) used such a relationship in the probabilistic
modelling.

Research at ARRB Group (Austroads 2004) has shown a nearly linear increase of corrosion rate against the
chloride content of concrete, expressed as a percentage of cement mass, in the range up to 6.3 kg Cl-/m3
concrete (1.53% and 1.97% of cement mass for 50 MPa and 30 MPa concrete, respectively). The corrosion
rate of steel in the 30 MPa concrete was greater than that in the 50 MPa concrete when compared at the
same chloride-to-cement ratios (Figure 4.4).

Austroads 2016 | page 39


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The dependence of the corrosion rate on chloride concentration can be used in predicting concrete
deterioration in the second stage, i.e. the corrosion rate increases with further accumulation of chloride ions
in the vicinity of the steel.

Figure 4.4: Relation between corrosion rate and chloride in concrete. Each point is the average corrosion
current density determined over nine bars during 500 days

The use of supplementary cementitious materials (SCMs) such as fly ash, slag and silica fume, in concrete
influence steel corrosion in different ways. The pozzolanic reaction consumes calcium hydroxide which
reduces the alkalinity of the pore solution. For example, Wiens et al. (1995) showed that 25% slag as cement
replacement resulted in a pH reduction in the pore solution of concrete, which was 11.7 at 730 days, and
was expected to decrease further with time. This may appear to increase the risk of steel corrosion, but the
drop in pH would increase the binding of chloride ions by cement paste, thereby reducing the diffusion of
chloride through concrete, and enhancing its corrosion resistance. Moreover, the pozzolanic reaction causes
pore refinement which physically reduces the rate of transport of chloride ions through the cement paste. The
results of Byfors (1987) showed that the overall effect of incorporation of appropriate amounts of silica fume
and fly ash in the concrete binder was to reduce the chloride diffusion rate into concrete through pore
refinement of the concrete microstructure, and this delayed corrosion initiation of the steel reinforcement.
However, it is known that incorporation of fly ash increases the chloride binding capacity of the binder, which
also reduced the chloride penetration rate through chemical effects.

Monteiro et al. (1985) made cement paste specimens, incorporating various chloride additions, with and
without inclusion of 16% silica fume as cement replacement. They found that specimens with silica fume
were more sensitive to chloride-induced corrosion. This does not imply that silica fume addition increased the
risk of steel corrosion in the concrete. It arose because the authors had already mixed the chloride into the
paste. The silica fume drastically reduced the pH, thereby increasing the Cl/OH ratio, which is conducive to
corrosion. As chloride was already incorporated in the specimen, this experiment cannot show anything
about the efficiency of silica fume in preventing the penetration of chloride ions.

On the other hand, the lower alkalinity resulting from SCM addition increases the electrolytic resistance, thus
the corrosion rate is expected to decrease due to the physical effect of a reduction in permeability. Moreover,
fly ash and slag increase the aluminate content of the paste, which chemically bind chloride ions and
significantly delay the rate of penetration towards the reinforcement. The use of blended cements
incorporating fly ash or slag is a well-known solution against chloride-induced corrosion. For example,
Thomas and Bamforth (1999), studying long-term field data showed that incorporation of fly ash and slag as
cement replacement in concrete may not affect the chloride diffusion coefficient in the short-term, but can
reduce it by an order of magnitude over a few years. Mohammed et al. (2004) showed that carbon steel bars
did not show any sign of corrosion after 35 years of exposure of a reinforced concrete that incorporated
blended slag cement. The results of Thomas et al. (2008) also showed that blended slag cements
incorporating 45% and 65% slag significantly improved the corrosion resistance in reinforced concrete
exposed to marine conditions.

Austroads 2016 | page 40


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Bouteiller et al. (2012) also found that replacement of Portland cement by ground granulated blast furnace
slag at the dosage rate of 65% reduced the penetration of chloride into the concrete, largely through
microstructural improvement. They found that the current international standards for assessing the corrosion
status of steel embedded in Portland cement concrete by electrochemical methods do not apply to concretes
made with blended slag cement. They suggested that rather than single threshold values, a range of values
should be used.

A recent investigation on the durability of very high strength concrete for marine conditions (Lim et al. 2016)
found that the incorporation of SCMs in concrete significantly increased the electrical resistance and
significantly reduced the chloride diffusion rate of concrete.

It is evident that the use of SCMs is an important method of minimising the risk of corrosion, provided that
the concrete incorporates sufficient amounts of SCM and is adequately cured.

4.5.4 Steel Corrosion in Carbonated Concrete

The carbonation process involves diffusion of CO2 gas into concrete pore spaces and dissolution in pore
solution to ultimately form carbonic acid, which reacts with Portlandite in concrete and neutralises its
alkalinity. This reaction reduces the pH in concrete, which is in excess of pH 13, to that prevailing at
equilibrium in saturated calcium carbonate in solution, e.g. pH about 9. At this pH, the passive layer of
reinforcing steel will be damaged and corrosion would be inevitable. Further ingress of CO 2 can worsen the
situation: the partial pressure of the atmospheric CO2 is about 3.5  10 atm, and the equilibrium solution at
-4

25 °C will have a pH value of 8.3, which would be damaging to the reinforcing steel in concrete, as the latter
is normally protected by the high alkalinity (pH > 13) of concrete.

Concrete carbonation is a slow process; thus, in field investigations of reinforcement corrosion, it is often
uncertain as to whether the cover concrete is carbonated. Recent field investigations carried out by ARRB
Group on areas of an old bridge, where the concrete was suspected to have been carbonated, showed that
the corrosion current density determined by a corrosion-rate meter was about 0.1 to 0.2 µA/cm 2 (equivalent
of corrosion rate of 1–2 µm/year). The carbonation front was found not to have reached the level of the
reinforcement bars.

Moreno et al. (2004) showed the corrosion rate of the steel in carbonated concrete (w/c = 0.55) to be below
0.03 µA/cm2 during the time in the carbonation chamber (4% CO2), and increased to 0.3 µA/cm2 when
moved into a high humidity chamber, which indicated that some humidity was needed for the progress of
corrosion. This is the amount of moisture needed for dissolving the CO 2 to form H2CO3, which then reacts
with Portlandite to form calcium carbonate (Figure 4.3).

Constantinou and Scrivener (1997) showed that the corrosion rate of steel bars embedded in a concrete
which was first carbonated in a 100% CO2 atmosphere, and then cured at 70% and 90% RH, was about 1
µA/cm2 which equates to 11 µm/year. Likewise, Law et al. (2003) showed that the corrosion current density
of steel was about 1.3 to 1.7 µA/cm2 in carbonated concrete, where the specimens had been exposed to
100% CO2 atmosphere for two weeks and then subjected to wet-dry cycles for about 1000 days.

The corrosion current density of 1 µA/cm 2 reported above is higher than that of steel bars in slabs made with
concrete containing moderately high contents of chloride (Figure 4.4). This may be due to the effect of
carbonation at 100% CO2 which could make the pore solution acidic (the solution pH would be 5.96 at partial
pressure of 1 atmosphere), encouraging depassivation of the embedded steel bars.

Extensive carbonation measurements conducted on UK structures (Hobbs 1998) showed that the
carbonation depth for similar concrete mixes varied and depended on exposure conditions and the degree of
humidity in the concrete, as well as on the composition of the binder. The onset of corrosion and its rate
would also be affected by these factors. Hobbs et al. (1998), showed that the data on carbonation depth (dc)
in the relationship dc = ktn does not follow the square root of the time relationship (i.e. dc = k t0.5), but that the
value of the time exponent (n) varies depending on the humidity, as shown in Table 4.2, reproduced from
that document. The time period to corrosion-induced cracking (tcr) is also provided by these authors.

Austroads 2016 | page 41


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 4.2: Influence of humidity on carbonation equation parameters and corrosion rate (CR)
RH (%) 80 90 95 98 100
CR (µm/year) 5 10 20 50 10
tcr (year) 20 10 5 2 10
n 0.415 0.317 0.256 0.216 0.187

Source: Hobbs (1998).

The estimated corrosion rates of steel in carbonated concrete of 10 µm/year at 100% RH, and 5 µm/year at
80% RH, are similar to the test results stated above. These authors suggested that a value of n = 0.4 better
fitted the carbonation data obtained from field structures of varying atmospheric relative humidity. The values
of n given above are based on curve fitting, and there is no theoretical basis for them. However, the case for
the t0.5, is based on diffusion theory.

4.5.5 Salt Scaling/Erosion

The surface concrete component standing in water can be scaled off and coarse aggregate exposed, which
is called erosion (or sometimes called water-wash). This is more likely to occur in flowing water where debris
and sand particles brought by the flow contribute to the abrasion. This effect reduces the cover thickness or
causes disintegration of concrete if left unrepaired. Formation of large amounts of efflorescence on the
concrete surface can also cause softening and scaling of the concrete surface.

Due to the fact that all cement hydration products are stable only at high alkaline conditions, e.g. pH > 11,
while the water is mostly neutral or acidic, the water flow would result in leaching of Ca(OH) 2 from concrete,
which in the long run will lead to decomposition of concrete, or at least the zone will have much less
resistance to chloride ingress.

It is essential that the water to cement ratio of concrete is low and the compaction is well conducted to
reduce the erosion. In addition, surface coating on concrete would improve the resistance of concrete to
water erosion.

4.5.6 Acid Water

When water contains acids, the erosion rate of concrete will increase due to dissociation of the cement
hydration products. In natural waters, the most commonly occurring acid is carbonic acid due to
decomposition of plants and other organic matter.

The aggressiveness of water depends on the pH and temporary hardness of water expressed as CaCO 3
(mg/L). When pH is below 6, the concentration of carbonic acid (H 2CO3) and the aggressive carbon dioxide
will be high, which will react with Ca(OH)2 to form CaCO3, which in turn will become dissolved Ca(HCO3)2 to
be in equilibrium with CO2 in water. A guideline for the aggressiveness of water is given in Table 4.3. A
survey in New Zealand concluded that when the contact with acid water was seasonal and the movement of
water was slow, the attack was less severe than that indicated by this table (Eglinton 1998).

Austroads 2016 | page 42


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 4.3: Swedish classification of natural waters containing carbon dioxide

Class of Temporary hardness CaCO3 Aggressive carbon dioxide Degree of aggressiveness to


water (mg/L) (mg/L) concrete
I > 35 < 15 Very slight
> 35 15–40
II Slight
3.5–35 < 15
> 35 40–90
III 3.5–35 15–40 Severe
< 3.5 < 15
> 35 > 90
IV 3.5–35 > 40 Very severe
< 3.5 > 15

Source: Eglinton (1998).

4.5.7 Alkali-aggregate Reaction

General overview

Alkali-aggregate reaction (AAR) is the reaction between the highly alkaline pore solution of concrete and
certain reactive phases in some aggregate types. The alkalinity of the pore solution develops as a result of
cement hydration and formation of a large amount of calcium hydroxide in the solid phase of the hydrated
cement. The solubility of Ca(OH)2 is such that it provides a pH of about 12.5 in the pore solution in contact
with it. However, as a result of other processes in cement hydration and ettringite (hydrated calcium tri-
sulfoaluminate) formation, the alkali sulfates (K2SO4 and Na2SO4) present in the cement are converted into
alkali hydroxides (NaOH and KOH) which remain in the pore solution and make it more highly alkaline. This
increases the pH of the pore solution to the range of 13 to 14.

The reactive components in the aggregate vary depending on the nature of the aggregate, and determine
the type of alkali-aggregate reaction that can occur in the concrete. Generally, alkali reactive constituents
include opaline materials, chalcedony, cristobalite, tridymite, highly strained quartz, cryptocrystalline and
microcrystalline quartz, acidic volcanic glass, amorphous silica, and synthetic siliceous glass. Aggregate
materials containing these constituents include glassy to cryptocrystalline intermediate to acidic volcanic
rocks, and metamorphic rocks such as some argillites, phyllites, greywacke, gneiss, schist, gneissic granite,
metadolerite, metabasalt, vein quartz, quartzite, sandstone, siltstone and chert. These can often, but not
always, be identified by experienced petrographers, using petrographic examination techniques such as
ASTM C295, or AS 1141.65.

Broadly, two types of AAR are recognised: alkali-carbonate reaction (ACR) and alkali-silica reaction (ASR).
In this report, AAR and ASR are used synonymously.

Alkali-carbonate reaction

Certain dolomitic limestones containing argillaceous minerals are susceptible to alkali and cause extensive
cracking when used in concrete of high alkali content. The reactions involved are between dolomite Ca,
Mg(CO3 )2 (Equation 10) and alkali hydroxide denoted by MOH (M = Na or K) (Equation 11).

Ca, Mg(CO3)2 + 2MOH = CaCO3 + Mg(OH)2 + M2 CO3 10

M2 CO3 + Ca(OH)2 = CaCO3 + 2MOH 11

Austroads 2016 | page 43


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The first reaction is called the dedolomitisation reaction, where dolomite breaks down to brucite, Mg(OH)2
and calcite CaCO3. In the second reaction, the alkali-carbonate from the first reaction reacts with Ca(OH) 2 in
the hydrated cement to form more calcite, and as a consequence, the alkali-hydroxide is regenerated to
participate in further reaction with the dolomite.

The dedolomitisation reaction in itself is not an expansive process. It has been suggested that the expansion
observed in the ACR results from two other phenomena. Firstly, the breakdown of dolomite crystals exposes
the dry clay minerals they include. Absorption of water by the latter in the confined zone of the reacted
crystals generates expansion through hydraulic pressure. Secondly, the rearrangement of dolomite into
calcite and brucite produces a physical configuration with microporosity that can absorb water and cause a
volume increase. Brucite crystallisation pressure may also be involved in the expansion. The contribution of
these phenomena to the overall expansion is not clear.

Alkali-silica reaction

This reaction occurs with a wide range of aggregates. Classical ASR is between highly unstable, non-
crystalline silica phases like opaline materials and the alkali hydroxides in the pore solution of concrete. This
produces a highly hydrated alkali-silica gel, which involves large volumetric increases (expansion) compared
to the volume of the original reactive grains, and causes build-up of considerable stresses in the concrete.
When the strains that result from the expansion stresses exceed the tensile strain capacity of the concrete,
the latter cracks and some stress is relieved by flow of the gel into the newly formed cracks. Further stress
build-up causes more cracking and widening of the existing cracks.

Opaline materials are highly reactive and cause significant damage to concrete, depending on the amount
present, the alkali content of the concrete, and moisture available to allow hydration of the gel. Therefore,
three factors are critical for deleterious ASR to occur:
1. adequate amount of reactive phase
2. adequate amount of alkali hydroxide
3. adequate amount of moisture.

Limiting any of the three factors, if possible and practicable, would prevent the reaction from developing into
harmful stages.

ASR is not limited to aggregates containing opaline materials. Aggregates made from several rock types that
contain both silica and silicate minerals can be reactive. The reactive phases in these rocks may be various
forms of silica, viz. opal, tridymite, cristobalite, microcrystalline and cryptocrystalline quartz; quartz in the
form of chalcedony; and highly strained quartz in metamorphic rocks, which usually also contain
microcrystalline quartz.

Previously, a distinction was made between the reaction of siliceous rocks, called alkali-silica reaction, and
that of complex silicate-bearing rocks, such as greywackes, called alkali-silicate reaction. However, this
distinction is now removed, and it is agreed that both are alkali-silica reaction with similar reaction products,
but different reaction rates.

The damage caused by AAR can vary from minor to severe cracking of concrete elements. The cracking can
enhance other deterioration processes by allowing pathways to the interior of the concrete for aggressive
agents.

Aggregates must be carefully evaluated for susceptibility to AAR before being used in the concrete so that
future damage to concrete structures is minimised. Details of testing for the identification of AAR are
discussed later in the report.

Once an aggregate is suspected to be prone to AAR, the use of appropriate amounts of silica-rich mineral
additives, or SCMs, of suitable quality can suppress the reaction and expansion. Careful cost-effective
management is required for existing AAR-affected structures, taking into account the present condition of the
structure, its expansion behaviour and remaining service life.

Austroads 2016 | page 44


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

AAR – a worldwide problem

The first case of AAR was discovered by Stanton (1940) in California, USA, and since then many cases have
been identified in a large number of countries, including Australia, and in a variety of structures. Since the
early 1970s, fourteen international conferences on alkali-aggregate reaction in concrete have been held in
different countries, the 10th conference having been held in Melbourne, Australia in 1996.

The literature on AAR is vast and probably around 3000 papers have been published in international
conference proceedings, as well as in some concrete durability conferences and also international journals.
Moreover, at least four books have been published on AAR by Hobbs (1988), Swamy (1992), West (1996)
and Blight and Alexander (2011). The latter book focuses on structural damage caused by AAR. Many
authors have investigated the effects of AAR on concrete and structures, and a summary of the main effects
was recently presented by Shayan (2016).

The published information deals with a wide range of issues, including mechanisms of AAR, factors
promoting AAR, detection/diagnosis of AAR in structures, characterisation, prevention/mitigation,
assessment of damage, prognosis, repair, strengthening and rehabilitation, and development of tests and
criteria for identification of reactive aggregates. Nevertheless, new aspects of AAR are continually emerging
which necessitate further investigations.

AAR in Australia

Australia was one of the pioneering countries involved in AAR research in the 1940s and 1950s, and the
excellent work of Harold Vivian and others in those years is well recognised worldwide. Idorn (1996) provided
a summary of Vivian’s work in those early years. However, no field cases of AAR were detected in Australia
until the early 1980s. Shayan (1995, 2003) and Shayan et al. (1996) presented the history of AAR in
Australia and summarised new developments in testing and field cases of AAR up to those times. Further
progress has been made since 1996 in the study of AAR, including detection, characterisation, mitigation
and rehabilitation. Shayan (2015) presented the current status of AAR and its mitigation in Australia.

Reactive aggregates in Australia and type of reaction

The aggregate types identified as reactive in the various Australian regions (either by field evidence or by
various test methods) are listed in Table 4.4.

Table 4.4: List of Australian reactive aggregates

State Source
New South Wales River sand, river gravel, quartzitic rocks, deformed granitic (gneissic) rock, dacite,
meta-greywacke, meta-argillite, hornfels, quartz gravel of gneissic origin, rhyodacite tuff.
Victoria Opal from Gippsland, scoria from Mount Noorat, several river gravels, quartz gravel of
gneissic origin, dacite, rhyodacite, sandstone, phyllite, hornfels, quartz from gold mine
tailings, deformed granitic (gneissic) rock.
Queensland Brisbane river sand, quartz-feldspar porphyry, Beerburrum trachyte, ignimbrite/rhyolite,
greywacke, quartzite, river gravel, basalt.
Western Australia Various deformed granitic rocks, meta-dolerite, quartzite, sandstone, various gravels, chert.
South Australia Oodnadatta surface gravel, Coober Pedy quartzite, meta-sedimentary rock (phyllite/slate).
Tasmania Derwent River sand and other siliceous coarse and fine aggregates, meta-quartzite and
deformed granitic rock, basalt.
Northern Territory Tennant’s Creek gravel (sandstone), Katherine River gravel (quartzite), Matarauka gravel
(quartzite), rhyolite.
Australian Capital Deformed granitic (gneissic) rock, acid porphyry, dacite, rhyodacite.
Territory

Austroads 2016 | page 45


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

No reactive, dolomitic carbonate rock, which would be prone to alkali-carbonate reaction (ACR), has been
detected so far in Australia. In recent years, the ACR mechanism has become controversial, and its
existence has been questioned by Katayama and Sommer (2008), who believe that alkali-silica reaction is
responsible for the observed reaction of these rocks. However, they have not proven that ACR does not
exist.

The reactive components of the aggregate types listed in Table 4.4 include either amorphous/or glassy silica,
strained quartz with stress lamellae and patchy extinction, microcrystalline quartz and perhaps some silicate
minerals such as altered feldspar. Shayan et al. (1992) stated that the nature of AAR is the same in these
rock types, and only the rate of reaction may vary. Therefore, the type of reaction that has occurred in
Australia allows no distinction between the form of silica or silicate, and is considered to be of the alkali-silica
reaction. The term AAR refers to this type of reaction in the present report.

4.5.8 Structures Affected by AAR in Australia

As mentioned earlier, field cases of AAR in Australia had remained undetected until their gradual discovery
in and after the 1980s. Table 4.5 presents a compilation of cases identified so far in Australia. The AAR-
affected structures are distributed in all geographical locations of Australia, and no region can be assumed to
be immune from it. A useful outcome of the diagnosis has been the collection of field performance data for
the aggregates concerned, which has been employed for the development of new test methods, through
correlation of field performance data with the results of various laboratory test methods. Development of
such a correlation is very important in generating confidence in the results of any test method proposed for
predicting the alkali reactivity of aggregates with unknown service record.

Table 4.5: Reported cases of AAR-affected structures in Australia up to 2003

Name or type of structure Aggregate Location/Year identified Reference


Upper Yarra Dam Sandstone/phyllite Victoria – 1980 Cole et al. (1981); Shayan
(1989a)
Canning Dam Deformed granitic Western Australia – 1993 Shayan, et al. (2000)
Dam Dacite Victoria – 1987 Shayan (1988a)
Water tank Hornfels North Victoria – 1992 Leamon and Shayan
(1996)
Dam Deformed quartz gravel Victoria – 1996 Shayan (1999)
Dam Phyllite/slate N.A. – 1997 1998, unpublished report*
Dam Deformed granitic N.A. – 1998 1998, unpublished report*
Two dams Deformed granitic N.A. – 1998 1998, unpublished report*
Dam Deformed granitic N.A. – 2002 2002, unpublished report*
Gordon Dam intake tower Quartzite Tasmania – 1992 Blaikie et al. (1996)
Cooling tower Quartzite Tarong, Queensland – Carse (1993)
1992
Bridges (95) Greywacke/ignimbrite/chert Queensland mid 1980s Carse (1988)
Load-out jetty Not reported Western Australia – 1992 Davies et al. (1996)
Causeway Bridge Deformed granite, Western Australia – 1983 Shayan and Lancucki
metadolerite (1987)
Ross and Shayan (1996)
Bridges (4) Metadolerite, deformed Western Australia – 1997 1997, unpublished report*
granite
Bridges (7) Quartz gravel Victoria – 1991–94 Shayan (1994a)
Bridges (2) Hornfels/schist Victoria – 1992 Shayan et al. (2003)
Culvert Hornfels Victoria – 1999 Shayan et al. (2003)
Building Sandstone Victoria – 1999 Shayan et al. (2003)

Austroads 2016 | page 46


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Name or type of structure Aggregate Location/Year identified Reference


Railway bridge Quartz gravel Victoria – 1999 1999, unpublished report*
Precast planks jetty Deformed granitic Queensland – 1999 1999, unpublished report*
approach
Railway bridges (2) Quartz gravel Victoria – 1996 1996, unpublished work*
Bridge Hornfels Nth New South Wales – Shayan et al. (1998)
1998
Bridge Hornfels Nth New South Wales – 1996, unpublished report*
1998
Deep Creek Bridge Quartz gravel Nth New South Wales – Shayan and Morris (2003)
2000
Bridge Quartz gravel Sth New South Wales – Shayan and Morris (2002)
2000
Bridges (4) Hornfels/greywacke Sth New South Wales – Shayan and Morris (2002)
2000
Bridges (4) Deformed granitic/sand- New South Wales – 2003 2003, unpublished*
stone/acid igneous/hornfels
Railway sleepers (1 000 Deformed granitic rock N.A. – 1990 Shayan and Quick (1992a)
000)
Bridge (4) Quartz gravel Victoria – 2005 Shayan and Andrews-
Phaedonos (2005)
Bridge (2) River gravel New South Wales – 2005 Shayan and Morris (2005)
Dam Deformed granite N.A. – 2005 Shayan and Grimstad
(2006)
Bridge Acid igneous rock N.A. – 2008 2009, unpublished report*
Weir Acid igneous N.A. – 2008 2009, unpublished report*
Bridge River gravel Como Bridge, Sydney – Shayan et al. (2008)
2008
Dam Acid porphyry N.A. – 2011 2012, unpublished report*
Bridge Meta-sediment Queensland – 2010 2010, unpublished report*
Dam Acid igneous & N.A. – 2010 2010, unpublished report*
sedimentary
Dam Quartzite N.A. – 2012 2012, unpublished report*
Bridge Quartzite N.A. – 2012 2012, unpublished report*
Several Wharves and Various Miscellaneous 2003–12, unpublished
bridges reports*
Bridge Hornfels, meta-sedimentary New South Wales – 2013 2014, unpublished report*
Bridge Gneissic quartz gravel Victoria – 2014 Shayan, Xu and Andrews-
Phaedonos (2015)
Bridge Schist and Gneissic rocks Victoria – 2015 Shayan, Xu and Andrews-
Phaedonos (2015)

* Entries
in Column 4 which are labelled with an asterisk relate to confidential reports by Dr A. Shayan, which have not
been published.

4.5.9 Diagnosis of AAR and Characterisation of Concrete

Two examples of structures affected by AAR in Australia and their cracking features are given in Figure 4.5
and Figure 4.6. Symptoms of the reaction within the concrete and the crack pattern can vary depending on
the age of concrete, type of aggregate, and severity and extent of reaction, as well as amount of
reinforcement or prestressing. Cracks are often parallel to the direction of prestressing.

Austroads 2016 | page 47


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 4.5: Map-cracking caused by AAR in a large bridge pylon (left) and in a support column of a water
diversion gate in a dam (right)

Figure 4.6: Parallel cracking caused by AAR in a prestressed railway sleeper

Figure 4.7 shows an example of the manifestation of AAR within the concrete. Such features are not always
obvious, particularly in the case of early or mild reaction. The reaction products are the cause of concrete
cracking.

Austroads 2016 | page 48


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 4.7: AAR manifested by reaction products around aggregate particles

The diagnosis of AAR consists of visual examination of the structure concerned, core drilling from concrete
elements, preparation of appropriate specimens for stereoscopic, petrographic and scanning electron
microscope (SEM) examinations; the latter are often associated with analytical capability such as energy-
dispersive X-ray (EDX) analysis. Not only AAR, but other secondary chemical reactions and microstructural
features of concrete are identified, and information on the constituents of concrete are obtained through
these examinations. Shayan (1994b) published an illustrated guide for the identification of AAR in Australian
concrete structures, which was intended to assist asset owners for this purpose.

Characterisation of the affected structure is then conducted by the following procedures:


 determination of the strength properties of the concrete
– compressive, tensile, bending strength and elastic modulus
 determination of the residual active alkali content in concrete
 determination of the residual expansion potential of the concrete.

These parameters are very important in determining the rehabilitation strategy of the affected structure.
Detailed analysis of these parameters is outside the scope of this report, although a brief account is
presented below.

There are many papers in the AAR literature from different countries, particularly those in the proceedings of
the past 14 International AAR Conferences, which describe diagnostic features of AAR in concrete and
characterisation of the affected concrete.

Characterisation of affected concrete

The visual, petrographic and SEM examinations can reveal the microstructural features of the affected
concrete and the extent of deterioration. The estimation of the residual strength properties and expansion
potential of concrete, and whether or not the deterioration process would continue are of great importance in
deciding on an appropriate repair methodology.

Austroads 2016 | page 49


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The structural effects of AAR have been discussed in The Institution of Structural Engineers (UK) Report
(1992), and by Clayton (1989), Hobbs (1998) and Swamy (1992), amongst other researchers. It is generally
accepted that the expansion and microcracking resulting from AAR adversely affects the engineering
properties of concrete. Particularly, the tensile and flexural strength and elastic modulus are significantly
reduced by AAR, but the compressive strength is less sensitive and is usually reduced to a lesser extent. It
has also been shown that applied restraint can suppress AAR expansion, and this can be used to advantage
in the repair of some AAR-affected elements.

It has been claimed by some researchers that AAR expansion increases the load capacity of the affected
elements by inducing some degree of prestress in the steel reinforcement. However, although a small
amount of expansion may have the alleged effect, excessive AAR expansion is definitely deleterious and
could even induce yielding of mild steel, and cause extensive concrete cracking and spalling. In addition to
the adverse strength effects, AAR cracking can cause significant additional durability problems, depending
on the exposure conditions. In some Australian structures, AAR cracking has had adverse effects on the
corrosion of steel reinforcement in the affected elements, such as bridge piers. In the latter case, the AAR
cracking has allowed penetration of aggressive salts into the concrete and complete deterioration of a
section of the concrete. Significant reductions (30–50%) in the strength properties of AAR-affected, precast,
prestressed concrete deck planks have been reported (Shayan & Morris 2002), and the entire deck has
recently been replaced because of AAR damage.

The residual expansion of the AAR-affected concrete can vary in a structure depending on the extent of
reaction and microcracking, residual alkali content, remaining reactive components and exposure conditions.
Where significant reaction has already taken place, and existing microcracking can accommodate newly
formed AAR gel, the residual expansion could be small. In Australia (Shayan 1999), these features were
exhibited in a severely affected dam wall, whereas concrete from a dry tunnel exhibited milder AAR
symptoms but larger residual expansion.

In Canning Dam (Shayan et al. 2000), the mild residual expansion of the dam wall (0.017%) was confined by
post-tensioning and anchorage of the dam wall into the bedrock. Other properties of concrete may need to
be determined, depending on the deterioration types and stage of deterioration of the concrete before
deciding on the repair technique. For instance, corrosion of reinforcement may be involved. A 20-year old
bridge in salty tidal water in Australia had a significant corrosion problem in addition to AAR (Shayan &
Morris 2003), which needed to be addressed in the rehabilitation of the bridge. Several other bridges,
including the Causeway Bridge in Perth, Western Australia, are in similar condition and need to be
individually characterised.

4.5.10 Test Methods for Detecting Reactive Aggregates

Developments of AAR test methods in Australia in the past five or six decades were summarised by Shayan
(1995a) up to that time. Until the early 1990s, Australia used the three test methods traditionally used in the
USA, i.e. the petrographic method (ASTM C295), the mortar bar test designated ASTM C227 (in Australia
AS 1141-38 – now withdrawn) and the quick chemical test designated ASTM C289 (in Australia AS 1141-39
– now withdrawn). An Australian standard is now available for petrographic examination (AS 1141.65).
Although both ASTM C-295 and AS 1141.65 are useful in identifying the aggregate constituents and their
textural features, they cannot be considered definitive for classing aggregates as non-reactive, particularly
for very fine-grained rock types. However, some reactive aggregates which contain known reactive
components can be classed as potentially reactive by petrographic methods. Nevertheless, some expansion
testing is required to quantify the level of reactivity.

Austroads 2016 | page 50


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In the 1980s, when several field cases of AAR damage were identified, it was recognised that these methods
failed to detect many of the slowly reactive aggregates, and accelerated test methods were developed
(Shayan et al. 1988; Shayan 1989b). It was then established that the results of the accelerated mortar bar
test (AMBT), which could be conducted within 24 days, correlated well with field performance of a number of
known aggregates (Shayan 1992). A concrete prism test, using a cement content of 410 kg /m3, was also
found to be satisfactory, except that it could not detect the reactivity of some slowly reactive aggregates.
Since then, the AMBT method has been adopted by a number of road agencies in Australia, such as Roads
and Maritime Test Method T363 in New South Wales, VicRoads Test Method RC376.03 in Victoria, and
Main Roads WA Test Method WA 624.1 in Western Australia. In Queensland, TMR has adopted a concrete
prism test using 50 °C curing temperature (Carse & Dux 1990).

Efforts since 1995 have involved consolidation of the Australian AMBT by testing many aggregates, including
those that have caused damage to major concrete structures. In the mid-1990s, ASTM C1260 was
introduced in North America as an AMBT based on experience with North American aggregates. This test
was based on the South African method (Oberholster & Davies 1986) and has less stringent acceptance
limits than the Australian AMBT (e.g. RMS T363).

Relatively recent work (Shayan & Morris 2001) has shown that:
1. ASTM C1260 produces higher mortar bar expansion than the Australian AMBT at a given age for reactive
aggregates, whereas the two methods produce the same expansion curves for the slowly reactive
aggregates. This has been attributed to the effect of the higher w/c ratio in the ASTM C1260 method (i.e.
higher porosity) on the penetration of alkali solution into the concrete and better access to the reactive
aggregate particles which are susceptible to alkali. The slowly reactive aggregates are less responsive to
this effect, as they are less susceptible to alkali.
2. Regardless of which procedure is used for the manufacture of mortar bar specimens, the reactive
aggregates can be detected by both the Australian AMBT and the ASTM C1260 methods, whereas the
slowly reactive aggregates are identified only by the Australian limit of 0.10% at 21 days.
3. The Australian AMBT is more appropriate for the slowly reactive aggregates which are not detected by
the ASTM C1260 limit of 0.20% at 14 days (Shayan & Morris 2001), nor by the corresponding Canadian
limit of 0.15% at 14 days.
4. The concrete prism test method cannot detect dangerous, slowly reactive aggregates which are detected
only by the expansion limit of 0.10% at 21 days in the Australian AMBT (Shayan 2001). It has been
suggested that the ASTM C1260 limit should be lowered to 0.08% at 14 days to enable it to detect the
slowly reactive aggregates, or it should adopt the 0.10% expansion at 21 days.

The above features are illustrated for representative reactive, slow reactive and non-reactive aggregates in
Figure 4.8.

The limit of the Australian AMBT of 0.10% at 21 days for slowly reactive aggregates was confirmed as an
appropriate durability performance measure for coarse aggregates (Shayan et al. 2003). A strong correlation
has been observed between the results of the AMBT and those of the concrete prism test (CPT) and field
performance of many aggregates (Shayan 1992; Shayan et al. 2003), as indicated by Figure 4.9. Moreover,
Shayan (2007) demonstrated that the limits of the Australian methods are more appropriate than those of
ASTM C1260 for the Australian slowly reactive aggregates. The latter would accept some dangerous slowly
reactive aggregates.

Supplementary investigations have shown that elevated temperatures (up to 80 °C) enhance the rate of the
reaction, but do not alter the nature of the AAR products (Shayan & Quick 1989).

In addition to the AMBT, some road agencies also make use of concrete prism tests (CPTs); for example,
the RMS T364 and VicRoads RC376.04 test methods, which are similar and are conducted at 38 °C, 100%
RH (CPT38). These methods are broadly similar to the ASTM C1293 concrete prism test and the Canadian
(CSA A23.2-27A and CSA A23.2-28A) test methods, although the aggregate grading and the test limits are
different from the Australian method. All these tests take at least one year to produce results.

Austroads 2016 | page 51


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

However, some slowly reactive Australian aggregates do not adequately respond to the CPT38 test, even
though these aggregates cause significant damage to field structures over many years after construction
(Shayan 2007). A modified version of CPT38, which is conducted at 60 °C (CPT60), was found to
unambiguously detect such aggregates within four months (Shayan et al. 2008), and was preferable to both
AMBT and CPT38 for these aggregates. It should be noted that the CPT60 test is under trial testing in
Europe and North America.

Based on research conducted on the Australian aggregates (Shayan & Morris 2001, Shayan 2007), it was
proposed by Shayan (2011) that the limits of the AMBT methods RMS T363 and VicRoads RC376.03, be
adopted in Australia, and that CPT38 be replaced with CPT60 in the Australian road agencies’ concrete
specifications for structural concrete.

Fortunately, a newly published AMBT in Australian Standard AS 1141.60.1-2014 employs the limits which
have been developed for RMS T363 and VicRoads RC376.03.

It is believed that application of the RMS T363 and VicRoads RC376.03 AMBTs, in combination with the
corresponding CPTs (RMS T364 and VicRoads RC376.04), would ensure that reactive aggregates would be
detected and not used in new structures without appropriate precautions, which is essential in achieving
service lives as long as 100 years for significant concrete structures.
BGT Gneiss

Figure 4.8: AMBT results for reactive dacite aggregate (top0.35


left), non-reactive basalt (top right) and two slowly
reactive gneissic granite aggregates (bottom) 0.3
MRD Rhyodacitethe two test methods indicated
by Gordonvale basalt
(%)

0.25
ExpansionExpansion

0.8 0.25
0.2
0.7 0.15 RTA
0.2
(%)

0.6
Expansion (%)

0.1
ASTM
0.5 0.15
0.05 RTA
0.4 0
Series
0.1 3ASTM
0.3 0 10 20 30 Series 40
RTA
0.2 0.05 4Series
Time (day)
0.1 ASTM 3
0 Series
0 Series
0 10 20 30 3 40 0 10 20 30 4 40
Series Time (day)
Time (day) 4 BGT Gneiss
CNN gneissic granite
0.35 0.3
0.3 0.25
Expansion (%)

UPY SAN
Expansion (%)

0.25 0.2
0.8 0.2
0.15 French Coarse 30% Exl70%
0.70.15 RTA RTA
0.6 0.1 0.1
Expansion (%)

ASTM 0.4
ASTM
0.50.05 0.350.05
0.4 0
Series Series
Expansion (%)

0.3 0
3 3
0.3 0 10 20 30 Series 40 0.25 0 10 20 30 Series 40
RTA
0.2 4 0.2 Time (day) 4
Time (day)
0.1 ASTM 0.15
0 Series k 0.1 RTA
0 10 20 30 3 40 0.05
Series ASTM
Time (day) 4 0
CNN gneissic 0granite 10 20 30 Series 40 MKR gneissic granite
3
0.3 0.25 Time (day) Series
4
0.25 0.2 Series
Expansion (%)

Expansion (%)

UPY SLT RTA


5
0.2 0.15
0.7 ASTM
0.15
0.6 0.1 Series
RTA French Fine
0.1 3
Expansion (%)

0.5 0.05 Series


ASTM 0.35 4
0.4 0.05 0
Series 0.3
0.3 0
Expansion (%)

3 0.25 0 10 20 30 40
0.2 0 10 20 30
RTA Series 40
0.2 Time (day)
Time (day) 4
0.1 ASTM 0.15
0 Series 0.1 RTA
0 10 20 30 3 40
Series 0.05 ASTM
Time (day) 4 Austroads 2016 | page 52
0
Series
0 10 20 30 3 40
MKR gneissic granite
Series
Time (day)
0.25 4
Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 4.9: Comparison of AMBT and CPT results for a large collection of aggregates from various projects

AMBT-CPT Correlation
0.2
CPT-1 year expansion (%)

0.15

0.1

0.05

0
0 0.2 0.4 0.6 0.8
AMBT-21 day expansion (%)
Test Results CPT-0.03% limit AMBT-0.10% limit AMBT-0.15% limit-sand

Unfortunately, the CPT method published as an Australian Standard (AS 1141.60.2-2014) contains a lower
alkali content that those of RMS T364 and VicRoads RC376.04 (i.e. 1.25% rather than 1.38% by mass of
cement) and would not detect some slowly reactive aggregates which are just detectable by using the higher
alkali content. In other words, the 1.25% alkali level is inadequate for identifying some slowly reactive
aggregates, and needs to be amended.

Mitigation of AAR

As mentioned earlier, three essential ingredients must combine in the concrete for AAR to take place, i.e.
sufficient amounts of reactive components, active alkali and moisture. Lack of any of the ingredients would
suppress AAR in the concrete, and these may be taken as the controlling factors in the mitigation of AAR.

Therefore, the mitigation of AAR could involve selecting an aggregate which is free of deleterious
components, although an aggregate may contain a small amount of a mildly reactive component without
causing deleterious expansions in concrete. This would be the safest approach. Another approach may be to
limit the amount of active alkali in the concrete, such that the amount of alkali is insufficient for causing
deleterious expansion. This approach may involve using low-alkali cement in the concrete, with low total
alkali content. The effectiveness of this approach is less certain as some aggregates can release alkali into
the concrete pore solution, especially when used as fine aggregate, and could enhance the total alkali
content to dangerous levels, resulting in excessive expansion and cracking of concrete.

Controlling the moisture content of concrete is the least certain approach as it may be possible only for thin
internal concrete elements in well-ventilated environments. It is often impossible to isolate the concrete from
moisture, such as in hydraulic structures or submerged elements. In large concrete elements, the moisture
content of concrete is often sufficient for deleterious AAR, and access to further environmental moisture
would then exacerbate the expansion and cracking.

Apart from the use of non-reactive aggregate, and in locations where the use of reactive aggregate is
unavoidable, the most reliable method of AAR mitigation is the use of appropriate supplementary
cementitious materials (SCMs) in the concrete formulation. Extensive literature exists on the utilisation of
traditional SCMs, such as fly ash, silica fume and blast furnace slag, for the suppression of AAR in concrete.
Natural pozzolans and meta-kaolinite have also been used elsewhere for this purpose.

In Australia, silica fume, low calcium fly ashes and blast furnace slag, have all been effective in suppressing
deleterious expansion of reactive aggregate in concrete (Shayan 1992, 1995b).

Austroads 2016 | page 53


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Chemical compounds such as lithium salts (Stokes 1996) or lithium-bearing glass (Stokes et al. 2000) are
not currently favoured in Australia for suppressing AAR expansion, as better and cheaper alternatives are
available; although original work on lithium salts dates back to the 1950s (McCoy & Caldwell 1951). For
existing structures, driving the lithium ions through the concrete, similar to the re-alkalisation process, has
been suggested as a method of reducing concrete expansion (Whitmore & Abbott 2000). Although this
technique may be effective for the cover concrete around the reinforcement bars, it would not affect the
centre of the element which would still expand and cause cracking in the concrete.

In recent years, glass powder, made from waste bottle glass, has been shown to be effective in suppressing
AAR expansion (Shayan & Xu 2004). Extensive laboratory research was conducted on the use of glass
powder as a pozzolanic material in concrete, leading to a field trial (Shayan & Xu 2006), which produced very
promising results. It was shown that glass powder was effective in suppressing AAR expansion in long-term
laboratory expansion tests. The high alkali content of the glass powder did not appear to cause expansion of
a reactive aggregate which was combined with low-alkali cement and glass powder, i.e. the large alkali
content of the glass powder did not contribute sufficiently to the soluble alkali content of concrete to cause
deleterious AAR expansion. It was also found that for a 40 MPa mortar mixture, the strength gain of mortar
containing silica fume and glass powder was comparable on the basis of equivalent cement content.

Rehabilitation of structures affected by AAR

Examples of major rehabilitation projects for AAR-affected structures can be found in the proceedings of the
international conferences on AAR held since the early 1990s. These include major bridge and dam
structures, and cover various aspects of repair and rehabilitation, such as surface coatings, replacement of
affected concrete, post-tensioning, prestressing, application of fibre composites, and slot cutting to relieve
AAR expansion pressure.

It is important to determine the current mechanical properties and residual AAR potential of the concrete
prior to any repair action, so these properties of concrete are known and can be taken into account in the
design of the rehabilitation. It is also important to monitor the behaviour of the repaired structure by using
embedded sensors, so that the effectiveness of the repair can be assessed. Torii et al. (2000) applied a new
concrete prestressing technique involving confinement of AAR-affected columns in a large-scale laboratory
model as well as in a bridge. Embedded sensors have shown the effectiveness of the technique in confining
the AAR expansion and cracking, and preventing strength reduction.

Wigum and Thorenfeldt (2004) showed that wrapping AAR-affected mortar cylinders effectively prevents
volume changes in the specimen. This was attributed to the fact that the elastic modulus of the AAR-affected
cylinders had severely deteriorated and the confining pressure from the wrapping was sufficient to contain
the volume changes. Mohamed et al. (2006) confined small non-reinforced cylinders of reactive and non-
reactive concretes and found that confinement by CFRP (uni-dimensional carbon fibre of HEXCEL of 200
g/m2) reduced the longitudinal strain by 21% and transverse stain by 75%. The difference may have arisen
because of the fibre direction, being more effective along the fibre direction (i.e. transverse direction on the
cylinders).

Several cases of repair to reinforced concrete structures in Japan, which were damaged due to AAR,
involved steel plating of the damaged elements (Seto et al. 2004; Nomura et al. 2004; Torii et al. 2004).
These structures have exhibited rupture of reinforcement bars in the areas of bends in the reinforcement.
Although it was originally thought that the AAR-induced stresses had ruptured the steel bars, later studies
showed that defects in the manufacturing of the bent reinforcement bars were responsible for the rupture,
which was magnified under the influence of AAR stresses (Torii et al. 2012).

Unlike North America, UK and Japan, research and practice on the rehabilitation of AAR-affected structures
is rather limited in Australia. The Australian cases of repair have included silane treatment (unpublished work
and Shayan 1995b), concrete jacketing (Carse 1996), patching and surface coating (Davies et al. 1996),
crack injection and surface coating (unpublished), and concrete replacement and post-tensioning (Shayan et
al. 2000). Badly deteriorated sections of retaining walls in a dam were recently replaced (unpublished).

Austroads 2016 | page 54


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In the past decade, the use of fibre composites, such as carbon fibre reinforced polymer (CFRP) products,
has been considered for the repair of AAR-affected structures. Shayan et al. (2008, 2012) used AAR-
affected columns with square cross-sections and circular sections, respectively, to investigate the
effectiveness of confinement by CFRP wrapping compared to conventional jacketing with reinforced
concrete. In the case of square section columns, the CFRP was one layer of 240 GPa elastic modulus. Both
types of jacketing reduced the rate of expansion, but did not suppress it to safe levels. The strain developed
in the CFRP did not exceed its elongation capacity. The reinforced concrete jacket developed vertical
cracking in the centre of one side. Given the highly reactive aggregate used, it is possible that the tensile
strain capacity of the CFRP could be exceeded over a few years.

In the case of circular section columns, up to three layers of CFRP with both 240 GPa and 640 GPa elastic
moduli were applied. The wrapping again reduced the expansion rate, and higher modulus CFRP was more
effective. Three layers did not improve the confining effect of CFRP. Early age application of the material
was more effective than later applications.

Similar results were obtained in a parallel study (Abdullah et al. 2012a) with respect to strain development in
wrapped columns, but another phase of the latter study (Abdullah et al. 2012b) also showed that CFRP
wrapping significantly enhanced the mechanical properties of the affected columns.

Lack of information on the long-term durability of the FRP systems under aggressive environmental
conditions is considered a barrier in the use of these products. Nevertheless, the use of such composite
systems is expected to increase in the future.

When AAR-affected concrete also suffers from chloride-induced corrosion, the repair becomes very
complicated as mitigation of corrosion by impressed current cathodic protection (ICCP) is able to aggravate
the AAR expansion.

This effect has been shown by Alkadhimi and Banfill (1996), Kuroda et al. (1996), Sergi and Page (1992),
Page et al. (1992), Page and Yu (1995) and Torii et al. (1996), for electrochemical processes including ICCP,
chloride extraction and re-alkalisation. Shayan (2000) found similar results for reactive Australian aggregates
based on measurement of linear expansion of reinforced concrete prisms subject to ICCP currents of 25, 50
and 75 mA/m2 for two years. From the axial expansion measurements, it appeared that the 25 mA/m 2 current
density had the largest effect on expansion. However, careful inspection of the same specimens at the age
of three years (Shayan & Xu 2001) found that considerable longitudinal cracking had occurred at the bottom
and side surfaces of the specimens, and that the degree of cracking and crack width increased at higher
current levels. The reason why earlier axial expansion measurements did not detect this effect was due to
the longitudinal nature of the cracks, which were parallel to the prisms’ axis and to the direction of expansion
measurement. The effects are, therefore, reflected largely in lateral expansion, and not in the axial expansion
of the concrete prisms.

Several aspects of AAR and its mitigation in Australian concrete structures are covered in the new version of
HB79 ‘Guidelines on minimising the risk of damage to concrete structures in Australia’ (SA HB 79:2015).

It is obvious that the best way of dealing with AAR is to use non-reactive aggregates by appropriately testing
the materials before use, or employing appropriate concrete mix formulations to suppress the aggregate
reactivity, if alternative aggregates are not available. In any case, repair/rehabilitation methodologies should
be provided as part of durability planning for the structures concerned with respect to AAR issues.

Austroads 2016 | page 55


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

4.5.11 Sulfate Attack

Sulfate attack on concrete is thought to be due to the formation of expansive compounds, such as gypsum
(CaSO4·2H2O), ettringite (CaO·Al2O3·3CaSO4·32H2O) and thaumasite (Ca3Si(OH)6(CO3)(SO4)·12H2O), in
hardened concrete through the reaction of sulfate ions with cement hydration products, largely Portlandite
Ca(OH)2, calcium aluminate hydrates (e.g. C4AH13), monosulfate (AFm) and calcium silicate hydrates, or the
CSH. This process can lead to disintegration of the cement hydration products, which changes the
microstructure of cement paste. Recent studies by Lothenbach et al. (2010) indicated that the expansion
involved in the sulfate attack arose from crystallisation pressure of ettringite in small pores, rather than a
change in the volume of solids.

External attack

Sulfate attack on concrete can be either external or internal. External attack relates to exposure of concrete
to sulfate-rich environments such as some soils and groundwaters, and in sewage pipes. The formation of
ettringite and gypsum is through the reaction of sulfate ions which penetrate the concrete from external
sources, and calcium and aluminium ions in the pore solution of concrete. The supply of calcium and
aluminium ions is mainly from the Ca(OH)2, C4AH13 and calcium mono-sulfoaluminate, which are
decomposed by the action of sulfate ions, leading to weakness of the cement matrix. In severe cases, the
calcium silicate hydrate phase may also react with sulfates and undergo disintegration. Calcium may also be
supplied from limestone powder, used as an additive to the cement, or from limestone aggregates (Khayat et
al. 2005). The reactions involved in the attack are described by Equation 12 and Equation 13.

Ca(OH) 2 + Na2 SO4 = Ca SO4 + 2NaOH 12

C4AH13 + 3CaSO4 + H2O = C3A . 3CaSO4 . 31H2O + Ca(OH)2 13

The external attack causes progressive disintegration of the concrete surface through these reactions. If
sulfate solutions can penetrate to the interior of the concrete through cracks or an open pore structure, then
formation of the reaction products in a confined space may cause large expansions.

The attack can be minimised by lowering the quantity of susceptible phases in the hydrated cement, which
are reactive to sulfate. Making the concrete less permeable, e.g. low water/cement ratio, would also help.

The amount of Ca(OH)2 in the concrete can be reduced by the addition of SCMs to the concrete mix, which
would improve its microstructural properties and improve permeability. The amount of aluminate phases can
be reduced in the concrete by the choice of so-called sulfate-resisting cements which are low in C3A
(3CaO·Al2O3) content. Khatri et al. (1997) showed that both the composition of paste and low permeability
were influential in reducing the sulfate-induced expansion of concrete or mortar specimens immersed in
soluble sulfate solutions.

When sulfate is in the form of magnesium-sulfate, e.g. in seawater, the reaction occurs as shown in
Equation 14.
Ca(OH)2 + Mg SO4 = Ca SO4 + Mg(OH)2 14

Because of the very insoluble nature of Mg(OH)2 (called brucite), this reaction continues to the right until all
of the Ca(OH)2 is consumed.

Austroads 2016 | page 56


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In seawater-induced sulfate attack, magnesium-sulfate attack is the most aggressive, which is due to the fact
that its content in sea water is higher than the other types of sulfates. Magnesium also reacts with the
cement hydration product (CSH) to produce an Mg-substituted variety, i.e. MSH, which may not have the
same binding properties as CSH. The formation of brucite occurs at a lower pH (pH = 10.7 at 20 °C), at
which most cement hydrates, i.e. Ca(OH)2 and CSH are not stable, so the reaction proceeds to completion
(Eglinton 1998). In the process, the cement hydration products lose calcium and silicate ions, which changes
the microstructure of cement paste.

Depending on the type of sulfate ions, by-products having no bonding strength would form, e.g. brucite
(Mg(OH)2) in the presence of MgSO4 brine (Bonen 1997). Thaumasite was found in some buried concrete;
for example, bridge foundations (Floyd & Wimpenny 2003).

The use of sulfate-resisting cements is essential for protection of the concrete against sulfate attack. Incorporation
of pozzolanic materials, such as fly ash, slag and silica fume into concrete, consumes the Ca(OH)2 in concrete,
leading to better resistance against sulfate attack (Giergiczny 1997; Plowman & Cabrera 1996; Prusinski &
Carrasquillo 1995; Yeginobali & Dilek 1995). As mentioned above, improved pore structure of concrete and the
resulting decrease in permeability also improves the sulfate susceptibility of the concrete. Xu et al. (1998)
conducted a detailed review of the mechanisms of sulfate attack, which includes additional details.

Table 4.6 shows the severity classification for exposure to sulfate, and possible preventive measures against
sulfate attack, recommended by ACI (2008). Table 4.7 shows the corresponding table for the UK (BS 8110-
1997, cited by Eglinton 1998; (now replaced by BS EN 1992-1-1:2004+A1:2014)). Other tables given below
show the sulfate aggressiveness of the various environments for concrete (Table 4.8 to Table 4.10). Types of
suitable cements recommended for such environments are given in Table 4.11.

Table 4.6: Requirements to protect against damage to concrete by sulfate attack from external sources of sulfate

Severity of Water-soluble Sulfate (SO4) in w/c by mass, Cementitious material


exposure sulfate (SO4) water, ppm max. requirements
No special No special
Class 0 exposure 0.00 to 0.10 0 to 150
requirements requirements
C 150 Type II or
Class 1 exposure > 0.10 and < 0.20 > 150 and < 1 500 0.50
equivalent
C 150 Type V or
Class 2 exposure 0.20 to < 2.0 1 500 to < 10 000 0.45
equivalent
C 150 Type V plus
Class 3 exposure 2.0 or greater 10 000 or greater 0.40
pozzolan or slag

Table 4.7: Recommendation of BS 8110-1997ǂ for concrete exposed to sulfates

SO3 in 1:2 SO3 in Cement


Severity of Total SO3
soil/ water water Type of cement recommended content
exposure in soil (%)
extract (g/L) (g/L) (kg/m3)
No special requirements 330
Moderate 0.2–0.5 1.0–1.9 0.3–1.2 OPC plus 25–40% pfa; OPC plus 70–90% GGBFS 310
SRPC or SSC 280
OPC plus 25–40% pfa; OPC plus 70–90% GGBFS 380
Severe 0.5–1.0 1.9–3.1 1.2–2.5
SRPC or SSC 330
Very severe 1.0–2.0 3.1–5.6 2.5–5.0 SRPC or SSC 370
Extreme > 2.0 > 5.6 > 5.0 SRPC or SSC with protective coatings 370

ǂNow replaced by BS EN 1992-1-1:2004+A1:2014.


Note: OPC = ordinary Portland cement, pfa = pulverised-fuel ash, GGBFS = ground granulated blast furnace slag, SRPC
= sulfate resisting Portland cement, and SSC = super sulfated cement.

Austroads 2016 | page 57


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The distribution of chemicals in aggressive soils can vary from place to place; for example, the soil analysis for
the job site of the new Gateway Bridge (0) showed the content of sulfate (expressed as SO4) of 300–1900
mg/kg, or SO3 content of 0.03–0.16% and magnesium content of 60–8400 mg/kg. The site also contained very
high levels of chloride (1390–15900 mg/kg), except for three locations which showed high acidity (pH < 5).

The above sulfate content in soil can be categorised as ‘moderate’ by BS 8110-1997 (now replaced by BS
EN 1992-1-1:2004+A1:2014), or Class 1 exposure to sulfate by ACI (2008). The high magnesium
concentration detected in this site suggests the possibility of magnesium-sulfate attack. The maximum
chloride content measured was nearly twice that in seawater, assuming that the soil water content was 50%.
Therefore, special material requirements or protective design will be necessary for the foundation piles in this
soil in order to ensure long-term durability.

Internal attack

Internal sulfate attack occurs when the ingredients of the concrete contain the sulfate contents required for
the above reactions, such as over-sulfated cements, or sulfate impurities in the aggregate or mix water.
Sometimes, iron sulfide minerals such as pyrite or pyrrhotite in the aggregate could oxidise to form sulfuric
acid and cause internal sulfate attack, but this is largely limited to the surface layers of concrete, where
oxygen is available for oxidation of the sulfides (Shayan 1988b). Other than the dissolution of the cement
paste, crystallisation of hydrated calcium sulfate or calcium alumino-sulfates in confined spaces can cause
expansion and disruption of the concrete element involved.

Delayed ettringite formation can also be considered a form of internal sulfate attack, details of which were
discussed earlier in the report.

Table 4.8: Sulfate aggressiveness to concrete

Aggressiveness
None Mild Moderate Severe Very severe
Water SO4 (ppm) < 200 200–600 600–3000 3000–6000 > 6000
Soil SO4 (%) < 0.2 0.2–0.6 0.6–1.2 > 1.2

Source: CEMBUREAU (1978).

Table 4.9: Recommendations for sulfate resistance

Aggressiveness
Mild Moderate Severe Very severe
Water SO4 (ppm) 0–150 150–1 500 1 500–10 000 > 10 000
Soil SO4 (%) 0.00–0.10 0.10–0.20 0.20–2.00 > 2.00
Cement type II, IP(MS), IS(MS) V V + pozzuolana or slag
w/c, max. 0.50 0.45 0.45

Source: ACI 201.2R-10 (American Concrete Institute 2008).

Austroads 2016 | page 58


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 4.10: Requirements for well-compacted cast in situ concrete, 140 mm to 450 mm in thickness, exposed on
all vertical faces to a permeable sulfate soil or fill

Concentration of sulfate and magnesium Minimum


Cement type Maximum
Soil or fill Groundwater Cement(1) content
Class By acid By 2:1
(Max. aggregate
extraction water/soil (g/l) (See Table 4.11 Water/cement
size 20 mm)
(%) extract (g/l) below) ratio
SO4 SO4 Mg SO4 Mg (kg/m3)
1 < 0.24 < 1.2 < 0.4 A–L Note (2) 0.65
> 0.24 A–G 330 0.50
classify on
2 1.2–2.3 0.4–1.4 H 280 0.55
basis of
2:1 extract I–L 300 0.55

H 320 0.50
3 2.3–-3.7 1.4–3.0
I–L 340 0.50
3.7–-6.7 < 1.2 3.0–6.0 < 1.0 H 360 0.45
4 I–L 380 0.45
3.7–6.7 > 1.2 3.0–6.0 > 1.0 H 360 0.45
> 6.7 < 1.2 > 6.0 < 1.0
5 As for Class 4 plus surface protection
> 6.7 > 1.2 > 6.0 > 1.0

1 Cement content includes supplementary cementing materials, such as fly ash and slag. To maintain the mortar
fraction, the cement contents given may be increased by 40 kg/m3 for 10 mm maximum nominal size aggregate and
may be decreased by 30 kg/m3 for 40 mm maximum nominal size aggregate (Table 8 of BS 5328:Part 1-1990, now
replaced by BS EN 206:2013).
2 Minimum value 275 kg/m3 for unreinforced concrete (BS 8110-1985 (now replaced by BS EN 1992-1-
1:2004+A1:2014) and BS 5328: Part 1-1990 (now replaced by now replaced by BS EN 206:2013); 300 kg/m3 (BS
8110-1985 (now replaced by BS EN 1992-1-1:2004+A1:2014) and maximum w/c 0.60 for reinforced concrete.

Source: BRE (1991).

Table 4.11: Types of cement

Code Type or combination Code Type or combination


A Portland cement (BS 12)(1) H Sulfate-resisting Portland cements (BS 4027,
now withdrawn)
B Portland blast furnace cements (BS 146)(1) I High-slag blast furnace cement (BS 4246)(1)
containing not less than 74% slag by mass of
nucleus
C High slag blast furnace cement (BS 4246)(1) J Combinations of Portland cements and blast
furnace slag containing not less than 70% slag
and not more than 90% slag by mass of slag
plus cement
D Combinations of Portland cements (BS 12(1)) and K Portland PFA cements containing not less than
blast furnace slag (BS 6699, now withdrawn) 26% PFA by mass of nucleus
E Portland PFA cements (BS 6588)(1) L Combination of Portland cements and PFA
containing not less than 25% PFA and not
F Combinations of Portland cements (BS 12)(1) and
more than 40% PFA by mass of PFA plus
PFA
cement
(BS 3892: Part 1)(2)
G Pozzolanic PFA cements (BS 6610-1991, now Combinations of Portland cements and various
withdrawn) pozzolanic materials including natural
pozzolanic materials

1 Now replaced by BS EN 197-1:2011.


2 Now replaced by BS EN 450-1:201.

Austroads 2016 | page 59


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

4.5.12 Acid Sulfate Soil

Acid sulfate soils contain iron sulfide minerals, predominantly the mineral pyrite, FeS2. In an undisturbed
state below the water table, acid sulfate soils are benign. However, if the soils are drained, excavated or
exposed to air by a drop in the water table, the sulfides react with oxygen to form sulfuric acid, as indicated
by the chemical reaction in Equation 15.

2 FeS2 + 7O2 → 2FeSO4 + 2H2SO4 15

Thomas et al. (2003) produced a technical manual for coastal acid sulfate soils, which states that 40 000
square kilometres of acid sulfate soil exist in coastal areas of Australia, and that potentially they contain one
billion tonnes of sulfuric acid. The acid leachate generated by the exposure of these soils (pH = 3.5) is very
harmful for the environment (kills plants and fish) and is highly corrosive for pipes and concrete structures.

Sulfur-reducing bacteria, which oxidise molecular hydrogen, while reducing sulfate to hydrogen sulfate (H2S)
and sulfides, can also give rise to the same conditions. They thrive under acidic and aerobic conditions within
the range of temperate ambient temperatures and where there is an availability of reduced sulfur and carbon
compounds (Czerewko et al. 2003).

This type of attack is most severe in sewer systems, where sulfur deposited on the concrete surface above
the sewage attracts microorganisms of the genus thiobacillus. Hormann et al. (1997) tested the deterioration
of several specimens in an accelerated test (simulated biogenic sulfuric acid bath of pH 3.5) and showed that
specimens made with high alumina cement had much higher resistance to attack compared to those made
with normal Portland cement.

Soil tests conducted at the site of the new Gateway Bridge showed that the soil sampled at the location of
Pier 7 piles had a pH of 4–5, which is classified as exposure condition C–C1 by AS 5100.5-2004 (Maunsell &
AECOM 2007). It should be noted that exposure Class C1 requires full isolation of the concrete surface from
the environment.

The acidic condition prevailing at acid sulfate soil sites may be too harsh even for concrete containing GGBS
(slag), although they are normally resistant to sulfate attack (Madrid et al. 1997; O’Connell et al. 2012).
Matthews (1992) studied the behaviour of concretes with total binder contents ranging from 180 to 500 kg/m 3
and fly ash contents of 20–40%, during five years of exposure to slow-flowing groundwater of pH 3.5 to 4.5.
It was concluded that the concrete made with high C3S cement showed more mass loss than its counterparts
of low C3S cements, and the incorporation of fly ash did not provide any benefit.

In fact, concrete is basic in nature and will disintegrate in strongly acidic solutions. The results of Khatri and
Sirivivatnanon (1997) showed that a ternary blend of high slag cement + silica fume was highly resistant to
sulfate attack. Cao et al. (1997) also showed that even when the sulfate solution was controlled at pH 3,
blended slag cement with more than 60% slag content performed satisfactorily.

However, it should be noted that these tests were relatively short-term (1–2 years), and that concrete made
with these types of binders may seriously deteriorate over several decades of contact with the highly
aggressive exposure conditions of acid sulfate soil. Thus, even the sulfate-resistant concrete elements could
need to be physically isolated from contacting the acidic soils, in order to prevent damage due to sulfate
attack.

The Building Research Establishment (BRE) in the UK, published a special digest in 2005, which addresses
the requirements for concrete in aggressive soils (BRE 2005). This document describes the various
aggressive chemical environments to which concrete elements can be exposed to, and recommends specific
requirements for concrete in such environments.

Austroads 2016 | page 60


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

A brief document was also published by Cement Concrete and Aggregates Australia (CC&AA) in 2011,
which described the requirements for sulfate-resisting concrete (CC&AA 2011) in buried and partially buried
concrete. It states that for buried condition the concrete must be made from sulfate-resisting cement (Type
SR) with a minimum cement content of 335 Kg/m3 and water to cement ratio of 0.50. For partially buried
conditions subjected to cycles of wetting and drying the recommended cement content was 415 kg/m 3 and
water to cement ration 0.40.

4.5.13 Physical Salt Attack

Physical salt attack (PSA) is a form of distress caused by the crystallisation of salts in pores below the drying
faces of concrete, i.e. in the layer affected by evaporation. Wetting and drying cycles exacerbate the PSA,
whereby the salt dissolves when concrete is wet (e.g. by rain water or groundwater) and crystallises when
water evaporates and the solution concentration rises above the saturation level of the salt.

The salts responsible for this attack on concrete include sodium sulfate, sodium carbonate and sodium
chloride (in descending order of aggressiveness). Haynes and Bassuoni (2011) reported significant surface
scaling damage on a concrete foundation wall by sodium sulfate, on a garage slab by sodium carbonate and
on concrete specimens partially immersed in 5% sodium sulfate solution. They attributed the damage caused
by PSA to phase transformations of the salt. For example, sodium sulfate undergoes phase transformation
from thenardite (Na2SO4) to mirabilite (Na2SO4·10H2O), involving a massive volume change and resulting
stresses which can disrupt the matrix of the hydrated cement at the surface zone of the concrete.

It is also known that sodium sulfate has unusual solubility characteristics in water, and that its solubility in
water rises more than ten-fold between 0 to 32 °C, where it reaches a maximum of 497 g/L (Linke & Seidell
1965). Thus, diurnal temperature changes may cause crystallisation-dissolution cycles of the salt. This
process is similar to the process of freezing-thawing. Mehta and Monteiro (2006) pointed out that PSA-
damaged concrete has the same appearance as surface scaling caused by freezing and thawing cycles.
Yang et al. (1997) showed that the same mitigation methods used against freezing-thawing attack could be
used for PSA, i.e. concrete with entrained air content of 3% showed remarkable reduction in scaling. It is
clear that to eliminate PSA, the concrete must include entrained air as well as have low permeability.

In some cases, such as the bridge deck soffit, poor drainage and inadequate detailing of the deck surface
can allow rain water to accumulate and seep through to the deck soffit, where it can evaporate. The
dissolved salts are then deposited in the outer skin of the soffit, where the crystallisation pressure causes
disintegration of the cement post. The solution to this problem is improved drainage and sufficient detailing of
the deck surface, followed by patch repair of the soffit.

This problem does not affect the bulk of the concrete, and its main adverse effect is reduction of concrete
cover thickness, which in turn allows the carbonation front to reach the reinforcement level faster.

4.6 Surface Coating Treatments for Attack Mitigation


A comprehensive document (Munger 1984) prepared by the US National Association of Corrosion Engineers
(NACE) details corrosion prevention by protective coatings for various materials, including reinforced
concrete. These documents and relevant experts should be consulted in the prevention of corrosion during
the maintenance phase of concrete bridges.

Surface treatments, using hydrophobic materials as admixtures, can effectively increase the resistance of
concrete to water penetration, and help seal the cracks already formed on the concrete surface. Tittarelli and
Moriconi (2011) demonstrated the durability properties of a hydrophobic concrete, containing an alkyl-
triethoxy-silane admixture at dosage of 2.2% by cement mass, and showed that the water absorption was
greatly reduced and corrosion resistance of galvanised steel bars improved, even in the concrete fabricated
with cracks.

Austroads 2016 | page 61


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In Norway (Standards Norway 2003), most of the off-shore concrete structures produced prior to 1980 had
an epoxy coating of 2 to 3 mm on the concrete shaft at the splash zone. The coating was applied at an early
age when the young concrete still had some porosity and suction ability, so a very good bonding between
the concrete and the coating was achieved. This coating was shown to provide an efficient barrier against
chloride penetration (15 years at the time of the inspection), although it was also observed that a thinner
coating was being damaged within two years (Gjørv 2009).

In the UK, the Highways Agency (2003) Design Manual for Roads and Bridges (BD 43/03) requires bridge
structures to be treated with tri-ethoxy-silane, a pore-lining impregnating solution, which allows the concrete
to lose moisture, but prevents water from getting into concrete. This treatment was used as a measure to
minimise the risk of AAR expansion.

At present, many materials are available and claimed to be suitable for the protection of concrete structures,
e.g. siloxane-based Sikagard® by SIKA Technology and fluorinated acrylic copolymer BridgeGuard® by
EPICURO.

Each type of material may be effective only for limited purposes. Dai et al. (2010) showed that sodium-
silicate-based pore blockers were found to be inefficient in preventing chloride penetration of concrete under
simulated marine exposures. The long-term efficiency of water repellent agents used for surface
impregnation was found to be highly dependent on the type of agent. Song and Shayan (1999) summarised
the various surface coating materials that could be used for corrosion prevention in reinforced concrete
structures.

In Australia, standards similar to BD 43/03, e.g. VicRoads (2009) Specification Section 686 Coating of
Concrete, have been developed for the impregnation of reinforced and prestressed concrete elements of
highway structures, using hydrophobic pore-lining impregnants, to help prevent and control chloride-induced
reinforcement corrosion from de-icing salts, and in marine environments.

However, the long-term durability of the surface coating materials used for these treatments, as well as the
extent of the protection offered by such treatments, may be a problem for structures with very long service
lives (Dai et al. 2010). These materials may need to be applied in repeat treatments.

Moradllo et al. (2012) tested the performance of surface coatings in the tidal zone of the Persian Gulf region
and showed that epoxy polyurethane and aliphatic acrylic materials were the most efficient coatings,
although the performance was time-dependent, and recoating at appropriate times was essential to maintain
their efficiency.

Selander (2010) reported that concrete impregnated by alkyl-alkoxy silanes had shown a much reduced
water penetration as well as chloride penetration, and the hydrophobic effect remained after some 10 to 15
years under field conditions in Sweden and the UK, although the material on the concrete surface was
damaged due to exposure to sunlight. Schueremans et al. (2007) showed that treatment of a concrete quay
wall with isobutyl triethoxy silane had decreased the chloride penetration compared to the untreated
concrete, after 12 years exposure to the splash zone. Both these papers mentioned that the resistance to
carbonation was not improved by the hydrophobic impregnation.

Another way of improving concrete resistance to aggressive environments is to modify the cement paste
micro-structure. Beaudoin et al. (2009) suggested that the interaction of some polymers, such as
hexadecyltrimethyl ammonium, methylene blue, polyethylene glycol, polyvinylalcohol, and polyacrylic acid
with the nanostructure of C–S–H could increase silicate polymerisation, which may improve volume stability
and promote the mechanical properties of such systems. Their resistance in chemically aggressive
environments is allegedly increased as the defects in the silicate structure are eliminated by the presence of
polymer molecules.

Wiktor and Jonkers (2011) tested the possibility of mixing bacterial spores and calcium lactate into concrete,
which produced CaCO3 and carbon dioxide upon receiving oxygen from the cracks. They showed that cracks
up to 0.4 mm width were healed by calcium carbonate precipitation.

Austroads 2016 | page 62


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Several of the techniques mentioned in Table 4.12 could be combined to ensure long-term durability of
concrete structures.

Table 4.12: The durability concerns of reinforced concrete

Component Exposure Durability problems Mitigation method


Deck slabs CO2, moisture, airborne Cl- Carbonation Anti-carbonation coating, corrosion
inhibitor
Beams and CO2, moisture, airborne Cl- Carbonation, chloride Low w/b, anti-carbonation coating
cross beams
Piles CO2, salty water splash Chloride ingress, erosion, Blended cement, low w/b, anti-
carbonation carbonation coating; using stainless
steel reinforcing
Aggressive soil Sulfate attack, salt attack Sulfate resistance cement, blended
cement, air entraining
Acidic soil Acid attack Protective coating
Pile-caps Seawater Thermal cracking, DFE, Blended cement; not too low w/b,
chloride contamination at protect the component at early age
early age, chloride ingress if possible
Abutment or Soil or seawater Thermal cracking, DFE, Blended cement; not too low w/b
abutment piles chloride ingress
Aggressive soil Sulfate attack, salt attack Sulfate resistance cement, blended
cement, air entraining
Acidic soil Acid attack Protective coating

Austroads 2016 | page 63


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

5. Prediction of Service Life


Prediction of service life can be made using the known time-dependent durability-related properties of
concrete, e.g. the transport properties, together with the knowledge of prevailing environmental conditions,
mechanical properties, and conditions and/or criteria that characterise each point of failure in the three
stages of service life, as defined in Section 2. This approach is known as the deterministic method, because
relevant properties are determined through tests which assign a single value to the property concerned, and
will result in a definitive service life prediction.

In reality, material properties and environmental conditions do not have single values and vary through a
range. For example, chloride concentration in concrete or the thickness of cover concrete can vary from
point-to-point in the same element or structure. Similarly, single value criteria established for points of failure,
based on previous experiences, may not be totally applicable and failure could occur within a range of the
values concerned. A new approach to service life prediction, termed the probabilistic approach, uses the
variation in relevant properties, which are assumed to follow a normal distribution, and evaluates the
probability that certain values could be exceeded at given times. Therefore, this approach determines the
probability of failure to occur during the service life.

In this section, firstly the deterministic approach is discussed for the three stages of service life, followed by a
brief description of the probabilistic approach.

5.1 Time to Onset of Potential Failure


Chloride penetration into concrete is governed by diffusion principles, where the mathematical expression for
one-dimensional diffusion in a semi-infinite space with constant boundary conditions is applicable. A critical
review of chloride diffusion, test methods and prediction of chloride penetration into concrete was presented
by Nilsson et al. (1996).

For structure members that are constantly exposed to a chloride source, e.g. seawater, the concentration of
the chloride in concrete, at time t and position x, i.e. C(t, x), is expressed by Equation 16:

x 16
C (t , x)  C0  [C S  C0 ][1  erf ( )]
4Dt
where

C0 = C(0, x) which is the initial concentration at time t = 0 in concrete

CS = concentration at the surface (x = 0)

D = the apparent diffusion coefficient

For structural members exposed to atmospheric conditions, where chloride ions are brought to the concrete
surface and gradually build up, the concentration at the surface, i.e. CS(t), is expressed by Equation 17
(Nilsson et al. 1996; Ann et al. 2009):

Austroads 2016 | page 64


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

 x x 
17
2
x 
C (t , x)  k t e 4 Dt  ( )erf ( )
 4 Dt 4 Dt 

Penetration of Chloride Ions into Concrete

When chloride ions penetrate into concrete, some of the ions react with various components of the hydrated
cement and are chemically combined with them, either in crystalline form or in adsorbed form on the walls of
pores in the concrete matrix. This is called the ‘bound’ chloride. A portion of the chloride ions remain in the
pore solution of concrete and are able to travel further into concrete, which is termed the ‘free’ chloride. The
sum of these forms, plus any native amount, comprises the total chloride content, which is also called the
acid-soluble chloride content.

Equation 16 was derived for ion diffusion in solution, which corresponds to the movement of free chloride
ions in a concrete pore solution. In practice, the total acid soluble chloride is determined, as well as the free
ions which are water-soluble. Nilsson (1993) showed that Equation 16 was applicable to the apparent
diffusion in concrete as a homogenous media, driven by the total chloride gradient, with an apparent
diffusion coefficient Da. Nilsson (1993) derived the relationship between D and Da as expressed by
Equation 18:

D 18
Da 
cb
psol (1  )
c f

where

psol = saturated pore volume

cb = concentration of bound chloride

cf = concentration of free chloride

cb / c f = binding capacity obtained from the chloride-binding isotherm

Indeed, field investigations carried out worldwide in the past several decades demonstrate that Equation 16
applies to the total chloride content profiles, and the apparent diffusion coefficients obtained vary between
0.1 and 10  10-12 m2/s.

The apparent diffusion coefficient depends on the type of cementitious material and its hydration level,
because of the chemical and physical binding of chloride by these ingredients, as well as the porosity of
concrete (Shayan & Xu 2002). Some supplementary cementitious materials (SCMs), such as fly ash and
blast furnace slag, have been found to be particularly beneficial in reducing the apparent diffusion coefficient.
On the other hand, carbonation of concrete will reduce the chloride-binding capacity (Larsen 1997), thus
increasing the rate of chloride diffusion. Jones et al. (1994) noted a two-fold increase in Da due to
carbonation.

The chloride content of the surface concrete, Cs, determined on field concrete, increases with the exposure
time. The results of Swamy et al. (1994) showed that the value of Cs was around 5% chloride by cement
mass for concrete samples with w/c ratios of 0.4 to 0.6, which had been submerged in seawater for five
years, and one concrete with a w/c ratio of 0.3–0.4 had a high Cs value of 8%.

Austroads 2016 | page 65


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Bolin et al. (1997) showed the maximum surface chloride content of a bridge pier in the Baltic sea was 6% by
cement mass, which is a very high value considering that chloride content in seawater is only 0.4%.
Laboratory tests by Lindvall (2007), using 5 g Cl -/L and 20 g Cl-/L solutions (simulating the salinity of the
Baltic Sea and Atlantic Sea), resulted in CS values of 2% and 4% by binder mass. It appears that in situ
concrete may be subjected to chloride enrichment by processes such as wetting and drying, which can
induce high local concentrations.

More detailed results on total chloride content in concrete exposed to various concentrations of NaCl
solutions can be found in papers on chloride binding isotherms (Baroghel-Bouny et al. 2012; Tang 1996).

Results of chloride tests conducted by ARRB Group in the past 10 years on various concrete structures in
Australia (mainly in Victoria, and some in New South Wales) show the Cs value for the submerged zone of
concrete elements to be around 1% by concrete mass, with a maximum value of 2.5% by concrete mass
(concrete grade 50 MPa) obtained from the tidal zone of a bridge pier. A survey conducted on five bridges in
Norway showed CS values of 1.25%, 0.6% and 0.2% by concrete mass at 0, 10 m and 15 m above sea level,
respectively (Maage et al. 1997), demonstrating the effects of splash and mist actions on penetration of
chloride into concrete. Note that the concrete facing the prevailing wind direction contained much less
chloride, which might be due to the effect of rain washing the concrete surface (Fluge 1997).

The first phase of service life is defined as the time when the critical (or threshold) chloride content of
concrete (Ccrit) is reached at the reinforcement level (i.e. change of the initial chloride content of concrete
(C0) to (Ccrit)). The depth at which the threshold chloride concentration is reached is obtained by solving
Equation 16, which yields Equation 19:

Cs  Ccrit 19
X C  erf 1 ( ) 4 Da t1
C s  C0

The first phase of the service life can be calculated by letting XC equal the cover depth and solving
Equation 16 for the time t1. Note the inverse of the error function can be empirically expressed as a
polynomial function of Cs, which facilitates calculation of the safety factor.

5.1.1 Diffusion Coefficient Variations – Dependence on Temperature and Time

It has been recognised that the value of chloride diffusion changes with temperature and time. The former is
a general physical phenomenon for all ions and the latter is due to the change of microstructure of concrete
with age. Both have been considered as important factors in service life prediction.

Temperature dependence

In the literature, e.g. Page et al. (1981), Ehlen (2008) and Gjørv (2009), the temperature dependence of the
diffusion coefficient was considered to follow that of the Arrhenius equation for the chemical reaction rate, i.e.
k = A·exp(-Ea/(RT)), where Ea is the activation energy and its value varies depending on the materials and
nature of the reaction. For example, the Life-365 model uses Ea = 35 kJ/mol, whilst Tang (1996) and Polder
and de Rooij (2005) assumed Ea = 40 kJ/mol, whereas Saetta et al. (1993) used 40 to 45 kJ/mol for water
cement ratios of 0.35 to 0.52. The arbitrary selection of Ea will introduce more uncertainty rather than making
the simulation more accurate.

Austroads 2016 | page 66


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In fact, for ion diffusion in liquid, the diffusion coefficient and temperature are described by the Stokes-
Einstein relation (Atkins 1987) (Equation 20):

k BT 20
D
6a
where

kB = Boltzman constant (1.38 x 10-23JK-1)

T = temperature (K)

a = diameter of the diffusing particle (sphere)

 = viscosity of water

The relative diffusion coefficient can be expressed by a simple relation (Equation 21):

D(T1 ) T1 2 21

D(T2 ) T21

Figure 5.1 demonstrates the variation in the relative diffusion coefficient and water viscosity as a function of
temperatures.

Figure 5.1: The water viscosity at different temperatures and diffusion coefficient relative to that at 20 °C

Austroads 2016 | page 67


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Time dependence

Cementitious materials in concrete continue to hydrate and the pore structure is continually refined with time.
This fact will lead to the reduction of the diffusion coefficient with time. It has been observed that the
diffusivity decreased drastically in the later age, as determined by accelerated test methods, e.g. using high
concentration salt solution by Poulsen (1997), and using electrical field by Tang (1996). Nevertheless, the
diffusion coefficient is used as a concrete quality control parameter, e.g. at 28 days.

The available service life prediction models consider that the decrease in the diffusion coefficient follows the
following empirical relation (Poulsen 1997) (Equation 22):

m 22
 t ref 
D(t )  Dref  
 t 

where

Dref = diffusion coefficient determined at the reference time (e.g. 28 days) tref

m >0 = age factor

The constant m is shown by Maage et al. (1997) to depend on the water/cement ratio, curing time and the
composition of the cementing materials (e.g. m = 0.4 for Portland cement, c.f. values in Table 5.2).

There has been a large variation in the period during which Equation 22 is valid. The Life-365 model applies
this equation up to 25 years. Thomas et al. (2011) (using a similar but different form), extended the period to
100 years, whereas the results shown by Tang (1996) indicate this effect may be very small after one year
for the specimens he tested. It would be logical to draw a parallel between the evolution of compressive
strength of concrete with time (which, practically, does not change after one year of curing) and that of the
chloride diffusion coefficient, because most of the pore refinement would take place within the first few years
of curing.

Nilsson et al. (1996) showed that Equation 22 can be incorporated into the diffusion equation, such that
Equation 16 can be written as Equation 23:

x 23
C (t , x)  C0  [CS  C0 ][1  erf ( )]
t'
4 D(t )dt
t0

where

t0 = time (age) at which the concrete is exposed to the environment

t’ = t0+t

Equation 16 allows using any form of the expression for D(t). If the expression in Equation 22 is used, then
the explicit form for the integral of diffusion coefficient with time is (Equation 24):

m 24
Dref   1 m
 t0  
1 m m
t'
 t0   t ref 
 D(t )dt  1  m  1  t '    t '   
 t'


t0  

Austroads 2016 | page 68


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Equation 23 (together with Equation 24) has been shown to match the chloride profiles of concrete exposed
to seawater after 1 year and 10 years (Tang & Gulikers 2007), and has been adopted in some service life
prediction models.

It may be pointed out that the exposure conditions of concrete in accelerated tests, NT Build 443 (Nordtest
1995) and NT Build 492 (Nordtest 1999), are different from that in normal seawater. It is not certain whether
the values of m, the age factor, and Dref, obtained by these tests would be the same as that for concrete
exposed to field conditions.

Song et al. (2009) conducted service life prediction of a tunnel structure placed on the seabed. They
determined the rate of chloride transport in concrete mixes identical to that used in the structure concerned,
and assumed the chloride threshold level for corrosion to be 0.69, based on 65 published data, to determine
the time-dependent chloride diffusion coefficient, which includes an ageing factor (e.g. Equation 22). Based
on Monte Carlo simulation, they calculated the probability of steel corrosion and its corresponding service
life. The service life was found to depend on the time dependency of chloride transport, and was equated to
31, 51, 85, and 147 years, for ageing factors of 0.1, 0.2, 0.3 and 0.4, respectively. Using the time-
independent chloride diffusion coefficient in the model indicated only 27 years of service life.

Andrade et al. (2011) found that the reduction in chloride diffusion over time paralleled increases in concrete
resistivity, and proposed an accelerated test to determine the value of m, the ageing factor. This involved
measuring the resistivity for concrete cured at 40 °C for 28 days and that of specimens cured at 25 °C for
one year, and applying the relationship presented in Equation 3a in Andrade et al. (2011).

It should be noted that some important factors, which enhance chloride penetration into concrete in the long-
term, do not appear to have been included in any of the available service life prediction models. One
example is that carbonation reduces the capacity of concrete to bind chloride (Larsen 1997; Nilsson et al.
1996), which can happen in the splash zone as well as in the atmospheric zone. Numerous field
investigations have demonstrated that the content of total (acid-soluble) chloride at the concrete surface was
much lower than that inside the concrete surface, which is believed to arise because carbonation released
the bound chloride from the concrete.

The carbonation and other surface deterioration issues such as lime-leaching, erosion, sulfate attack (Jarrah
et al. 1995) and concrete cracking, will reduce the effective cover in resisting the chloride penetration. This
effect needs to be included in service life prediction models.

5.1.2 Carbonation Rate

The carbonation process is essentially a diffusion process which is driven by the atmospheric CO 2
concentration, and which has been relatively constant, being 0.03% by volume of the air. The carbonated
zone has a lower pH and the depth of carbonation closely corresponds to the neutralisation zone, so it can
be detected by certain pH indicators such a phenolphthalein.

Möller (1994) determined the profiles of CO2 and calcium content in a concrete specimen and found that the
maximum degree of carbonation was about 80% (i.e. the mass ratio of CO 2/CaO = 44/56). The ratio at the
carbonation front (indicated by the colour change of phenolphthalein) was about 40%, as shown in
Figure 5.2. This indicates that the amount of CO2 in the concrete is proportional to the depth measured by
the pH indicator, and that the zone ahead of the carbonation front has also carbonated to some extent.
Therefore, the steel in this zone may also be prone to corrosion.

Austroads 2016 | page 69


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The square-root-of-time law has been found to be generally adequate for describing the carbonation depth
(dc) of concrete as a function of time (Equation 25):

d C  kC t ` 25

where

kC = carbonation rate constant

t = time

Figure 5.2: Carbonation degree and the depth measured by phenolphthalein pH indicator

Source: Möller (1994).

Furman (2008) used a modified version of Equation 25 for the carbonation depth, which is used in the CEB
(Comité Européen du Béton) model, in which the influence of factors such as age, moisture content of the
concrete, mix design and curing parameters are taken into account, as expressed in Equation 26. This
equation includes an age factor (n), because it was considered that the carbonation rate of concrete would
decrease with time in a similar manner to that of chloride diffusivity.

n 26
t 
d C  k k1k 2t  0 
t 

where

k = carbonation coefficient (mm/√year)

k1 and empirical constants describing the effect of moisture content and curing on the
=
k2 resistance to carbonation

t0 = the age when testing was done (year)

t = exposure time (years)

Austroads 2016 | page 70


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

However, the carbonation rate varies as the extent of concrete carbonation increases, which means that
these equations do not apply to all cases of carbonation.

As mentioned in Section 4.5.4, extensive UK studies (Hobbs et al. 1998) have shown that the carbonation
depth of concrete, where the relative humidity was around 80%, was proportional to t0.4 rather than √t; and at
relative humidity of 98%, the carbonation depth was proportional to t0.2.

Sisomphon and Franke (2007) studied the carbonation rate of laboratory specimens that contained large
volumes of pozzolanic materials such as fly ash and slag. To account for variations of the carbonation rate
constant (k) with the age of concrete, the humidity and the amount of carbonation that has already occurred
(so that the rate can be considered constant), they chose a value of n = 0.40 and found that the relationship
dc = k t0.4 best fitted their field data.

Steffens et al. (2002) developed a numerical carbonation model to predict the risk of carbonation-induced
corrosion damage to concrete structures, but this is a complicated model which is not easy to use in practice.
Further studies may be needed to unify the current approaches and establish the corrosion rate of steel in
carbonated concrete.
Kondo et al. (1968), on the consideration that the CO2 diffused into concrete would be totally
bound by the concrete at the reaction front, and that the CO2 gradient is CCO2/x, where x is the
distance from the concrete surface to the reaction front, derived the following Equation 27:

CCO2 27
x 2D t
Cb

where

D = diffusion coefficient of CO2 in concrete

CCO2 = CO2 concentration in air

Cb = the capacity of concrete to bind CO2

This formula was also applied by many other investigators (Goodbrake 1978; Hamada 1968; Marques &
Costa 2010; Monteiro et al. 2012; Papadakis et al. 1991; Parrot & Hong 1991; Smolczyk 1968; Tuutti 1982).

Empirically, a convenient way to express the carbonation rate is its relation with concrete strength or the
water to cement ratio, i.e. the carbonation rate is low when the strength grade of concrete is high (Burden
2006; Collepardi et al. 2004; Hamada 1968; Ho & Lewis 1987; Smolczyk 1968). Results of an investigation
on two Australian bridges (Shayan & Xu 2012) demonstrate this relation (Figure 5.3).

Monteiro et al. (2012) conducted a statistical analysis of carbonation data collected on concrete structures up
to 99 years old, and showed that the mean carbonation rate constant for concrete with mean strength of 53
MPa was 2.85  0.70 mm/year. The rate was 1 to 9 mm/year for various concretes with water to binder
ratios of 0.4 to 0.5, which reflects the overall influence of binder type, binder content and w/c ratio. Based on
the data in Figure 5.3 (Shayan & Xu 2012), the average carbonation rate for 50 MPa concrete, made using
ordinary Portland cement, is about 2.3 mm/year, which is in agreement with the data of Monteiro et al.
(2012).

Austroads 2016 | page 71


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 5.3: Example of relationship between carbonation depth and concrete strength for deck slabs, beams
and piers of two bridges constructed in 1940 and 1976

7
Carbonation rate (mm/Year0.5)

6
5
4
3
2
1
0
20 25 30 35 40 45 50 55 60
Concrete Compressive Strength (MPa)

Source: Based on data from Shayan and Xu (2012).

The carbonation rate in cracked concrete and at construction joints is faster than in sound concrete, and the
variation is larger. Ann et al. (2010) surveyed a concrete bridge (30 MPa grade) after 18 years of exposure to
air, and showed that the carbonation depth was 11.6 mm, 24.7 mm and 17.4 mm for sound concrete,
cracked concrete (crack width 0.1 to 0.2 mm) and at construction joints, respectively.

As indicated by Equation 27, the carbonation rate increases with the concentration of CO 2 in air but
decreases with the cement binding capacity. It is evident that concretes which incorporate pozzolanic
materials would have a lower resistance to carbonation (Marques & Costa 2010; Sasatani et al. 1995),
because the pozzolanic reaction consumes Ca(OH) 2, which is the major contributor to the binding of CO2 in
concrete. Therefore, this type of concrete should be made with lower water to binder ratio and higher total
cementitious material content to increase its resistance to carbonation.

Note that in the diffusion theory, the carbonation rate is proportional to the CO 2 in air (e.g. Xu & Shayan
1999; Austroads 2000b), whereas formulae such as those given in Equation 26 and Equation 27 imply that
the rate is proportional to √CO2. Thus, the prediction as to how exactly the increase in CO2 content of air
would affect the concrete carbonation is uncertain. More research appears to be necessary to explain the
discrepancy and give the correct prediction.

Neves et al. (2012) used a semi-probabilistic model for service life prediction of reinforced concrete
structures subjected to CO2-induced corrosion. This was based on accelerated carbonation test results
obtained under various exposure conditions and carbonation depth derived from Fick’s first law of diffusion.
The authors defined a maximum acceptable level of deterioration and developed a model which uses a
safety factor associated with a reliability index to define the service life of the structure concerned.

The increase in the atmospheric CO2 concentration over the service life of a new structure with a long design
life, particularly if it is substantially more than 100 years, would mean that the carbonation rate would
increase. Increases in the concentration of other air pollutants such as SO 2 could also further enhance
concrete carbonation (Skenderovic & Lomic 1992). Therefore, a carbonation rate constant higher than that
based on the present data should be used for service life prediction.

Stewart et al. (2012) have predicted a greater than 16% increase in the probability of carbonation-induced
corrosion damage to reinforced concrete structures due to the predicted increase in atmospheric CO2
(Figure 5.4) and changes in climate in Australia by the year 2100.

Austroads 2016 | page 72


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 5.4: Predicted low, median and high estimates of CO2 concentration

Source: Stewart et al. (2012).

In Australia, the prediction of carbonation-induced damage could be improved if existing data on carbonation
of concrete bridges (and other structures) could be collected and analysed with respect to age of structures,
concrete formulations used, and moisture condition of the concrete and the environment. This would be
valuable for service life prediction, and would form the basis for the evaluation of future carbonation data with
respect to carbonation risk in the future.

5.2 Time to Detectable Failure Due to Steel Corrosion

5.2.1 Time to Corrosion-induced Cover Cracking

A number of models have been proposed in the past for the estimation of crack formation due to rebar
corrosion (Ahmad 2003; Xu & Shayan 1999). However, most of the models are either empirical in nature,
containing only a few but not all the essential parameters, e.g. concrete properties, cover thickness,
reinforcing bar size, corrosion depth, or that the relationship between parameters is expressed using
empirical constants which cannot be used in general.

Liu (1996) proposed a more general model by considering the steel bar – concrete cover configuration as a
thick-walled cylinder, which included all the necessary parameters. However, the mathematical expression
derived for this model indicates that corrosion of a smaller bar would generate larger pressure than a larger
bar, which contradicts the general observation of greater expansion and cracking caused by larger bars. In
addition, the derivation of the relationship between concrete strength and the pressure caused by rust
formation uses a thin-walled configuration, despite the use of a thick-wall configuration in the overall model.
Both these inconsistent aspects need to be reconsidered in Liu’s model. Further models were published by
other investigators, e.g. Bhargava et al. (2005), Cusson et al. (2011) and Li et al. (2005), but they used a
similar analysis to that of Liu (1996) with respect to these two main points.

Chen and Xiao (2012) used a fracture mechanics approach to analyse crack initiation, propagation, and
crack width, in the concrete around corroding steel bars. They considered that the critical crack front divides
the thick-walled cylinder into two zones, i.e. uncracked concrete in front of it and cracked concrete behind,
and derived a set of equations for crack width, as well as other mechanical parameters of the cracked
concrete at different stages. Pan and Lu (2012) presented a finite element analysis for stochastic modelling
of reinforced concrete cracking in a mixed localised, uniform pattern and quantified it as a transient
displacement boundary condition. These models are promising, although they lack simplicity and their
applicability needs to be carefully verified.

Austroads 2016 | page 73


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Based on Liu’s model, Xu and Shayan (forthcoming) derived an equation to correlate the depth of steel
corrosion with cracking of the cover concrete. The critical corrosion depth, crit, is expressed by Equation 28:

ft C (C   ) b 2  a 2 28
 crit  ( 2  )
Eeff  b  a2

where

ft = concrete tensile strength

Eeff = effective modulus of elasticity (considering the creep of concrete)

 = Poisson’s ratio

Eeff = E/(1+)

=
 creep factor

C = cover depth

=
 rebar diameter

a = /2

b = C+ a

Equation 28 shows that the critical corrosion depth increases with cover depth, i.e. larger amounts of
corrosion products are needed to crack a thicker cover concrete, while it decreases with the bar size
(Figure 5.5). It also indicates that high strength concrete may not be much superior to lower strength
concrete in resisting rust-induced pressure, because the creep factor of low strength concrete is high, and in
effect, the ratio of ft /Eeff is similar to that of high strength concrete.

This equation has been proven to be applicable to corrosion-induced concrete cracking in concrete slabs of
various cover thicknesses and containing various sizes of steel bars (Austroads 2004).

Figure 5.5: Dependence of critical corrosion depth on cover depth and bar diameters

Austroads 2016 | page 74


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

During the corrosion period, the corrosion rate will increase with further build-up of the chloride at the steel
bar surface. The relationship between the corrosion rate and the chloride content can be considered to be
linear in a certain concentration range, and the corrosion depth is expressed by Equation 29.

t2 29
  k corr,t1t 2  C (t , X C )dt
t1

where

kcorr, t1-t2 = rate of corrosion (µm/year/[Cl])

t1 = the time when the chloride content at steel depth (XC) reaches Ccrit

the time when the chloride content has increased to a level which has resulted in
t2 = the critical corrosion depth, which caused cracking of the cover concrete, i. e.
(t2) = crit

If the relationship between the corrosion rate and chloride content is not linear in the whole range of C(t1, XC)
to C(t2, XC), then this range can be divided into several smaller intervals in which the relation is linear, and
the calculation can be performed using Equation 30.

 4  z 3 t  z 3 t z 5 t  z 5 t z 7 t  z 7 t  30
  kcorr CS (t 2  t1 )  (CS  C0 )  zt 2t 2  zt1t1  t 2 2 t1 1  t 2 2 t1 1  t 2 2 t1 1  
  3 30 210  

where

t1 =
time to corrosion initiation

t2 =
time to cracking
=
zt2 XC/(4Dt2)
=
zt1 XC/(4Dt1)

Equation 30 is valid when zt2 < zt1 < 1.25. For larger z values, see the derivation of this equation in Appendix
E.

In recent years, ARRB Group has conducted an Austroads sponsored corrosion study on 32 MPa and
50 MPa concrete strength grades in the chloride content range of [Cl] = 0.25 to 1.5% by cement mass, and
at two temperatures of 20 °C and 40 °C (Austroads 2004). The results are plotted in Figure 4.4 and show
that the average corrosion rate of steel in the 50 MPa concrete was icorr (µA/cm2) = 0.127·C(t, x), which is
equal to kcorr (µm/year) = 1.47·C(t, x). The case study of time to cracking for 24 mm diameter steel bars in 50
MPa concrete is illustrated in Figure 5.6. It shows that the time to cracking since the initiation of corrosion ( t
= t2 – t1) substantially increases with the increase in cover thickness. For example, the total time to cracking
for a cover of 125 mm is 250 years, which is more than 150 years longer than that of a 75 mm cover.

Austroads 2016 | page 75


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 5.6: Time to crack formation in cover concrete due to carbon steel corrosion; bar size 24 mm, concrete
grade 50 MPa, Ccrit = 0.4% of cement mass

It should be noted that this calculation has not considered the long-term effect of chloride binding by the
cementitious material. The binding capacity of the cover concrete could get incrementally exhausted over a
long period of exposure, and for each increment, the effective boundary will move into the concrete, which
may have the effect of faster corrosion than calculated.

In addition, the effect of temperature on the corrosion rate should be taken into consideration when
predicting the long-term durability. Stewart et al. (2012) have predicted a 6 °C temperature increase for the
Sydney area during this century (Figure 5.7), which would enhance the rate of steel corrosion in reinforced
concrete bridges in aggressive environments.

Figure 5.7: Predicted median temperatures and low and high estimates for Sydney, using CSIRO Mk3.5 climate
model for the A1FI emission scenario

Source: Stewart et al. (2012).

Austroads 2016 | page 76


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Melchers (2006) tested the corrosion rate of structural steel samples in Coffs Harbour seawater. The initial
corrosion rate was found to increase exponentially with an increase in the seawater temperature, and
followed the Arrhenius equation k = A·exp(-Ea/(RT)). The activation energy calculated from the carbon steel
corrosion data was Ea = 37kJ·mol-1 (Figure 5.8). The results indicated that the rate of corrosion at 26 °C
increased by 40% compared to that at 20 °C.

Therefore, service life prediction models should consider the higher corrosion rate at the higher temperatures
predicted during the next century.

Figure 5.8: Left: Initial corrosion rate of steel in seawater; Right: Arrhenius plot for the carbon steel data

Source: Melchers (2006).

5.3 Time to Functional Failure Due to Steel Corrosion


The length of time to functional failure depends on the definition of minimum acceptable performance, i.e.
when n% of the functional capacity remains, as illustrated earlier in Figure 3.1. It can also be defined as the
repair threshold, i.e. when m% (m = 1 – n) of the structure or certain structural components show severe
damage, such that remedial actions should be taken. Other examples are for flexural members, where
functional failure could be defined as when the load safety factor reaches a certain value, or in physical form,
such as when the steel section loss reaches, say 20%.

Weyers et al. (1994) proposed that the end of functional service life of untreated bridge decks and bridge
substructures be quantitatively defined as the point when the level of damage is 5.8–10.0% of the whole
deck, or 9.3–13.6% of the worst damaged lane.

In this phase, cracks in the cover concrete, caused by steel corrosion, will allow easier access of seawater or
moisture to the steel surface, and will further increase the corrosion rate as if the steel concerned is in
contact with seawater. For example, Melchers (2006) showed that steel samples exposed to seawater at
Coffs Harbour for two years (average temperature of 22 °C) had a 0.34 mm loss in diameter, which is a high
rate of corrosion. Hwang et al. (1994) showed that the corrosion rate of steel bars in beams subjected to
load-induced cracking under artificial marine conditions was about 1 mm/year (evaluated by non-destructive
test (NDT) methods).

Steel corrosion results in section loss of reinforcing bars and reduces the bond strength between the
concrete and the steel, which can lead to the reduction in beam stiffness (Vidal et al. 2007). The mechanical
behaviour of the components may also be altered as a result of the corrosion, e.g. the ductility could
decrease (Berto et al. 2009).

Austroads 2016 | page 77


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Xia et al. (2011) showed that both the stiffness and shear capacity of beams decreased as the corrosion
level increased, although the loss in stiffness was insignificant when the applied load was less than 20–30%
of its ultimate load. The shear capacity decreased with the increase in corrosion, in such a way that stirrup
failure, instead of concrete crushing, became the main failure mode when corrosion was severe. They also
showed that when steel corrosion was 20%, the average width of cracks was 0.2 mm and the beam shear
capacity loss was about 10%.

Hariche et al. (2012) showed that a steel corrosion of 6% in beams induced a 10–15% reduction in the
flexural strength, depending on the configuration of the steel bars, and that at 20% steel corrosion the beam
capacity was reduced to 50%. Note that in these tests, the steel corrosion was induced by impressed current
which could not produce the same amount of rust as that in slow corrosion processes.

With further corrosion, the concrete would exhibit significant signs of distress probably associated with
severe damage, although the structural members may still remain functioning for a period of time. Khayat et
al. (2005) showed some marine structures constructed in 1901 and 1928 still had reasonable strength and
elastic modulus values.

Stewart and Mullard (2007) presented the results of a spatial time-dependent reliability analysis of a
reinforced concrete bridge deck, which showed that the quality of concrete and threshold for repair greatly
influenced the life-cycle cost.

The time interval between crack formation and structural failure may be difficult to predict. For the sake of
safe operation, the bridge should be repaired when severe damage to structural elements is observed.

At present, some rule-of-thumb criteria are used to define functional failure. The Road Structures Inspection
Manual (Vicroads 2011) defines condition state 3, at which the component requires remedial actions, as the
condition in which section loss of reinforcing steel bar may reach 20%, and medium cracking may have
occurred. The section loss is not defined in the manual, and presumably refers to the sectional area instead
of steel bar diameter.

Criteria for functional failure need to be clearly defined based on structural analysis, which would then allow
the time to functional failure to be estimated with some confidence.

5.4 Probability of Failure


The probabilities of failure can be estimated in two ways, deterministic modelling and probabilistic modelling.
The former uses single-point estimations, where each uncertain variable within a model is assigned a best
guess estimate, to depict scenarios (such as best, worst, or most likely case) for each input variable. This
approach has most commonly been used for the case of concrete deterioration due to chloride-induced
corrosion. In this context, deterministic modelling uses the mean values of the relevant parameters (cover,
apparent surface chloride concentration, effective diffusion coefficient, carbonation parameter) to arrive at a
predicted service life.

The probabilistic approach, such as Monte Carlo simulations, samples the probability distribution for each
variable to produce hundreds or thousands of possible outcomes. The results are analysed to obtain
probabilities for the occurrence of different outcomes. It uses the principles of limit states and structural
reliability. In structural terms, the failure of an element occurs when the load effect or stress (S) exceeds the
resistance (R).

The probability of failure is defined as Pr{G(X) < 0}, where G(X) or G(x1, x2, …, xn) is the limit state function
describing the capacity to resist stress. This concept is described as G = R–S, depending on the variables
involved x1, … xn; and as long as G(x) > 0, the component or structure is considered to operate safely, i.e.
when resistance (R) is greater than stress (S). For example, until the chloride content of concrete (S) at the
reinforcement depth (R) is below the corrosion initiation threshold, corrosion damage would not occur.

Austroads 2016 | page 78


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Considering that the resistance (R) and stress (S) are independent and both are normally distributed with
mean and standard deviation of R, R and S, S, respectively, the probability of failure (Pf) can be
expressed as a ‘reliability index’ (), as in Equation 31:

 G ( X )  G 0  G    G  31
Pf  Pr{G( X )  0}  Pr       (  )
 G G   G 

where

(…) = the standard normal distribution function with a mean of zero and a unit standard
deviation

 = G/G, defined as the reliability or safety index (Robinson 1998)

For the R–S system, the reliability index is expressed as Equation 32:

R  S 32

 R2   S2

The safety index can be calculated according to the limit state function formula. For example, the R is the
cover thickness, the S is the depth of critical amount of chloride penetration, and G = 0 is the limit state when
the steel corrosion is possible.

In general, the probability of failure is expressed by the n-dimensional integral shown in Equation 33:

Pr{g ( X )  0}   X  f X ( x1 ,..., xn )dx1...dxn 33

where

fX(x) = the joint density for X = (X1, … Xn)

The analytical solution to a system with many variables may not be possible to obtain. There are a number of
approaches to approximate the probability of failure. The most commonly used were the mean value first
order moment (MVFOSM) and the first order reliability method (FORM); both are based on the reliability
index and use the linear part of the Taylor series expansion to obtain the approximate mean and variance
(Robinson 1998). The linear part of the mean and variance of a function having two independent variables is
expressed by Equation 34:

E[G ( X )]  G ( X ) X  34

2 2
 G ( X )   G ( X ) 
V [G ( X )]   
2
   x22  
 
x1
 x 1  X   x 2  X 

For the process of chloride-induced corrosion, x1 = CS and x2 = D; and X =  is the mean value of these
parameters. In practice, the coefficient of variation (Cv) is known or can be tested, thus the standard
deviation  = Cv·. A detailed expression of the mean value method with examples for concrete structure is
given by the DuraCrete technical report (DuraCrete 1999).

Austroads 2016 | page 79


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

McGee (2001) applied the concept of probabilistic modelling to the corrosion problem of Tasmanian bridge
stock, induced by concrete carbonation and chloride penetration to the depth of steel reinforcement, and
generated graphs of the probability of corrosion for bridges of various ages and various distances from the
coast.

More accurate estimation can be made by computing a higher order of the Taylor series expansion of the
limit state function around the design point, e.g. second order reliability method (SORM). However, the
computation will become much more complicated and more time consuming.

Saassouh and Lounis (2012), using FORM and SORM to estimate chloride diffusion through cover concrete,
demonstrated the sensitivity of the variables involved in the process. They showed that SORM resulted in
slightly lower probability of failure compared to FORM, but the difference was insignificant. The results for the
same case study by Monte Carlo simulation (MCS), at 100 000 iterations, was almost identical to that by
SORM. Their results indicate that when the coefficient of variation was relatively low (less than 50%), then
FORM was sufficiently accurate, whilst it was more efficient compared to MCS.

A number of researchers (Biondini & Frangopol 2009; Ferreira 2010; Marques & Costa 2010; Saassouh &
Lounis 2012) have used probabilistic modelling for predicting the time-dependent probability of corrosion of
reinforced concrete structures, related to carbonation and chloride penetration through the concrete cover.

The probability of component or structure failure can be solved by inputting parameters directly assessed by
experts or obtained from statistical data and empirical functions, without definite relationships existing
between the various resistance functions (for example, when several independent deterioration mechanisms
simultaneously occur in the same area of a component, and each of them is a time-dependent process). As
reviewed by Frangopol et al. (2004), the expected deterioration of a concrete structure at time t is often
proportional to a power law (Equation 35):

v(t ) 35
E ( X (t ))   at b
u

where

v(t) > 0 = shape function

u>0 = scale parameter

X(t) = continuous-time stochastic process with independent gamma-distributed


increments

a >0 and = constants


b >0

For example, b = 0.5 is for diffusion-controlled deterioration, b = 2 is for sulfate attack (Ellingwood & Mori
1993), and b = 1/8 is for time-related creep (Çinlar et al. 1977).

The function X(t) in Equation 35 follows the gamma distribution, as shown by Equation 36:

Ga(x| v, u) = uvxv-1exp(-ux)/(v), for x  0 36

The variance of the distribution is given by Equation 37:

v(t ) 37
Var ( X (t )) 
u2

Austroads 2016 | page 80


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

For a system involving several variables, the computing itself would introduce errors and make the results
somewhat unreliable. More advanced computing techniques have been developed to make more realistic
estimations, including the Rackwitz-Fiessler algorithm (Liu & Der Kiureghian 1991; Rackwitz & Fiessler 1978;
Val et al. 1996), Monte Carlo simulation and directional sampling (Bentz 2003; Melchers 1999; Nie &
Ellingwood 2000; Val et al. 1996).

However, the use of this algorithm was not without problems, as pointed out by Marques and Costa (2010),
in that the convergence between two methodologies, i.e. the safety factor and probabilistic approaches in the
estimation of service life, compared to the target life defined in the prescriptive specification, still needed to
be improved.

Bertolini et al. (2011) showed significant differences between the field inspection results on concrete
carbonation and those calculated by a probability model (FIB 2006). The doubtful reliability of the inspection
data, uncertainty in the model parameters, or in the measurement of parameters chosen, e.g. the
carbonation resistance and cover, were considered to be among the reasons for the discrepancy. The latter
study indicated that generic models may not be applicable for reliable prediction of service life.

It can be pointed out that the standard deviations used in these models were calculated from the coefficient
of variation, which is often an assumed value; whereas in the field investigation, the standard deviation
depends on the number of samples taken, and can be much greater than that assumed in the models.

Nevertheless, the safety index () can provide an estimation of the probability of the onset of potential failure.
By definition, a larger  means a smaller probability of failure. A failure probability of 10%, corresponding to a
safety index of  = 1.28, has been specified for new structures by some standards, e.g. the Norwegian Code
(Gjørv 2009), states that the safety index will decrease with time (Figure 5.9). The safety index is related to
the probability of failure, as shown in Table 5.1 (DuraCrete 1999).

Table 5.1: Relationship between safety index and probability of failure

- 1.3 2.3 3.1 3.7 4.2 4.7 5.2 5.6


Probability of failure 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8

Source: DuraCrete (1999).

At present, there are a number of probabilistic models for calculation of a concrete structure safety index,
e.g. Ehlen (2008), DURACON (Ferreira et al. 2004) and DuraCrete (1999), which mainly focus on calculation
of the corrosion initiation, i.e. time when the chloride content in concrete at the reinforcement depth reaches
the critical value. This approach is adopted because it is often considered that the next stages of corrosion,
i.e. the crack development and propagation phases, are relatively rapid. However, many structures can
perform satisfactorily even 20 years after initial crack development. At this stage, detailed engineering
inspection combined with structural analysis, taking into account the current conditions of the concrete cover
and the reinforcement, are required in order to verify the load capacity of specific elements in a given
structure.

Austroads 2016 | page 81


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Figure 5.9: The principles of a type-dependent reliability analysis

Source: Gjørv (2009).

The effect of other deterioration, such as carbonation and scaling, on the chloride penetration rate was not
considered by any of the available models. The fact that carbonation greatly reduces the chloride-binding
capacity of concrete will allow faster diffusion through the carbonated zone of concrete. This is relevant to
the atmospheric zones of marine structures but not to the submerged zones. In the latter case, after a very
long period of immersion in salt water, the capacity of the concrete to bind chloride will be exhausted, after
which the corrosion rate would increase in the remaining period of service life. This should be considered in
the assessment of functional service life of relevant structures.

Reliable prediction requires information on environmental loading, and concrete quality and structure
information. For example, the input parameters required by DURACON (Gjørv 2009) are:
 environmental loading
– chloride load, CS
– temperature, T (°K)
 concrete quality
– apparent diffusion coefficient, which depends on time and temperature, and which further requires:
- diffusion coefficient determined at an early age, e.g. starting at 28 days, Dt0
- the age factor, m
- time and temperature history during the exposure period
– critical chloride content, Ccrit, which depends on steel composition and structure type
 concrete cover thickness.

Due to the fact that the diffusion coefficient measured at an early age may be considerably different from that
at a later age, and may vary with temperature during service life, variability in the estimation of these
parameters in the actual service environment is unavoidable. The currently applied formulae in the service
life models contain uncertainties and need to be modified or changed for accurate predictions (Table 5.2). As
mentioned earlier, significant discrepancies in the results of Bertolini et al. (2011), with respect to their field
inspection data compared to the outcome of modelling, originated largely from such uncertainties.

Austroads 2016 | page 82


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

For chloride-induced corrosion, the most critical problem is the diffusion coefficient. The values used in the
models for a given structure are often determined by the NT Build 443 Test Method (Nordtest 1995) at
different ages, or those obtained from the chloride profiles obtained in field investigations of similar existing
nearby structures. In some cases, such as for the Brighton Bypass and Sorrell Bridge in Tasmania, the value
of the long-term chloride diffusion coefficient has simply been assumed to be 10% of the 28-day value. This
assumption may introduce considerable error in the prediction of service life for some structures.

As another example, the m values used by Life-365 were based on the chloride profiles of concretes
obtained from field investigations, including the data published by Bamforth (1999) on 18 concrete blocks
(cement content around 300 kg/m 3) which were exposed in the seawater splash zone for up to eight years. It
is likely that m values for concrete mixes that contain larger amounts and different types of binder would be
different from that used in Life-365, which would again generate errors in the prediction of service life.

Moreover, the NT Build 443 Test Method uses a chloride concentration of 100 g Cl-/litre (about 2.82 M) as
the immersion environment, which is very different from the chloride concentration in seawater (0.6 M).
Therefore, the diffusion coefficient determined by the NT Build 443 Test Method will be much higher than
that achieved for the same concrete were it to be exposed to seawater (Xu & Shayan 2008). The correlation
between the diffusion coefficients determined under accelerated tests and under normal exposure conditions
need to be established for various concrete strength grades and binder types.

In addition to the values of each parameter taken for the simulation, the coefficient of variation plays a vital
role in computing the probability of failure. Typical values published in the literature are presented
inTable 5.3.

Saassouh and Lounis (2012) analysed the sensitivity of the parameters used in service life prediction (i.e.
when the chloride content at the reinforcing steel surface reaches the threshold level). They remarked that
an increase in the coefficient of variation (up to 40%) of any parameter increases its importance factor, and
that the cover depth variation (cover depth of 70 mm) has the highest impact on the safety index, followed by
the diffusion coefficient, maximum chloride content at the concrete surface, and threshold chloride content.

In all the present models, the coefficients of variation used were not larger than 20%, whereas much higher
values were observed in real structures, e.g. Lindvall (2007) showed that the coefficient of variation ranged
from 8 to 33% for CS and 9 to 35% for D.

It is clear that in order to achieve the service life predicted by the models, the tolerances and/or standard
deviations specified for different parameters must be controlled and achieved.

Austroads 2016 | page 83


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 5.2: Time or temperature-dependent variables used in current service life prediction models

Dependent
Parameter Current formula Alternative and/or comment
variable
Diffusion Temperature U  1 1   T ref
coefficient D(T )  Dref exp     D(T ) 
Tref 
 R  Tref T 
Note: the viscosity of seawater is
Life-365 and DURACON higher than that of pure water
Value of U is around 40 kJ/mol (Stanley & Batten 1969).
Diffusion Time  t ref
m
 , t < 25 years (Life-365) Dref is determined by accelerated
coefficient D(t )  Dref   test (NT Build 443). This test uses
 t  2.8 M chloride solution; the result of
m this test may be applied to a very
 t ref 
D(t )  Dref   , t > 25 years (Life-365) high chloride environment, but not
 25  directly to normal seawater
diffusion.

 28 
m
  28  m  Expression of D(t) used in Concrete
D(t )  D28day    Dult 1     Works model (Riding 2007), cited
 t    t  
  by Thomas et al. (2011).
t in days; Dult = ultimate diffusivity Dult = D(t = 36 500), i.e. at 100 years

Age factor Cement type Life-365 Model: m = 0.20, for OPC The ‘m’ value chosen by different
m = 0.20 + 0.4(%FA/50 + %SG/70), FA < 50%, models can vary by 2-fold. These
SG,70% empirical values may depend on
DSF = DOPC·epx(–0.165SF), SF <15% composition of cements and the
supplementary materials. These
DURACON model: m = 0.40,  = 0.08, for OPC values need to be established for
m = 0.50,  = 0.10, for blast furnace cements local materials.
m = 0.60,  = 0.12, for fly ash cements
m = 0.48 for Dutch slag cement Polder and de Rooij (2005).
Chloride at Time Cs = 0.8% (mass of concrete) for marine splash Life-365
surface zone Comment: Actual values should be
Cs = 0.1% /year until 1.0% (mass of concrete) for determined and applied.
marine spray zone

Austroads 2016 | page 84


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 5.3: Typical values for mean and coefficient of variation for different parameters

D28-day Age factor Ccrit CS kc XC


Reference
(10-12 m2/s) (m) (% cement) (% cement) (mm/year) (mm)
Gjørv (2009) 6.0, cv = 0.11 0.40, cv = 0.40, 5.5, cv = 0.24 — 70, cv = 0.09
(Cement I 0.20 cv = 0.25 3.5, cv = 0.23 90, cv = 0.07
with 10% SF) 1.5, cv = 0.33 120, cv = 0.05
13.3, cv = 0.40, cv = 0.40, ( = 6 mm)
0.06 0.20 cv = 0.25
(Cement I)
Enright and cv = 0.05– — cv = 0.05– cv = 0.05–0.15 — cv = 0.05–0.30
Frangopol (1998) 0.20 0.15
Collins et al. (2001) cv = 0.20 — cv = 0.17 cv = 0.15 — cv = 0.13–0.25
Bentz (2003) cv = 0.25 — cv = 0.20 cv = 0.30 — cv = 0.08
Maunsell & AECOM 2.0, cv = 0.25 0.5 (65% 0.06, 0.65, cv = 0.30 — 150 (C-steel)
(2007) BFS), 0.36 cv = 0.20 55 (S-steel)
(20% FA), cv = 0.20
0.4 (25%
FA); cv =
0.25, (t < 30
years)
Ann et al. (2010), — — — — 2.73, 57.9, cv = 0.22
survey data for cv = 0.21 53.6, cv = 0.21
sound concrete, 5.79, 61.8, cv = 0.23
concrete with 0.1 cv = 0.15
mm cracks and 4.10,
construction joints cv = 0.16

There is no reasonable prediction for the time of detectable failure, t2, in Figure 3.1 by any of the available
models. Life-365 assumes a propagation period (defined as the period between the initiation time and the
end of service life) to be six years based on the results of Weyers (1998). This assumption does not seem to
have been applied elsewhere. Some other predictions (such as those for most structures listed in Table 2.2)
assumed a propagation period of 20 years before functional damage to structures occurs. Therefore, the
models have used an initiation period of ‘design life minus 20 years’. This means that the safety index
calculated by the models applies only to the first phase of the deterioration process, i.e. the initiation time or
onset of potential failure.

It should be noted that not all components of a bridge have a 100-year design life, and that some agencies
require replaceable components such as elastomeric bearings and deck wearing surfaces to have service
lives of 35 years and 20 years, respectively; e.g. as per required design lives of bridge components for
Gerringong Upgrade, presented in Appendix F (Roads and Maritime Services 2010). The 100-year design
life applies largely to concrete elements, particularly those which are difficult to replace.

Austroads 2016 | page 85


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

6. Tests for The Evaluation of Durability


To ensure long-term durability beyond 100 years and to enable accurate prediction of the service life of
concrete structures, verification tests need to be performed on the concrete ingredients, as well as the
proposed concrete formulations. Quality control tests are also required to be applied during the construction
phase. The latter tests should include those that determine the transport properties of the concrete, including
the chloride diffusion coefficient, resistance to carbonation and sulfate attack, and the rate of wetting after a
period of drying (sorption rate).

It is also important to verify the corrosion properties of stainless steel in extreme conditions, i.e. a very high
chloride content environment, as mild steel would not be suitable for use in such harsh conditions.

6.1 Resistance of Concrete to Chloride Penetration


The chloride diffusion coefficient can be calculated from the chloride penetration profile, which is obtained by
in situ testing of concrete, as well as by using accelerated test methods on concrete specimens. The former
results in the apparent diffusion coefficient, and the latter can be used for quality control of concrete.

Results obtained from accelerated tests may also provide information on the actual diffusion coefficient if
appropriate comparison data were available.

6.1.1 Apparent Chloride Diffusion Coefficient of Concrete Exposed to Seawater

This can be determined by analysing a concrete sample taken from the actual structure or from newly made
concrete by the ponding test (ASTM C1543).

This method requires sampling by: 1) impact drilling and collecting concrete powder samples at the required
depth interval for chloride analysis; and 2) core drilling of concrete in the field and then layer-wise precision
milling of the drilled core in the laboratory and collection of the concrete powder for chloride analysis.

The powder sample is dried, dissolved in nitric acid, boiled and filtered. The chloride concentration in the
solution is then determined by chemical analysis. The most convenient and accurate method of chemical
analysis for this purpose is potentiometric titration.

The chloride content profile of the concrete is fitted to the Fick’s second law of diffusion, i.e. Equation 16, by
trials of CS and D values, until the best fit is obtained. Note that the CS and D values obtained by this method
are linked, i.e. a higher CS will result in a lower D, and vice versa.

Note also that the in situ sampling may need a number of replicates from the same exposure condition in
order to get a reliable mean value and standard deviation. The ponding test method recommends doing the
sampling at exposure ages of 3, 6, and 12 months, and 12-month intervals for longer periods, although the
short exposure periods (to 3% NaCl solution at 23 °C) may not reveal very useful data.

Austroads 2016 | page 86


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

6.1.2 Accelerated Tests for Chloride Diffusion Coefficient

Currently, there are two standard accelerated test methods for determining the chloride diffusion coefficient:
NT Build 443 (Nordtest 1995 or ASTM C1556), and the NT Build 492 (Nordtest 1999).

The NT Build 443 method uses a high chloride concentration environment of 2.82 mole/L (a solution of 165 g
NaCl/L, being equivalent to 100 g Cl/L) to accelerate the diffusion process. Concrete specimens are cured
for 28 days, followed by immersion in the salt solution for 35 days, after which the chloride ingress profile is
analysed in the same way as that for the apparent diffusion coefficient.

The value of CS by this test is usually about 1% by dry mass of concrete, which is similar to that obtained
from the in situ concrete exposed to seawater for many years. The fact that the chloride concentration of this
test is 4.5 times larger than that of seawater indicates that the value of D obtained by this method would not
be the same as that of in situ concrete exposed to seawater.

It should also be noted that the duration of immersion is relatively long compared to the young age of the
concrete specimen. The test nominally determines the 28-day diffusion coefficient, whereas the age of the
specimen is 61 days at the time of testing for chloride ingress. The age difference may affect the value of the
chloride diffusion coefficient; for example, if the age factor is m = 0.4, then it can be shown that the diffusion
coefficient at the end of the test period is about 72% of that at the start.

The NT Build 492 method uses the voltage difference across an electrical field to accelerate chloride
penetration. The concentration of the chloride solution is 2 mole/L, the voltage difference is 30 volts across
50 mm (specimen thickness), and the duration is 24 hours. The test method lists the actual electrical field
strength and duration, which can vary according to the initial current (I 0), as shown in Table 6.1.

Table 6.1: Test voltage and duration (NT Build 492)

Initial current (mA) with 40 V Applied voltage after adjustment


Test duration (hours)
potential difference (V)
I0 < 5 60 96
5 < I0 < 10 60 48
10 < I0 < 15 60 24
15 < I0 < 20 50 24
20 < I0 < 30 40 24
30 < I0 < 40 35 24
40 < I0 < 60 30 24
60 < I0 < 90 25 24
90 < I0 < 120 20 24
120 < I0 < 180 15 24
180 < I0 < 360 10 24
360 < I0 10 6

At the conclusion of the test, the specimen is split to expose the interior section. A silver nitrite solution (0.1
M) is sprayed on the newly fractured surface, which will react with chloride ions to form AgCl; the latter
turning to a white colour in the zone where sufficient chloride ions have penetrated. This layer will turn black
later, on exposure to air. The thickness of the white zone is measured and recorded as Xd. The chloride
concentration at the interior boundary of the white zone is about 0.07 mole /L.

Austroads 2016 | page 87


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

The diffusion process in this test is a non-steady-state migration (nssm), and the respective diffusion
coefficient is termed Dnssm and is calculated using the following equations:

RT xd   xd 38
Dnssm 
zFE t
RT  2c 
 2 ·erf 1 1  d 
zFE  c0 

where

R = gas constant, R = 8.31441 JK-1mol-1

T = temperature (K)

z = valency of chloride z=1

F = Faraday constant, F = 9.64846  104 Cmol-1

E = electrical field strength, e.g. 30 V/0.05 m

c0 = 2 mole/L

cd = 0.07 mole/L

The NT Build 492 standard gives a convenient formula by assuming  = 0.0238 for the basic test condition.

Note that Dnssm is the diffusion coefficient of the free chloride ions, because the ions move rapidly due to the
effect of the electrical gradient, such that the residency time for the chloride ions to be adsorbed by the
cement paste is too short, resulting in rapid penetration of the free chloride ions (Xu & Shayan 2008). Thus,
the chloride binding capacity (expressed by Equation 18), which varies for different binders, also needs to be
taken into account to understand the total chloride status of the concrete.

It should also be noted that Equation 38 is not valid for small penetration depth, for example, at and near xd -
xd = 0. The correct general approach for this type of migration is to apply the diffusion equation with a flow
zFE
D
of velocity V , where V = RT in Equation 39.

c0  x  Vt x  Vt  39
V
x
C ( x, t )  erfc ( )  e D erfc ( )
2 4 Dt 4 Dt 

6.2 Other Transport Properties of Concrete


The performance of concrete subjected to many aggressive environments is, to a large extent, a function of
the penetrability of the pore system. The rate of liquid and gas penetration into the concrete depends on the
concrete pore volume and the pore structure. The VPV (volume of permeable voids) test and sorptivity test
are used to determine these properties and are considered to reflect the durability properties of concrete.

Austroads 2016 | page 88


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

6.2.1 Apparent Volume of Permeable Voids

The Australian Standard Test Method AS 1012.21-1999 (which is similar to ASTM C642-06) estimates the
maximum amount of water that can be absorbed by an oven-dried concrete. The specimens are dried to
constant mass, then immersed in water until the absorption stops, and then boiled under water for five hours.
The test results are the water absorption, which is determined from the second step, and the VPV, which is
the difference in the volume of water absorbed between the second and third steps. The VPV is expressed
as a percentage by volume of concrete and is known as the VPV value of concrete.

The pore volume varies depending on concrete grade, compaction and binder composition. VicRoads uses
this test in the specification of structural concrete (Section 610) and has established the thresholds for
acceptance of concrete shown in Table 6.2.

Table 6.2: VicRoads requirements for VPV for various grades of structural concrete

Cementitious Water / VPV tested at age 28 days (%)


Concrete material content cementitious
grade material ratio Test cylinders Test cylinders
minimum (kg/m3) Test cores
(max.) (vibration) (rodding)
VR330/32 330 0.50 14 15 17
VR400/40 400 0.45 13 14 16
VR450/50 450 0.40 12 13 15
VR470/55 470 0.36 11 12 14

Source: VicRoads (2010a): Table 610.061 and 610.071.

Note that in this test, the standard test cylinder or core is sawn into four smaller segments and the test is
performed on each segment, but the average is reported as the overall VPV value of the specimen. Although
the VPV value is meant to reflect the pore structure of the hydrated cement paste, as a result of the sawing,
the coarse aggregate particles are directly exposed to water, thus the VPV value is influenced by water
absorption of the coarse aggregate.

Experience at ARRB shows that the VPV obtained for exactly the same concrete mix design was influenced
by the type of coarse aggregate, to the extent that one did not comply with the requirements of Table 6.2 and
the other did. Therefore, in some cases, a VPV slightly higher than that required by Table 6.2 may not
indicate poor durability, because the discrepancy may arise from sawing of the aggregate used, whereas in
concrete structures, the aggregate particles are totally covered by the cement paste.

6.2.2 Water Sorption Test

The water sorption test is mainly for evaluating the sorption properties of concrete surfaces, which in field
situations are exposed to the open air and can dry out prematurely. This would lead to a more porous layer
at the concrete surface, adversely affecting its quality. The rate of water absorption by concrete, or depth of
water penetration into concrete after a period of soaking, following the RMS T362 Test Method (Roads and
Traffic Authority 2001) can reflect the loss in the quality of the concrete surface. The rate of water absorption,
which is governed by the capillary structure of the concrete, also reflects the resistance of concrete to liquid
penetration. The water sorption of sawn surfaces can also be tested (ASTM C1585-11).

In this test method, the specimens are exposed to a drying condition for a specified period before testing,
which is different from the VPV test. The two commonly used test methods in Australia and their definitions
for sorptivity are given in Table 6.3.

ASTM C1585 specifies two sorptivity values: S1, the initial absorption (e.g. 0 to 6 hours), and S2 for later
times. There are no guidelines as to how these values are indicative of concrete durability.

Austroads 2016 | page 89


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Note that in testing concrete, the plot of penetration depth against t is often shown to be curvilinear rather
than a straight line, as predicted by the capillary action theory. This discrepancy indicates that more research
is needed to clarify the mechanism.

The Roads and Maritime Specification B80 (Roads and Maritime Services 2012b) specifies requirements for
concrete when tested in accordance with the RMS T362 Test Method, as shown in Table 6.4.

Table 6.3: Sorptivity tests

Standard Specimen and curing Pre-conditioning and time Test condition and properties tested
RMS T362 Concrete prism coated 50  5% RH, 23  2 °C Immersed in water for 24 hours (6 hours
with curing compound Exposure: A and B1: 21 days for A1 and B1 exposure class), and
and exposed to breaking the specimen to measure the
Exposure: B2: 28 days
simulate field condition depth of water penetration by dusting the
Exposure: C: 35 days exposed surface with methylene blue dye.
ASTM Concrete cylinder or 80  3% RH, 50  2 °C for 3 Once surface is in contact with water,
C1585-11 drilled core days measure the mass change at intervals
then in sealed container at 23° and calculate the ‘sorptivity’, S (mm/s)
 2 °C for 15 days or more defined as the volume of water per
exposed area per square root of time.
(to obtain an internal RH 50 to
70%)

Table 6.4: Durability requirements for concrete

Minimum Maximum water to Maximum sorptivity penetration depth (mm)


Exposure cement cement ratio (determined according to RMS T362)
classification
(kg/m3) (by mass) Portland cement Blended cement
A 320 0.56 35 35
B1 320 0.50 25 25
B2 370 0.46 17 20
C 420 0.40 N/A 8

Source: Roads and Maritime Services (2012b): Table B80.6.

A practical use of the sorptivity data may be for the estimation of seawater absorption by concrete exposed
to drying and wetting cycles, which will enhance the chloride penetration. The significance and extent of this
effect are yet to be quantified, although Sirivivatnanon and Khatri (2011) have correlated sorptivity to chloride
penetration resistance of concrete.

6.2.3 Resistance to Carbonation

Carbonation depth of concrete is determined by spraying a solution of phenolphthalein pH indicator onto the
newly exposed concrete surface and measuring the depth of the colourless layer that develops at the
concrete surface. Concrete below this layer is alkaline and develops a pink colour on the fracture surface.
The test can be performed on in situ concrete by core extraction from the structure, or in drill holes.

The resistance of concrete to carbonation can be evaluated by accelerated testing using high CO 2
concentration, e.g. 3% CO2 which is 100 times higher than the CO2 concentration in the ambient
atmosphere. However, the carbonation rate recorded by different researchers has varied widely, probably
due to differences in the properties of the different concretes used, and a clear conclusion relating the
carbonation rate to CO2 concentration has not been established. This may arise, at least partly, because the
carbonation of concrete produces water, which blocks the path for the CO 2 to penetrate deeper. This effect
would be more prominent for the carbonation at higher CO2 concentration condition, e.g. 100% CO2.

Austroads 2016 | page 90


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Nevertheless, Sisomphon and Franke (2007) used a simplified relationship between the carbonation rates
under accelerated and ambient conditions, in the form of the relationship:

Kacc / Kamb = √(Cacc/Camb), where K and C refer to the carbonation rate and CO2 concentration under
accelerated and ambient conditions. For the example given above, the carbonation rate for exposure to 3%
CO2 concentration would be: Kacc / Kamb = √(3/ 0.03) = √100 = 10, based on Equation 40, i.e. the carbonation
rate for accelerated exposure under 3% CO2 concentration would be 10 times faster than the rate under
ambient conditions.

Kacc / Kamb = √(Cacc/Camb) 40

where

carbonation rate, Kacc is the carbonation rate under accelerated conditions (3%
K = CO2 concentration) and Kamb is the carbonation rate under ambient conditions
(0.03% CO2 concentration)

C = CO2 concentration, Cacc = 3% and Camb = 0.03%

Although the theoretical process of concrete carbonation cannot satisfactorily be described by the published
empirical formulae currently available, the latter together with the accelerated testing are useful in comparing
the carbonation resistance of different types of concrete.

6.3 Tests for Preventable Potential Durability Problems


As discussed in the earlier sections, some deterioration problems can develop due to the use of unsound or
unsuitable materials in concrete and can be prevented at the time of manufacturing the concrete by using
alternative sound materials. These problems are often exhibited many years after construction of the
structures concerned, and are very costly to manage. Since there are no cures to ‘correct’ built-in problems
after completion of the structure, it is essential that the concrete ingredients and/or concrete mixes are tested
for potential durability problems before the mix design is approved.

6.3.1 Alkali-aggregate Reactivity

This problem was addressed in some detail in Section 4.5.7, where the list of reactive aggregates detected in
Australia is presented together with the list of structures identified to date to have been affected by AAR.

It is essential that reactive aggregates are identified, and their use managed through appropriate concrete
mix design procedures. It has been established that the test methods listed in Table 6.5 can identify a wide
range of reactive aggregates, from slowly-reactive aggregates to highly-reactive aggregates. The proposed
aggregates need to be assessed by some of these tests well before the construction dates, and relevant
management strategies established.

Austroads 2016 | page 91


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Table 6.5: Test methods for assessment of AAR

Storage
Test method Specimens Criterion for reactive aggregate
conditions
In 1 M Expansion at 21-day ≥ 0.10%, coarse
RMS T363, Mortar bars (2.5 x 2.5 mm cross-section), aggregate
NaOH
RC 376.03 cement-to-sand ratio = 4:9, and w/c ca
(VicRoads) 0.4
solution at Expansion at 21-day ≥ 0.15%, fine
80 °C aggregate
RMS T364, Concrete prisms (75 x 75 mm cross-
At 38 °C and
RC 376.04 section), w/c ca 0.42, cement 420 kg/m3, Expansion at 1 year ≥ 0.03%
100% RH
(VicRoads) alkali 1.38% Na2Oeq

CPT60 Concrete prisms (75 x 75 mm cross- Expansion at 4 months ≥ 0.03%


At 60 °C and
section), w/c ca 0.42, cement 420 kg/m3,
Rilem AAR-4 100% RH Shayan et al. (2008)
alkali 1.25% Na2Oeq

When the use of reactive aggregate is unavoidable, mitigation of AAR would involve incorporation of
sufficient amounts of SCMs in the concrete mix, as stated earlier. This would serve the dual purpose of
countering both AAR and DEF expansion problems.

6.3.2 Resistance to Sulfate and Acid Attacks

The mechanisms of sulfate and acid attacks are discussed in Section 4.5.11 and Section 4.5.13,
respectively. In exposure environments where these attacks are likely, appropriate sulfate-resisting binders
need to be selected and used in concrete mix designs which are suitable for sulfate-rich environments.
Australian Standards method AS 2350.14-2006, which involves expansion measurement of mortar bars
exposed to a 5% Na2SO4 solution for 16 weeks, can be used for the assessment of binders. Blended
cements containing large amounts of SCMs, such as fly ash or slag, are suitable as binders for concrete to
resist sulfate and acid attacks, including those containing limestone aggregate. Approaches such as those
used by Khatri et al. (2009) and Sirivivatnanoon and Lucas (2011) would be appropriate for the development
and assessment of concrete mixes proposed for sulfate-rich environments.

6.4 Corrosion of Stainless Steels


Various aspects of corrosion of carbon steel are relatively well-established. Details in relation to the
performance of stainless steel in concrete are given in Section 4.3.3. Although some stainless steel varieties
have performed well in reinforced concrete exposed to marine conditions for some 70 years, their use in
structures with much longer service lives beyond 100 years needs to be investigated. Particularly, situations
such as when thin cover thicknesses are specified for atmospheric zones need attention because thin covers
will likely be completely carbonated after 100 years, and the embedded stainless steel may then be exposed
to high contents of chloride originating from airborne sources, such as sea spray. It should be noted that the
capacity of cement paste to bind chloride ions is severely diminished due to carbonation, allowing for faster
ingress of chloride into concrete.

Moreover, cheaper, low-Ni varieties of stainless steel are new, without a long history of use in harsh
conditions. The performance of these varieties needs to be investigated in normal and carbonated concretes
at chloride contents in the range of 1–6% by cement mass; the latter being the equivalent of 1% chloride by
mass of concrete, which has been encountered in many old bridges.

In addition, the performance of low-Ni stainless steels need to be investigated under exposure conditions of
high acidity and sulfate content, such as those prevailing in acid sulfate soils, as well as high chloride
contents of marine environments.

Austroads 2016 | page 92


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

7. Summary and Conclusions


A detailed literature review has been conducted on the various deterioration mechanisms and durability
issues that affect the service life of reinforced concrete structures. This includes information from sources
within Australia and overseas, including published literature, standards, codes and jurisdiction reports.

Data collected by various jurisdictions on existing bridges have shown that they started to deteriorate after
only 30 years, due partly to inadequate standards and specifications and partly to lack of knowledge
regarding the durability issues involved. The main purpose of this report is to identify critical gaps in the
existing information regarding performance of concrete bridges in aggressive environments, so that such
issues could be addressed, which would enable design service lives of 100 years to be achieved.

At present, road agencies engage durability specialists to develop detailed durability plans for specific major
bridge structures that are expected to last over 100 years. Examples are provided of such bridges
constructed in some jurisdictions, and the issues that were addressed in each case.

A common feature of the specifications, all the durability plans, the Australian Standard AS 5100.5-2004, and
the international literature is the use of blended cements for combating various deterioration problems and
achieving long service lives for reinforced concrete structures. Potentially dangerous deterioration
mechanisms, such as alkali-aggregate reaction (AAR) can be avoided.

The existing information shows that realising service lives of beyond 100 years in aggressive environments
requires some of the following actions, where possible:
 increase in concrete cover thickness (not always practical)
 use of appropriate protective surface coatings
 isolation of the concrete from the aggressive environment by appropriate encasement
 use of corrosion inhibitors
 installation of cathodic protection or cathodic prevention
 use of corrosion-resistant materials such as appropriate varieties of stainless steel
 use of high performance concrete mixes
 neutralisation of acidic soils.

Of course, careful supervision of all aspects of design and construction phases is of paramount importance.

An issue that has been noted with respect to delayed ettringite formation (DEF), is that most agency
specifications and AS 5100.5-2004 allow sulfate contents in the cement of 5% SO3 by mass of cement,
which equates to over 8% gypsum in the cement, and is considered to be too high. It is suggested that the
allowable SO3 content of cement be reduced to a maximum of 3.0%.

The literature review identified some gaps in the available knowledge, which should be addressed through
future research efforts, including:
 uncertainties in terms of accuracy of the chloride diffusion coefficient of concrete, which need to be better
clarified through further research
 uncertainties in models for the prediction of time to the appearance of detectable damage and time of
failure
 the long-term tolerance of concrete and stainless steel reinforcing materials to high chloride contents,
particularly for the new, low-Ni varieties of stainless steel for which the chloride tolerance is not well-
established and requires further research.

Austroads 2016 | page 93


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

7.1 Materials for Experimental Work (Austroads 2016)

7.1.1 Concrete Materials

Based on the review of existing information on the durability plans for various agency bridge structures
located in aggressive environments, a number of concrete mixes were identified, as shown in Table 7.1.

Table 7.1: Various concrete mix designs used for various bridge structures

1 2 3 4 5 6 7 8
Ingredient VS40 VS50
VR400/40 S40
Pump Pump
Cement 400 320 260 320 240 360 333 480
Fly ash – 107 125 150 40 30 31.5 100
Slag – – 110 132 100 90 85.5 –
Coarse aggregate 1150 1020 1190 1185 1050 1030 1110 1160
Sand 690 636 655 533 870 755 605 515
Water 180 184 172 174 160 180 180 190
Water reducer L/m3 – 1.28 0.72 0.9 – – 2.5 –
Superplasticiser (kg/m3) 1.9 – – 2 – – – –
Superplasticiser (L/m3) 1.6 0.43 1.16 – 2.47 3.12 – 3.2
Air entraining agent – – 2 2 – – – –
Total binder 400 427 495 602 380 480 450 580
Water/binder ratio 0.45 0.43 0.35 0.29 0.42 0.38 0.40 0.33
Sand/total aggregate ratio 0.38 0.385 0.36 0.31 0.45 0.42 0.35 0.31
Target slump (mm) 80 100 < 200 < 200 150 150 – 80
Tested slump (mm) 81 130 180 180 – – – 180
Comp. strength 1d – 12 – 15.5 Not available (N/A)
(MPa)
3d – – 35.5 – N/A
7d 42.7 47 46.2 47.0 N/A
28d 55.4 56 52.9 71.5 N/A
56d – – – 80.5 N/A
90d 63.8 – – – N/A

Mix 1 is the plain cement concrete used in the experimental work as reference 40 MPa concrete.

Mix 2 is the same 40 MPa concrete but with fly ash replacement (not to be used in Phase 2).

Mixes 3 and 5 are triple blended 40 MPa concrete, but are too weak for a 100-year service life, although Mix
3 may pick up strength at later ages.

Mix 4 is a triple blend 50 MPa concrete but is too high in cementitious content (may not be an economical
mix).

Mixes 6–8 have been used by road agencies but test results were not provided. Note that Mix 8 has only
17% fly ash/mass of binder and may not be adequate for durability in the long-term.

In consultation with the Austroads Project Manager, it has been agreed that Mix 1 and Mix 6 be used for
Phase 2 trials.

Austroads 2016 | page 94


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

7.1.2 Steel Component

Three types of steel have been selected for the experimental phase of the project, being 500 MPa mild steel
(carbon steel), stainless steel of high-Ni content (Grade 316), and the newer low-Ni content (LDX 2101)
varieties. The mild steel is chosen for reference, and the stainless steels for high performance in aggressive
exposure conditions. Lower grades of stainless steel would not be suitable for such applications.

Parameters included in the experimental phase are:


 concrete type
 steel type
 aggregate type (e.g. for AAR work)
 exposure condition (level of chloride contamination, with consideration of content)
 exposure for sulfate resistance
 combined carbonation and chloride contamination.

Durability issues such as AAR are dealt with at the time of selection of concrete materials.

Austroads 2016 | page 95


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

References
Abdullah, S, Shayan, A & Al-Mahaidi, R 2012a, ‘Strain monitoring of CFRP wrapped RC columns damaged
by alkali aggregate reaction’, International conference on alkali-aggregate reaction, 14th, Austin, Texas,
USA, 2012, University of Texas, Austin, USA.

Abdullah, S, Al-Mahaidi, R & Shayan, A 2012b, ‘Experimental investigation of CFRP confined columns
damaged by alkali aggregate reaction’, International Journal of Integrated Engineering, vol. 4, no. 2, pp.
49–52.

AbiGroup 2011, ‘Hunter Expressway’, prepared for Roads and Maritime Services, Sydney, NSW.

Ahmad, S 2003, ‘Reinforcement corrosion in concrete structures, its monitoring and service life prediction: a
review’, Cement and Concrete Composites, vol. 25, no. 4-5, pp. 459–71.

Aldridge, L, Collins, F, Gates, W, Garcez, E, Islam, J, Huang, L, Ong, M & Fernado, K 2013, ‘Service life of
concrete during chloride ingress’, Proceedings of the 25th biennial conference of Concrete Institute of
Australia, Gold Coast, Queensland, 16-18 October 2013.

Alexander, M & Thomas, M 2015, ‘Service life prediction and performance testing: current developments and
practical applications’, Cement and Concrete Research, vol. 78, no. 2, pp. 155–64.

Alkadhimi, TKH & Banfill, PFG 1996, ‘The effects of electrochemical re-alkalisation on alkali-silica expansion
in concrete’, Proceedings of the 10th international conference on alkali-aggregate reaction in concrete,
18-23 August 1996, Melbourne, Australia, pp. 637–44.

Alonso, C, Andrade, C, Castello, M & Castro, P 2000, ‘Chloride threshold values to depassivate reinforcing
bars embedded in a standardized OPC mortar’, Cement and Concrete Research, vol. 30, no. 7, pp.
1047–55.

American Concrete Institute 2001, Protection of metals in concrete against corrosion, ACI 222R-01, ACI,
Farmington Hills, MI, USA.

American Concrete Institute 2003, Design and construction practices to mitigate corrosion of reinforcement
in concrete structures, ACI 222.3R-03, ACI, Farmington Hills, MI, USA.

American Concrete Institute 2008, Guide to durable concrete, ACI 201.2R-10, ACI, Farmington Hills, MI,
USA.

ANCON 2016, Stainless steel reinforcement, Ancon Building Products, Sydney, NSW, viewed 8 July 2016,
<http://www.ancon.com.au/products/stainless-steel-reinforcement>.

Andrade, C & Page, CL 1986, ‘Pore solution chemistry and corrosion in hydrated cement systems containing
chloride salts: a study of cation specific effects’, British Corrosion Journal, vol. 21, no. 1, pp. 49–53.

Andrade, C, Castellote, M & D’Andrea, R 2011, ‘Measurement of ageing effect on chloride diffusion
coefficients in cementitious matrices’, Journal of Nuclear Materials, vol. 412, no. 1, pp. 209–16.

Andrews-Phaedonos, F, Shayan, A & Xu, A 2011, ‘Extending the service life of concrete bridges: corrosion
monitoring of Lynch’s Bridge over Maribyrnong River: early coating provides enhanced performance’,
Concrete Institute of Australia, 25th biennial conference, Perth, West Australia, Concrete Institute of
Australia, Sydney, NSW.

Austroads 2016 | page 96


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Andrews-Phaedonos, F, Shayan, A & Xu, A 2013, ‘Use of corrosion monitoring sensors to monitor the in-situ
performance and intervention needs of reinforced concrete structures’, Concrete Institute of Australia,
26th biennial conference, Brisbane, Queensland, Concrete Institute of Australia, Sydney, NSW.

Angst, U, Elsener, B, Larsen, CK & Vennesland, O 2009, ‘Critical chloride content in reinforced concrete: a
review’, Cement and Concrete Research, vol. 39, no. 12, pp. 1122–38.

Angst, U & Vennesland, O 2007, Critical chloride content: state of the art, report SBF BK A07037, COIN
Concrete Innovation Centre, SINTEF Building and Infrastructure, Trondheim, Norway.

Ann, KY, Ahn, JH & Ryou, JS 2009, ‘The importance of chloride content at the concrete surface in assessing
the time to corrosion of steel in concrete structures’, Construction and Building Materials, vol. 23, no. 1,
pp. 239–45.

Ann, KY, Jung, HS, Kim, HS, Kim, SS & Moon, HY 2006, ‘Effects of calcium nitrite-based corrosion inhibitor
in preventing corrosion of embedded steel in concrete’, Cement and Concrete Research, vol. 36, no. 3,
pp. 530–35.

Ann, KY, Pack, SW, Hwang, JP, Song, HW & Kim, SH 2010, ‘Service life prediction of a concrete bridge
structure subjected to carbonation’, Construction and Building Materials, vol. 24, no. 8, pp. 1494–1501.

Apostolopoulos, CA & Koutsoukos, PG 2008, ‘Study of the corrosion of reinforcement in concrete elements
used for the repair of monuments’, Construction and Building Materials, vol. 22, no. 7, pp. 1583–93.

Atkins, PW 1987, Physical chemistry, 3rd edn, Oxford University Press, Oxford, UK.

Austroads 2000a, Monitoring of steel corrosion in concrete, AP-T06-00, Austroads, Sydney, NSW.

Austroads 2000b, Service life prediction of reinforced concrete structures, AP-T07-00, Austroads, Sydney,
NSW.

Austroads 2002, Management of concrete bridge structures to extend their service life, AP-T11-02,
Austroads, Sydney, NSW.

Austroads 2004, Concrete durability: development of models to predict and extend the service life of
concrete bridges, AP-R245-04, Austroads, Sydney, NSW.

Austroads 2016, ‘Realising 100-year design life of bridge structures in an aggressive environment:
experimental work’, AP-T317-16 Austroads, Sydney, NSW.

Ballina Bypass Alliance 2008, ‘Ballina Bypass B80: concrete specification for Ballina Bypass based on RTA
B80’, project no. BBA-DU2402-SP01, final design, prepared for Roads and Maritime Services, Sydney,
NSW.

Bamforth, PB 1999, ‘The derivation of input data for modeling chloride ingress from eight year UK coastal
exposure trials’, Magazine of Concrete Research, vol. 51, no. 2, pp. 87-96.

Banora Point Upgrade Alliance 2011, ‘Durability report’, project no. D/00092RTA, prepared for Roads and
Maritime Services, Sydney, NSW.

Baroghel-Bouny, V, Wang, X, Thiery, M, Saillio, M & Barberon, F 2012, ‘Prediction of chloride binding
isotherms of cementitious materials by analytical model or numerical inverse analysis’, Cement and
Concrete Research, vol. 42, no. 9, pp. 1207–24.

Bautista, A, Blanco, G & Velasco, F 2006, ‘Corrosion behaviour of low-nickel austenitic stainless steels
reinforcements: a comparative study in simulated pore solutions’, Cement and Concrete Research, vol.
36, no. 10, pp. 1922–30.

Austroads 2016 | page 97


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Beaudoin, JJ, Raki, L & Alizadeh, R 2009, ‘A 29Si MAS NMR study of modified C–S–H nanostructures’,
Cement and Concrete Composites, vol. 31, no. 8, pp. 585–90.

Bellezze, T, Malavolta, M, Quaranta, A, Ruffini, N & Roventi, G 2006, ‘Corrosion behaviour in concrete of
three differently galvanised steel bars’, Cement and Concrete Composites, vol. 28, no. 3, pp. 246–55.

Bentour, A, Diamond, S & Berke, NS 1997, Steel corrosion in concrete: fundamentals and civil engineering
practice, E & FN Spon, London, UK.

Bentz, EC 2003, ‘Probabilistic modeling of service life for structures subjected to chlorides’, ACI Materials
Journal, vol. 100, no. 5, pp. 391–7.

Berke, NS & Hicks, MC 1998, ‘Long-term corrosion performance of epoxy-coated steel and calcium nitrate’,
Corrosion 98: NACE international conference, San Diego, California, NACE, Houston, Texas, USA.

Bertolini, L, Bolzoni, F, Pastore, T & Pedeferri, P 1996, ‘Behaviour of stainless steel in simulated concrete
pore solution’, British Corrosion Journal, vol. 31, no. 3, pp. 218–22.

Bertolini, L, Gastaldi, M, Pastore, T & Pedeferri, MP 2000, ‘Corrosion behaviour of stainless steels in chloride
contaminated and carbonated concrete’, Internationale Zeitschrift fur Bauinstandsetzen und
Baudenkmalpfledge, pp. 273–92.

Bertolini, L & Gastaldi, M 2009, ‘Corrosion resistance of austenitic and low-nickel duplex stainless steel bars’,
Eurocorr 2009, 6-9 September, Nice, France.

Bertolini, L & Gastaldi, M 2011, ‘Corrosion resistance of low-nickel duplex stainless steel rebars’, Materials
and Corrosion, vol. 62, pp. 120–9.

Bertolini, L, Lollini, F & Redaelli, E 2011, ‘Durability design of reinforced concrete structures,’ Construction
Materials, vol. 164, no. CM6, pp. 273–82.

Berto, L, Vitaliani, R, Saetta, A & Simioni, P 2009, ‘Seismic assessment of existing RC structures affected by
degradation phenomena’, Structural Safety, vol. 31, no. 4, pp. 284–97.

Bhargava, K, Ghosh, AK, Mori, Y & Ramanujam, S 2005, ‘Modeling of time to corrosion-induced cover
cracking in reinforced concrete structures’, Cement and Concrete Research, vol. 35, no. 11, pp. 2203–18.

Biondini, F & Frangopol, DM 2009, ‘Lifetime reliability-based optimization of reinforced concrete cross-
sections under corrosion’, Structural Safety, vol. 31, no. 6, pp. 483–9.

Blaikie, NK, Bowling, AJ & Carse, A 1996, ‘The assessment and management of alkali silica reaction in the
Gordon River Power Development intake tower’, Proceedings of the 10th international conference on
AAR, Melbourne, Australia, pp. 500–7.

Blight, GE & Alexander, MG 2011, Alkali-aggregate reaction and structural damage to concrete, Taylor &
Francis, London, UK.

Blue Water 2008, ‘Sydney desalination project: report, durability plan final design’, report no. WTW0155-G-0-
RP-1002-5 (revision 5), Blue Water.

Bolin, JO, Lindbladh, L & Nilsson, PA 1997, ‘The management of the Öland Bridge: a challenge to the bridge
owner’, International conference repair of concrete structures: from theory to practice in a marine
environment, Svolvaer, Norway, Norwegian Road Research Laboratory, pp. 69–79.

Bonen, D 1997, ‘The microstructure of concrete subject to high-magnesium and high-magnesium sulfate
brine attack’, International congress on the chemistry of cement, 10th, Gothenburg, Sweden, Amarkai AB
and Congrex, paper 4iv022.

Austroads 2016 | page 98


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Bouteiller, V, Cremona, C, Baroghel-Bouney, V & Maloula, A 2012, ‘Corrosion initiation of reinforced


concretes based on Portland or GGBS cements: chloride contents and electrochemical characterisation
versus time’, Cement and Concrete Research, vol. 42, pp. 1456–67.

Boyd, WK & Tripler, AB 1968, ‘Corrosion of reinforced steel bars in concrete,’ Materials Protection, vol. 7, no.
10, pp. 40–7.

BRE 1991, Sulphate and acid resistance of concrete in the ground, BRE digest 363, Building Research
Establishment, UK.

BRE 2005, BRE special digest 1: concrete in aggressive ground, Building Research Establishment, Watford,
UK.

Broomfield, JP 2007, Corrosion of steel in concrete: understanding, investigation and repair, 2nd edn, Taylor
& Francis, New York, NY, USA.

Burden, D 2006, The durability of concrete containing high levels of fly ash, PCA R&D serial no.2989,
Portland Cement Association, Skokie, IL, USA.

Byfors, K 1987, ‘Influence of silica fume and fly ash on chloride diffusion and pH values in cement paste’,
Cement and Concrete Research, vol. 17, pp. 115–30.

Camden Valley Way Upgrade C2C Alliance 2011, ‘Final durability plan report’, design lot no: DU.001, report
prepared for Roads & Maritime Services, Sydney, NSW.

Cao, HT, Bucea, L, Raf, A & Yozghatlian, S 1997, ‘The effect of cement composition and pH of environment
on sulfate resistance of Portland cements and blended cements’, Cement and Concrete Composites, vol.
19, no. 2, pp.161–71.

Carse, A 1988, ‘Field survey of concrete structures suffering alkali-silica reaction distress’, Proceedings of
the 2nd Australia/Japan workshop on durability of reinforced concrete structures, 28-30 November 1988,
CSIRO, Highett, Victoria, pp. 2.67-2.82.

Carse, A & Dux, PF 1990, ‘Development of an accelerated test on concrete prisms to determine their
potential for alkali-silica reaction’, Cement and Concrete Research, vol. 20, no. 6, pp. 869–74.

Carse, A 1993, ‘The identification of ASR in the concrete cooling tower infrastructure of the Tarong Power
Station’, Construction and Building Materials, vol. 7, no. 2, pp. 117-9.

Carse, A 1996, ‘The asset management of a long bridge structure affected by alkali-silica reaction’,
Proceedings of the 10th international conference on alkali-aggregate reaction in concrete, Melbourne,
Australia, 18-23 August 1996, pp. 1025–32.

Carse, A & Bell, A 2004, ‘Design of Queensland road infrastructure for high risk environments’, ISCO 13th
International Soil Conservation Organisation conference, Brisbane, Australia, paper no. 1018, 7 pp.

CEMBUREAU 1978, Use of concrete in aggressive environments, recommendations, CEMBUREAU Paris,


France.

Cement Concrete & Aggregates Australia 2011, Sulfate-resisting concrete, technical note 68, CC&AA,
Sydney, NSW.

Chamberlin, WP, Irwin, RJ & Amsler, DE 1977, Waterproofing membranes for bridge deck rehabilitation,’
FHWA-NY-77-59-1, Department of Transportation & Federal Highway Administration, Washington, DC,
USA.

Austroads 2016 | page 99


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Chauveau, E & Demelin, B 2007, ‘New lean duplex stainless steel rebar: pitting corrosion resistance and
galvanic coupling behavior, International conference on concrete under severe conditions: environment
and loading, 5th, Tours, France, Laboratoire Central des Ponts et Chaussées, Paris, France, pp. 1859–
66.

Chaussadent, T, Nobel-Pujol, V, Farcas, F, Mabille, I & Fiaud, C 2006, ‘Effectiveness conditions of sodium
monofluorophosphate as a corrosion inhibitor for concrete reinforcement, Cement and Concrete
Research, vol. 36, pp. 556–61.

Chauveau, E & Demelin, B 2007, ‘New lean duplex stainless steel rebar: pitting corrosion resistance and
galvanic coupling behavior’, in Toutlemonde, F, Sakai, K, Gjørv, OE & Banthia, N (eds), Proceedings of
the 5th international conference on concrete under severe conditions: environment and loading,
Laboratoire Central des Ponts et Chaussées, Paris, France, vol. 2, pp. 1859–66.

Chen, HP & Xiao, N 2012, ‘Analytical solutions for corrosion-induced cohesive concrete cracking’, Journal of
Applied Mathematics, vol. 2012, 25 pp.

Çinlar, E, Bažant, ZP & Osman, E 1977, ‘Stochastic process for extrapolating concrete creep’, Journal of the
Engineering Mechanics Division, vol. 103, no. EM6, pp. 1069–88.

Clayton, N 1989, ‘Structural performance of ASR affected concrete’, Proceedings of the 8th ICAAR, Kyoto,
Japan, pp. 671–6.

Clear, KC 1976, Time-to-corrosion of reinforcing steel in concrete slabs: volume 3: performance after 830
daily salt applications, report no. FHWA-RD-76–70, Federal Highway Administration, Washington, DC,
USA.

Cole, WF, Lancucki, CJ & Sandy, MJ 1981, ‘Products formed in an aged concrete’, Cement and Concrete
Research, vol. 11, no. 3, pp. 443–54.

Collepardi, M 1997, ‘A holistic approach to concrete damage induced by delayed ettringite formation’, Mario
Collepardi symposium on advances in concrete science and technology, 5th CANMET/ACI international
conference on superplasticisers and other chemical admixtures in concrete, Rome, Italy, ACI, Farmington
Hills, USA, pp. 373–95.

Collepardi, M, Collepardi, S, Ogoumah Olagot, JJ & Simonelli, F 2004, ‘The influence of slag and fly ash on
carbonation of concrete’, 8th CANMET/ACI international conference on fly ash, silica fume, slag, and
natural pozzolans in concrete, ACI, Farmington Hills, MI, USA, pp. 483–94.

Collins, F, McIntyre, M, Goelst, A, Pitt and Sherry & McGee, R 2001, ‘Durability by design approaches to
specifying durable concrete for the Sorell Causeway Bridge,’ Proceedings of the 20th biennial conference
of the Concrete Institute of Australia, Perth, Western Australia, pp. 501–8.

Comite Euro-International du Beton 1992, Design guide for durable concrete structures, 2nd edn, Thomas
Telford Publishers, UK.

Concrete Institute of Australia 2001, CIA performance criteria for concrete in marine environments, CIA
Z13-2001, CIA, Sydney, NSW.

Concrete Institute of Australia 2014a, Durability planning, CIA Z7/01-2014, CIA, Sydney, NSW.

Concrete Institute of Australia 2014b, Good practice through design, concrete supply and construction, CIA
Z7/04-2014, CIA, Sydney, NSW.

Concrete Institute of Australia 2015, Performance tests to assess concrete durability,


http://infostore.saiglobal.com/store/Details.aspx?ProductID=1811333, CIA Z7/07-2015, CIA, Sydney,
NSW.

Austroads 2016 | page 100


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Concrete Society 1996, Developments in durability design and performance-based specification of concrete,
special publication CS109, The Concrete Society, Camberley, UK.

Connal, J & Berndt, M 2009, ‘Sustainable bridges: 300 year design life for second Gateway Bridge’,
Austroads bridge conference, 7th, 2009, Auckland, New Zealand, Convention Management, Auckland,
NZ, paper no. 0057, 16 pp.

Constantinou, AG & Scrivener, KL 1997, ‘Corrosion of steel in carbonated concrete’, International congress
on the chemistry of cement, 10th, Gothenburg, Sweden, Amarkai AB and Congrex, paper 4iv084, 8 pp.

Cortec Corporation 2011, Migrating corrosion inhibitors from grey to green, Cortec, Saint Paul, MN, USA,
viewed 8 July 2016, <http://www.CortecMCI.com>.

Cox, RN & Oldfield, JW 1996, ‘The long-term performance of austenitic stainless steel in chloride
contaminated concrete’, in Page, CL, Bamforth, PB & Figg, JW (eds), Corrosion of reinforcement in
concrete construction, SCI Special Publication 183, pp. 662–9.

Cramer, SD, Covino, BS, Bullard, SJ, Holcomb, GR, Russell, JH, Nelson, FJ, Laylor, HM & Soltesz, SM
2002, ‘Corrosion prevention and remediation strategies for reinforced concrete coastal bridges’, Cement
and Concrete Composites, vol. 24, no. 1, pp. 101–17.

Cusson, D, Lounis, Z & Daigle, L 2011, ‘Durability monitoring for improved service life predictions of concrete
bridge decks in corrosive environments’, Computer-Aided Civil and Infrastructure Engineering, vol. 26, no.
7, pp. 524–41.

Cutler, CP, Cochrane, DJ & Jenkinson, DC 1998, Stainless steel reinforcing bars: corrosion and prevention
1998 proceedings, Australasian Corrosion Association, Hobart, Tas.

Czerewko, MA, Cross, SA, Dumelow, PG & Saadvandi, A 2003, ‘Assessment of pyritic Lower Lias mudrocks
for earthworks’, Proceedings of the Institution of Civil Engineers, Geotechnical Engineering, vol. 164,
issue GE2, pp. 59–77.

Dai, JG, Akira, Y, Wittmann, FH, Yokota, H & Zhang, P 2010, ‘Water repellent surface impregnation for
extension of service life of reinforced concrete structures in marine environments: the role of cracks’,
Cement and Concrete Composites, vol. 32, no. 2, pp. 101–9.

Darwin, D, Browning, JA, Locke, CE & Nguyen, TV 2007, Multiple corrosion protection systems for reinforced
concrete bridge components, report no. FHWA-HRT-07-043, Federal Highway Administration, McLean,
Virginia, USA.

Davies, MJS, Grace, WR, Green, WK & Collins, FG 1996, ‘Assessment and management of a marine
structure affected by ASR’, Proceedings of the 10th ICAAR, Melbourne, Australia, pp. 1018–24.

De Rincón, OT, Pérez, O, Paredes, E, Caldera, Y, Urdaneta, C & Sandoval, I 2002, ‘Long-term performance
of ZnO as a rebar corrosion inhibitor’, Cement and Concrete Composites, vol. 24, no. 1, pp. 79–87.

Deng, M & Tang, M 1994, ‘Formation and expansion of ettringite crystals’, Cement and Concrete Research,
vol. 24, no. 1, pp. 119–26.

Department of Infrastructure, Energy and Resources 2006, Bridgeworks specification B10: supply of
concrete, DIER, Hobart, Tas.

Department of Planning, Transport and Infrastructure 2011, Specification part 320: supply of concrete, DPTI,
Adelaide, SA.

Department of Transport and Main Roads 2009, Guideline for the preparation of road structure durability
plans, DTMR, Brisbane, QLD.

Austroads 2016 | page 101


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Department of Transport and Main Roads 2010, Main Roads technical standard: concrete, MRTS70, DTMR,
Brisbane, QLD.

Diamond, S 1994, ‘Combined alkali silica reaction: secondary ettringite problems, a current assessment’,
Concrete across borders international conference, 1994, Odense, Denmark, Danish Concrete
Association, Copenhagen, Denmark, pp. 191–204.

Diamond, S 1996, ‘Delayed ettringite formation: process and problems’, Cement and Concrete Composites,
vol. 18, no. 3, pp. 205–15.

Diamond, S & Ong, S 1994, ‘Effects of added alkali hydroxides in mix water on long-term SO42-
concentrations in pore solution’, Cement and Concrete Composites, vol. 16, no. 3, pp. 219–26.

DuraCrete 1999, ‘The European Union – Brite EuRam III, DuraCrete – probabilistic performance based
durability design of concrete structures’, document BE95-1347/R0, Gouda, The Netherlands.

Eglinton, M 1998, ‘Resistance of concrete to destructive agents’, in Hewlett, PC (ed), Lea’s chemistry of
cement and concrete, 4th edn, Arnold, London, UK, pp. 299–343.

Ehlen, MA 2008, Life-365 service life prediction model and computer program for predicting the service life
and life-cycle costs of reinforced concrete exposed to chlorides, National Ready Mixed Concrete
Association, Silver Spring, MD, USA, viewed 8 July 2016,
<www.nrmca.org/research/Life365v2UsersManual.pdf>.

Ellingwood, BR & Mori, Y 1993, ‘Probabilistic methods for condition assessment and life prediction of
concrete structures in nuclear power plants’, Nuclear Engineering and Design, vol. 142, no. 2-3, pp. 155–
66.

Enright, MP & Frangopol, DM 1998, ‘Probabilistic analysis of resistance degradation of reinforced concrete
bridge beams under corrosion’, Engineering Structures, vol. 20, no. 11, pp. 960-71.

Erlin, B (ed) 1999, ‘Ettringite: the sometimes host of destruction’, ACI convention, 1997, Seattle, Washington,
ACI International, Farmington Hills, MI, USA, pp. 265.

Famy, C 1999, ‘Expansion of heat-cured mortars’, PhD thesis, University of London, UK.

Famy, C, Scrivener, KL, Atkinson, A & Brough, AR 2002, ‘Effects of an early or a late heat treatment on the
microstructure and composition of inner C-S-H products of Portland cement mortars’, Cement and
Concrete Research, vol. 32, no. 2, pp. 269–78.

Ferreira, M, Årskog, V, Jalali, S & Gjørv, OE 2004, ‘Software for probability-based durability analysis of
concrete structures’, in Oh, BH, Sakai, K, Gjørv, OE & Banthia N (eds), Proceedings vol 1: fourth
international conference on concrete under severe conditions: environment and loading, Seoul National
University and Korea Concrete Institute, Seoul, South Korea, pp. 1015–24.

Ferreira, RM 2010, ‘Optimization of RC structure performance in marine environment’, Engineering


Structures, vol. 32, no. 5, pp. 1489–94.

FIB 2006, Model code for service life design, bulletin no. 34, International Federation for Structural Concrete,
Lausanne, Switzerland.

Flint, GN & Cox, RN 1988, The resistance of stainless steel partly embedded in concrete to corrosion by
seawater, Magazine of Concrete Research, vol. 40, no. 142, pp. 13–27.

Floyd, M & Wimpenny, DE 2003, ‘Procedures for assessing thaumasite sulfate attack and adjacent ground
conditions at buried concrete structures’, Cement and Concrete Composites, vol. 25, no. 8, pp. 1077–88.

Austroads 2016 | page 102


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Fluge, F 1997, ‘Environmental loads on coastal bridges’, International conference on repair of concrete
structures, 1997, Svolvær, Norway, Norwegian Road Research Laboratory, Oslo, Norway, pp. 89-98.

Force Institute 1999, Corrosion aspects of galvanic coupling between carbon steel and stainless steel in
concrete, Arminox, Viborg, Denmark.

Frangopol, DM, Kallen, MJ & van Noortwijk, JM 2004, ‘Probabilistic models for life-cycle performance of
deteriorating structures: review and future directions’, Progress in Structural Engineering and Materials,
vol. 6, no. 4, pp. 197–212.

Freire, L, Carmezim, MJ, Ferreira, MGS & Montemor, MF 2011, ‘The electrochemical behaviour of stainless
steel AISI 304 in alkaline solutions with different pH in the presence of chlorides’, Electrochimica Acta,
vol. 56, no. 14, pp. 5280–9.

Furman, S 2008, ‘Durability planning for new infrastructure projects’, Corrosion & prevention: Australasian
Corrosion Association conference, 48th, 2008, Wellington, New Zealand, ACA, Kerriemuir, Vic, 12 pp.

García-Alonso, MC, González, JA, Miranda, J, Escudero, ML, Correia, MJ, Salta, M & Bennani, A 2007,
‘Corrosion behaviour of innovative stainless steels in mortar’, Cement and Concrete Research, vol. 37,
no. 11, pp. 1562–9.

Geiker, M, Rostam, S & Vincentsen, LJ 1997, ‘Next generation bridge management’, International
conference on repair of concrete structures, 1997, Svolvær, Norway, Norwegian Road Research
Laboratory, Oslo, Norway, pp. 555–66.

GHD 2002, ‘Sorell Causeway bridge: report on detailed durability design’, internal report, GHD, Melbourne,
Vic.

GHD 2009, ‘VECT JV report for Brighton Bypass 85% durability plan’, internal report, GHD, Melbourne, Vic.

Giergiczny, Z 1997, ‘Sulphate resistance of cements with mineral admixtures’, International congress on the
chemistry of cement, 10th, 1997, Göteborg, Sweden, paper 4iv019.

Gjørv, OE 2009, Durability design of concrete structures in severe environments, Taylor & Francis, London,
UK.

Glasser, FP, Damidot, D & Atkins, M 1995, ‘Phase development in cement in relation to the secondary
ettringite problem’, Advances in Cement Research, vol. 7, no. 26, pp. 57–68.

Goni, S & Andrade, C 1990, ‘Synthetic concrete pore solution chemistry and rebar corrosion rate in the
presence of chlorides’ Cement and Concrete Research, vol. 20, no. 4, pp. 525–39.

Goodbrake, CJ 1978, ‘Reaction of beta di-calcium silicate and tri-calcium silicate with carbon dioxide and
water’, PhD thesis, University of Illinois Urbana-Champaign, USA.

Goodwin, PD, Frantz, GC & Stephens, JE 2000, Protection of reinforcement with corrosion inhibitors, phase
II, report JHR00-279, Connecticut Department of Transportation, Rocky Hill, CT, USA.

Gouda, VK 1970, ‘Corrosion and inhibition of reinforcing steel: I: immersed in alkaline solutions’, British
Corrosion Journal, vol. 5, no. 5, pp. 198–203.

Gouda, VK & Halaka, WY 1970, ‘Corrosion and corrosion inhibition of reinforcing steel: II: embedded in
concrete’, British Corrosion Journal, vol. 5, no. 5, pp. 204–8.

Hamada, M 1968, ‘Neutralization (carbonation) of concrete and corrosion of reinforced steel’, in Kyokai, S
(ed), Proceedings of the 5th international symposium on chemistry of cement, Tokyo, 1968, Japan
Cement Association, Tokyo, Japan.

Austroads 2016 | page 103


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Hansson, CM & Sorensen, B 1988, ‘The threshold concentration of chloride in concrete for initiation of
reinforcement corrosion’, in Berke, N & Whiting, V (eds), Corrosion rates of steel in concrete, ASTM STP
1065, ASTM International, West Conshohocken, PA, USA, pp. 3–16.

Hariche, L, Ballim, Y, Bouhicha, M & Kenai, S 2012, ‘Effects of reinforcement configuration and sustained
load on the behaviour of reinforced concrete beams affected by reinforcing steel corrosion’, Cement and
Concrete Composites, vol. 34, no. 10, pp. 1202–9.

Hausmann, DA 1967, ‘Steel corrosion in concrete: how does it occur?’ Materials Protection, vol. 6, no. 11,
pp. 19–23.

Haynes, H & Bassuoni, MT 2011, ‘Physical salt attack on concrete’, Concrete International, vol. 33, no. 11,
pp. 38–42.

Heinz, D & Ludwig, U 1986, ‘Mechanism of subsequent ettringite formation in mortars and concrete after
heat treatment’, International congress on the chemistry of cement, 8th, Rio de Janeiro, Brazil,
Conference Secretariat, Rio de Janeiro, Brazil, vol. 5, pp. 189–94.

Heinz, D, Ludwig, U & Rüdiger, I 1989, ‘Delayed ettringite formation in heat treated mortars and concretes’,
Concrete Precasting Plant and Technology, vol. 56, no. 11, pp. 56–61.

Hertfort, D, Thomas, MDA, Scrivener, K & Kurdowski, W 2003, ‘A discussion of the paper: “Role of delayed
release of sulphates from clinker in DEF” by Weislaw Kurdowski’, Cement and Concrete Research, vol.
33, no. 3, pp. 455–8.

Hewitt, J & Tullmin, M 1994, ‘Corrosion and stress corrosion cracking performance of stainless steel and
other reinforcing bar materials in concrete’ in Swamy, RN (ed), Corrosion and corrosion protection of steel
in concrete: proceedings of the international conference, Sheffield Academic Press, Sheffield, UK, pp.
527–39.

Highways Agency 1987, 'Early thermal cracking of concrete' in Design manual for roads and bridges, volume
1, section 3, BA 24/87, Highways Agency, Department for Transport, London, UK.

Highways Agency 2003, ‘The impregnation of reinforced and prestressed concrete highway structures using
hydrophobic pore-lining impregnants’ in Design manual for roads and bridges, volume 2, section 4, part 2,
BD 43/03, Highways Agency, Department for Transport, London, UK.

Hime, WG & Marusin, SL 1999, ‘Delayed ettringite formation: many questions and some answers’, in Erlin, B
(ed), Ettringite, the sometimes host of destruction, American Concrete Institute, Farmington Hills, MI,
USA, pp. 199–206.

Ho, DWS & Lewis, VE 1987, ‘Carbonation of concrete and its prediction’, Cement and Concrete Research,
vol. 17, no. 3, pp. 489–504.

Hobbs, DW 1988, ‘Alkali silica reaction in concrete’, Thomas Telford Publishers, London, UK.

Hobbs, DW (ed) 1998, Minimum requirements for durable concrete: carbonation- and chloride-induced
corrosion, freeze-thaw attack and chemical attack, British Cement Association, Crowthorne, UK.

Hobbs, DW 1999, ‘Expansion and cracking in concrete associated with delayed ettringite formation’, in Erlin,
B (ed), Ettringite, the sometimes host of destruction, American Concrete Institute, Farmington Hills, MI,
USA, pp. 159–81.

Hobbs, DW, Marsh, BK, Matthews, JD & Petit, S 1998, ‘Minimum requirements for concrete to resist
carbonation-induced corrosion’, in DW Hobbs (ed), Minimum requirements for durable concrete:
carbonation- and chloride-induced corrosion, freeze-thaw attack and chemical attack, British Cement
Association, Crowthorne, UK.

Austroads 2016 | page 104


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Hope, BB & Ip AKC 1987, ‘Chloride corrosion threshold in concrete,’ ACI Materials Journal, vol. 84, no. 4, pp.
306–14.

Hormann, K, Hofmann, F-J & Schmidt, M 1997, ‘Stability of concrete against biogenic sulfuric acid corrosion
– a new method for deterioration’, International congress on the chemistry of cement, 10th, 1997,
Gothenburg, Sweden, Amarkai, Gothenburg, Sweden, paper 4iv038.

Houska, K & Holsing, T 2012, ‘LDX 2101 and 2205 duplex stainless steel: proper cleaning of mill scale’,
FDOT-FHWA corrosion-resistant rebar (CRRB) seminar, 17 July 2012, Tampa, Florida, USA, Florida
Department of Transportation, viewed 8 July 2016,
<http://www.dot.state.fl.us/statematerialsoffice/structural/meetings/crrb/2_millscale.pdf>.

Hurley, MF & Scully, JR 2002, Chloride threshold levels in clad 316L and solid 316LN stainless rebar, Center
for Electrochemical Science and Engineering, University of Virginia, Charlottesville, VA, USA, viewed 8
JULY 2016, <http://www.virginia.edu/cese/research/paper02224.pdf>.

Hurley, MF & Scully, JR 2006, ‘Threshold chloride concentrations of selected corrosion resistant rebar
materials compared to carbon steel’, Corrosion, vol. 62, no. 10, pp. 892-904.

Hussain, SE, Rasheduzzafar, SE, Al-Mussallam, A & Al-Gahtani, AS 1995, ‘Factors affecting threshold
chloride for reinforcement corrosion in concrete’, Cement and Concrete Research, vol. 25, no. 7, pp.
1543–55.

Hwang, CL, Chen, JC & Chiou, IJ 1994, ‘Concrete cracking and corrosion of steel bar’ in Swamy, RN (ed),
Corrosion and corrosion protection of steel in concrete, Sheffield Academic Press, Sheffield, UK, pp.
310–23.

Idorn, GM 1996, ‘Systematic ASR expertise – Australian research 1940s to 1958, in Shayan, A (ed), Alkali-
aggregate reaction in concrete: proceedings of the 10th international conference on AAR, Melbourne,
Australia, pp. 15–26.

Jarrah, NR, Al-Amoudi, OSB, Maslehuddin, M, Ashiru, OA & Al-Mana, AI 1995, ‘Electrochemical behaviour
of steel in plain and blended cement concretes in sulphate and/or chloride environments’, Construction
and Building Materials, vol. 9, no. 2, pp. 97–103.

Johansen, V & Thaulow, N 1999, ‘Heat curing and late formation of ettringite’ in Erlin, B (ed), Ettringite, the
sometimes host of destruction, American Concrete Institute, Farmington Hills, MI, USA, pp. 47–64.

Johansen, V, Thaulow, N & Idorn, GM 1994, ‘Expansion reactions in mortar and concrete’ (in English),
Zement-Kalk-Gips, vol. 5, pp. 150–4.

Johnstone, JR & Glasser, PF 1992, ‘Carbonation of portlandite single crystals and portlandite in cement
paste’, International congress on the chemistry of cement, 9th, 1992, New Delhi, India, National Council
for Cement and Building Materials, New Delhi, vol. 5, pp. 370–6.

Jones, MR, McCarthy, MJ & Dhir, PK 1994, ‘Chloride ingress and reinforcement corrosion in carbonated and
sulphated concrete’ in Swamy, RN (ed), Corrosion and corrosion protection of steel in concrete, Sheffield
Academic Press, Sheffield, UK, pp. 365–76.

Justnes, H 2004, ‘Preventing chloride-induced rebar corrosion by anodic inhibitors: comparing calcium nitrate
with calcium nitrite’, Conference on our world in concrete and structures, 29th, 2004, Singapore, CI-
Premier, Singapore, 16 pp.

Justnes, H 2006a, Calcium nitrate as a multifunctional concrete admixture, Norwegian Society of Graduate
Technical and Scientific Professionals, Norway, viewed 8 July 2016,
<https://www.scribd.com/doc/311682442/Calcium-Nitrate-as-a-Multifunctional-Concrete-Admixture>.

Austroads 2016 | page 105


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Justnes, H 2006b, Corrosion inhibitors for reinforced concrete, ACI special publication SP 234-04,
American Concrete Institute, Farmington Hills, MI, USA, pp. 53–70.

Katayama, T & Sommer, H 2008, ‘Further investigation of the mechanisms of so-called alkali-carbonate
reaction based on modern petrographic techniques’, International conference on alkali-aggregate reaction
in concrete, 13th, 2008, Trondheim, Norway, Quik Scan, The Netherlands, pp. 850–60.

Kayyali, OA & Haque, MN 1995, ‘The ratio of Cl-/OH- in chloride contaminated concrete: a most important
criterion’, Magazine of Concrete Research, vol. 47, pp. 235–42.

Kelham, S 1996, ‘The effect of cement composition and fineness on expansion associated with delayed
ettringite’, Cement and Concrete Composites, vol. 18, no. 3, pp. 171–9.

Kelham, S 1997, ‘Effects of cement composition and hydration temperature on volume stability of mortar’,
International congress on the chemistry of cement, 10th, 1997, Gothenburg, Sweden, Amarkai,
Gothenburg, Sweden, vol. 4.

Kelham, S 1999, ‘Influence of cement composition on volume stability of mortar’, in Erlin, B (ed), Ettringite,
the sometimes host of destruction, American Concrete Institute, Farmington Hills, MI, USA, pp. 27–45.

Kempsey Bypass Alliance 2010, ‘Upgrading the Pacific Highway: durability plan’, KBA-REP-2G00-GE068A-
FD-03, prepared for Roads and Maritime Services, Sydney, NSW.

Khatri, RP, Sirivivatnanon, V & Yang, JL 1997, ‘Role of permeability in sulphate attack’, Cement and
Concrete Research, vol. 27, no. 8, pp. 1179–89.

Khatri, R, Lee, L, Canceri, J & Chirgwin, G 2009, ‘Designing for a 100 year life in a high chloride and acid-
sulfate environment’, Concrete Institute of Australia conference, 24th, 2009, Sydney, New South Wales,
Australia, paper no. 5c-2, Concrete Institute of Australia, Rhodes, NSW, 10 pp.

Khayat, KH, Tagnit-Hamou, A & Petrov, N 2005, ‘Performance of concrete wharves constructed between
1901 and 1928 at the Port of Montréal’, Cement and Concrete Research, vol. 35, no. 2, pp. 226–32.

Klemm, W & Miller, FM 1999, ‘Internal sulfate attack: distress mechanism at ambient and elevated
temperatures?’ in Erlin, B (ed), Ettringite, the sometimes host of destruction, American Concrete Institute,
Farmington Hills, MI, USA, pp. 81–8.

Knudsen, A & Skovsgaard, T 1999, ‘Ahead of its peers’, Concrete Engineering International,
August/September, pp. 58–61.

Knudsen, A, Jensen, FM, Klinghoffer, O & Skovsgaard, T 1998, ‘Cost-effective enhancement of durability of
concrete structures by intelligent use of stainless steel reinforcement,’ International conference on
corrosion and rehabilitation of reinforced concrete structures, 1998, Orlando, Florida, USA, Turner-
Fairbank Highway Research Center, McLean, VA, USA, 15 pp.

Kompen, R 1997, ‘What can be done to improve the quality of new concrete structures’, International
conference: repair of concrete structures from theory to practice in a marine environment, Svolvær,
Norway, Norwegian Road Research Laboratory, Oslo, pp. 529–36.

Kondo, R, Daimon, M & Akiba, T 1968, ‘Mechanisms and kinetics on carbonation of hardened cement’,
International symposium on chemistry of cement, 5th, 1968, Tokyo, Japan, Cement Association of Japan,
Tokyo, vol. III, pp. 402–9.

Kupwade-Patil, K, Cardenas, HE, Gordon, K & Lee, LS 2012, ‘Corrosion mitigation in reinforced concrete
beams via nanoparticle treatment’, ACI Materials Journal, vol. 109, no. 6, pp. 617–26.

Kurdowski, W 2002, ‘Role of delayed release of sulphates from clinker in DEF’, Cement and Concrete
Research, vol. 32, no. 3, pp. 401-7.

Austroads 2016 | page 106


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Kuroda, T, Nishibayashi, S & Bian, Q 1996, ‘Study of alkali-aggregate reactions in electrical fields’,
Proceedings of the 10th international conference on alkali-aggregate reaction, 18-23 August 1996,
Melbourne, Australia, pp. 645-52.

Lambert, P, Page, CL & Vassie, PRW 1991, ‘Investigation of reinforcement corrosion. electrochemical
monitoring of steel in chloride contaminated concrete’, Materials and Structures, vol. 24, no. 5, pp. 351–8.

Larsen, CK 1997, ‘Effect of temperature, carbonation and drying and wetting on chloride uptake in concrete’,
International conference: repair of concrete structures from theory to practice in a marine environment,
Svolvær, Norway, Norwegian Road Research Laboratory, Oslo, pp.153–62.

Law, DW, Cairns, JJ, Millard, SG & Bungey, JH 2003, ‘Evaluation of corrosion loss of steel reinforcing bars in
concrete using linear polarisation resistance measurements,’ International symposium on non-destructive
testing in civil engineering, 2003, Berlin, Germany, viewed 8 July 2016,
<www.ndt.net/article/ndtce03/papers/p015/p015.htm>.

Lawler, JS & Kruass, PD 2011, ‘35-year field performance of epoxy-coated reinforcing bars’, Concrete
International, vol. 33, no. 11, pp. 29–37.

Lawrence, BL, Myers, JJ & Carrasquille, RL 1999, ‘Premature concrete deterioration in Texas Department of
Transportation precast elements’, in Erlin, B (ed), Ettringite, the sometimes host of destruction, American
Concrete Institute, Farmington Hills, MI, USA, pp. 141–58.

Lawrence, CD 1999, ‘Long-term expansion of mortars and concretes’, in Erlin, B (ed), Ettringite, the
sometimes host of destruction, American Concrete Institute, Farmington Hills, MI, USA, pp. 105–23.

Leamon, RJ & Shayan, A 1996, ‘Alkali-aggregate reaction in a concrete water storage tank’, Proceedings of
the 10th international AAR conference, Melbourne, Australia, 18-23 August 1996, pp. 235–42.

Lee, SK, Hart, WH & McIntyre, JF 1996, ‘Accelerated testing of epoxy coated reinforcing steel- part II:
ambient temperature aqueous exposure and electrochemical impedance spectroscopy’, in Page, CL,
Bamforth, PB & Figg, JW (eds), Corrosion of reinforcement in concrete construction, SCI special
publication 183, pp. 642–53.

Li, CQ, Lawanwisut, W, Zheng, JJ & Kijawatworawet, W 2005, ‘Crack width due to corroded bar in reinforced
concrete structures’, International Journal of Materials and Structural Reliability, vol. 3, no. 2, pp. 87–94.

Li, L & Sagues, AA 2001, ‘Chloride corrosion threshold of reinforcing steel in alkaline solutions: open-circuit
immersion tests’, Corrosion, vol. 57, no. 1, pp. 19–28.

Lim, TYD, Teng, S, Bahador, SD & Gjørv, O 2016, ‘Durability of very-high-strength concrete with
supplementary cementitious materials for marine environments, ACI Materials, vol. 113, January-
February, pp. 95-103.

Lindvall, A 2007, ‘Chloride ingress data from field and laboratory exposure: influence of salinity and
temperature’, Cement and Concrete Composites, vol. 29, no. 2, pp. 88–93.

Linke, WF & Seidell, A 1965, Solubilities: inorganic and metal-organic compounds: a compilation of solubility
data from the periodical literature, 4th edn, Van Nostrand, Washington, DC, USA.

Liu, PL & Der Kiureghian, A 1991, ‘Optimization algorithms for structural reliability’, Structural Safety, vol. 9,
no. 3, pp. 161–77.

Liu, Y 1996, ‘Modelling the time-to-corrosion cracking on the cover concrete in chloride contaminated
reinforced concrete structures’, PhD thesis, Virginia Polytechnic Institute and State University,
Blacksburg, Virginia, USA.

Austroads 2016 | page 107


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Liu, Y & Weyers, RE 1998, ‘Modelling the time-to-cracking in chloride-induced reinforced concrete
structures’, ACI Materials, vol. 95, no. 6, pp. 675–81.

Lothenbach, B, Bary, B, Le Bescop, P, Schmidt, T & Leterrier, N 2010, ‘Sulfate ingress in Portland cement’,
Cement and Concrete Research, vol. 40, no. 8, pp. 1211–25.

Maage, M, Helland, S & Carlsen, JE 1997, ‘Service life prediction of concrete in marine environment’,
International conference: repair of concrete structures from theory to practice in a marine environment,
Svolvær, Norway, Norwegian Road Research Laboratory, Oslo, pp. 177–87.

Madrid, J, Diez, JM, Goñi, S & Macias, A 1997 ‘Durability of ordinary Portland cement and ground granulated
blast furnace slag cement in acid medium’, International congress on the chemistry of cement, 10th,
1997, Gothenburg, Sweden, Amarkai, Gothenburg, Sweden, paper 4iv040,
9 pp.

Main Roads Western Australia 2013, Specification 820: concrete for structures, MRWA, Perth, WA.

Manning, DG 1996, ‘Corrosion performance of epoxy-coated reinforcing steel: North American experience,’
Construction and Building Materials, vol. 10, no. 5, pp. 349–65.

Mansfeld, F 1981, ‘Recording and analysis of AC Impedance data for corrosion studies,’ Corrosion, vol. 37,
no. 5, pp. 301–7.

Markeset, G, Rostam, S & Klinghoffer, O 2006, Guide for the use of stainless steel reinforcement in concrete
structures, project report 405, Norwegian Building Research Institute (BYGGORSK, SINTEF), Oslo,
Norway.

Marques, PF & Costa, A 2010, ‘Service life of RC structures: carbonation induced corrosion: prescriptive vs.
performance-based methodologies’, Construction and Building Materials, vol. 24, no. 3, pp. 258–65.

Marusin, SL 1995, ‘Deterioration of railway ties in the USA’, CANMET/ACI international workshop on alkali-
aggregate reaction in concrete, 1995, Dartmouth, Nova Scotia, Canada, Canada Centre for Mineral and
Energy Technology & American Concrete Institute, Farmington Hills, MI, USA, pp. 243–55.

Matthews, JD 1992, ‘The resistance of PFA concrete to acid ground waters,’ International congress on the
chemistry of cement, 9th, 1992, New Delhi, India, National Council for Cement and Building Materials,
New Delhi, India, vol. 5, pp. 355–62.

Maunsell & AECOM 2007, ‘Durability report on Gateway Bridge upgrade project’, contract report no. GUP-
RP-G-BR-050004 rev 3, Maunsell & AECOM, Australia.

McDonald, DB & Pfeifer, DW 1995, ‘Epoxy-coated bars: state-of-the-art,’ Regional conference on concrete
durability in the Arabian Gulf, 2nd, 1995, Saudi Arabia, Bahrain Society of Engineers, vol. 2.

McDonald, DB, Sherman, MR, Pfeifer, DW & Virmani, YP 1995, ‘Stainless steel reinforcing as corrosion
protection’, Concrete International, May, pp. 65–70.

McCoy, WJ & Caldwell, AG 1951, ‘New approach to inhibiting alkali-aggregate expansion’, Journal of the
American Concrete Institute, vol. 22, no. 9, pp. 693–706.

McGee, RW 2001, ‘On the service life modelling of Tasmanian concrete bridges, PhD thesis, Department of
Civil Engineering, University of Tasmania.

Mehta, PI 2000, ‘Sulfate attack on concrete: separating myths from reality’, Concrete International, August,
pp. 57–61.

Mehta, PK & Monteiro, PJM 2006, Concrete: microstructure, properties and materials, 3rd edn, McGraw-Hill,
New York, NY, USA, pp. 135.

Austroads 2016 | page 108


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Meland, I, Justnes, H & Lindgård, J 1997, ‘Durability problems related to delayed ettringite formation and / or
alkali aggregate reactions’, International congress on the chemistry of cement, 10th, 1997, Gothenburg,
Sweden, Amarkai, Gothenburg, Sweden, paper 4iv064.

Melchers, RE 1999, Structural reliability analysis and prediction, 2nd edn, Wiley, Chichester, UK.

Melchers, RE 2006, ‘Recent progress in the modeling of corrosion of structural steel immersed in seawaters’,
Journal of Infrastructure Systems, vol. 12, no. 3, pp. 154–62.

Miah, S & Hinton, B 2007, ‘A corrosion related reliability centred maintenance pilot program for the royal
Australian Air Force C-130 J-30 Hercules aircraft’, TRI-service corrosion conference, Denver, CO, USA,
Department of Defence, Washington DC, 14 pp.

Michaud, V & Suderman, R 1999, ‘Solubility of sulfates in high SO3 clinkers’, Erlin, B (ed), Ettringite, the
sometimes host of destruction, American Concrete Institute, Farmington Hills, MI, USA, pp. 15–25.

Michaud, V, Nonat, A & Sorrentino, D 1997, ‘Experimental simulation of the stability of ettringite in alkali silica
solutions, produced by alkali-silica reaction in concrete’, International congress on the chemistry of
cement, 10th, 1997, Gothenburg, Sweden, Amarkai, Gothenburg, Sweden, vol. 4, paper no. 65.

Mohamed, I, Ronel, S & Curtil, L 2006, ‘Influence of composite materials confinement on alkali-aggregate
mechanical behavior’, Materials and Structures, vol. 39, no. 4, pp. 479–90.

Mohammed, TU, Hamada, H & Yamaji, T 2004, ‘Concrete after 30 years of exposure: part II: chloride ingress
and corrosion of steel bars’, ACI Materials Journal, vol. 101, no. 1, pp. 13–8.

Möller, JS 1994, Measurement of carbonation in cement based materials, report no. P-93:11, Department of
Building Materials, Chalmers University of Technology, Gothenburg, Sweden.

Monteiro, I, Branco, FA, de Brito, J & Neves, R 2012, ‘Statistical analysis of the carbonation coefficient in
open air concrete structures’, Construction and Building Materials, vol. 29, April, pp. 263–9.

Monteiro, PJM, Gjørv, OE & Mehta, PK 1985, ‘Microstructure of the steel-cement paste interface in the
presence of chloride’, Cement and Concrete Research, vol. 15, no. 5, pp. 781–84.

Monticelli, C, Criado, M, Fajardoc, S, Bastidas, JM, Abbottoni, M & Balbo, A 2014, Corrosion behaviour of a
low Ni austenitic stainless steel in carbonated chloride-polluted alkali-activated fly ash mortar, Cement
and Concrete Research, vol. 55, no. 1, pp. 49–58.

Moradllo, MK, Shekarchi, M & Hoseini, M 2012, ‘Time-dependent performance of concrete surface coatings
in tidal zone of marine environment’, Construction and Building Materials, vol. 30, May, pp. 198–205.

Moreno, EI, Cob, E & Castro-Borges, P 2004, ‘Corrosion rates from carbonated concrete specimens’,
Corrosion NACExpo 2004: annual conference and exposition, 59th, New Orleans, USA, National
Association of Corrosion Engineers, Houston, TX, USA, paper no. 04439.

Moser, R 2007, Mass concrete: CEE8813A material science of concrete, PowerPoint presentation, viewed 8
July 2016, <people.ce.gatech.edu/~kk92/massconcrete.pdf>.

Munger, CG 1984, Corrosion prevention by protective coatings, National Association of Corrosion Engineers,
Houston, TX, USA.

Neves, R, Branco, FA & de Brito, J 2012, ‘A method for the use of accelerated carbonation tests in durability
design’, Construction and Building Materials, vol. 36, November, pp. 585–91.

Nie, J & Ellingwood, BR 2000, ‘Directional methods for structural reliability analysis’, Structural Safety, vol.
22, no. 3, pp. 233–49.

Austroads 2016 | page 109


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Nilsson, L-O 1993, ‘The effect of non-linear chloride binding on chloride diffusivities and chloride penetration:
a theoretical approach’, presentation, Nordic mini-seminar on chloride ingression in concrete structures,
1993, Gothenburg, Sweden.

Nilsson, L-O, Poulsen, E, Sandberg, P, Sørensen, HE & Klinghoffer, O 1996, HETEK: chloride penetration
into concrete: state of the art: transport process, corrosion initiation, test methods and prediction models,
report no. 53, Danish Road Directorate, Denmark.

Nomura, N, Kakio, T, Matsuda, Y & Nishibayashi, S 2004, ‘Investigation and repair process of fractured
reinforcements due to ASR, International conference on alkali-aggregate reaction in concrete, 12th, 2004,
Beijing, China, International Academic Publishers, Beijing, China,
pp. 1271–6.

Nordtest 1995, Concrete, hardened: accelerated chloride penetration, Nordtest method NT Build 443,
Nordtest, Norway.

Nordtest 1999, Concrete, mortar and cement-based repair materials: chloride migration coefficient from non-
steady-state migration experiments, Nordtest method NT Build 492, Norway.

Nurnberger, U, Beul, W & Onuseit, G 1993, ‘Corrosion behaviour of welded stainless reinforcing steel in
concrete’, Otto-Graf -Journal, vol. 4, pp. 225–59.

Nurnberger, U 1996, ‘Corrosion behaviour of welded stainless reinforced steel in concrete’, in Page, CL,
Bamforth, PB & Figg, JW (eds), Corrosion of reinforcement in concrete construction, SCI special
publication 183, pp. 623–9.

O’Connell, M, McNally, C & Richardson, MG 2012, ‘Performance of concrete incorporating GGBS in


aggressive wastewater environments’, Construction and Building Materials, vol. 27, no. 1,
pp. 368–74.

Oberholster, RE & Davies, G 1986, ‘An accelerated method for testing the potential alkali reactivity of
siliceous aggregates’, Cement and Concrete Research, vol. 16, pp. 181–9.

Oberholster, RE, Maree, H & Brand, JHB 1992, ‘Cracked prestressed concrete railway sleepers: alkali-silica
reaction or delayed ettringite formation’, International conference on alkali-aggregate reaction in concrete,
9th, London, UK, The Concrete Society, London, UK,
pp. 739–49.

Ormellesse, M, Berra, M, Bolzoni, F & Pastore, T 2006, ‘Corrosion inhibitors for chloride induced corrosion in
reinforced concrete structures’, Cement and Concrete Research, vol. 36, no. 3,
pp. 536–47.

Østnor, TA & Justnes, H 2011, ‘Anodic corrosion inhibitors against chloride induced corrosion of concrete
rebars’, Advances in Applied Ceramics, vol. 110, no. 3, pp. 131–6.

Page, CL, Short, NR & El Tarras, A 1981, ‘Diffusion of chloride ions in hardened cement pastes’, Cement
and Concrete Research, vol. 11, no. 3, pp. 395–406.

Page, CL, Sergi, G & Thompson, DM 1992, ‘Development of alkali-silica reaction in reinforced concrete
subjected to cathodic protection’, Proceedings of the 9th international conference on alkali-aggregate
reaction in concrete, London, UK, vol. 2, pp. 774-81.

Page, C & Yu, SW 1995, ‘Potential effects of electrochemical desalination of concrete on alkali-silica
reaction’, Magazine of Concrete Research, vol. 47, no. 170, pp. 23-31.

Pan, T & Lu, Y 2012, ‘Stochastic modelling of reinforced concrete cracking due to nonuniform corrosion:
FEM-based cross-scale analysis’, Journal of Materials in Civil Engineering, vol. 24, no. 6, pp. 698–706.

Austroads 2016 | page 110


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Papadakis, VG, Vayenas, CG & Fadis, MN 1991, ‘Physical and chemical characteristics affecting the
durability of concrete’, ACI Materials Journal, vol. 88, no. 2, pp. 186–96.

Parrot, L & Hong, CZ 1991, ‘Some factors influencing air permeation measurement in cover concrete’,
Materials and Structures, vol. 24, no. 144, pp. 403–8.

Pérez-Quiroz, JT, Terán, J, Herrera, MJ, Martínez, M & Genescá, J 2008, ‘Assessment of stainless steel
reinforcement for concrete structures rehabilitation’, Journal of Constructional Steel Research, vol. 64, no.
11, pp. 1317–24.

Petrov, N & Tagnit-Hamou, A 2003, ‘Is microcracking really a precursor to DEF and consequent expansion?’
CANMET/ACI international conference on recent advances in concrete technology, 6th, June 2003,
Bucharest, Romania, Canada Centre for Mineral and Energy Technology & American Concrete Institute,
Farmington Hills, MI, USA, pp. 405-20.

Petterson, K 1992, Corrosion threshold value and corrosion rate in reinforced concrete, CBI report 2:92,
Swedish Cement and Concrete Research Institute, Stockholm, Sweden.

Petterson, K 1994, ‘Chloride threshold value and the corrosion rate in reinforced concrete’, in RN Swamy
(ed), Corrosion and corrosion protection of steel in concrete, Sheffield Academic Press, Sheffield, UK, pp.
461–71.

Petterson, K 1996, ‘Factors influencing chloride induced corrosion of reinforced concrete’, in Sjostrom, C
(ed), Durability of building materials and components: 7: prediction, degradation and materials,
(International conference on durability of building materials and components, 7th, 1996, Stockholm,
Sweden), vol. 1, Chapman & Hall, London, UK, pp. 334–41.

Pfeifer, DW, Landgren, JL & Zoob, A 1987, Protective systems for new prestressed and substructure
concrete, report no. FHWA/RD-86/193, Federal Highway Administration, Washington, DC, USA.

Plowman, C & Cabrera, JG 1996, ‘The use of fly ash to improve the sulphate resistance of concrete, Waste
Management, vol. 16, no. 1–3, pp. 145–9.

Polder, RB & de Rooij, MR 2005, ‘Durability of marine concrete structures: field investigations and
modelling’, HERON, vol. 50, no. 3, pp. 133–53.

Poulsen, E 1997, ‘Four-parametric description of marine exposure and concrete’s response to its chloride
intensity’, International conference: repair of concrete structures from theory to practice in a marine
environment, Svolvær, Norway, Norwegian Road Research Laboratory, Oslo,
pp. 189–99.

Pritchard, R 2011, ‘Conformity assessment for critical construction products for bridge infrastructure
projects’, Austroads bridge conference, 8th, 2011, Sydney, New South Wales, Australia, AP-G90/11,
Austroads, Sydney, NSW, 13 pp.

Prusinski, JR & Carrasquillo, RL 1995, ‘Using medium- to high-volume fly ash blended cements to improve
the sulphate resistance of high-lime fly ash concrete’, CANMET/ACI International conference on fly ash,
silica fume, slag, and natural pozzolans in concrete, 5th, Milwaukee, Wisconsin, publication no. SP-153,
American Concrete Institute, Detroit, MI, USA, pp. 43–65.

Rackwitz, R & Fiessler, B 1978, ‘Structural reliability under combined random load sequences’. Computers
and Structures, vol. 9, no. 5, pp. 489–94.

Ramachandran, VS 1984, Concrete admixtures handbook, Noyes Publications, Park Ridge, NJ, USA.

Randström, S, Almén, M, Petterson, R & Adair, M 2010, ‘Reproducibility of critical chloride threshold levels
for stainless steel reinforcement’, International conference on structural faults and repair, 13th, 2010,
Edinburgh, Scotland, Engineering Technics, Edinburgh, Scotland.

Austroads 2016 | page 111


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Rasheeduzzafar, FHD, Bader, MA & Khan, MM 1992, ‘Performance of corrosion resisting steels in chloride-
bearing concrete’, ACI Materials Journal, vol. 89, no. 5, pp. 439–48.

Riding, KA 2007, ‘Early age concrete thermal stress measurement and modelling’, PhD thesis, University of
Texas at Austin, USA.

Riding, KA, Poole, JL, Schindler, AK, Juenger, MCG & Folliard, KJ 2008, ‘Quantification of effects of fly ash
type on concrete early-age cracking’, ACI Materials Journal, vol. 105, no. 2, pp. 149–55.

Roads and Maritime Services 2010, ‘Princes Highway Gerringong upgrade Mount Pleasant to Toolijooa
Road: appendix 13: structural performance and design requirements’, RMS, Sydney, NSW.

Roads and Maritime Services 2012a, ‘Gerringong upgrade durability report, Mount Pleasant to Toolijooa
Road: appendix 13: structural performance and design requirements’, RMS, Sydney, NSW.

Roads and Maritime Services 2012b, Guide to QA specification B80: concrete works for bridges, 2nd edn,
RMS, Sydney, NSW.

Roads and Traffic Authority 2001, Interim test for verification of curing regime: sorptivity, RTA test method
T362, RTA, Sydney, NSW.

Roads and Traffic Authority 2005, Guidelines for the management of acid sulfate materials: acid sulfate soils,
acid sulfate rock and monosulfidic black ooze, version 1, RTA/Pub. 05.032, RTA, Sydney, NSW.

Robinson, D 1998, A survey of probabilistic methods used in reliability, risk and uncertainty analysis:
analytical techniques 1, report SAND98-1189, Sandia National Laboratories, Albuquerque, NM, USA.

Ross, I & Shayan, A 1996, ‘Alkali-aggregate reaction in Western Australia: investigations on the Causeway
bridge and some aggregate sources’, Proceedings of the 10th ICAAR, Melbourne, Australia, 18-23
August 1996, pp. 257–64.

Rostam, S 2005, ‘Design and construction of segmental concrete bridges for service life of 100 to 150 years,
American Segmental Bridge Institute convention, 2005, Washington DC, American Segmental Bridge
Institute, Buda, TX, USA, 27 pp.

Saassouh, B & Lounis, Z 2012, ‘Probabilistic modeling of chloride-induced corrosion in concrete structures
using first- and second-order reliability methods’, Cement and Concrete Composites, vol. 34, no. 9, pp.
1082–93.

Saetta, AV, Scotta, RV & Vitaliani, RV 1993, ‘Analysis of chloride diffusion into partially saturated concrete’,
ACI Materials Journal, vol. 90, no. 5, pp. 441–51.

Safehian, M & Ramezanianpour, AA 2013, ‘Assessment of service life models for determination of chloride
penetration into silica fume concrete in the severe marine environmental condition’, Construction and
Building Materials, vol. 48, pp. 287–94.

Sagüés, AA & Powers, RG 1997, ‘Corrosion and corrosion control of concrete structures in Florida: what can
be learned?’, International conference: repair of concrete structures from theory to practice in a marine
environment, Svolvær, Norway, Norwegian Road Research Laboratory, Oslo, pp. 49–58.

Sagüés, AA & Zayed, AM 1989, ‘Corrosion of epoxy-coated reinforcing steel in concrete: phase I and II’,
South Florida University for the Florida Department of Transportation & Federal Highway Administration,
USA.

Sahu, S, Clark, BA & Lee, RJ 1998, ‘Delayed ettringite formation and the mode of concrete failure’ in Cohen,
M, Mindess, S & Skalny, JP (eds), Materials science of concrete: the Sidney Diamond symposium:
special volume, American Ceramic Society, Westerville, OH, USA, pp. 379-94.

Austroads 2016 | page 112


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Saraswathy, V & Song, H-W 2007, ‘Improving the durability of concrete by using inhibitors,’ Building and
Environment, vol. 42, no. 1, pp. 464–72.

Sarkar, SL & Xu, A 1993, ‘Hydration and properties of fly ash concrete’, in Ghosh, SN, Sarkar, SL & Harsh, S
(eds), Mineral admixtures in cement and concrete: progress in cement and concrete, vol. 4, ABI Books,
New Delhi, India, pp. 174–225.

Sasatani, T, Torii, K & Kawamura, M 1995, ‘Five-year exposure test on long-term properties of concretes
containing fly ash, blast-furnace slag, and silica fume’, CANMET/ACI International conference on fly ash,
silica fume, slag, and natural pozzolans in concrete, 5th, Milwaukee, WI, USA, publication no. SP-153,
American Concrete Institute, Detroit, MI, USA, pp. 283–96.

Schiessl, P & Breit, W 1996, ‘Local repair measures at concrete structures damaged by reinforcement
corrosion’, in Corrosion of reinforcement in concrete construction: international symposium on corrosion
of reinforcement in concrete construction, 4th, 1996, Cambridge, UK, SCI, Cambridge, UK, pp. 525–34.

Schueremans, L, Van Gemert, D & Giessler, S 2007, ‘Chloride penetration in RC-structures in marine
environment: long term assessment of a preventive hydrophobic treatment, Construction and Building
Materials, vol. 21, no. 6, pp. 1238-49.

Scott, JF & Duggan, CR 1987, ‘Potential new test for alkali-aggregate reactivity’ in Grattan-Bellew, PE (ed),
Concrete alkali-aggregate reaction, Noyes Publications, Park Ridge, NJ, USA, pp. 319–23.

Scranton Gillette Communications 2002, ‘Willing to bend: defying tradition, ODOT goes with alternative
stainless steel rebar in a seismic zone’, Roads and Bridges, vol. 40, no. 5, pp. 34–7.

Scrivener, K & Lewis, M 1997, ‘A microstructural and microanalytical study of heat cured mortars and
delayed ettringite formation’, International congress on the chemistry of cement, 10th, 1997, Gothenburg,
Sweden, Amarkai, Gothenburg, Sweden, paper 4iv061, 8 pp.

Scrivener, K & Skalny, J (eds) 2002, ‘Internal sulfate attack and delayed ettringite formation’, International
RILEM workshop, September 2002, Villars, Switzerland, RILEM proceedings PRO 35, RILEM, Bagneux,
France.

Scrivener, KL & Taylor, HFW 1993, ‘Delayed ettringite formation: a microstructural and microanalytical
study’, Advances in Cement Research, vol. 5, no. 20, pp. 139–46.

Selander, A 2010, ‘Hydrophobic impregnation of concrete structures: effects on concrete properties’, PhD
thesis, Department of Civil and Architectural Engineering, Royal Institute of Technology, Stockholm.

Sergi, G & Page, CL 1992, ‘The effects of cathodic protection on alkali-silica reaction in reinforced concrete’,
TRRL contract report no. 310, Crowthorne, UK, 53 pp.

Seto, K, Nishizono, T, Mikata, Y, Maeda, S & Miyagawa, T 2004, ‘Maintenance for ASR damaged road
viaduct’, International conference on alkali-aggregate reaction in concrete, 12th, 2004, Beijing, China,
International Academic Publishers, Beijing, China, pp. 1277–82.

Shayan, A & Ivanusec, I 1996, An experimental clarification of the association of delayed ettringite formation
and alkali-aggregate reaction, Cement and Concrete Composites, vol. 18, no. 3,
pp. 161–70.

Shayan, A & Lancucki, CJ 1987, ‘Alkali-aggregate reaction in the Causeway Bridge, Perth, Western
Australia’, Proceedings of the 7th ICAAR: concrete alkali aggregate reactions, Noyes Publications, NJ,
USA, pp. 392–7.

Shayan, A & Morris, H 2005, ‘Combined deterioration problems in a coastal bridge in NSW, Australia’, Asian
Journal of Civil Engineering (Building & Housing), vol. 6, no. 6, pp. 477–93.

Austroads 2016 | page 113


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Shayan, A & Morris, H 2006, ‘Deterioration of precast, prestressed concrete piles in marine environment: a
case study’, Concrete Plant & Precast Technology, vol. 72, no. 1, pp. 38–47.

Shayan, A & Quick, GW 1992a, ‘Microscopic features of cracked and uncracked concrete railway sleepers’,
ACI Materials Journal, vol. 89, no. 4, pp. 348–61.

Shayan, A & Quick, GW 1992b, ‘Relative importance of deleterious reactions in concrete: formation of AAR
products and secondary ettringite’, Advances in Cement Research, vol. 4, no. 16,
pp. 149–52.

Shayan, A & Quick, GW 1994, ‘Alkali-aggregate reaction in concrete railway sleepers from Finland’, in
International conference on cement microscopy, 16th, Richmond, Virginia, USA, International Cement
Microscopy Association, Duncanville, TX, USA, pp. 69–79.

Shayan, A & Xu, A 2001, ‘Some electrochemical effects of CP on concrete cracked due to AAR and
contaminated by chloride ions’, Proceedings of corrosion & prevention 2001, Australasian Corrosion
Association, Newcastle, NSW, paper 65, 10 pp.

Shayan, A & Xu, A 2002, ‘An overview of literature on the chloride binding capacity of concrete and influence
on the chloride diffusion’, contract report RC2107, ARRB Transport Research, Vermont South, Vic.

Shayan, A & Xu, A 2003, ‘Relationship between rust growth, cover thickness and concrete cracking’,
Australasian Corrosion Association conference, 43rd, 2003, ACA, Melbourne, Vic, vol. 2, paper no. 96, 13
pp.

Shayan, A & Xu, A 2005, ‘Stainless steel reinforcement to increase durability in concrete structures’,
Australian small bridges conference, 2005, Sydney, New South Wales, Australia, Hallmark Conferences
and Events, Brighton, Vic, 18 pp.

Shayan, A & Xu, A 2006, ‘Prediction and progress of cracking of cover concrete caused by chloride-induced
corrosion of steel reinforcement’, Austroads bridge conference, 6th, 2006, Perth, Western Australia,
Austroads, Sydney, NSW, 14 pp.

Shayan, A & Xu, A 2012, ‘Condition survey of two bridges over Wimmera River in Horsham’, contract report
005428, ARRB Group, Vermont South, Vic.

Shayan, A 1985, ‘Warping of precast, white concrete panels’, Cement and Concrete Research, vol. 15, no. 2,
pp. 245–52.

Shayan, A 1988a, ‘Alkali aggregate reaction in a 60-year-old dam in Australia’, International Journal of
Cement Composites and Lightweight Concrete, vol. 10, no. 4, pp. 259–66.

Shayan, A 1988b, ‘Deterioration of a concrete surface due to the oxidation of pyrite contained in pyritic
aggregates’, Cement and Concrete Research, vol. 18, no. 5, pp. 723–30.

Shayan, A 1989a, ‘Re-examination of AAR in an old concrete’, Cement and Concrete Research, vol. 19, pp.
434-42.

Shayan, A 1989b, ‘Experiments with accelerated tests for predicting alkali-aggregate reaction’, Proceedings
of the 8th international conference on AAR, Kyoto, Japan, pp. 321–6.

Shayan, A 1992, ‘Prediction of alkali-reactivity of some Australian aggregates and correlation with field
performance’, ACI Materials Journal, Jan-Feb, pp. 13–23.

Shayan, A 1993a, ‘Reply to M.A. Adams on ‘Microscopic features of cracked and uncracked concrete railway
sleepers’’, ACI Materials Journal, vol. 90, pp. 284–7.

Austroads 2016 | page 114


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Shayan, A 1993b, ‘Alkali reactivity of deformed granitic rocks: a case study’, Cement and Concrete
Research, vol. 23, pp. 1229–36.

Shayan, A 1994a, ‘Alkali-aggregate reaction: a bridge management problem for road authorities’, ARRB
conference, 17th, 1994, Gold Coast, Queensland, Australia, ARRB Group, Vermont South, Vic, vol. 17,
no. 4, pp. 71-85.

Shayan, A 1994b, ‘An illustrated guide to the identification of alkali-aggregate reaction in concrete
structures’, CSIRO technical report TR94/2, CSIRO, Melbourne, Vic.

Shayan, A 1995a, ‘Developments in testing for AAR in Australia’, CANMET/ACI international workshop on
alkali-aggregate reaction in concrete, 1995, Dartmouth, Nova Scotia, Canada, Canada Centre for Mineral
and Energy Technology & American Concrete Institute, Farmington Hills, MI, USA, pp. 139–52.

Shayan, A 1995b, ‘Behaviour of precast, prestressed concrete railway sleepers affected by AAR’, 5th
CANMET/ACI international conference on fly ash, silica fume, slag & natural pozzolans in concrete (RN
Swamy symposium), Milwaukee, Wisconsin, USA, 4–9 June, 1995, pp. 35–56.

Shayan, A 1999, ‘Characterisation of AAR-affected concrete from a dam structure for rehabilitation
purposes’, Proceedings of the international conference on infrastructure regeneration and rehabilitation,
June 1999, Sheffield University, Sheffield, UK, pp. 777–87.

Shayan, A 2000, ‘Combined effects of alkali-aggregate reaction AAR) and cathodic protection currents in
reinforced concrete’, Proceedings of the 11th international conference on alkali aggregate reaction in
concrete, June 2000, Quebec City, Canada, pp. 229–38.

Shayan, A 2001, ‘Validity of accelerated mortar bar test methods for slowly reactive aggregates - comparison
of test results with field evidence’, Concrete in Australia, June-August 2001,
pp. 24–6.

Shayan, A 2003, ‘AAR in Australia and recent developments’, International seminar on road construction
materials, October 2003, Kanazawa, Japan.

Shayan, A 2007, ‘Field evidence for inability of ASTM C 1260 limits to detect slowly reactive Australian
aggregates’, Australian Journal of Civil Engineering, vol. 3, no. 1, pp, 13–26.

Shayan, A 2011, ‘Aggregate selection for durability of concrete structures’, Construction Materials, vol. 164,
pp. 111–21.

Shayan, A 2015, ‘The current status of AAR in Australia and mitigation measures’, Concrete in Australia, vol.
41, no. 2 pp. 44–51.

Shayan, A 2016, ‘Effects of AAR on concrete and structures’, Proceedings of the Institution of Civil
Engineers - Construction Materials, Vol. 169, Issue 3, pp. 145–153.

Shayan, A & Andrews-Phaedonos, F 2005, ‘Investigation of the causes of cracking in some VicRoads
bridges in East Gippsland, and estimation of their repair costs’, Concrete 05: proceedings of the 22nd
biennial conference of the Concrete Institute of Australia, 17-19 October 2005, Melbourne, 14 pp.

Shayan, A & Grimstad, J 2006, ‘Deterioration of a hydroelectric concrete gravity dam and its
characterisation’, Cement and Concrete Research, vol. 36, pp. 371–83.

Shayan, A & Morris, H 2001, ‘A comparison of RTA T363 and ASTM C-1260 accelerated mortar bar test
methods for detecting reactive aggregates’, Cement and Concrete Research, vol. 31, no. 4, pp. 655-63.

Shayan, A & Morris, H 2002, ‘Cracking in precast, prestressed deck planks in two bridges and rehabilitation
options’, ACI Materials Journal, vol. 99, March–April, pp. 165–172.

Austroads 2016 | page 115


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Shayan, A & Morris, H 2003, ‘Durability investigation of Deep Creek bridge, Northern NSW’, ARRB Transport
Research conference, 21st, 2003, Cairns, Queensland, Australia & Road Engineering Association of Asia
and Australasia (REAAA) conference, 11th, 2003, Cairns, Queensland, Australia, ARRB Transport
Research, Vermont South, Vic, 22 pp.

Shayan, A, Diggins, RG, Ivanusec, I & Westgate, PL 1988, ‘Accelerated testing of some Australian and
overseas aggregate for alkali aggregate reactivity’, Cement and Concrete Research, vol. 18, no. 6,
pp. 843–51.

Shayan, A, Green, WK & Collins, FG 1996, ‘Alkali- aggregate reaction in Australia’, in Shayan, A (ed), Alkali-
aggregate reaction in concrete: proceedings of the 10th international conference on AAR, Melbourne,
Australia, 18-23 August 1996, pp. 85–92.

Shayan, A, Morris, H & Doolan, T 1998, ‘Investigation of cracking of precast concrete bridge piles
submerged in tidal river water In northern NSW’, ARRB Transport Research conference, 19th, 1998,
Sydney, New South Wales, Australia, ARRB Transport Research, Vermont South, Vic, pp. 118–41.

Shayan, A & Xu, A 2004, ‘Value-added utilisation of waste glass in concrete’, Cement and Concrete
Research, vol. 34, pp. 81–9.

Shayan, A & Xu, A 2006, ‘Performance of glass powder as a pozzolanic material in concrete: a field trial on
concrete slabs’, Cement and Concrete Research, vol. 36, pp. 457–68.

Shayan, A, Wark, RE, & Moulds, A 2000, ‘Diagnosis of AAR in Canning Dam, characterisation of the affected
concrete and rehabilitation of the structure’, Proceedings of the 11th international conference on AAR,
June 2000, Quebec City, Canada, pp. 1383–92.

Shayan, A, Xu, A & Al-Mahaidi, R 2004, ‘Condition assessment of a reinforced concrete jetty structure, its
load capacity and suggested rehabilitation strategy’, Austroads bridge conference, 5th, 2004, Hobart,
Tasmania, Australia, Austroads, Sydney, NSW, 16 pp.

Shayan, A, Xu, A & Andrews-Phaedonos, F 2003, ‘Development of a performance measure for durability of
concrete bridges’, Proceedings of the 21st biennial conference: concrete in the third millennium’,
Brisbane, Australia, Concrete Institute of Australia, vol. 2, pp. 739–57.

Shayan, A, Xu, A & Hii, A 2006, ‘Causes of deterioration of precast bridge piles: an experimental study’,
Austroads bridge conference, 6th, 2006, Perth, Western Australia, Austroads, Sydney, NSW, 14 pp.

Shayan, A, Xu, A & Olasiman, R 2008, ‘Factors affecting the expansion and cracking of model bridge piles in
seawater, and the effects of mechanical confinement’, International conference on alkali-aggregate
reaction in concrete, 13th, 2008, Trondheim, Norway, Quik Scan, The Netherlands, pp. 1196–1209.

Shayan, A, Xu, A & Salamy, R 2012, ‘Confinement of AAR expansion in cylindrical reinforced columns by
CFRP wrapping’, International conference on alkali-aggregate reaction in concrete, 14th, 2012, Austin,
Texas, USA.

Shayan, A, Xu, A & Tagnit-Hamou, A 2004, ‘Effects of cement composition and temperature of curing on
AAR and DEF expansion in steam-cured concrete’, International conference on alkali-aggregate reaction
in concrete, 12th, 2004, Beijing, China, International Academic Publishers, Beijing, China, pp. 773–88.

Shayan, A & Quick, GW 1989, ‘Microstructure and composition of AAR products in conventional standard
and new accelerated testing’, Proceedings of the 8th international conference on AAR, Kyoto, Japan, pp.
475–82.

Shayan, A, Quick, GW, Lancucki, CJ & Way, SJ 1992, ‘Investigation of greywacke aggregates for alkali-
aggregate reactivity’, Proceedings of the 9th international AAR conference, London, UK, pp. 958–79.

Austroads 2016 | page 116


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Shayan, A, Quick, GW & Lancucki, CJ 1993, ‘Morphological, mineralogical and chemical features of steam-
cured concretes containing densified silica fume and various alkali levels’, Advances in Cement
Research, vol. 5, no. 20, pp. 151–62.

Sirivivatnanon, V & Khatri, R 2011, ‘Testing and specifying chloride resistance of concrete’, Austroads bridge
conference, 8th, 2011, Sydney, New South Wales, Australia, Austroads, Sydney, NSW, pp. 472–87.

Sirivivatnanon, V & Lucas, G 2011, ‘Specifying sulfate-resisting concrete’, Austroads bridge conference, 8th,
2011, Sydney, New South Wales, Australia, Austroads, Sydney, NSW, pp. 454–71.

Sisomphon, K & Franke, L 2007, ‘Carbonation rates of concretes containing high volume of pozzolanic
materials’, Cement and Concrete Research, vol. 37, no. 12, pp. 1647–53.

Skenderovic, B & Lomic, G 1992, ‘Effects of air-pollution on chemistry and dynamics of concrete
carbonation’, International congress on the chemistry of cement, 9th, 1992, New Delhi, India, National
Council for Cement and Building Materials, New Delhi, India, vol. 5, pp. 383–8.

Smolczyk, HG 1968, ‘Discussion to Hamada’’, International symposium on chemistry of cement, 5th, 1968,
Tokyo, Japan, Cement Association of Japan, Tokyo, vol. 3, pp. 369–84.

Song, H-W, Pack, S-W & Ann, KY 2009, ‘Probabilistic assessment to predict the time to corrosion of steel in
reinforced concrete tunnel box exposed to sea water’, Construction and Building Materials, vol. 23, no.
10, pp. 3270–8.

Song, G & Shayan, A 1999, Corrosion prevention of reinforced concrete structures, research report ARR
332, ARRB Transport Research, Vermont South, Vic.

Song, G & Shayan, A 1998, Corrosion of steel in concrete: causes, detection and prediction: a state of the
art review, review report 4, ARRB Transport Research, Vermont South, Vic.

Sørensen, B, Jensen, PB & Maahn, E 1990, ‘The corrosion properties of stainless steel reinforcement’, in
Page, CL, Treadaway, K & Bamforth, PB (eds), Corrosion of reinforcement in concrete: international
symposium on corrosion of reinforcement in concrete construction, 1990, Warwickshire, UK, Elsevier
Applied Science, New York, NY, USA, pp. 601–10.

Standards Norway 2003, Concrete structures: design and detailing rules, NS3473, Standards Norway,
Lysaker, Norway (in Norwegian) [withdrawn standard].

Stanley, EM & Batten, RC 1969, ‘Viscosity of sea water at moderate temperatures and pressures’, Journal of
Geophysical Research, vol. 74, no. 13, pp. 3415–20.

Stanton, TE 1940, ‘Expansion of concrete through reaction between cement and aggregate’. Proceedings of
the American Society of Civil Engineers, December, vol. 66, no. 10, pp. 1781–1811.

Stark, J & Bollmann, K 1997, ‘Ettringite formation: a durability problem of concrete pavements’, International
congress on the chemistry of cement, 10th, 1997, Gothenburg, Sweden, Amarkai, Gothenburg, Sweden,
vol. 4, paper no. 62.

Stark, J, Bollmann, K & Seyfarth, K 1992, ‘Investigation into delayed ettringite formation in concrete,’
International congress on the chemistry of cement, 9th, 1992, New Delhi, India, National Council for
Cement and Building Materials, New Delhi, vol. 5, pp. 348–54.

Stark, J & Seyfarth, K 1999,’ Ettringite formation in hardened concrete and resulting destruction’ in Erlin, B
(ed), Ettringite: the sometimes host of destruction, SP-117, American Concrete Institute, Farmington Hills,
MI, USA, pp. 125–40.

Steffens, A, Dinkler, D & Ahrens, H 2002, ‘Modelling carbonation for corrosion risk prediction of concrete
structures’, Cement and Concrete Research, vol. 32, no. 6, pp. 935–41.

Austroads 2016 | page 117


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Stewart, MG & Mullard, JA 2007, ‘Spatial time-dependent reliability analysis of corrosion damage and the
timing of first repair for RC structures’, Engineering Structures, vol. 29, no. 7, pp. 1457–64.

Stewart, MG, Wang, X & Nguyen, MN 2012, ‘Climate change adaptation for corrosion control of concrete
infrastructure’, Structural Safety, vol. 35, March, pp. 29–39.

Stokes, DB 1996, ‘Use of lithium to combat alkali silica reactivity’, Proceedings of the 10th international
conference on AAR, Melbourne, Australia, pp. 862–67.

Stokes, DB, Thomas, MDA & Shashiprakash, SG 2000, ‘Development of lithium-based material for
decreasing ASR-induced expansion in hardened concrete’, Proceedings of the 11th international
conference on AAR, Quebec City, Canada, pp. 1079–87.

Stratfull, RF, Jurkovich, WJ & Spellman, DL 1975, ‘Corrosion testing of bridge decks,’ Transportation
Research Record, no. 539, pp. 50–9.

Swamy, RN (ed) 1992, The alkali silica reaction in concrete, Blackie & Son Ltd, London, UK.

Swamy, RN, Hamada, H & Laiw, JC 1994, ‘A critical evaluation of chloride penetration into concrete in
marine environment’, in Swamy, RN (ed), Corrosion and corrosion protection of steel in concrete,
Sheffield Academic Press, Sheffield, UK, pp. 404–19.

Tagnit-Hamou, A & Petrov, N 2004, ‘A new method for evaluating the risk of DEF’, Cement, Concrete and
Aggregates, vol. 26, no. 2, 6 pp.

Tang, L 1996, ‘Chloride transport in concrete: measurement and prediction’, PhD thesis, Chalmers University
of Technology, Gothenburg, Sweden.

Tang, L & Gulikers, J 2007, ‘On the mathematics of time-dependent apparent chloride diffusion coefficient in
concrete’, Cement and Concrete Research, vol. 37, no. 4, pp. 589–95.

Tang, SW, Yao, Y, Andrade, C & Li, ZJ, 2015, ‘Recent durability studies on concrete structures’, Cement and
Concrete Research, vol. 78, pp. 143–54.

Tarcutta Hume Alliance 2009, ‘Design criteria report: durability assessment and design life’, THA-R-20-
GE007A-DD-03, prepared for Roads and Maritime Services, Sydney, NSW.

Tepponen, P & Eriksson, BE 1987, ‘Damages in concrete railway sleepers in Finland’, Nordic Concrete
Research, vol. 6, pp. 199-209.

The Institution of Structural Engineers 1992, Structural effects of alkali-silica reaction: technical guidance on
the appraisal of existing structures, Institution of Structural Engineers, London, UK.

Thomas, BP, Fitzpatrick, RW, Merry, RH & Hicks, WS 2003, Coastal acid sulfate soil management
guidelines, Barker Inlet, SA, Acid Sulfate Soil Technical Manual, version 1.2, CSIRO Land and Water,
Glen Osmond, SA.

Thomas, M 1996, ‘Chloride thresholds in marine concrete,’ Cement and Concrete Research, vol. 26, no. 4,
pp. 513–9.

Thomas, MDA & Bamforth, PB 1999, ‘Modelling chloride diffusion in concrete: effect of fly ash and slag’,
Cement and Concrete Research, vol. 29, no. 4, pp. 487–95.

Thomas, MDA, Bremner, T & Scott, ACN 2011, ‘Actual and modelled performance in a tidal zone,’ Concrete
International, vol. 33, no. 11, pp. 23–8.

Austroads 2016 | page 118


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Thomas, MDA, Matthews, JD & Haynes, CA 1990, ‘Chloride diffusion and reinforcement corrosion in marine
exposed concretes containing PFA’, in Page, CL, Treadaway, K & Bamforth, PB (eds), Corrosion of
reinforcement in concrete: international symposium on corrosion of reinforcement in concrete
construction, 1990, Warwickshire, UK, Elsevier Applied Science, New York, NY, USA, pp. 198–212.

Thomas, MDA, Mukherjee, PK, Sato, JA & Everitt, MF 1995, ‘Effect of fly ash composition on thermal
cracking in concrete’, CANMET/ACI International conference on fly ash, silica fume, slag, and natural
pozzolans in concrete, 5th, Milwaukee, WI, USA, publication no. SP-153, American Concrete Institute,
Detroit, MI, USA, pp. 81–98.

Thomas, MDA, Scott, A, Bremner, T, Bilodeau, A & Day, D 2008, ‘Performance of slag concrete in marine
environment’, ACI Materials Journal, vol. 105, no. 6, pp. 628–34.

Tittarelli, F & Moriconi, G 2011, ‘Comparison between surface and bulk hydrophobic treatment against
corrosion of galvanised reinforcing steel in concrete’, Cement and Concrete Research, vol. 41, no. 6, pp.
609–14.

Torii, K, Kawamura, M, Matsumoto, K & Ishii, K 1996, ‘Influence of cathodic protection on cracking and
expansion of the beams due to alkali-silica reaction’, Proceedings of the 10th international conference on
AAR, 10th, 1996, Melbourne, Victoria, Australia, CSIRO, Melbourne, Vic, pp. 653-60.

Torii, K, Kumagai, Y, Okunda, Y, Ishii, K & Sato, K 2000, ‘Strengthening method for ASR-affected concrete
piers using prestressing steel wire’, Proceedings of the 11th international conference on AAR, Quebec
City, Canada, pp. 1225–33.

Torii, K, Sannoh, C, Kubo, Y & Ohashi, Y 2004, ‘Serious damages of ASR affected RC bridge piers and their
strengthening techniques’, International conference on alkali-aggregate reaction in concrete, 12th, 2004,
Beijing, China, International Academic Publishers, Beijing, China, pp. 1283–88.

Torii, K, Prasetia, I, Minato, T & Ishii, K 2012,’The feature of cracking in prestressed concrete bridge girders
deteriorated by alkali-silica reaction’, Proceedings of the 14th ICAAR, May 2012, Austin, Texas, USA, 10
pp.

Treadaway, KWJ 1978, ‘Durability of steel in concrete’, in Proceedings of a symposium on corrosion of steel
reinforcements in concrete construction, Society of Chemical Industry Materials Preservation Group,
London, UK, 15 February 1978, pp. 1–14.

Treadaway, KWJ, Cox, RN & Brown, BL 1989, ‘Durability of corrosion resisting steels in concrete’,
Proceedings of the Institution of Civil Engineers, vol. 86, pp. 305–31.

Tuutti, K 1982, Corrosion of steel in concrete, report no. CBI Forskning 4/82, Swedish Cement and Concrete
Research Institute, Stockholm, Sweden.

Val, D, Bljuger, F & Yankelevsky, D 1996, ‘Optimization problem solution in reliability analysis of reinforced
concrete structures’, Computers and Structures, vol. 60, no. 3, pp. 351–5.

Val, DV & Stewart, MG 2003, ‘Life-cycle cost analysis of reinforced concrete structures in marine
environments’, Structural Safety, vol. 25, no. 4, pp. 343–62.

VicRoads 2009, Coating of concrete, specification section 686, VicRoads, Kew, Vic.

VicRoads 2010a, Structural concrete, specification section 610, VicRoads, Kew, Vic.

VicRoads 2010b, Guide to the assessment, maintenance and rehabilitation of concrete bridges, technical
bulletin TB 51, VicRoads, Kew, Vic.

VicRoads 2011, Road structures inspection manual, VicRoads, Kew, Vic.

Austroads 2016 | page 119


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Vidal, T, Castel, A & François, R 2007, ‘Corrosion process and structural performance of a 17 year old
reinforced concrete beam stored in chloride environment’, Cement and Concrete Research, vol. 37, no.
11, pp. 1551–61.

Way, SJ & Shayan, A 1989, ‘Early hydration of Portland in water and sodium hydroxide solutions:
composition of solutions and nature of solid phases’, Cement and Concrete Research, vol. 19, no. 5, pp.
759–69.

West, G 1996, ‘Alkali-aggregate reaction in concrete roads and bridges’, Thomas Telford, London, UK.

Weyers, RE 1998, ‘Service life model for concrete structures in chloride laden environments’, ACI Materials
Journal, vol. 95, no. 4, pp. 445–53.

Weyers, RE, Fitch, MG, Larsen, EP, Al-Qadi, IL, Chamberlain, WP & Hoffman, PC 1994, Concrete bridge
protection and rehabilitation: chemical and physical techniques: service life estimates, report no. SHRP-
S-668, Strategic Highway Research Program, Washington, DC, USA.

Whitmore, D & Abbott, S 2000, ‘Use of an applied electric field to drive lithium ions into alkali-silica reactive
structures’, Proceedings of the 11th international conference on AAR, Quebec City, Canada, pp. 1089–
98.

Wiens, U, Breit, W & Schiessl, P 1995, ‘Influence of high silica fume and high fly ash contents on alkalinity of
pore solution and protection of steel against corrosion’, CANMET/ACI International conference on fly ash,
silica fume, slag, and natural pozzolans in concrete, 5th, Milwaukee, WI, USA, publication no. SP-153,
American Concrete Institute, Detroit, MI, USA, pp. 741–61.

Wigum, BJ & Thorenfeldt, E 2004, ‘Sheets of carbon fibre reinforced polymers (CFRP) as a repair material in
order to strengthen and repair concrete damaged by alkali-aggregate reaction’, International conference
on alkali-aggregate reaction in concrete, 12th, 2004, Beijing, China, International Academic Publishers,
Beijing, China, pp. 1289–98.

Wiktor, V & Jonkers, HM 2011, ‘Quantification of crack-healing in novel bacteria-based self-healing


concrete’, Cement and Concrete Composites, vol. 33, no. 7, pp. 763–70.

Xia, J, Jin, W & Li, L 2011, ‘Shear performance of reinforced concrete beams with corroded stirrups in
chloride environment’, Corrosion Science, vol. 53, no. 5, pp. 1794–1805.

Xie, P & Beaudoin, JJ 1992, ‘Mechanism of sulfate expansion I: thermodynamic principle of crystallisation
pressure’, Cement and Concrete Research, vol. 22, no. 4, pp. 631–40.

Xu, A & Shayan, A 1999, ‘A review of state road authorities corrosion damage survey of concrete bridges:
phase II, part 2: review of approaches to service life prediction’, contract report RC7007AA, ARRB
Transport Research, Vermont South, Vic.

Xu, A & Shayan, A 2008, ‘Evaluation of chloride diffusion coefficient in concrete by the rapid chloride
penetration tests: problems and observations’, ARRB conference, 23rd, 2008, Adelaide, South Australia,
Australia, ARRB Group, Vermont South, Vic, 15 pp.

Xu, A & Shayan, A 2016, ‘Relationship between reinforcing bar corrosion and concrete cracking’, ACI
Materials Journal, vol. 113, no. 1, pp. 3–12.

Xu, A, Shayan, A & Baburamani, P 1998, Test methods for sulfate resistance of concrete and mechanism of
sulfate attack, review report no. 5, ARRB Transport Research, Vermont South, Vic.

Yang, Q, Wu, X & Huang, S 1997, ‘Concrete deterioration due to physical attack by salt crystallization’,
International congress on the chemistry of cement, 10th, 1997, Gothenburg, Sweden, Amarkai,
Gothenburg, Sweden, paper 4iv032, 5 pp.

Austroads 2016 | page 120


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Yeginobali, A & Dilek, FT 1995, ‘Sulfate resistance of mortars containing silica fumes as evaluated by
different methods’, CANMET/ACI International conference on fly ash, silica fume, slag, and natural
pozzolans in concrete, 5th, Milwaukee, WI, USA, publication no. SP-153, American Concrete Institute,
Detroit, MI, USA, pp. 795–813.

Zhang, Z, Olek, J & Diamond, S 2002, ‘Studies on delayed ettringite formation in heat cured mortars II:
characteristics of cement that may be susceptible to DEF’, Cement and Concrete Research, vol. 32, no.
11, pp. 1737–42.

Zoob, AB, Le Claire, PJ & Pfeifer, DW 1985, ‘Corrosion protection tests on reinforced concrete with solid
stainless steel reinforcing bars for Joslyn stainless steels’, Wiss, Janney, Elstner Associates, Northbrook,
IL, USA.

ASTM & International Standards

ASTM A775 / A775M-07b, Standard specification for epoxy-coated steel reinforcing bars.

ASTM A955 / A955M-12e1, Standard specification for deformed and plain stainless-steel bars for concrete
reinforcement.

ASTM C227-03, Standard test method for potential alkali reactivity of cement-aggregate combinations
(mortar-bar method).

ASTM C289-07, Standard test method for potential alkali-silica of aggregates (chemical method).

ASTM C295 / C295M-12, Standard guide for petrographic examination of aggregates for concrete.

ASTM C1260: 2014, Standard test method for potential alkali reactivity of aggregates (mortar-bar method).

ASTM C1543-10a, Standard test method for determining the penetration of chloride ion into concrete by
ponding.

ASTM C642-06, Standard test method for density, absorption, and voids in hardened concrete.

ASTM C1556-11a, Standard test method for determining the apparent chloride diffusion coefficient of
cementitious mixtures by bulk diffusion.

ASTM C1585-11, Standard test method for measurement of rate of absorption of water by hydraulic-cement
concretes.

ASTM C1202-12, Standard test method for electrical indication of concrete's ability to resist chloride ion
penetration, West Conshohocken, United States.

ASTM C1293-08, Standard test method for determination of length change of concrete due to alkali- silica
reaction.

British/ European Standards

BS EN 197-1:2011, Cement: composition, specifications and conformity criteria for common cements.

BS EN 206:2013, Concrete: specification, performance, production and conformity.

BS EN 1992-1-1:2004+A1:2014, Eurocode 2: Design of concrete structures. General rules and rules for
buildings.

BS EN 206:2013, Concrete: specification, performance, production and conformity.

BS EN 450-1:201, Fly ash for concrete: definition, specifications and conformity criteria.

Austroads 2016 | page 121


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Canadian Standards

CSA A23.2-27A, Standard practice to identity degree of alkali-reactivity of aggregates and to identity
measures to avoid deleterious expansion in concrete.

CSA A23.2-28A, Standard practice for laboratory testing to demonstrate the effectiveness of supplementary
cementing materials and lithium-based admixtures to prevent alkali-silica reaction in concrete.

Australian Standards

AS 1012.21-1999, Methods of testing concrete: determination of water absorption and apparent volume of
permeable voids in hardened concrete.

AS 1141.60.1-2014, Method for sampling and testing aggregates: method 60.1: potential alkali-silica
reactivity: accelerated mortar bar method.

AS 1141.60.2-2014, Method for sampling and testing aggregates: method 60.2: potential alkali-silica
reactivity: concrete prism method.

AS 1141.65-2008, Method for sampling and testing aggregates: method 65: alkali aggregate reactivity:
qualitative petrological screening for potential alkali-silica reaction.

AS 2350.14-2006, Methods of testing Portland, blended and masonry cements: length change of cement
mortars exposed to sulfate solution.

AS 3600-2009, Concrete structures.

AS 3972-2010, General purpose and blended cements.

AS 5100.5-2004, Bridge design: concrete.

AS/NZS 3582.1-2016, Supplementary cementitious materials: part 1: fly ash.

AS/NZS 3582.2-2016, Supplementary cementitious materials: part 2: slag: ground granulated blast-furnace.

AS/NZS 3582.3-2016, Supplementary cementitious materials: part 3: amorphous silica.

RMS T363, 2001, Accelerated mortar bar test for AAR assessment.

SA HB 79:2015, Alkali aggregate reaction: guidelines on minimising the risk of damage to concrete
structures in Australia.

VicRoads RC376.03, 2013, Accelerated mortar bar test: alkali-silica reactivity of aggregate.

WA 624.1, 2012, Potential alkali - silica reaction by accelerated mortar bar test: alkali-silica reactivity of
aggregates.

Austroads 2016 | page 122


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Appendix A Comparison of Requirements for Concrete Mix Design and


Durability by Four Jurisdictions

Table A 1: Comparison of requirements for concrete mix design and durability by four jurisdictions

Issue MRWA VicRoads Roads and Maritime TMR


Concrete mix Contractor to provide for each class All concrete shall be special class as in AS Based on durability provision. To All concrete mixes to have a
design of concrete specified. 1379. Durability requirements for condition U. achieve target strength fc. md greater minimum 20% fly ash by mass of
Contractor responsible for mix to meet than cx equal to f c min + M control. cement.
strength, durability and other requirements. M = margin for variation; f c min is the Contractor to submit mix six weeks
greater of spec. min. 28-day strength, or before the operations. Info on class
min. 28-day strength for relevant of concrete, supplier, mix design,
durability provision – see later. Hold target slump and strength needed
points apply. Maximum compressive for each class of concrete. Quarry
strength shall not exceed 80 MPa. certificate required for aggregate.
Class of S 35 RC substructure, non- Grade of concrete indicating cement content Depends on durability provision and Class denoted as x MPa/Y, where
concrete marine and specified 28-day strength are given. exposure classification. x = characteristic compressive
Details of all the materials are required, strength and Y maximum aggregate
S 35/10 Precast parapet panels
Certificates needed. size in mm.
(10 mm max. aggregate
size) VR 330/32VR min. cement/ MPa Reinforced concrete > 25 MPa.
VR 400/40 Concrete class noted on drawings.
S 40 Normal RC superstructure
VR 450/50
S 50 Special RC superstructure
VR 470/55
S 50M RC substructure, marine

Austroads 2016 | page 123


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Issue MRWA VicRoads Roads and Maritime TMR


Corresponding F’c Cement Strength Cement Exp. Cement Max. Min. fc Exp. Min. Strength
target Class (MPa) W/C kg/m3 (MPa) W/C kg/m3 class kg/m3 W/C MPa class cement W/C fc (MPa)
strength, W/C
S35 42 0.50 350 3 7 28 A 320 0.56 25 B1 320 0.56 32
ratio and
minimum S40 42 0.45 350 15 24 32 0.50 330 B1 320 0.50 32 B1 390 0.46 40
cement
content S50 48 0.43 400 18 30 40 0.45 400 B2 370 0.46 40 C 450 0.40 50
S50M 58 0.40 420 23 40 50 0.40 450 C 420 0.40 50 Minimum target strength, f’t to be
calculated
OPC for all but S50M, in which 25 45 55 0.36 470 U Special cases
blended cement is used. f’t = f’c + 1.65s; where s = standard
deviation for the class of concrete;
35% OPC – 8% replaced by S.F.
s  0.08 f’c. If strength records not
65% slag – 8% replaced by S.F.
available, then 0.12 f’c < s < 0.20 f’c.
Trial mix  Exemption may be granted  Two cylinders per mix tested for comp.  Contractor to provide all details on  Contractor to provide approved
based on previous results. strength. One cylinder tested for VPV at intended use of nominated mix. aggregate to be used (hold point).
 Trial mix needed for all grades of 28 days.  Materials data, not older than  3 days’ notice for each trial mix
concrete.  If VPV requirement is not met, then the 12 months to be provided. (witness point).
 Slump test AS 1012.3: ± 15 mm contractor must modify mix and test again.  Describe method of control of AAR.  Min. vol. of mix = 25% capacity of
of agreed slump for the class of  Water may be added within the limit prior  Give materials quantities in the mix. mixer.
concrete. to discharge, but this is recorded and  Trial mix to be tested for slump,
 Target strength, etc.
Max. slump 100 mm. slump measured after it. workability and strength.
 Control and accuracy of batching.
 When super-plasticiser is used  ID of all materials to be recorded.  For strength gain, a minimum of 3
max. slump before using it = 100  Control of aggregate moisture
 Concrete must be discharged within 60 cylinders per age to be cast in
mm. content. Control of variation of comp.
minutes from the start of mixing. addition.
strength.
 9–21 comp. strength cylinders to  Each mix tested for slump, comp. strength,  Mix is approved when:
be tested to check the strength  Minimum – maximum slump.
air entrainment, VPV, drying shrinkage, - slump range is achieved
requirements. soluble salts.  Method of curing.
- mean 28-day strength  target
 Mean 28-day strength shall not  Expected air temperatures, min./max. strength
be less than target strength, and  For durability provision A: - mix conforms to AAR test
70% of specimens with strength
- Max sorptivity depth, comp. Q458.
± 3 MPa from the mean.
strength and ‘mix report’.  Early approval can be given if 7-
 Batching and mixing  AS 1379.
 Max. allowed variations to nominated day strength  0.8 target strength,
mix: but 28-day strength must be
- Cement 3%, aggregate 5%, water achieved.
3%, admix 5%, all percentages are  Approved mix must not be
by mass. changed.
 Each batch to be tested
separately.

Austroads 2016 | page 124


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Issue MRWA VicRoads Roads and Maritime TMR


Temperature Concrete temp < 32 °C at 5 °C < air temp. < 35 °C. Concrete temp. > 10 °C. 10 °C < concrete temp. < 35 °C.
of delivered placement. 10 °C < concrete temp. > 32 °C. Concrete temp. > 32 °C for precasting, Take precautions if air temp. >
concrete Ambient temp < 38 °C at the time Cold and hot weather concreting addressed. 5 °C < concrete temp < 35 °C. 35 °C.
placement is completed. Temperature difference across any element Do not place concrete if air temp.
not > 20 °C. is > 45 °C at or 2 hours after
placement.
Consistency Slump test  AS 1012.3  AS 1012.13, clause 5.2 AS 1379 before Concrete slump shall not exceed 200 Slump  Q451A, AS 1012.8 related
No water shall be added to and after superplasticiser. mm. to f’c as follows:
concrete if too dry for satisfactory Do slump test on each batch of concrete and Concrete slump before addition of For f’c 20, 25, 32, 40, 50 MPa, pump
placement. visual inspection of mix – both to be superplasticiser shall not be less than and tremie concrete, slump should
recorded. 40 mm. be 70–90, 60–100, 50–100, 80–100
Water may be added at discharge point, mm. Hold point on slump.
max. 9 kg/m3 provided specified W/C not Wet mixes can be rejected.
exceeded. Tolerances are as given in Table 6
of AS 1379.
Strength of Sampling  AS 1012, not less than Each sample for comp. strength: Tests on hardened concrete: Cylinders to be cured as in
delivered 3 specimens to be cast. 3 cylinders for reinforced, prestressed or pre- Strength  AS 1012 AS 1012.8.
concrete Compliance conditions: tensioned concrete. Not moved for 18 hrs after casting.
Shrinkage AS 1012.13
Greater than 95% of nominated 3 or 5 cylinders for post- tensioning, if applied Sulfate and chloride Stored in lime water, close to 27 °C.
strength is achieved by all the after or before 28 days. Capping may be in rubber, but
Trial mix report according to AS 1012.2.
specimens and mean 28-day 2 cyl. min. to be tested at 28 days. sulphur capping is needed if 3
strength at 28 days for any group of Concrete cracking – no crack width to
1 cylinder min. to be tested for 7-day strength specimens give mean strength <
exceed 0.05 mm.
4 consecutive samples  nominated and other requirements for prestressed and 0.5(f’t + f’c). For matched set of
strength. post-tensioned concrete. samples, strength tolerance is
If mix does not comply or new Compliance conditions: 2–3 MPa for 2 or 3 samples,
materials are introduced, then new respectively.
Average strength not less than that specified,
trial mixes shall be prepared and Conditions for rejection of mix:
and each test not less than 90% of that
tested. Either of:
specified.
Cores from completed work may be tested  sample strength < 0.9 f’c. OR
3 cores  AS 1012.14. average of 10 samples in 4 weeks
less than 0.5(f’c + f’t) or
 standard deviation exceeds that
used for calculation of f’t.

Austroads 2016 | page 125


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Issue MRWA VicRoads Roads and Maritime TMR


Curing Curing to start immediately after AS 1012 Curing depends on durability Curing procedure to be submitted 2
final set. requirements and exposure class. weeks prior to concrete pour.
Min. curing time for structural conc. Steam-cured concrete – specimens for Curing period (days) Curing to start within 1 hour of form
7 days. testing to be the same as the S.C. element. removal and continue 14 days for
Min. curing time for S50M 14 days. slabs and 7 days for other surfaces.
Curing times are: Exp. SL Blended Blended
Concrete surface  permanently Class Cement BCF/FA SF Other options for curing are given.
wet during curing.
GP VR 330/32, 6–7 days; GB 8–9 days A 7 7 NA
GP VR 400/40, 5–6 days; GB 7–8 days B1 7 7 NA
GP VR 450/55, 5 days; GB 7 days. B2 7 14 7
C NA 14 7
U Special cases
Steam Curing Air temp. rise 24 °C/h, temperature Essentially similar to MRWA requirements. Similar to MRWA and VicRoads, but Max. temperature to be maintained
under steam cover < 75 °C; temp. differential with ambient < 40 °C for 6 hours for piles, girders, decks
maintain at 75 ± 5 °C. permitted (too high). and kerb units.
Rate of cooling < 30 °C. Remove For durability provision B  Temp (not Air temp. rise not more than
steam cover when ambient less than 60 °C) × t (hours)  350. 24 °C/hr.
temperature differential < 30 °C. Max. air temp. 70 °C, max. concrete
Temperature differential between temp. 75 °C. Element exposed when
adjacent parts < 10 °C. Achieve 7- difference with air temp. < 20 °C.
day strength equivalence. Expose Transfer of prestress when unit
elements when temp. differential cools to 60 °C.
with ambient < 30 °C. Note: Surface
temp. is measured not concrete
temperature.

Austroads 2016 | page 126


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Issue MRWA VicRoads Roads and Maritime TMR


Durability Intent is to provide durable concrete In addition to AAR and other tests, VPV Control of AAR. See earlier. No concrete to be placed until the
for permanent works. values are specified as follows: Durability provision A (Performance) and aggregate has been approved.
No specific action. Grade of Test cylinders durability provision B (Methods). Conform to AS 2758.1 and AAR test
concrete In both, for exposure condition C Q458.
Vibrated Rodded Cores
concrete shall be made with blended Min. 20% fly ash in concrete
VR 330/32 14 15 17 cement. Silica fume shall not be used for No other durability test specified.
exposure conditions A and B1, for which
VR 400/40 13 14 16
a min. cement content of 230 kg/m3 shall
VR 45050 12 13 15 be used.
Concrete decks and slabs in any
VR 470/55 11 12 14
condition shall not have S.F. in the
blended cement. Wet curing specified for
concretes containing S.F.
Exp. Cement W/C Sorptivity
Class Min. max. depth
kg/m3 (mm)
OPC
GPB
A1 320 0.56 35 35
B1 320 0.50 25 25
B2 370 0.46 17 20
C 420 0.40 NA 8
Provision B: No sorptivity needed, but
other requirements are as stated
earlier.

Austroads 2016 | page 127


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Issue MRWA VicRoads Roads and Maritime TMR


Drying Not specifically covered. One sample per trial mix is tested. Max. shrinkage strain (s) Not specifically covered.
Shrinkage Refers to AS 1379. 3 specimens are made according to Refers to AS 1379.
Drying period
AS 1012.13. Drying shrinkage strain < 750 x
10-6 after 56 days. Exp. Class 3 weeks 8 weeks
A 570 800
B1, B2 500 700
C 430 600
U Special cases

Table A 2: Comparison of requirements for Materials and Tests to be complied with by four jurisdictions

Materials MRWA VicRoads Roads and Maritime TMR


Cement GP AS 3972 GP, GB AS 3972 Shrinkage limited (SL) cement used GP or GB AS 3972.
Slag AS/NZS 3582.2 Fly ash AS/NZS 3582.1 AS 3972. Portland cement  60% of total
S.F. AS/NZS 3582.3 Slag AS/NZS 3582.2 Blended AS 3972 for GB. cement for all mixes.
Fly ash not used S.F. AS/NZS 3582.3 Acceptable composition of blended Cement alkali < 0.6% Na2O eq.
cements given. Fly ash AS/NZS 3582.1.
Acceptable composition of blended cements Only five graded fly ash to be
given. used:
total alkali content: < 2%
active alkali: < 0.5%.
Slag AS 3582.2:
total alkali content: < 1%
active alkali content: < 0.5%.
S.F. not used in decks or slabs.
Chemical AS 1478 requirements, to be AS 1478 requirements as in AS 1379. AS 1478 requirements, No CaCl2. Written approval needed AS
Admixture used to improve workability, not No CaCl2 or ca-formate. Corrosion inhibitor: Ca-Nitrite, 30% 1478.
to reduce cement content. Shall No air entraining agent unless approved, solution, dosage, 9 kg solids/m3. No CaCl2, alkali contribution to
not affect the quality of concrete then, < 6% air content. concrete < 0.2 kg/m3.
and reinforcement steel.
If AEA used, then air content <
6%.

Austroads 2016 | page 128


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Materials MRWA VicRoads Roads and Maritime TMR


Aggregate Fine aggregate  natural sand AS 2758.1. AS 2758.1 plus other requirements. Fine aggregate > 50% natural
AS 2758.1. Crusher fines maximum 20% of sand. Water absorp. coarse aggr. < 2.5%. sand.
Water absorption < 1.5%. Coarse aggregate: Water absorp. fine aggr. < 2.5%. Coarse aggregate:
Coarse aggregate – not graded. < 5% unsound rock Water absorp. slag aggr. 6%. Quarry assessment procedure.
Shall be single sized crushed < 10% unsound and marginal. Wet/dry strength variation Flakiness Q201A.
river gravel or igneous rock. Flakiness < 38%, grading AS 2758.1. AS 1141.22. Water absorption < 2.5% Q214B.
Water absorption. < 2.5%. AAR restrictions. Non-plastic by AS 1289.3. 10% fines – Q214.
AS 2758.1 Exposure Class C. Manufactured sand allowed. Wet/dry strength Q205A, B, C.
AAR compliance Clause 2.5, B80.
AAR Contractor shall demonstrate Petrographic  ASTM C295. Petrographic: ASTM C295. AS 2758.1,
that no AAR will take place. AAR code of practice RC 500.16. Allows rejection of reactive material In addition, Test Method Q458
Concrete alkali content < AMBT expansion (21 days): without further testing. concrete mixes with 20% of an
2.5 kg/m3 by AMBT: RMS T363 Method. approved fly ash are deemed to
< 0.15% for sand
Report No. TR3 Part 2, NZ C & CPT: RMS T364 Method. comply with AAR requirements.
< 0.10% for aggregate.
CA. Aggregate class: Aggregate used with written
If expansion > 0.10% in 21 days, then all the approval of superintendent.
following shall apply: non-reactive – no action
AS 1478 – no CaCl2.
alkali content < 2.8 kg/m3, use blended slowly reactive – use low-alkali
cement or blended cement Alkali contribution < 0.2 kg/m3.
cement
reactive – use other aggregates or Air content if A.E.A used < 6%.
expansion of mortar bar containing
aggregate + GB combination < 0.10% in 21 use blended cement and test by T364
days method.
other aggregate tests have been identified,
the frequency of testing has been given.
Soluble salts Water: Water  AS 1379 Water  AS 1379. AS 1379.
< 1500 ppm by WA method AS 1012 – Part 20, SO3 < 5% by concrete AS 1012.20. Not separately specified.
CH/4, mass. SO3 < 5% by mass of concrete.
British Standard 3148-1959. Max. chloride content: Cl  0.3–0.8 kg/m3 depending on
0.1% prestressed concrete exposure and element type.
0.15% reinforced
0.03% post-tension grout.

Austroads 2016 | page 129


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Appendix B Threshold Values for Steel Corrosion

These data for carbon steel are those presented by Hurley and Scully (2002).

B.1 Carbon Steel

Table B 1: Chloride threshold concentration to initiate corrosion on carbon steel

Free Cl- Total Cl- Cl-/OH-


Reference Condition and test specials Environment Test
(% cement) (% cement) (mole ratio)
Hausmann (1967) Solution simulating concrete Solution 0.60 Corrosion potential
Gouda (1970) Solution simulating concrete Solution 0.35 Corrosion potential
Goni and Andrade (1990) Alkaline solution Solution 0.25–0.8 Corrosion rate
Gouda and Halaka (1970) Mortar suspensions OPC 2.42 Anodic polarisation
BFSC 1.21
Concrete with added Cl OPC 3.04
BFSC 1.01
Non pre-cleaning bars OPC 0.6
Petterson (1992, 1994, Cements with high to low alkali content Mortars 2.5–6.0 Corrosion rate
1996) 80% RH 0.6–1.8 1.7–2.6
100% RH 0.5–1.7 1.7–2.6
Concrete in 10% NaCl water Concrete 1.8–2.9
Andrade and Page (1986) Cement with added chloride OPC 0.15–0.69 Corrosion rate
BFSC 0.12–0.44
Hansson and OPC mortar, external chloride 100% RH Corrosion rate,
Sorenson (1988) 50% RH 0.6–1.4 potentiostatic
Lambert et al. (1991) Concrete, external chloride Concrete 3.0 Corrosion rate

Austroads 2016 | page 130


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Free Cl- Total Cl- Cl-/OH-


Reference Condition and test specials Environment Test
(% cement) (% cement) (mole ratio)
Kayyali and Haque (1995) Med. strength concrete Concrete with added 1.15 0.6 assumed
High strength chloride 0.85
High strength + supplement 0.80
High strength + supplement + fly ash 0.45
Hussain et.al. (1995) C3A = 2.43% Concrete 0.14 0.35 0.3 assumed
C3A = 7.59% 0.17 0.62
C3A = 14.0% 0.22 1.00
Schiessl and Breit (1996) Concrete with added Cl Concrete exposed Concrete 0.5–1 Macrocell currents
to Cl 1–1.15
Thomas et al. (1990) Concrete at marine exposure Concrete 0.5 Visual observation,
mass loss
Thomas (1996) Fly ash= 0%, marine exposure Concrete 0.70 Mass loss
Fly ash = 15%, marine 0.65
Fly ash = 30% marine 0.50
Fly ash = 50% marine 0.20
Hope and Ip (1987) Concrete slabs with added Cl to various Concrete 0.1–0.19 Corrosion rate, visual
exposure conditions observation, mass loss
Li and Sagues (2001) Simulated concrete pore solution pH = 12.6 0.25–0.7 Corrosion potential and
pH = 13.5 1.05–1.25 polarisation resistance
Alonso et al. (2000) Simulated concrete pore solution Solution 0.25–0.8
Mortar 0.39–1.06 1.24–3.08 1.17–3.98

Austroads 2016 | page 131


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

B.2 Stainless Steel

Table B 2: Chloride threshold concentration for stainless steel

Stainless steel type and test Chloride threshold of stainless Cl-/OH-


Reference Environment Test
medium steel (%) (mole ratio)
Cox and Oldfield 302, 315, 316 steel in concrete Concrete > 3.2% by cement mass 0.60 Visual inspection for 22 years
(1996)
Bertolini et al. 410, 304L, 316L, 2304 steel in In chloride solutions of pH 2.0, 5.0, 5.5, and >10% in solution, 1–4 Corrosion rate, potentiostatic
(1996) Ca(OH)2 based solution =9 respectively >2 held at +200 mV (vs. SCE)
pH = 12.6 >8
pH = 13.9 >8
Bertolini et al. 304LN, 316LN, 2304, LDX2101, in In chloride solution of 6.5–10, >10, 7.5–8, 3.5–6% in Potentiostatic as above
(2009) chloride solutions pH = 12.6 solution, respectively
Hurley and Scully 316LN, LDX 2101 In chloride solution of 2.8–11.29 and 1.1–1.4% in As above
(2006) pH = 12.6 solution, respectively
Randström, et al. 304L, 316L, LDX 2101, 2304 in In chloride solution of 2.8–10.6, 5.7–6.8, 5.7–10.6, 4.6– As above
(2010) chloride solutions pH = 12.8–13.3 10.6% in solution, respectively
Houska and 304, 316, LDX 2101, 2304 cast in Concrete with 2.6% 90% probability that the chloride Note: chloride tolerance of 316,
Holsing (2012) concrete chloride by cement mass, threshold of 316, LDX 2101 and LDX 2101 and 2304 are similar
and concrete with up to 2304 will exceed 5.7%, and the and better than that of 304. The
10% chloride by cement value for 304 is about 2% critical chloride threshold
mass decreased by 50% at 40 °C.

Austroads 2016 | page 132


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Appendix C Exposure Conditions of Gateway


Bridge Site

C.1 Chemical Composition of the Brisbane River Water


The following results are cited from the Maunsell and AECOM (2007) report.

Table C1: Test results of Brisbane River water


Issue Unit Upstream Upstream Downstream Downstream
Bicarbonate Alkalinity mg CaCO3 /L 160 150 150 160
Carbonate Alkalinity mg CaCO3 /L < 10 < 10 < 10 < 10
Total Alkalinity mg CaCO3 /L 160 150 150 160
Chloride mg /L 17 000 17 000 18 000 18 000
Sulfate mg /L 2 000 2 000 2 100 2 100
Nitrate (as Nitrogen) mg N /L 0.33 0.27 0.23 0.24
Electrical Conductivity µS/cm 33 000 33 000 34 000 34 000
pH – 7.8 7.8 7.9 7.9
Total Dissolved Solids mg /L 32 000 32 000 33 000 33 000
Calcium mg /L 360 370 380 380
Magnesium mg /L 1 200 1 200 1 200 1 200
Potassium mg /L 350 360 380 380
Sodium mg /L 9 400 9 700 10 000 10 000

Source: Maunsell and AECOM (2007).

C.2 Soil Analysis Results


The following results are cited from the Maunsell and AECOM report (2007).

Table C2: Results of soil analysis


Chloride Sulfate Magnesium Resistivity
Sample ID Location pH
(mg/kg) (mg/kg) (mg/kg) (·cm)
BH TP2 0-0.2 Pier 8 8.2 10 300 940 9 900
BH TP2 5.0-5.7 Pier 8 9.0 1 390 450 2 280 290
BH TP2 10.0-10.7 Pier 8 8.8 8 090 1 900 7 370 92
BH TP2 15.0-15.7 Pier 8 8.8 4 990 810 3 360 150
BH TP2 20.0-20.7 Pier 8 8.6 4 510 900 3 110 140
BH TP2 25.0-25.7 Pier 8 8.6 10 300 1 830 8 400 71
P-NGB-87 0.0-1.0 Pier 7 4.7 90 40 < 10 –
P-NGB-87 1.0-2.3 Pier 7 4.6 90 40 < 10 –
P-NGB-87 2.3-2.7 Pier 7 5.0 190 30 < 10 –
BH-NGB-50 5-5.5 Pier 5 8.5 15 900 1 580 320 –
BH-NGB-50 10-10.5 Pier 5 8.8 5 450 730 60 –
BH-NGB-50 15-15.5 Pier 5 8.0 6 370 890 70 –

Source: Maunsell and AECOM (2007).

Austroads 2016 | page 133


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Appendix D PSTR Requirements for Design Life


The following requirements are cited from the Maunsell and AECOM (2007) report.
Bridge (durability) 300 years
Bridge (loadings) 100 years
Bridge bearings 100 years
Expansion joints Refer to TS90 in PSTR Appendix 6
Deck wearing surface 20 years
Drainage systems 100 years
Abutment support facing structure 100 years
Safety screen and fencing 50 years
Handrails and balustrades 100 years

D.1 PSTR Requirements for Concrete Mix


The following requirements are nominated in PSTR Appendix 31.3, cited from the Maunsell and AECOM
report (2007):
 minimum 40 MPa concrete strength
 minimum 20% fly ash
 mix requirements for concrete in potential acid and sulfate soil.

Austroads 2016 | page 134


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Appendix E Calculation of Accumulated


Corrosion
If the steel corrosion rate is proportional to the chloride concentration in concrete, then the accumulated
corrosion depth with time can be calculated by the integral of the concentration with time (Equation A1):

t2 A1
  k corr,t1t 2  C (t , X C )dt
t1

In Equation A1, kcorr is the corrosion rate (µm/year/Cl precents of cement), t1 is the time when the chloride
content at cover depth (XC) reaches Ccrit, and t2 is the time when the corrosion depth reaches the critical
depth defined by Equation 29; i.e. (t2) = crit. C(t, x) is the concentration of chloride expressed by Fick’s
second law of diffusion:

x A2
C (t , x)  C0  [CS  C0 ][1  erf ( )]
2 Dt

The integral can be calculated by approximation of the error function using the Taylor expansion described
as follows. The error function expressed by the Taylor expansion is given by Equation A3:

2  1 z3 1 z5 1 z7 n 1 1 z 2 n1  A3
erf ( z )        
(n  1)! 2n  1
z ... ( 1)
  1! 3 2! 5 3! 7

For the diffusion theory, z = XC/(4Dt). A1 is an alternating function which quickly converges when z is small.
When the nth term is < 1, if the terms after the nth term are omitted, the residual will be smaller than the (n +
1)th term. So it is sufficient to calculate a few first n terms. The integral against time for the error function part
is given by Equation A4:

t2 A4
4  
t2
z3 z5 z7
t1 erf ( z ) dt  

 
z   
3 2!5  3 3!7  5
 ...t 
  t1

It can be verified that when z < 1, it is adequate to take the first three terms as the approximation or the first
four terms for z < 1.25, the error incurred is less than 0.5%. Thus:

t2 t2 A5
   C (t , X C )dt   [C0  (CS  C0 )(1  erf ( z )]dt
t1 t1

4  zt32t 2  zt31t1 zt52t 2  zt51t1 zt72t 2  zt71t1 


 C S (t 2  t1 )  (C S  C0 )  t2 2
z t  z t    
 
t1 1
3 30 210 

Austroads 2016 | page 135


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

In Equation A6, z2 = XC/(4Dt2) and z1 = XC/(4Dt1). For larger z, more terms in Equation A3 should be
selected by the following criterion:

z 2 n1 A6
1
(n  1)!((2n  1)
z 2 n1  (n  1)!(2n  1)

Austroads 2016 | page 136


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Appendix F Minimum Design Life of Bridges and


Bridge Sub-Assets (as per Roads
and Maritime (2012a) – Gerrinong
Upgrade)

Bridge element Minimum


design life
(Years)
Bridge bearings, non-replaceable 100
Elastomeric bearings (laminated) 35
Pot bearings 35
Expansion joints – structural components 100
Expansion joints – rubbers 40
Deck wearing surface 20
Abutment support structure 100
Bridge barriers – concrete elements 100
Handrails and balustrades 50
Bridge barriers, guard railing – metal elements 50
Safety screens and fencing – metal elements 20
Approach slabs 100
Bridge deck waterproofing under asphalt surfacing, provided the waterproofing is not removed
during future asphalt replacement works, and provided it is not exposed to the atmospheric 100
environment in the future
Scour protection 50
Protective coatings 15
Drainage systems – accessible for refurbishment, including those supported beneath bridge decks 50
Drainage systems – inaccessible for refurbishment, i.e. scupper pipes encased in concrete 100

Source: Roads and Maritime Services (2010, 2012a).

Austroads 2016 | page 137


Realising 100-year Bridge Design Life in an Aggressive Environment: Review of the Literature

Austroads 2016 | page 138

You might also like